You are on page 1of 86

Licensed copy from CIS: qacentral, HBK Contracting Company W.L.L., 25/06/2016, Uncontrolled Copy.

A single copy of this document is licensed to

qacentral

On

25/06/2016

This is an uncontrolled copy. Ensure use of the


most current version of the document by searching
the Construction Information Service.
Licensed copy from CIS: qacentral, HBK Contracting Company W.L.L., 25/06/2016, Uncontrolled Copy.
BRE
Garston, Watford, WD25 9XX 2001
Licensed copy from CIS: qacentral, HBK Contracting Company W.L.L., 25/06/2016, Uncontrolled Copy.

Modelling degradation
processes affecting
concrete

Keith Quillin

BRE Centre for Concrete Construction

constructing the future


Prices for all available
BRE publications can be
obtained from:
CRC Ltd
151 Rosebery Avenue
London, EC1R 4GB
Licensed copy from CIS: qacentral, HBK Contracting Company W.L.L., 25/06/2016, Uncontrolled Copy.

Tel: 020 7505 6622


Fax: 020 7505 6606
email:
crc@construct.emap.co.uk

BR 434
ISBN 1 86081 531 6

© Copyright BRE, 2001


First published 2001

Printed from supplied


camera-ready copy.

BRE is committed to
providing impartial and
authoritative information
on all aspects of the built
environment for clients,
designers, contractors,
engineers, manufacturers,
occupants, etc. We make
every effort to ensure the
accuracy and quality of
information and guidance
when it is first published.
However, we can take no
responsibility for the
subsequent use of this
information, nor for any
errors or omissions it may
contain.

Published by
Construction Research
Communications Ltd
by permission of
Building Research
Establishment Ltd

Requests to copy any


part of this publication
should be made to:
CRC Ltd
Building Research
Establishment
Bucknalls Lane
Watford, WD25 9XX

BRE material is also published quarterly on CD Construction Research Communications

Each CD contains BRE material published in the current CRC supplies a wide range of building and construction
year, including reports, specialist reports, and the related information products from BRE and other highly
Professional Development publications: Digests, respected organisations.
Good Building Guides, Good Repair Guides and
Information Papers. Contact:
post: CRC Ltd
The CD collection gives you the opportunity to build a 151 Rosebery Avenue
comprehensive library of BRE material at a fraction of London, EC1R 4GB
the cost of printed copies.
fax: 020 7505 6606
As a subscriber you also benefit from a 25% discount on phone: 020 7505 6622
other BRE titles. email: crc@construct.emap.co.uk
website: www.constructionplus.co.uk
For more information contact:
CRC Customer Services on 020 7505 6622
Contents
Licensed copy from CIS: qacentral, HBK Contracting Company W.L.L., 25/06/2016, Uncontrolled Copy.

Executive summary iv
1 Introduction 1
1.1 Background 1
1.2 Durability and service life 1
2 Use of models in service life design 3
2.1 Background 3
2.2 Probabilistic approaches 3
2.3 Intended service period design 5
2.4 Lifetime design 6
2.5 Lifetime safety factors 6
3 Causes of concrete deterioration 7
3.1 Approaches to service life design 7
3.2 Factors affecting reinforced concrete service life 8
4 Models for corrosion in concrete 9
4.1 Data input requirements for models for corrosion of steel in concrete 11
4.2 Predicting the time to corrosion initiation 11
4.3 Modelling the propagation period 24
4.4 The effect of temperature on corrosion initiation and propagation 28
4.5 The effect of cracks 29
5 Models for other degradation mechanisms 30
5.1 Sulfate attack 30
5.2 Freeze-thaw 30
5.3 Models for alkali-silica reaction 32
6 Basic protection strategies for reinforced concrete 33
7 Analysis of BRE data 37
7.1 Carbonation 37
7.2 Chloride ion diffusion coefficients 49
7.3 BRE data on corrosion rates 53
8 Conclusions 55
9 Further reading 57
10 References 58
Annexe A Other models for carbonation
Annexe B Additional analyses of BRE carbonation and permeability data

iii
EXECUTIVE SUMMARY
A number of advances have been made in recent years in the development of techniques that
facilitate both the economic design of new concrete structures, and the planned maintenance and
Licensed copy from CIS: qacentral, HBK Contracting Company W.L.L., 25/06/2016, Uncontrolled Copy.

refurbishment of existing structures. Such techniques are aimed principally at the prevention of
premature deterioration, mainly due to the corrosion of the steel reinforcement and, given the wide
range of protection strategies available, the selection of the most appropriate and cost effective means
of achieving the required service life.

Models for predicting time to failure for concrete are increasingly being used as tools for service life
design and prediction. The corrosion of reinforcement, induced by carbonation or by chloride ingress,
is perhaps the major durability problem affecting concrete structures in the UK. These two
degradation processes, together with abrasion, are also the most suitable for service life modelling.
Other forms of concrete degradation are best avoided through choice of materials or through the use
of measures such as protective coatings although models have been developed in some cases.

Reinforced concrete deterioration is generally considered to follow a 2-stage process, with a corrosion
initiation stage, in which the reinforcement is protected by the cover concrete, and a propagation
stage, in which corrosion proceeds. Chloride ingress in saturated concrete is generally modelled
nd
using Fick’s 2 law of diffusion, although other processes such as wick action and convection are
important where the concrete is periodically dry. The inputs for these models can either be estimated
from measurements on actual structures or predicted using empirical relationships derived from
laboratory and field concretes. Carbonation leads to a reduction in the pH of the concrete pore
solution. Carbonation depth generally shows a square root relationship with time. The rate depends
on the physical properties of the concrete and its pH buffering capacity, which is in turn dependent on
the cement content. Corrosion is initiated when the carbonation front reaches the reinforcement.

One of the main problems with the development and application of models for reinforced concrete
deterioration is the large number of variables involved. These include:

• The composition of the concrete and its components.


• The physical properties of the concrete.
• The curing conditions employed.
• The range of exposure conditions.
• The use of protective measures such as coatings and corrosion inhibitors.

There are a considerable amount of data, primarily from laboratory samples and field trials, which can
be used in calibrating and validating models. However, the large number of variables involved means
that the data are often very limited in important areas. The reliability of models based on these data is
consequently limited. There is therefore a need for more data from experimental work (laboratory
work and data from real structures) in determining:

• The effects of different cement compositions and curing regimes on the transport properties of the
concrete.
• The effects of different exposure conditions (including the effects of changes in relative humidity,
short-term application of deicing salts rather than constant marine exposure, the combined effects
of deterioration processes).
• The rate of the corrosion process and its dependence on measurable parameters such as
environmental conditions and properties of the concrete.
• The effects of different protection and repair strategies.
• The effects of ageing on the transport properties of the concrete.
• The application of models derived primarily from laboratory samples to real structures.

This report considers current approaches to modelling degradation mechanisms that can affect
concrete and how these can be used in service life design. Models based on BRE data have been
developed for carbonation and chloride ingress and are compared with those given in the literature.

iv
1 INTRODUCTION
1.1 Background
Licensed copy from CIS: qacentral, HBK Contracting Company W.L.L., 25/06/2016, Uncontrolled Copy.

Considerable advances have been made in recent years in the development of techniques
that facilitate both the economic design of new concrete structures, and the planned
1
maintenance and refurbishment of existing structures . Such techniques are aimed at the
prevention of premature failure due to the corrosion of the steel reinforcement and, given
the wide range of protection strategies available, the selection of the most appropriate
means of achieving the required service life (see below).

There are a number of approaches that can be used to predict the service life of an element
or structure including:

• The use of knowledge and experience acquired from laboratory and field trials to make
semi-quantitative predictions.
• Estimates based on the performance of similar materials in a similar environment.
• The use of accelerated testing
• Modelling degradative processes
• Combinations of these
All of these approaches have advantages and disadvantages. The use of experience and
knowledge acquired from laboratory and field testing, and from the performance of existing
structures may be adequate for structures with a short service life in environments that are
not too severe. However, the approach may be less suitable where long service lives are
required in new or aggressive environments or where new materials have been used.
Accelerated testing can, in principle, be used to predict service life under in-service
conditions provided the degradation proceeds by the same mechanism in both cases and an
acceleration factor can be established. However, the lack of long-term in service data that
can be used to calibrate the tests is a problem.

Models for predicting time to failure for concrete are increasingly being used as tools for
service life prediction. This report summarises models of the degradation processes
affecting concrete.

At present service life modelling approaches are most appropriate for carbonation and
chloride-induced corrosion, and for abrasion. Models for other processes leading to concrete
degradation have been developed but to a lesser degree than those for corrosion.

1.2 Durability and service life

BRE Digest 455 Corrosion of steel in concrete - service life design and prediction 1
addresses issues relevant to the service life design of new concrete structures and the
prediction of residual service life of existing structures. It seeks to provide an overview of
service life design and whole life costing including:

• Current approach to durability design


• Effect of construction quality on service life
• Whole life costing and sustainability
• Service life and criticality
• Assessment of exposure environment
• Selection of concrete parameters
• Estimating time to failure

1
• Protection strategies for steel in concrete
• Durability Audit Procedures

There are a number of definitions of service life of an element of structure. Service life can
Licensed copy from CIS: qacentral, HBK Contracting Company W.L.L., 25/06/2016, Uncontrolled Copy.

2
be defined as:

The period of time after manufacturing or installation during which all essential
properties meet or exceed minimum acceptable values, when routinely maintained.

This definition doesn’t necessarily imply that the structure shall not be fit for purpose at the
end of that period, or that it will continue to be serviceable for that period without regular and
adequate inspection and maintenance. It also does not imply that the structure will continue
to be used for the same purpose.
3
PrEN1990, Basis of Design, defines the ‘design working life’ as :

The assumed period for which a structure or part of it is to be used for its
intended purposes with anticipated maintenance but without major repair
being necessary.

The service life of the whole structure will be limited by the degradation of the non-
replaceable components. Individual components making up a structure will have different
expected service lives. Structural members are expected to perform their intended function
for at least the lifetime of the building whereas other components will need to be repaired or
replaced during the service life of the whole structure. It designing an element or structure
for a given service life it is necessary to define:

a. What constitutes unacceptable performance (the serviceability limit state)


b. What probability of this occurring should be considered in the design.

According to the Concrete Society report, Developments in durability design & performance-
4
based specification of concrete the end of the design working life should be taken as the
formation of cracks at the surface of approximately 0.1 – 0.2 mm due to corrosion.
However, the appropriate stage would depend on the nature of the particular structure. In
5
some cases of chloride attack reinforcement corrosion does not result in cracking and so
an alternative (more conservative) criterion for the end of the design working life would be
required. The end of the design working life could be taken as:

• The point at which corrosion is initiated.


• The first appearance of cracking (visible with magnification).
• Cracking visible to the naked eye.
• First spalling.
• Excessive deflection.
• Collapse under the design loading.

2
2 Use of models in service life design
2.1 Background
Licensed copy from CIS: qacentral, HBK Contracting Company W.L.L., 25/06/2016, Uncontrolled Copy.

There are two ways in which analytical models could be used in the design process:

1. In individual designs with the models developed to an extent to which they can
confidently be used in this respect. This approach requires:
• detailed knowledge of loads pertaining to the particular environment
• agreed limiting performance criteria
• values for ’modifying’ factors to cover issues such as site workmanship, quality of
materials etc.

2. As simplified generic guidance on the relative merits of alternative options in cases


where the investigations needed to obtain a detailed knowledge of the environmental
loads are not feasible. Such an approach would, for example, allow the designer to
compare the use of additional protection methods with the use of additional cover and to
assess the cost/benefit ratio.

6,7
Approaches and concepts used in structural design can be applied to durability design .
The requirement for performance during the service life can be written using the following
simplified limit state function:

R(t) – S(t) ≥ 0

I.e. the resistance R, must be greater than or equal to the load, S, at all times up to the
required or target service life, tg. In the case of structural design the resistance is assumed
to be constant and the load, although fluctuating, is often assumed to be characterised by a
single value. Partial safety factors are then used to take account of variability and
uncertainties, leading to the design values. The limit states can refer to structural safety,
serviceability, functionality, comfort, aesthetics etc. The nature of R and S will depend on
a
the serviceability limit state being used .

2.2 Probabilistic approaches

Service life predictions can be made using single or mean values for input parameters.
However, factors such as environment and the physical properties of the materials can vary
greatly. The resistance, load and service life are therefore unlikely to be known with
certainty and are better represented by distributions rather than just by mean values. In
principle service life can then be predicted in probabilistic (or stochastic) terms. These
concepts are illustrated by Figure 1. Alternatively appropriate safety factors, or a safety
margin, would need to be introduced to take account of variability and uncertainties in the
calculation as discussed in Section 2.5.

a
In durability terms the limit state is stage in the limit state degradation process chosen to mark the end of the
service life). For example if the onset of corrosion is used as the limit state:
• For carbonation-induced corrosion then the cover depth could be used as R and the carbonation depth
used as S.
• For chloride-induced corrosion the corrosion threshold could be used as R and the chloride concentration
at the reinforcement could be used as S.
Similarly other serviceability limit states, such as the onset of cracking can, at least in principle, be used.
Models of the time dependence of carbonation depth and chloride ingress, and for subsequent corrosion and
time to cracking are discussed in Section 4.3.2.

3
The top half of this figure shows the load, S(t), increasing and the resistance, R(t), constant
over time. Both are, at any given time, represented by distributions. The dashed lines
represent their mean values over time. The mean service life is then given by the
Licensed copy from CIS: qacentral, HBK Contracting Company W.L.L., 25/06/2016, Uncontrolled Copy.

intersection between these two mean value lines. However, there is a finite probability of
failure at earlier ages represented by the overlap between the two distributions. The
probability of failure increases with time as shown in the lower half of Figure 1. The
acceptable probability of failure occurring within the target service life will depend on the
nature of the structure and the clients needs. Where the potential consequences of failure
are large (and the accepted failure probability correspondingly small) the mean service life
b
could be substantially greater than the target service life (although this will also depend on
the nature of the distributions for R and S).

Examples of the use of probabilistic methods in durability design are given in RILEM 14,
6
Durability design of concrete structures .

R,S
Service period design

R(t)
S(t)

Distributions of
R(t) and S(t)

Time
Probability of failure, Pf

Mean service life Cumulative probability

Lifetime design

Target service life


Accepted failure probability

Time
8
Figure 1 Service period design and lifetime design (adapted from Siemes & Rostam )

Where a probabilistic approach is used assumptions must be made about the form of
distributions. Service life is often assumed to follow a log-normal distribution, i.e. the service
life is distributed normally on a logarithmic time scale, with a peak in the probability of failure
in a given year, followed by a gradual decrease towards zero as the service life increases.

b
RILEM 14 defines the target service life as: the required service life imposed by general rules, the client or
the owner of the structure.

4
The resistance and the load may be treated as being normally distributed. Cover to
9
reinforcement can be treated as having an exponential distribution
8
Durability design can be presented in two (theoretically equivalent) ways :
Licensed copy from CIS: qacentral, HBK Contracting Company W.L.L., 25/06/2016, Uncontrolled Copy.

• Intended service period design


• Lifetime design

These approaches use the same information and will lead to the same result.

2.3 Intended service period design

In this approach the limit state may not be reached within the target service life. This
approach is comparable with that used in conventional structural design and is summarised
in the top part of figure 1. The concept can be represented by the formula:

Pf,t = P{R(t) - S(t) <0}T[ Ptarget = Φ(-β)

where:

Pf,t is the probability of failure of the structure within the target service life, tg.
Ptarget is the accepted maximum value of the probability of failure.
Φ is the standard normal distribution function
β is the reliability index. It is the number of standard deviations from the mean of a
normal distribution outside which the area under the curve represents the probability
of failure (this is dependent on the consequences of failure and will be high if the
potential consequences of failure are large).

If the onset of corrosion is used the determining factor in the service life is the time taken for
the chloride ion concentration at the reinforcement to exceed a minimum threshold value or
for the carbonation front to reach the reinforcement. The following expressions can be
10
used :

For chlorides:

Pf,t = {Cx(t) - Cth <0}t [ 10-2

For carbonation:

Pf,t = {Xc(t) – Xf <0}t [ 10-2

where Cx is the chloride level at the reinforcement and Cth is the threshold level for
corrosion, Xc is the carbonation depth and Xf is the cover to reinforcement. Pf,t is the
-2
probability of corrosion initiation within the target service life, tg. A value of 10 has been
used in the equations.

In these cases the load, S, is equivalent to the concentration of chloride at the reinforcement
(or the depth of carbonation) and the resistance, R, is equivalent to the chloride threshold
level (or the cover to reinforcement).

5
2.4 Lifetime design

In this approach (summarised in the bottom half of Figure 1) the reliability of the structure is
related to the probability that the target service life will be exceeded. In this case the
Licensed copy from CIS: qacentral, HBK Contracting Company W.L.L., 25/06/2016, Uncontrolled Copy.

probability of failure is given by the expression:

Pf = P{L [ tg}[Ptarget = Φ(-β)

i.e. the actual service life, L, must be greater than or equal to the target service life, tg
where:
Pf,t is the probability of failure of the structure within the target service life. For the
-5
ultimate limit state (collapse) Eurocode 1 gives a probability, Ptarget, of 7 x 10 for a
service life of 50 years. However, much higher levels of risk may be tolerated where
the consequences of failure are less severe.
Ptarget is the accepted maximum value of the probability of failure.
Φ is the standard normal distribution function
β is the reliability index. It is the number of standard deviations from the mean of a
normal distribution outside which the area under the curve represents the probability
of failure (this is dependent on the consequences of failure and will be high if the
potential consequences of failure are large).

2.5 Lifetime safety factors

Where the application of probabilistic methods is not possible a lifetime safety factor could
6
be considered . The design service life, td, would then be given by:

td = γttg

where tg is the required or target service life and γt is the lifetime safety factor.

The safety factor, and consequently the design service life, will increase as the probability of
failure that is considered to be acceptable decreases. These issues are discussed in detail
in RILEM 14 Durability design of concrete structures6.

6
3 Causes of concrete deterioration
Causes of degradation of reinforced concrete are summarised in Table 1 together with the
suitability of service life design concepts in addressing them. Service life design concepts
Licensed copy from CIS: qacentral, HBK Contracting Company W.L.L., 25/06/2016, Uncontrolled Copy.

are most appropriate for chloride and carbonation-induced corrosion, and for physical
deterioration processes.

Table 1 Degradation processes and approach to service life design/prediction

Degradation Factors Approach


process
Carbonation- Mix design Time dependence of depth of attack - Service
induced corrosion Cover depth life prediction. Could use tanking/membranes
Exposure conditions or non-reactive materials etc.
Workmanship
Chloride-induced Mix design Time dependence of depth of attack –
corrosion Cover depth Service life prediction
Exposure conditions Could use tanking/membranes or non-
Workmanship reactive materials, cathodic protection etc.
under very severe conditions.
Internal chemical Quality of Avoidance through choice of materials, limit
processes components curing temperature, design for converted
Curing regime strength etc.
Mix design
External chemical Mix design, cement Processes are time dependent. However,
attack type, quality of selection of non-reactive materials or use of
components membranes etc. may be the best approach.
Physical processes Mix design, cement Can, in principle, use service life design for
and aggregate type, abrasion
Freeze/thaw Mix design, air Avoidance through mix design – inhibit
entrainment process through use of air entrainment.
Salt crystallisation Inhibit reaction

3.1 Approaches to service life design

Table 1 shows that a service life modelling approach is mainly applicable for carbonation
and chloride-induced corrosion and for abrasion. It may be more appropriate to avoid or
limit other degradation processes through choice of appropriate mix designs or the use of
protection techniques. At present models for deterioration processes are, in general,
formulated at the ‘meso’ level – i.e. the concrete is considered to be a continuum.

Models for the time dependence of the processes leading to the corrosion of reinforcement
in concrete have been developed, allowing time to failure to be estimated under a range of
exposure conditions. Models of processes leading to the degradation of the concrete itself
(such as sulfate attack, frost attack, ASR) have also been developed although to a lesser
extent than those leading to reinforcement corrosion.

7
3.2 Factors affecting reinforced concrete service life

The service life of a reinforced concrete structure depends on a number of parameters.


Licensed copy from CIS: qacentral, HBK Contracting Company W.L.L., 25/06/2016, Uncontrolled Copy.

1. The location of the structure and its expected environmental conditions.


The environment to which the concrete will be exposed is a key factor in designing for a
given service life. Relevant factors include:
• The general environment conditions.
• The specific location and orientation of the concrete surface being considered and its
exposure (exposed, sheltered or semi-sheltered) to prevailing winds and rainfall.
• Localised conditions such as surface ponding, exposure to surface run-off and spray,
aggressive agents, regular wetting, condensation etc.

2. Conceptual and structural design, detailing, intended use and level of maintenance.
Good conceptual and detailed design is vital in producing durable concrete structures
and elements. They can ensure that the effects of water on the structure are minimised
to prevent the penetration of water (and aggressive species contained in it) at
vulnerable areas and contribute to overall quality by ensuring that structures and details
are easy to build. Appropriate maintenance in use is also important.

3. Materials specification and properties


It is important that the materials used are appropriate to the environment and the
requirements of the structure. Cover to reinforcement is a key factor in determining the
time to the onset of corrosion as is the type of reinforcement used. The cement
component establishes a high pH within the concrete and acts as a chemical buffer to
CO2 attack. It can also affect the chloride binding capacity of the concrete. The cement
type influences the buffering capacity and the permeability of the concrete and therefore
can have a major impact on durability. For example the buffering capacity of a PC
concrete will be greater than that of a pfa concrete (for equal total cement contents).
However, the rate of ingress of CO2 into well-cured pfa concrete will be lower than that
for PC concrete due to the structure of the cement paste. The fineness of the cement
may also be an important factor. For a given w/c ratio the porosity of the paste (and
potentially the permeability) will tend to be higher for a coarse cement than for a fine
cement. The water: cement ratio used in the production of concrete is a key factor in
determining the service life. It affects the permeability of the concrete and therefore the
rate at which aggressive agents can penetrate the concrete.

4. Appropriate execution
The execution phase will be crucial in determining whether the structure or element will
meet its required service life. There have been numerous examples of durability
problems due to poor concrete, inadequate cover to reinforcement, chlorides in the
concrete or combinations of these. The quality of workmanship is influenced by the
quality of design and detailing. Some designs are difficult to build, even with
reasonable standards of workmanship on site. Communication and supervision are key
factors in ensuring good quality. Issues that need to be addressed include:
• Curing conditions and duration .
c

• Achieving appropriate compaction


• Achieving the specified depth of cover.
c
These impact on the extent to which the cementitious component of the concrete has hydrated and
therefore on the permeability of concrete (and potentially the buffering capacity) especially for
blended cement concretes. Concrete durability may be increased if it is allowed to dry out before
exposure (assuming that it has been adequately cured). A layer of calcite can then form at the
surface, blocking the pores and reducing permeability of the surface layer. However, rapid drying can
increase the permeability, possibly through the formation of fine shrinkage cracks.

8
4 Models for corrosion in concrete
Corrosion in concrete is often depicted as a two-stage process. First there is an initiation
period during which carbonation or chloride ingress occurs, starting at the concrete surface
Licensed copy from CIS: qacentral, HBK Contracting Company W.L.L., 25/06/2016, Uncontrolled Copy.

and progressing in a time-dependent manner into the element. This is followed by a


propagation period from the onset of corrosion to the durability limit state. Approaches to
modelling carbonation and chloride ingress, and the subsequent corrosion stage are
discussed below.
Time t t 1 2

Lifetime or time before repair

Propagation tp
Initiation ti

Controlled by rate of
CO2 or Cl- penetration
Load-bearing capacity

Cor
rosi
on

Appearance of cracking

Loss of 10%
reinforcement

11
Figure 2: Corrosion model - after Tuutti (1982).

The design service life, td, would be given by:

t d = ti + tp (1)

The relative lengths of the initiation and propagation periods depend on the exposure
conditions, cover and the characteristics of the concrete. In saturated concrete oxygen
transport to the steel surface will be negligable and consequently the propagation period will
be very long. In dry concrete the initiation period will be very long as the mechanisms for
chloride ion transport and carbonation are dependent on the availability of water or moisture.
Corrosion is most likely to occur in concrete that is alternately wetted and dried, or, for
carbonation-induced corrosion, at intermediate relative humidities.

Models for the initiation and propagation stages are discussed below. These models are
semi-empirical. The reliability of the predictions made using them is heavily dependent on
data for calibration collected from real structures, field trials and laboratory experiments.

The effects of combined degradation processes (such as carbonation and chloride ingress)
12
have not yet been seriously addressed from a modelling perspective . However, reliability
will improve as more data are collected.

9
BOX 1. Corrosion of steel in concrete
Licensed copy from CIS: qacentral, HBK Contracting Company W.L.L., 25/06/2016, Uncontrolled Copy.

The corrosion of steel reinforcement arises through local structural and/or compositional variations
within the concrete. Consequently some areas of the steel become positively charged (anodes) and
others negatively charged (cathodes). The anodic reaction can be written:

Fe → Fe 2 + + 2e −
The cathodic reaction is:

H 2 O + 12 O 2 + 2e − → 2OH −

Iron therefore dissolves at the anode with the corrosion product being deposited nearby. The corrosion
product occupies a volume several times that of the parent metal and its formation creates internal
stresses that result in cracking and delamination of the concrete. The circuit is established through the
movement of electrons within the metal and of metal and hydroxyl ions through the pore solution. The
key factors in the ongoing corrosion are:

• The availability of oxygen.


• The presence of moisture on the surface of the reinforcement and in the adjacent concrete.

If the pH in the electrolyte surrounding the steel is above about 10.5 a uniform protective surface oxide
layer forms on the metal surface. The rate of corrosion under these circumstances is insignificant. In
Portland cement concretes the pH is maintained at levels of at least 12.6 due to the presence of
significant amounts of Ca(OH)2. The presence of sodium and/or potassium salts can increase the pH
further. Atmospheric carbonation and the effect of ingressing or cast-in chloride ions can break down
the protective surface layer leading to the initiation of reinforcement corrosion. These processes are
time dependent and, as such provide a basis for predicting the service life of the structure or element.

The ingress of carbon dioxide can depassivate the reinforcement by removing hydroxide from the pore
solution:
2+ -
Ca(OH)2 τ Ca + 2OH .
- 2-
CO2 + 2OH τ CO3 + H2O.
2+ 2-
Ca + CO3 τ CaCO3

The reduction in [OH−] will also lead to the dissolution of other hydrates. The pH consequently falls to
a level lower than the 10.5 needed to maintain the passive oxide layer on the reinforcement allowing
corrosion to occur.

The presence of oxygen and sufficient quantities of free chloride ions in the pore water of concrete can
produce reinforcement corrosion even in highly alkaline conditions as discussed in Part 1 of
62
Digest 444 . Chlorides may be present in the components used to make the concrete at the time of
casting (cast-in) or enter the concrete from the environment after placing (ingressing). Principal
external sources of chloride in the UK are de-icing salts, seawater and marine conditions more
generally. Elsewhere airborne salt and saline ground waters can be a problem.

10
4.1 Data input requirements for models for corrosion of steel in concrete

The predictions made using models need to be reliable but cautious and the choice of
appropriate input parameters is therefore crucial. The following factors will influence the
Licensed copy from CIS: qacentral, HBK Contracting Company W.L.L., 25/06/2016, Uncontrolled Copy.

time to the onset of chloride-induced corrosion in concrete:

• Depth of cover
• Concrete mix design and materials
• Curing time and conditions
• Execution
• Exposure environment including relative humidity, proximity to source of chlorides,
• Concentration of chloride at the surface of the concrete
• Chloride binding
• Chloride threshold level (which depends on a number of factors including cement type,
relative humidity, type of reinforcement, use of corrosion inhibitors and surface coatings
etc).
• The use and impact of additional protection strategies such as coatings, corrosion
inhibitors etc.

The following factors will influence the time to the onset of carbonation-induced corrosion:
• Depth of cover
• Concrete mix design and materials
• Execution.
• Curing time and conditions
• Exposure environment including relative humidity, atmospheric CO2 concentration.
• The use of additional protection strategies such as coatings.

The factors that will influence the rate of corrosion will depend on:
• The relative humidity at the reinforcement
• The availability of oxygen
• The chloride ion concentration at the reinforcement.

In assessing the remaining service life of existing structures it may be possible to measure
parameters (such as depth of carbonation, chloride profiles, corrosion rates etc.) that can be
used directly in the models discussed here. However, for new structures these parameters
will not be available and will need to be predicted based on the factors listed above. Some
of these issues are discussed in the following sections.

4.2 Predicting the time to corrosion initiation

4.2.1 Chloride ingress


Chloride solutions, arising from exposure to a marine environment, airborne salts or the
application of deicing salts, can enter the concrete through the following mechanisms:
• Diffusion
• Capillary suction (where the concrete is not fully saturated).
• Hydrostatic pressure
• Ingress through cracks and joints.

Diffusion will be the main mechanism for chloride ingress in situations where the concrete
can be assumed to be fully saturated. Chloride ingress through diffusion is commonly

11
modelled using Fick’s second law. Choice of appropriate boundary conditions gives the
following:

  x 
Licensed copy from CIS: qacentral, HBK Contracting Company W.L.L., 25/06/2016, Uncontrolled Copy.

C x − C b = (C s − C b ) × 1 − erf  
 (Equation A)
 2(Dt )
1/ 2
 

Where:
Cx = chloride at depth x
Cs = chloride content at surface of the concrete
Cb = background chloride content (from the mix ingredients)
D = apparent chloride diffusion coefficient
t = exposure period

The apparent diffusion coefficient, D, is the only unknown in theory. However, the surface
chloride level, Cs, is difficult to measure and may vary with time (see Section 4.2.2). The
usual approach is therefore to fit a curve with Cs and D as the independent variables.

Equation A can be used to estimate D for an existing structure by calculating a best fit curve
for measured chloride concentration data and extrapolating back to the surface to obtain Cs.
These values can then be used in residual service life calculations. For new structures the
diffusion coefficient and surface chloride level need to be related to the mix design, curing
conditions and exposure condition. A significant amount of work has been done to do this
using data for laboratory concretes and real structures. Measurements using trial mixes and
the actual structure could verify the values obtained.

Apparent diffusion coefficients may fall with time due to the ongoing hydration of the
cementitious component. There is very little change beyond the first 12 months for PC
concretes but there can be a marked drop for pfa and ggbs concretes. The time
dependence of the apparent diffusion coefficient can be modelled using a relationship of the
13
form :
n
 t 
D = D0 ×  
 t0 

Where D and D0 are the apparent diffusion coefficients after t and a reference period t 0
64
respectively and n is a parameter dependent on the cement type . For PC concretes n is
typically about -0.25. Respective values for pfa and ggbs concretes are about -0.7 to -0.6
although these may depend on the actual composition of the cement. Apparent diffusion
coefficients, and consequently rates of chloride ingress, for blended cements will therefore
fall more rapidly over time than for PC concretes.

4.2.2 Surface chloride level


The concentration in the concrete surface may be greater than that in the surrounding
environment. Wetting and drying cycles can lead to a build-up of salts at or near the
surface of the concrete as can physical and chemical absorption of chlorides and the
inhomogeneous nature of the concrete. The surface chloride level may be higher for pfa
and ggbs concretes due to enhanced binder capacity and higher binder volume. For
structures affected by deicing salts the surface chloride level will depend on the rate and
frequency of application, and the location of the structure in relation to the source of the
chlorides.

12
The surface chloride level has also been found to increase with time in submerged
14
concretes . There is evidence to suggest that the time dependence is either linear or
15
proportional to the square root of time. The following equations have been suggested for
these two cases:
Licensed copy from CIS: qacentral, HBK Contracting Company W.L.L., 25/06/2016, Uncontrolled Copy.

 x2   x   x  − x 2 / 4 Dt 
C x = kt 1 + erfc  − 
1/ 2 
e
1/ 2  
 2 Dt   2(Dt )   (Dt )  

and

 2 x π  x  
Cx = k t e − x / 4 Dt −  erfc 
1 / 2  
  2 Dt  2(Dt )  

A logarithmic expression for the time dependence of Csa has also been proposed:

Csa(texp) = Aln(texp + ∆tini) + B

where texp is the exposure time, ∆tini is an initial binding time and A and B are constants.
16
According to Bamforth the surface level remains approximately constant over time for
concretes in marine spray and splash zones with the steady state value being attained
relatively early in the lifetime of the concrete. However, there are significant variations in the
steady state value over the surface of the concrete. Concrete type also has a significant
effect. The surface chloride level is generally higher for pfa and ggbs concretes due to
enhanced chloride binding capacity and higher cement volume.
17
Meljbro has derived expressions that are based on modifications to the classical solution of
Fick’s law and take account of time dependency of the apparent diffusion coefficient and the
surface chloride level.
18
The diffusion coefficient follows an Arrhenius-type dependence on temperature . Activation
energies are reported to be in the range 32 – 45 KJ/mole for cement pastes. The rate
limiting processes are believed to involve surface interaction between the chloride ions and
the hydrates or changes in pore structure with temperature.

The corrosion process itself is also dependent on temperature. It is often assumed that
o
increasing the temperature by 10 C doubles the rate of corrosion. However, it has been
19
suggested that the actual ratio is about 1.6. Raising the temperature of the concrete itself
during the curing phase can also increase the rate of diffusion, as the pore structure
becomes coarser.

The following graphs give examples of predicted chloride ingress over time for nominal PC
and pfa concretes where the effects of the time dependency of the diffusion coefficient are
illustrated.

13
Chloride level (wt% cement) 5

4
Licensed copy from CIS: qacentral, HBK Contracting Company W.L.L., 25/06/2016, Uncontrolled Copy.

3 1 year
10 years
2 100 years

0
0 20 40 60
Depth/mm

Figure 3. Predicted chloride profiles for a C30 PC concrete assuming a diffusion coefficient
-11 2
after 1 year of 10 m /s and n = -0.26.

5
Chloride level (wt% cement)

3 1 year
10 years
2 100 years

0
0 20 40 60
Depth/mm

Figure 4. Predicted chloride profiles for a C30 15%pfa concrete assuming a diffusion
-12 2
coefficient after 1 year of 4 x 10 m /s and n = -0.7.

4.2.3 Chloride binding


The measured chloride levels in concrete are the total amounts present. However, only the
free chloride, i.e. that in solution, affects the corrosion of the reinforcement. A significant
proportion of the chloride is actually present in the cement matrix, primarily as
chlorocarboaluminates such as C4ACl.Hx although it can also substitute for anions in other
cement hydrates. The solubility product for this phase can be written as follows (assuming
an ideal solution):

2+ 4 - 2 - - 5
Ksp = [Ca ] .[Al(OH)4 ] .[Cl ].[OH ]

14
Precipitation of this phase will only occur if the solubility product is exceeded. On increasing
the Cl content of the cement the amount in solution will level off when the solubility constant
is reached and then increase as the buffering capacity of the other alumina-containing
phases is exhausted. The amount of bound chloride will therefore depend on the
Licensed copy from CIS: qacentral, HBK Contracting Company W.L.L., 25/06/2016, Uncontrolled Copy.

concentration in solution, the cement composition, the presence of carbonates and sulfates
and the solubility properties of the chlorocarboaluminates and other AFm phases.
20
According to Sergi et al the relationship between free and bound chlorides can be
represented by the Langmuir adsorption isotherm:

αC Cl
S Cl =
1 + βC Cl

where SCl is the concentration of bound chlorides and CCl is the concentration in the pore
solution. α and β are constants for a given cement but will depend on the binding capacity
of the cement and the presence of cast-in chlorides.

The use of apparent diffusion coefficients in modelling chloride ingress effectively takes
chloride binding into account via the assumed or measured coefficient. This approach
appears to be the most appropriate at present due to the limited data for predicting the free
chloride level. However, the use of the Langmuir isotherm could allow the effects of
different cement types to be modelled through the use of the free chloride: hydroxide ratio at
the reinforcement rather than the specification of a simple threshold level as is the case with
the apparent diffusion coefficient method.

4.2.4 Modelling chloride ingress in structures subject to wetting and drying cycles
In situations where the structure is subject to wetting and drying cycles, such as highway
structures and in the marine splash zone, the concrete can be considered to have ‘inner’
and ‘outer’ zones. In the ‘inner’ zone the concrete is assumed to be close to full saturation
and the predominant transport mechanism is likely to be diffusion. The ‘outer’ zone is likely
to be affected by the wetting and drying cycles and transport mechanisms other than
diffusion will contribute to chloride ingress. It may also be subject to carbonation and
leaching due to rain. The depth of the outer zone will depend on the environment and the
properties of the concrete. However, although models for multi-mechanistic transport in
partially saturated concrete are described in the literature their use requires detailed
knowledge of site-specific conditions and a wide range of material properties that are not
usually available to the engineer at the design stage.

Chloride ingress into dry concrete exposed to chloride solutions is initially likely to be rapid
and to occur through absorption. Up to 20mm of quality concrete can be penetrated in a
21
few hours through this mechanism . The factors that influence rates of absorption into
concrete are not well understood. Penetration by absorption is likely to be low in high quality
concrete and where frequent wetting occurs and where subsequent drying is slow. It may
be higher in poor quality concrete that is wetted infrequently and where subsequent drying is
22
rapid . The depth to which chloride can penetrate through absorption is limited to the depth
at which the pores remain dry. The chloride ion concentration remains approximately
constant with depth and equal to the value at the surface. Therefore absorption effectively
reduces the cover depth by an amount equal to the depth to which absorption takes place.

Where the concrete subsequently remains wet for long periods diffusion will become the
predominant mechanism, however, where the concrete is subjected to a further long period
of drying followed by further exposure to chlorides absorption will again become important.

15
Even where the concrete is not fully saturated it is often assumed that long term chloride
ingress can still be modelled using equation A. Chloride transport will occur during and
shortly after wet periods. The apparent diffusion coefficient will depend on the extent to
which the concrete is able to dry out following periods of exposure. Where the concrete
Licensed copy from CIS: qacentral, HBK Contracting Company W.L.L., 25/06/2016, Uncontrolled Copy.

remains relatively wet (for example where water is retained in the structure due to the effect
of a failed coating) diffusion coefficients and surface chloride levels are similar to those
found for marine exposure. Where the concrete is able to dry out apparent diffusion
coefficients are much lower. Wetting and drying cycles will also affect the surface chloride
concentration. The surface concentration in exposed concrete affected by deicing salts is
likely to reach a steady state (though with seasonal variations) due to the effects of leaching
due to rain and diffusion of chlorides towards the concrete surface. A steady increase is
likely to occur for sheltered concrete as the effects of leaching will be greatly reduced.
67
The Hetek programme gives the following information for concrete exposed to chlorides:

1. For sheltered concrete


• The surface chloride level will increase with time and can reach 0.5% by mass
concrete
• The surface chloride level will decrease linearly with height above road level and
logarithmically with distance from the road.
• The surface chloride level will have a maximum value on the leeward face
• The apparent diffusion coefficient will be significantly lower than in wet concrete. It
has been suggested that assuming a constant value for Da x t should be used for
modelling purposes.

2. For exposed concrete


• The surface chloride level can reach 2% by mass concrete and a steady state is
likely to have been reached after a few seasons.
• The surface chloride level will decrease linearly with height above road level and
logarithmically with distance from the road.

Departures from Fick’s law can also arise from carbonation of the concrete surface. The
chloride level in the surface layer may also fluctuate due to changes in the external
environment. Carbonation will reduce the capacity of the cement paste to bind chlorides
and consequently lead to a reduced chloride content in the surface layer. Free chlorides
can also be washed out from the surface layer. Subsequent back-diffusion could also
occur. These factors can lead to a peak in the chloride level at a depth of several mm into
the concrete rather than at the surface itself.
23
It has been suggested that models based on Fick’s law could be modified for application in
conditions where the concrete is not saturated. The proposed modifications include:

• Using appropriate correction factors to reduce apparent diffusion coefficients


• Modifying ageing factors for the diffusion coefficient.
• Assuming that the depth of the zone affected by changes in the environment can be
empirically related to the w/c ratio. The depth of this zone is used in modifying the
predicted chloride profiles.
• The effect of carbonation can be taken into account by allowing the depth of the cover
zone to vary with time in accordance with models for carbonation depth against time.
• Experimentally derived maximum surface chloride levels at the cut-off depth can be used
for design purposes.

16
Other models for chloride ingress into partially saturated concrete, for penetration due to
pressure-induced flow and diffusion, and for wick action have been proposed. Some of
them are briefly discussed in references 48 and 28.
Licensed copy from CIS: qacentral, HBK Contracting Company W.L.L., 25/06/2016, Uncontrolled Copy.

4.2.5 Chloride threshold levels


The minimum total chloride content, or the chloride threshold level, required for corrosion to
occur has been the subject of a considerable amount of debate over recent years. A value
of 0.4% by weight of cement is often used. However, it is clear that it cannot be represented
by a single value. It is dependent on a number of factors such as:
• the water / cement ratio
• cement type
• pH of the pore solution
• capacity of the cement paste to bind the chloride ions
• whether the chloride ions are ‘cast-in’ or ingressing from an external source
• the exposure conditions
• measurement difficulties (especially for free chloride and hydroxide near the
reinforcement)
• local variations in chloride concentration at the reinforcement

Literature threshold levels typically range from 0.17 to over 2 % by weight of cement, with
the probability of corrosion increasing with the chloride concentration. Values for the
threshold level for ingressing chlorides in uncarbonated concrete in marine environments
24
(taken from Brown ) are given in Table 2.

Table 2. Chloride threshold levels from Brown

Chloride ion concentration Probability of corrosion


(wt% cement)

<0.4 negligible
0.4-1.0 possible
1.0- 2.0 probable
>2.0 certain
25
Table 3 shows values given by Vassie . These are consistent with those of Brown, but
show that there is a significant risk of corrosion for chloride ion contents below 0.4%.

Table 3. Risk of corrosion at different threshold levels from Vassie

Chloride ion range (wt% cement) % analyses showing corrosion

0.2 2
0.2 – 0.3 22
0.3 – 0.5 23
0.5 – 1.0 32
1.0 – 1.5 64
1.5 76

17
26
Tuutti gives graphs showing the dependence of the threshold level on parameters such as
the w/c ratio, the environment and the use of additions. Table 4 summarises the data in
these graphs.
Licensed copy from CIS: qacentral, HBK Contracting Company W.L.L., 25/06/2016, Uncontrolled Copy.

Table 4 Chloride threshold levels (From Tuutti)

Chloride by weight of cement (%)


w/c Exposure condition PC 20% pfa 5% silica 10% silica
fume fume
0.3 Below sea water 2.2 1.4 1.6 1.2
Marine splash zone 1.0 0.5 0.6 0.4
Deicing salt splash zone 0.8 0.4 0.5 0.3
Above splash zone 1.2 0.5 0.7 0.5
0.4 Below sea water 2.0
Marine splash zone 0.8
Deicing salt splash zone 0.6
Above splash zone 1.0
0.5 Below sea water 1.5 0.7 1.0 0.6
Marine splash zone 0.6 0.3 0.4 0.2
Deicing salt splash zone 0.4 0.2 0.3 0.2
Above splash zone 0.6 0.3 0.4 0.2

Based on these data Tuutti has calculated the relative time to the onset of corrosion for
different concrete mix designs, assuming a marine environment. These are given in
Table 5.

Table 5 Relative corrosion initiation periods (from Tuutti)

Composition w/c = 0.5 w/c = 0.4 w/c = 0.3


100% PC 0.35 1.00 3.2
95% PC + 5% SF 0.25 0.75 4.5
90% PC + 10% SF 0.15 0.75 1.25
80% PC + 20% pfa* 0.1 0.6 0.9

Note: according to this analysis pfa concrete should be less resistant to chloride-induced corrosion
than PC concrete. This is not substantiated by experimental data from BRE or other sources.
In many cases low strength (nominally 25 MPa) fly ash concretes have been found to perform
55
better than higher strength (45 MPa) PC concrete .
27
Nilsson et al have given design threshold levels for various exposure conditions (these
data are similar to Tuutti’s and are presumably taken from the same source). These are
summarised in Table 6.

18
Table 6. Suggested design values for chloride threshold levels (black steel) in various
Nordic exposure zones (wt% chloride at reinforcement)

Concrete submerged marine splash de-icing salt atmospheric zone


Licensed copy from CIS: qacentral, HBK Contracting Company W.L.L., 25/06/2016, Uncontrolled Copy.

type zone zone splash zone marine/de-icing


w/c 0.50
100 % PC 1.5 % 0.6 % 0.4 % 0.6 %
5 % SF 1.0 % 0.4 % 0.3 % 0.4 %
10 % SF 0.6 % 0.2 % 0.2 % 0.2 %
20 % pfa 0.7 % 0.3 % 0.2 % 0.3%
w/c 0.40
100 % PC 2.0 % 0.8 % 0.6 % 0.8 %
5 % SF 1.5 % 0.5 % 0.4 % 0.5 %
10 % SF 1.0 % 0.3 % 0.2 % 0.3 %
20 % pfa 1.2 % 0.4 % 0.3 % 0.4 %
w/c 0.30
100% PC 2.2 % 1.0 % 0.8 % 1.0%
5% SF 1.6 % 0.6 % 0.5 % 0.6 %
10% SF 1.2 % 0.4 % 0.3 % 0.4 %
20 % pfa 1.4 % 0.5 % 0.4 % 0.5 %

27
According to Nilsson et al chloride threshold levels vary extensively in field exposed
concrete exposed to the air, as a consequence of the varying microclimate at the steel
surface. As a consequence the chloride threshold level depends on the cover thickness and
on the physical bonding between concrete and reinforcement. The chloride threshold levels
are only valid for “macro crack free” concrete with a maximum crack width of 0.1 mm and a
minimum cover of 25 mm. The data are not valid for calculations of the initiation time in
cracked concrete with crack widths > 0.1 mm.

Carbonation can reduce the threshold level by reducing the pH and the chloride binding
capacity of the cement paste. When the carbonation front reaches the reinforcement,
corrosion can commence in the absence of chlorides as the protective oxide layer breaks
down.

It is not clear what account the above threshold values take of the potential influence of
cracking upon corrosion risk. Pitting corrosion is of particular concern because of the
associated localised loss in cross-sectional area of reinforcement, which can be very severe
in some circumstances.
28 - -
Haussmann has shown that the probability of anodic reaction is dependent on the Cl /OH
ratio. This has been reported to have a value of about 0.61 (this is reported to be valid for
steel exposed to solution – the interface between reinforcement and concrete may have an
29
extra protective effect). Tang and Nilsson report a value of about 1.1. The w/c ratio is
also a factor. The threshold level rises as the w/c ratio increases (this is probably due to an
increase in pH at lower w/c ratios and may not be a factor in carbonated concrete, although
the lower permeability will inhibit the corrosion process). Leaching can reduce the OH-
concentration.

The very limited data on free chloride levels required to initiate corrosion are in the range
30 29
0.14 – 1.8 mol/litre . However, according to Tang & Nilsson a chloride content of 0.4% by
weight of cement corresponds to a free chloride concentration of about 0.9 - 1 g/l (0.025 –
0.028 mol/l). This corresponds to a [Cl]/[OH] ratio of about 0.6 in saturated lime. The

19
threshold limit of 0.4 may only hold therefore at pH’s of about 12.4 (i.e. when the alkali
hydroxides have been removed).
55
Thomas and Matthews have suggested the following dependence of the threshold level on
Licensed copy from CIS: qacentral, HBK Contracting Company W.L.L., 25/06/2016, Uncontrolled Copy.

the pfa content of the concrete based on an analysis of BRE data:


- d
pfa content/% Threshold Cl /weight% total cement
0 0.7
15 0.65
30 0.5
50 0.2
31
Dhir et al have suggested, based on experiments using cast-in chlorides, that there is no
specific chloride level at which corrosion is initiated. Rather the corrosion rate increases
gradually as the chloride level increases and that some corrosion activity will occur if free
chlorides are present at the steel interface. They state that there is very little difference in
the water-soluble chloride content required to initiate corrosion in equivalent strength PC
and pfa concrete. However, the rate at which the chloride ions are transported to the site
once corrosion has been initiated is important. The diffusion coefficient is therefore an
important parameter. They have assessed the risk of corrosion based on the diffusion
coefficient and the water soluble chloride content (expressed as weight % of cement) as
shown in Table 7.

Table 7 Corrosion risk – (from reference 31)

Corrosion intensity
Low (0.1 µA/cm ) Medium (1.0 µA/cm ) High 10.0 µA/cm )
2 2 2
-12 2
D (10 m /s) Water-soluble chloride content(wt % cement)
<20 1.0 - -
20-50 0.3 0.5 2.0
>50 0.2 0.4 0.6

32
It has been suggested that there is no specific chloride level at which corrosion is initiated.
Rather the corrosion rate increases gradually as the chloride level increases and that some
corrosion activity will occur if free chlorides are present at the steel interface.

The Hetek report has summarised available information on the effects of cement type and
exposure environment on the threshold level of cement and has derived the following
equation for design threshold levels:

C cr = k cr ,env × (1.2 − eqv( w / c) cr )

where kcr,env is a constant representing the environment and eqv(w/c) is the w/c ratio
‘corrected’ to account for the contributions made by the additions silica fume and pfa given
by:

eqv(w/c) = (mass water)/((mass cement) + (mass addition) x e)

d
i.e. the PC + pfa content of the concrete

20
where

Marine Marine de-icing


submerged splash/atmosphere splash
Licensed copy from CIS: qacentral, HBK Contracting Company W.L.L., 25/06/2016, Uncontrolled Copy.

kcr,env 2 0.75 0.5

and

Efficiency silica fume pfa


factor
e -3.5 -1

Reinforcement corrosion due to cast-in chlorides can be controlled by minimising chloride


levels in the mix (BS5328-1 defines recommended limits as discussed in Part 1 of
62 33
Digest 444 ). Currie and Robery have given guidance on the risk of corrosion due to
cast-in chlorides in PC concretes in structures representing 3 different age groups: 25, 40
and a projection for 60 years old. The risk of corrosion is estimated in relation to the
environment (damp or dry) and the alkalinity of the concrete (carbonated or uncarbonated).
The risk of corrosion increases with the chloride ion content, the age of the structure,
whether the depth of carbonation has reached the reinforcement and whether the
62
environment is wet or dry. These guidelines are summarised in Part 2 of Digest 444 .

The probabilistic nature of the chloride threshold level is reflected in Figure 5 which is
62
adapted from BRE Digest 444 Part 2 . The figure shows the probability that corrosion
initiation will have occurred due to ingressing chlorides in uncarbonated concrete. The
graph is based on available data from a range of sources and includes values taken from
laboratory tests and actual structures. The central line represents the average. The outer
areas are intended to represent a normal distribution reflecting the fact that measured
values from different sources show a large degree of scatter. The nature of the materials
used and the exposure environment affect the risk of corrosion.

Values for a 1% chloride ion concentration at the reinforcement are shown as an example.
the figure shows that corrosion is likely to have commenced in about 60% of cases.
However, the actual figure may vary between 30% and 90% depending on other factors
such as the environment and the nature of the materials. As Figure 5 is based only on the
limited available data it may need to be modified as knowledge and experience increase.

21
Licensed copy from CIS: qacentral, HBK Contracting Company W.L.L., 25/06/2016, Uncontrolled Copy.

Figure 5. Estimated risk of corrosion associated with ingressed chloride in the absence
of carbonation (from reference 62). See main text for explanation

22
4.2.6 Carbonation
The rate of carbonation generally follows an approximately square root relationship with
time. Figure 6 shows carbonation depth against time profiles for a selection of laboratory
o
concretes stored indoors at 65% RH and 20 C.
Licensed copy from CIS: qacentral, HBK Contracting Company W.L.L., 25/06/2016, Uncontrolled Copy.

30

25
Carbonation depth/mm

20

15
C35 PC 1-day cure,

C35 PC 3-day cure


10
C45 PC 1-day cure

5 C35 50% pfa 1-day


cure
C35 50% pfa 3-day
cure
0
0 500 1000 1500 2000 2500 3000 3500 4000
Time/days

Figure 6. Carbonation depth against time for laboratory concretes stored indoors

The following equation is often used for carbonation depth, possibly after an initial period
with a higher carbonation rate:

½
dc = Cct
e
where dc is the average carbonation depth in mm, t is the time in years and Cc is a rate-
determining constant determined by the permeability of the concrete (which is in turn
dependent on the mix design, curing conditions and level of workmanship), the buffering
capacity of the cement and the atmospheric concentration of CO2.
nd
The equation is based on experimental results (although solutions of Fick’s 2 law give a
similar dependence on time) and is a simplification. It assumes, for example, that the
porosity, permeability and moisture content of the concrete and the atmospheric CO2
concentration are constant over time (or can be represented by single values).

This model of carbonation can be readily used for existing structures to predict the
remaining service life of existing concrete structures based on measured carbonation
depths. However, there are problems in relating carbonation depth to either the type of
cement and the mix design, or to measurable properties of the concrete at early ages.
34
Parrott has shown that the depth of carbonation after 18 months for concretes prepared
under identical conditions but with different cements (Portland cements, Portland limestone

e
Measured carbonation depths can vary significantly, even for comparatively small laboratory specimens.
Factors such as the permeability of the aggregate and the extent of cracking may influence the depth of
carbonation.

23
cement and slag cements varies considerably (between about 7 and 20mm after a 1 day
cure).

The power value in the above equation can range considerably about the value of 0.5. BRE
Licensed copy from CIS: qacentral, HBK Contracting Company W.L.L., 25/06/2016, Uncontrolled Copy.

data show values of between 0.3 and 0.6 in many cases and in some cases ranging up to 1
(see Annexe B). Its value varies depending on the humidity, the time for which the concrete
has been cured under controlled conditions prior to exposure and possibly other factors.
35
Other authors also report similar variations and this has been assigned to the fractal
growth of CaCO3 in pores. Parrott has produced tables based on experimental data where
the value of the power function varies (see Annexe A).

Most models of carbonation are one-dimensional and therefore underestimate the extent of
carbonation in cases where attack is 2-dimensional, such as at the corners of buildings.
Some carbonation models are discussed in Annexe A.

The rate of carbonation in wet concrete is virtually zero and consequently where intermittent
wetting and drying occur the carbonated zone will not extend beyond the maximum drying
out depth. As the frequency of the wetting and drying cycle increases the drying out depth,
and consequently the maximum depth of carbonation, decreases as the concrete dries out
more slowly than it wets. Models for carbonation of concrete subject to wetting and drying
cycles have been developed (for example that by Schiessl). However, they are likely to
prove difficult to use in practical situations due to the nature of the input requirements.

4.3 Modelling the propagation period

For both chloride- and carbonation-induced corrosion the initiation stage of the corrosion
model has been developed to a greater extent than the propagation stage. However, there
are several ways in which the propagation stage can be addressed in service life models.
One approach is to assume that the onset of corrosion marks the end of service life. This
would be valid in cases where experience indicates that corrosion propagation would be
rapid or where no corrosion can be tolerated. Another approach is to assume that the
corrosion stage lasts for a fixed period after the onset of corrosion.

If the duration of the propagation period is to be predicted there are 3 practical approaches:

1. Predict values for the corrosion rate as a function of the exposure class or controlled
parameters.
2. Measure the corrosion rate in test specimens (design phase) or in-situ (existing
structures).
3. Use an empirical relationship incorporating parameters representing effects known to
govern the process such as the chloride content, galvanic effects, formation of corrosion
products, oxygen availability, resistivity of the concrete, exposure environment and the
effects of ageing. Models have been developed to take these parameters into account.
However, there is not sufficient data to allow these to be validated at the present time.

4.3.1 Prediction of corrosion rate based on environment and controlled parameters


The corrosion of steel reinforcement arises through local structural and/or compositional
variations within the concrete as discussed in Section 3.1. The corrosion product occupies
a volume several times that of the parent metal and its formation creates internal stresses
that result in cracking and delamination of the concrete. The rate of corrosion once the
initiation period has finished depends on a number of factors. The key factors in the

24
ongoing corrosion are the availability of oxygen and whether or not the surface of the
reinforcement remains wet. A current density of 1.00 µA/cm is equivalent to a rate of metal
2

loss for a circular reinforcing bar of 11.57µm/y. The diameter decrease is therefore given
36
by :
Licensed copy from CIS: qacentral, HBK Contracting Company W.L.L., 25/06/2016, Uncontrolled Copy.

φ (t ) = φ (i ) − 0.023I corr t

where φ(t) is the bar diameter at time t, φ(i) is the initial reinforcement diameter (both in mm)
and Icorr is the corrosion current.

The corrosion current is largely controlled by:


1. The resistivity of the concrete.
2. The availability of oxygen at the cathode.
3. The temperature.

Both the resistivity and the oxygen availability are dependent on the moisture level in the
concrete. The resistivity falls as the moisture level increases and, consequently the rate of
corrosion increases. The resistivity is also dependent on the level of dissolved salts in the
concrete and the continuity of the pore system. However, at higher moisture levels the
availability of oxygen becomes the factor most influencing the rate of corrosion and the rate
falls as the O2 permeability of the concrete drops. The rate of corrosion may also fall with
time at a given relative humidity as a consequence of the thickening corrosion product layer
and the consequent reduced diffusion rate.

Corrosion rates for carbonated concretes given in the literature vary widely. This may arise
through a combination of factors including the following:
• The presence of chloride ions (these will increase the rate of corrosion in carbonated
concrete).
• The type of cement
• Whether concrete or mortar specimens are used
• Conditions (laboratory or real structures)
• The method of measurement
• Temperature

For carbonation-induced corrosion the rate of corrosion has been related empirically to the
37
relative humidity. Parrott has given data (expressed as loss in diameter of steel
reinforcement) for carbonation-induced corrosion rates as shown in Table 8.

Table 8. Corrosion rate for carbonation-induced corrosion (from Reference 37)

Relative humidity (%) Corrosion rate (µm/year)


40 0.3
50 0.3
60 2
70 2
80 5
90 10
95 20
98 50
100 10

25
These corrosion rates relate to the reduction of diameter of the reinforcement. They are
also upper ‘limits’ based on an analysis of the available literature (there is a 10% chance of
them being exceeded). Table 9 shows a separate set of corrosion rates for carbonation-
induced corrosion.
Licensed copy from CIS: qacentral, HBK Contracting Company W.L.L., 25/06/2016, Uncontrolled Copy.

Table 9. Corrosion rate for carbonation-induced corrosion64

Relative humidity µm/year)


Corrosion rate (µ
<50% 0.1
>50%<60% 0.2
>60%<70% 0.3
>70%<80% 0.5
>80%<90% 1.0
>90%<90% 5.0
>98% 0.1

Similarly the rate of chloride-induced corrosion can vary widely. Table 10 gives corrosion
rates for different exposure classes in EN206 (taken from reference 42).

Table 10. Corrosion rates for different exposure classes given in EN206

Exposure classes Vcorr [µm/year] W t [-]


Mean Std Dev Mean Std Dev
0 No risk of carbonation 0 - 0 -
Totally carbonated
XC1 Dry 0 - 0 -
XC2 Wet, rarely dry (unsheltered) 4 3 1
XC3 Moderate humidity (sheltered) 2 1 0.5
XC4 Cyclic wet-dry (unsheltered) 5 3 0.75
Chloride-induced corrosion
XD1 Wet rarely dry 4 3 1
XD2 Cyclic wet-dry 30 20 0.75
XS1 Airborne sea water 30 20 0.5
XS2 Submerged Not expected except bad concrete/low cover
XS3 Tidal zone 70 40 1

Vcorr is the corrosion rate


W t is the fraction of the year during which the moisture state applies such that corrosion is
active.

The table assumes that the concrete grades recommended in prEN206 are used. Other
38
data have suggested that the corrosion rate for concrete which is highly contaminated with
chlorides can exceed 1100 µm/yr in concrete that is highly contaminated with chlorides.
Where pitting occurs the maximum rate is about 4 to 8 times greater than the average for
39
general corrosion .

The rate of chloride-induced corrosion increases with the chloride ion content. A
relationship of the form:
bCx
corrosion rate = ae

26
64
has been assumed in a model developed by Bamforth et al where Cx is the chloride
concentration at the reinforcement and the parameters a and b are dependent on the
exposure environment. Pitting corrosion is reported to increase the rate of corrosion by a
Licensed copy from CIS: qacentral, HBK Contracting Company W.L.L., 25/06/2016, Uncontrolled Copy.

factor of up to 10. This model uses a nominal factor of 5 to reflect the effects of pitting on
the duration of the propagation period after corrosion has been initiated.

It has been suggested that the rate of corrosion observed on site are substantially lower
40
than those in the laboratory, even under the most severe conditions . Corrosion rates
associated with carbonation measured on site rarely exceed 1 µA/cm whereas chloride-
2

induced corrosion rates can significantly exceed 1 µA/cm . This may be due to the fact that
2

surveys are normally done in dry conditions when the rate of corrosion is comparatively low;
41
carbonated concrete dries out more rapidly than chloride contaminated concrete .

Corrosion in the presence of chlorides can lead to pitting and, if this occurs, the use of
visible cracks to specify the end of the service life cannot be used. When the availability of
oxygen is not the controlling factor corrosion can occur very rapidly in these conditions. It
has been suggested that the propagation period under these conditions should be taken as
a small nominal value of between 0 – 10 years.

4.3.2 Amount of corrosion needed to cause cracking


The corrosion process leads to a decrease in the diameter of the steel reinforcement
together with the production of high volume oxide. The generation of oxide leads to
cracking when the tensile strength of the concrete is surpassed. These cracks start to form
at the interface between the bar and the concrete and propagate radially until they reach the
36
surface of the concrete . The time for cracking to occur once corrosion has been initiated
will depend on the corrosion rate and the amount of corrosion required to cause cracking.
The duration of the propagation stage is determined by the definition of the end of service
life. If the onset of corrosion is used then the propagation stage will not be a consideration.

The end of the design life may be taken as the point at which visible cracks (0.1 –0.2 mm)
37
are first apparent. This is equivalent to 100µm of corrosion . This value makes no
allowance for bar diameter, geometry. If this point is taken as the end of service life then
the following simple equation can be used for the duration of the propagation stage:

tp =100/Cr

Where tp is the duration of the propagation stage and Cr is the corrosion rate in µm/year. If
the mean relative humidity is known, corrosion rate data could be read from Table 1 and
used to predict the duration of the propagation stage. Corrosion rate data from in test
specimens from existing structures can also be used.

More sophisticated (but still empirical) models have been developed. The following
42
expression has been used to relate the amount of corrosion per year to cause cracking to
the cover/bar diameter ratio and the tensile splitting strength of the concrete:

X c = 83.8 + 7.4(c / φ ) − 22.6 f st

where
Xc is the loss of steel in microns
c is the depth of cover in mm
φ is the bar diameter in mm
fst is the splitting tensile strength in MPa

27
The fact that the amount of corrosion per year required to produce cracking decreases as
the tensile strength increases reflects the fact that the corrosion product can more easily be
accommodated in a poor quality concrete – this is the critical factor rather than the strength
of the concrete. The tensile strength can be related to the cube compressive strength using
Licensed copy from CIS: qacentral, HBK Contracting Company W.L.L., 25/06/2016, Uncontrolled Copy.

one of a number of empirical relationships. These relationships tend to be of the form:

f st = k ( f cu ) n
43
Neville has suggested respective values of 0.3 and 2/3 for k and n for a wide range of
data.
64
The following expression has been used for modelling purposes :

f st = 0.108{ f cu [1 + 0.399 log( f cu )] − 5.334} 0.722

where fcu is the cube compressive strength.

Andrade et al36 give the following (Table 11) expected times needed to cause cracking in a
3
PC concrete with a cement content of 350kg/m using a w/c ratio of 0.5. The rebar was 16
mm in diameter and the sample dimensions were 15 x 15 x 38 cm. The tensile strength was
3.55 MPa on average. Generalised attack, as in carbonation-induced corrosion, is
assumed. Higher localised penetration rates may occur in the case of chloride induced
corrosion.

Table 11 Expected time to cracking (from Reference 36)

Time to cracking (years)


Corrosion rate (µm per year) Crack width: 0.05 – 0.1 mm Crack width: 0.2 – 0.3 mm
(20µm cross section loss (100-150µm cross section
required) loss required)
1 20 >100
10 2 10-15
100 0.2 1-2
1000 0.02 0.1-0.2

The table shows that the time period between depassivation and the appearance of visible
cracks may be very small and, under these conditions, the use of the initiation period as the
service life is valid. However, the formation of crack widths of 0.3 mm may take much
longer.

4.4 The effect of temperature on corrosion initiation and propagation

Temperature influences the rate of both the initiation and propagation stages of corrosion:

• The chloride diffusion coefficient has been reported to follow an Arrhenius-type


44
dependence on temperature . Activation energies are reported to be in the range 32 –
45 KJ/mole for cement pastes. However, bulk diffusion tests have suggested that there
is little or no temperature dependence.
• The chloride threshold has been reported to fall with temperature.
• Carbonation is generally treated as being independent of temperature.

28

o
It is often assumed that increasing the temperature by 10 C doubles the rate of
45
corrosion. However, it has been suggested that the actual ratio is about 1.6.

46
The threshold level is also temperature dependent. Benjamin and Sykes and give the
following relationship:
Licensed copy from CIS: qacentral, HBK Contracting Company W.L.L., 25/06/2016, Uncontrolled Copy.

Ct = C20(1-0.0333(T-20))
o
where Ct is the threshold level at temperature t and C20 is the threshold level at 20 C.
However, again this has not been considered in the current model.

4.5 The effect of cracks

Cracks in concrete can increase the rate of penetration of aggressive agents such as
carbon dioxide and chlorides and consequently reduce the service life of the structure or
12 62
element . The formation of cracks in reinforced concrete is discussed in Part 1 of Digest
444. The main factors affecting corrosion at cracks are crack width, frequency, orientation
with respect to the reinforcement, whether the cracks are static or dynamic, the exposure
environment and the quality and depth of the cover concrete.

There is an inverse relationship (with a large amount of scatter) between the crack width
and the time to the onset of corrosion for reinforced concrete in which the crack is
perpendicular to the reinforcement. For normal crack widths corrosion initially propagates
more rapidly in wider cracks. Total reinforcement loss would be expected to increase with
48
the frequency of intersecting cracks .

Present codes and standards seek to limit the probability of unacceptable deterioration at
cracks by setting out rules for good practice for reinforcement detailing and limiting
permissible crack widths.

The models discussed here assume that the concrete is effectively homogeneous (although
the data used in the calibration of the models will have derived from cracked concrete) and
therefore do not specifically address the effects of cracking.

29
5 Models for other degradation mechanisms
5.1 Sulfate attack
Licensed copy from CIS: qacentral, HBK Contracting Company W.L.L., 25/06/2016, Uncontrolled Copy.

47
Atkinson and Hearne have developed a model to predict the effects of sulfate-containing
groundwaters on the service life of concrete. In this model cracking and delamination is
assumed to occur as a result of stresses induced by sulfate ingress (through diffusion) and
subsequent expansive reaction (leading predominantly to the formation of the mineral
ettringite) with aluminate-containing cement phases. As the concrete cracks and
delaminates new surfaces are exposed. The depth of degradation increases linearly with
time and depends on the concentrations of sulfate ions and aluminates, the rate of diffusion
and reaction and the fracture energy of the concrete.

X Spall Eβ 2 c 0 C E Di
R= =
t Spall αγ (1 − ν )

Where:
PC SRPC
R degradation rate, 1.06mm/y 0.70mm/y
XSpall critical thickness of the reaction zone at 1.6mm 3.9mm
spalling
tSpall time to the onset of spalling 1.5y 5.6y
E Young’s modulus of the concrete 20GPa
-6 3 -1
β linear strain caused by 1 mole of sulfate 1.8x10 m mol
3
reacted in 1m
-3
c0 external sulfate concentration 0.01Moldm
3 3
CE concentration of reacted sulfate as 369mol/m 236 mol/m
ettringite.
-12 2
DI intrinsic diffusion coefficient. 10 m /s
α roughness factor for fracture path 1
2
γ fracture surface energy of concrete 10Jm
ν Poisson’s ratio. 0.3

Some of the values require to be determined by laboratory experiments. Others are ‘typical’
values as values for specific concrete types and mix designs cannot be readily derived.
Atkinson and Hearne have used this model to predict degradation rates for Portland cement
and sulfate-resisting Portland cement concretes in 0.01M sulfate solutions. The
assumptions and results are summarised in the above table.

Time to failure models for delayed ettringite formation (an internal sulfate attack process)
and the thaumasite form of sulfate attack have not been developed to any degree.

5.2 Freeze-thaw

The mechanism of frost attack on concrete is not well understood. Water may freeze in:
• Capillary pores
• Voids and defects
• Porous aggregate

It may occur at the surface (leading to scaling) and within the concrete (leading to
deterioration of the physical properties of the bulk concrete) and can arise through variations
in the environment and/or the application of deicing salts.

30
The mechanisms and processes responsible for frost attack are not well understood.
Susceptibility is affected by the composition, porosity, moisture content and age of the
concrete, the environment of exposure (including temperature cycles, wet/dry cycles and the
presence and type of salt).
Licensed copy from CIS: qacentral, HBK Contracting Company W.L.L., 25/06/2016, Uncontrolled Copy.

A 2-stage process has been suggested for internal freeze-thaw-induced deterioration. The
initiation period is the time taken for a critical degree of saturation (which can be determined
experimentally) to be reached. Concrete is assumed to be highly resistant to frost action
where its degree of capillary saturation (Scap) remains below the critical level (Scr). Scap
depends on the properties of the concrete and the environment of exposure.

Where Scap exceeds the critical value the risk of damage is substantial and is likely to occur
following the first freeze-thaw cycle.
48
The following model has been suggested :

F= Scr - Scap = Scr – d – elog(t.Y)

where d and e are constants that can, in principle, be derived from laboratory experiments
and Y is a correction factor to allow for intermittent wetness in practice. Initiation is likely to
occur when F < 0.
48
The following model has been suggested for predicting the time to a certain level of frost
damage.

t CR = (WCR .n) /(k1 .k 2 .Wn .x.n k 3 )

where:

W CR is the critical degree of scaling


tCR is the time taken to reach a the critical level
n is the number of freeze thaw cycles in a laboratory test.
x is the number of freeze thaw cycles in a year for the concrete in the environment in which
it will be used.
W n is the actual degree of scaling per cycle
k1k2k3 are correction factors to take account of the fact that the relationship between the
amount of scaling and the number of freeze/thaw cycles is affected by various concrete and
environmental parameters and need not be linear.
6
The following general equation has been proposed as a model for the strength reduction at
the surface of a concrete element due to frost attack:

   d n 
f ck (d ) = f ck 1 − 1 −    
   H  
  

Where:

fck(d) is the characteristic compressive strength of concrete at depth d.


fck = the characteristic compressive strength of undamaged concrete
d = depth
H = depth of influence
n is an index related to the number of freeze-thaw cycles or time.

31
For structural design the loss in strength at the edge could be treated as a decrease in the
design strength or as a reduction in the cross sectional area of the concrete.
Licensed copy from CIS: qacentral, HBK Contracting Company W.L.L., 25/06/2016, Uncontrolled Copy.

Other models for freeze-thaw-induced damage are summarised in reference 48. However,
models for freeze-thaw induced damage have not been calibrated or validated due to the
absence of appropriate input data. As stated in Table 1 the approach taken to durability
design in the context of freeze-thaw attack is therefore to avoid the problem using measures
such as:
• Air entraining admixtures
• Low w/c ratio concrete
• Frost-resistant aggregate
• High strength concrete
• Keeping the concrete dry

5.3 Models for Alkali-silica reaction

Alkali silica reaction (ASR) occurs where alkalis in the pore solution react with siliceous
constituents of the aggregate to form an expansive gel which, on imbibition of water, can
lead to cracking of the concrete.

The concept of initiation and propagation can be applied to ASR. However, as a result of
the complexity of the process and the number of contributory factors suitable models have
not yet been developed or verified. As stated in Table 1 the approach taken to durability
design in the context of ASR is therefore to avoid the problem through the use of
appropriate mix designs etc.

RILEM 14 gives brief discussions of modelling surface deterioration due to weathering and
6
leaching and the abrasion of concrete by ice .

32
6 Basic protection strategies for reinforced concrete
Good conceptual and detailed design is vital in producing durable concrete structures and
elements. They can ensure that the effects of water on the structure are minimised to
Licensed copy from CIS: qacentral, HBK Contracting Company W.L.L., 25/06/2016, Uncontrolled Copy.

prevent the penetration of water (and aggressive species contained in it) at vulnerable areas
and contribute to overall quality by ensuring that structures and details are easy to build.
Good workmanship is also crucial, as is the maintenance of the structure in use.

There are 2 main approaches that can be used in improving the durability of concrete
structures and elements and, depending on the circumstances, there are a number of
options within each:

Design out
The objective of design out approaches is to avoid prevent deterioration either by isolating
the element or structure from an aggressive environment or by using non-reactive materials.
Examples include:

• The exclusion of water and aggressive species contained in it through the use of an
enveloping cladding system, continuous structures to avoid joints etc.

• The use of non-reactive materials such as non-ferrous reinforcement and sulfate-


resisting cements. Degradation processes, in which the concrete itself deteriorates, such
as sulfate attack and alkali aggregate reaction, can in many cases be prevented through
appropriate materials selection.

• The use of cathodic prevention to avoid reinforcement corrosion.

Provide resistance
In this approach the ability of the concrete to resist the environment is controlled. There are
2 options:

• Single-layer protection in which parameters such as the depth of cover and the
permeability of the concrete, are used to provide an appropriate service life.

• Multi-layer protection in which additional measures are used in addition to good quality
concrete and appropriate cover depths. Examples include coatings, corrosion inhibitors
and alternative reinforcement.

These techniques are generally complimentary and the use of combinations of materials,
design and construction options, may be appropriate for high criticality structures with long
required service lives, and in aggressive environments where the use of concrete quality
and cover thickness may not give the required service life. The effects of these techniques
on durability and the approach to modelling their impact on service life are summarised in
Table 12.

33
Licensed copy from CIS: qacentral, HBK Contracting Company W.L.L., 25/06/2016, Uncontrolled Copy.

Table 12 Summary of protection strategies for reinforced concrete

Technique Potential beneficial effects on durability Approach to modelling


High quality • Reduced permeability reduces rate of chloride and CO2 ingress. High cement Through use of appropriate values for parameters in
concrete content provides pH buffer against carbonation. models discussed above.
Use of pfa • Substantially reduced chloride diffusion coefficient in the long term (initial rate of Through use of appropriate values for parameters in
chloride ingress may be comparatively rapid). May reduce chloride threshold level. models discussed above.
• May lead to increased rate of carbonation when compared with PC concrete on basis
of equal w/c ratio. Insignificant difference when compared on basis of equal strength
grade.
Use of ggbs • Substantially reduced chloride diffusion coefficient in the long term (initial rate of Through use of appropriate values for parameters in
chloride ingress may be comparatively rapid). models discussed above.
• May lead to increased rate of carbonation when compared with PC concrete.
Silica fume (SF) • At low w/c ratios substantially reduced chloride diffusion coefficient in the long term Through use of appropriate values for parameters in
(initial rate of chloride ingress may be comparatively rapid). May reduce chloride models discussed above (limited data).
threshold level.
• No significant effect on the rate of carbonation up to 10% addition levels when
compared with PC concrete of equivalent grade. Higher carbonation rates at higher
addition levels.
Metakaolin • Substantially reduced chloride diffusion coefficient. Through use of appropriate values for parameters in
• Higher rate of carbonation than PC concrete with the same w/c ratio. models discussed above (limited data).
Plasticising • Reduction in w/c ratio reduces all environmental penetration including chloride Using data relating transport properties to (reduced)
admixtures ingress and carbonation. w/c ratio.
Superplasticisers • As plasticisers only more effective
Integral • Can reduce capillary suction and hence chloride ingress. No significant effects on Insufficient data to allow inclusion in models.
waterproofers diffusion expected. Best used where hydrostatic pressures are low. No significant
effect on carbonation.
Pore blockers • Reduced build up of surface chloride. Reduced chloride penetration under pressure. Insufficient data to allow inclusion in models.
No effect on chloride diffusion. No significant effect on carbonation.
Fibres • Lead to indirect improvements in durability through improved quality of surface zone No data to allow inclusion in models. Possible
concrete. modification of safety factors?

34
Licensed copy from CIS: qacentral, HBK Contracting Company W.L.L., 25/06/2016, Uncontrolled Copy.

Table 12 Summary of protection strategies for reinforced concrete (continued)


Technique Effects on durability Approach to modelling
Controlled • Chloride diffusion coefficients reduced to between 40% and 70% of concrete cast Treated as increased cover based on estimation of
permeability using conventional formwork. depth affected and reduced diffusion coefficient
formwork • Increased resistance to carbonation due to improved quality at surface. based on limited literature data. Surface chloride
level reduced.
Permanent • Permanent formwork can be used to provide a durable outer skin for the structure. Treated as effective increase in cover based on
formwork Improved chloride and carbonation resistance (depending on nature of materials thickness and diffusion coefficient.
used) as structural concrete is sheltered from environment.
Surface barrier • Dependent on the nature of the coating protection of new structures against chloride Extension to initiation period based on nature and
coatings ingress and carbonation can be effective for long periods. Coatings must be applied demonstrated performance of the coating.
correctly and regularly maintained. Coatings may increase the rate of chloride ingress
into the concrete once the protective effect has been lost by maintaining the concrete
in a relatively wet state. The apparent diffusion coefficient under these conditions is
higher than in an equivalent dry concrete49.
Corrosion • No effect on chloride ingress but effectively raise chloride threshold level, increasing Threshold level assumed to increase linearly with
inhibitors the time to the onset of corrosion and therefore increasing the durability of reinforced addition level based on experimental data. Safety
concrete in chloride-rich environments. factor may be incorporated.
• No effect on carbonation-induced corrosion.
Coated steel • Where chloride levels are not excessive (up to 1% by weight of cement1) and where Epoxy coating assumed to increase propagation
the pH is below 13.3 galvanising acts as an effective barrier to chloride-induced stage by 20 years in one US model (Life-365).
corrosion. It also provides effective protection against carbonation-induced corrosion.
• Where the coating remains intact epoxy coatings can provide good resistance to Galvanising assumed to increase the threshold to 1%
chloride induced corrosion. However, where the coating is scratched or damaged by weight of cement. Assumed to prevent
corrosion can occur; anodic activity can be intense where uncoated rebar is also carbonation-induced corrosion.
present to act as the cathode1.
• Epoxy coatings act as effective barrier to carbonation-induced corrosion.
Stainless steel • Highly effective means of preventing corrosion in most conditions and where high Threshold increased by a factor of about 10 to 3% by
initial costs can be justified. weight of cement for type 316.
Non-ferrous • Not prone to corrosion. Some resins may not be durable in alkaline conditions. No data on durability of materials. Assume no
reinforcement corrosion occurs.
Impressed • Prevents chloride-induced corrosion. Maintains alkalinity around reinforcement and Assume no corrosion if conditions adequately
current cathodic the stability of the passive oxide film. maintained.
protection
Sacrificial anodes • As impressed current technique. No data

35
Service life assessment is an important factor in the maintenance and repair of existing
structures as discussed in Part 2 of Digest 444. The approach taken is similar to that for
new structures and the models discussed below can readily be applied. Depth of
carbonation and chlorides, and corrosion rates measured on the actual structure mean that
Licensed copy from CIS: qacentral, HBK Contracting Company W.L.L., 25/06/2016, Uncontrolled Copy.

means that the models can be applied with greater reliability than for service life design
where these values have to be predicted based on existing data.

36
7 ANALYSIS OF BRE DATA
7.1 Carbonation
Licensed copy from CIS: qacentral, HBK Contracting Company W.L.L., 25/06/2016, Uncontrolled Copy.

BRE data are available showing depth of carbonation over time for concretes made using
different cement types, cement contents, w/c ratios etc. A range of different curing regimes
has also been used. A database has been assembled using the available information and
has been analysed to determine the feasibility of using BRE data to test and develop simple
models for predicting the rate of carbonation for different concretes. The parameters that
could influence the rate of carbonation are summarised in Table 13.

Table 13. Parameters affecting the rate of carbonation

Parameter Influence Comments


Controlled Variables
Cement type Affects buffering capacity Permeability generally falls with increasing pfa
and physical parameters content up to at least 50%. The trend is more
marked in well-cured concretes.
W/c ratio Affects physical parameters, Carbonation rate and permeability generally fall
may influence amount of with w/c ratio – other parameters being held
hydrated cement present and constant;
hence buffering capacity
Cement Affects buffering capacity Very little data are available concerning the
content effects of cement content on the rate of
carbonation.
Aggregate May lead to through This is being studied under a separate study at
type aggregate carbonation. BRE.
Cover Directly determines initiation Models need to give the depth of cover required
time for a given mix design for a given service life as an output.
Random Variables (curing)
Curing time Very significant effects on A single parameter such as the compressive
permeability strength at 28 days or the air permeability could
be used as a measure of the ‘quality’ of the
concrete prior to exposure.
RH of cure High RH is likely to give a
better quality concrete.
T of cure Ambient cured concretes are
likely to have better quality
than those cured at
significantly elevated
temperatures or at low
temperatures.
Random variables (environmental)
[CO2] Rate of carbonation .Most models assume a square root relationship
increases with [CO2] with [CO2], presumably following Fick’s law.
Temperature Arrhenius relationship
RH Maximum carbonation at Can be modelled using empirical relationships.
about 65% RH

The following graph (Figure 7) shows the distribution of carbonation depths (normalised to
0.5
20 years by extrapolation using a t relationship – see below) for BRE data for a wide
range of mix designs and cement types.

37
Carbonation depth normalised to 20 years
for all available data
Licensed copy from CIS: qacentral, HBK Contracting Company W.L.L., 25/06/2016, Uncontrolled Copy.

35
30
Frequency

25
20
15
10
5
0
12
16
20
24
28
32
36
40
44
48
52
0
4
8

Depth/mm

Figure 7 Distribution in carbonation depths after normalisation to an age of 20 years for


all available BRE data

The graph shows that there is a wide range of values. The scatter arises from:
• Differences in cement type and content
• Differences in concrete mix parameters
• Differences in curing regime
• Inter-sample variability.

If the skew in the distribution is ignored and a normal distribution assumed the mean and
standard deviation are 30.70 mm and 12.68 mm respectively. If a correction for cement
content is included (discussed below) the mean and standard deviation are 23.96 and 7.60
mm respectively (Figure 8).

Carbonation depth at 20 years


(normalised and corrected) for all
available data
50

40
Frequency

30
20

10
0
0 6 12 18 24 30 36 42 48

Depth/mm

Figure 8 Carbonation depth normalised to 20 years and corrected for cement content

38
For service life design and prediction the depth of carbonation needs to be related to one or
more of the following:

1. Mix design parameters such as w/c ratio, cement content, aggregate content etc.
Licensed copy from CIS: qacentral, HBK Contracting Company W.L.L., 25/06/2016, Uncontrolled Copy.

2. Curing conditions.
3. The use of additional protective measures such as coatings.
4. Environmental conditions.
5. Measurable parameters such as carbonation depth at early ages, oxygen permeability,
compressive strength etc.

For service life design it should, ideally, be possible to predict the rate of carbonation from
the mix design and the exposure conditions, probably backed up with trial mixes or early age
measurements. However, relating the carbonation depth to a physical parameter (such as
compressive strength or permeability) that is measurable at comparatively early ages may
eliminate much of the variability due to mix design and cure. Parrott has suggested that air
permeability at 18 months would be a good measure. However, permeability has been
measured only on a limited number of BRE samples and only 28-day data are available.

A plot of carbonation depth (normalised to 20 years) against the measured 28 day oxygen
50
permeability for a range of PC and PC/pfa concretes (from BR216 ) are summarised in
Figure 9.

39
Normalised 20-year carbonation depth against O2 permeability for PC & PC/pfa concretes

60
Licensed copy from CIS: qacentral, HBK Contracting Company W.L.L., 25/06/2016, Uncontrolled Copy.

50
y = 11.152x + 9.0383
R2 = 0.7771
Carbonation depth/mm

40

30

20

10

0
0 0.5 1 1.5 2 2.5 3
-16 2 0.5
(O2 permeability/10 m )

Figure 9 Graph showing relationship between carbonation depth and O2 permeability for
Portland cement and PC/pfa concretes using data taken from BR216 (indoor
exposure only).

34
Parrott has found that the carbonation depth (based on a wider range of data) is
empirically related to the permeability, k, raised to the power of 0.4 rather than to k itself.
f 0.5
However, diffusion theory suggests that it should be related to k and this has been used
in Figure 6. The normalised carbonation depth has been corrected for variations in buffering
capacity arising from the use of different cement contents and pfa levels using the term:

(Portland cement content + n x pfa content)/300 (a)

where n is a parameter reflecting the contribution to buffering capacity made by the pfa.
The relationship between permeability and carbonation depth, d, at 20 years for BR216 data
can therefore be expressed as:

d = (11.2k0.5 +9.0)(t/20)0.5/C (b)

Where

C = (PC content)/300 (c)

where pfa has been assumed to have a buffering capacity of 0 and t is the age in years.
However, when the same data are normalised to different ages different lines of best fit are
0.5
obtained. These lines are not related by a factor of t as would be expected from equation
(b), although this can be used as an approximate value, as the intercept is not zero. The
parameters for the lines of best fit are given in the following table where the line of best fit is:

d x C = mk0.5 + n (d)

where m and n are parameters as given in Table 14.

f
Assuming that the diffusion coefficient for the movement of CO2 through the concrete is linearly
related to the permeability

40
Table 14. Carbonation rate parameters derived from BRE data
Licensed copy from CIS: qacentral, HBK Contracting Company W.L.L., 25/06/2016, Uncontrolled Copy.

time/years m n standard
deviation/mm
1 3.38 0.58 0.76
5 6.15 3.59 1.55
10 8.22 5.85 2.28
20 11.15 9.04 3.37
30 13.40 11.49 4.21
50 16.97 15.37 5.55
70 19.97 18.53 6.65
100 23.52 22.51 8.03
200 32.79 32.60 11.55

The relationships between m, n and σ and time are approximately:

m = 3.165t0.49

n = 1.47t0.59

σ = 0.71t0.52

The relationship between d and k can then be written:

d = (3.165t0.49 x k0.5 +1.47t0.59)/C (e)

The relationship between carbonation depth and oxygen permeability given by Parrott is:

d = ak0.4tn/c0.5

Where a is a coefficient equal to 64, k is the oxygen permeability at 18 months, t is the time,
n is a power exponent close to 0.5 and c is the calcium oxide content. This equation differs
from that used in the above graphs in that the cement content has been used instead of the
square root of the CaO content (for BRE data a better correlation is obtained using the
cement content rather than its square root). The non-zero intercept in the graph and
equation derived from BRE data presumably arises from changes in the permeability arising
from drying on exposure.

Assuming the BRE data is representative of the broader picture this analysis could allow the
distribution of carbonation depths at a given age to be predicted for concrete subject to
indoor exposure, based on the known cement content and the measured permeability at 28
days. Additional factors need to be built in to address external exposure. A safety factor
reflecting the difference between 100 mm cubes and real structures is also required. Both
the mean value and the standard deviation would be expected to vary on placing these
concretes in external sites where the temperature and relative humidity are likely to vary as
well as the CO2 level.

41
7.1.1 Outdoor carbonation
Outdoor sheltered carbonation depths for BRE samples are, other factors being equal, lower
than those obtained for equivalent samples exposed indoors. The following graph plots
indoor and outdoor carbonation depths (normalised to 20 years and corrected for cement
Licensed copy from CIS: qacentral, HBK Contracting Company W.L.L., 25/06/2016, Uncontrolled Copy.

content) for concretes that have received otherwise identical treatments:

Outdoor against indoor carbonation depths corrected for


buffering capacity (normalised to 20 years in both cases)

60
Indoor carbonation depth

50

40
(mm)

30

20

10

0
0 5 10 15 20 25 30 35 40
Outdoor carbonation depth (mm)

Figure 10. Outdoor (sheltered) against indoor carbonation depths for BRE concretes

Figure 10 shows that indoor carbonation is greater than outdoor carbonation. However, the
g
line of best fit does not pass through the origin . A plot of outdoor (sheltered) carbonation
depth against 28-day oxygen permeability is given in following Figure 11.

Outdoor carbonation depth (normalised to 20 years)


against Oxygen permeability
35
Carbonation depth/mm

30
25
20
15
10
5
0
0 0.5 1 1.5 2 2.5 3
-16 0.5
(Permeability/10 )

Figure 11 Carbonation depth (normalised to 20 years) against O2 permeability for BRE


concretes exposed outdoors (data taken from BR216).

g
This may be related to RH and drying effects at early ages arising from differences in exposure
conditions.

42
The equation for the line of best fit is

d = (7.95k0.5 +3.44)
Licensed copy from CIS: qacentral, HBK Contracting Company W.L.L., 25/06/2016, Uncontrolled Copy.

7.1.2 Variability in relative humidity, carbon dioxide level and temperature


51
Wierig has derived a relationship between the carbonation rate and the relative humidity of
the environment in which the concrete is exposed:

Krel = -0.000508(RH)2 + 0.0556(RH) - 0.4472 (f)

where Krel is the carbonation rate relative to that at 65% RH.

This expression could be used in conjunction with the expression derived for indoor
carbonation at time t as a function of the measured permeability:

0.49 0.5 0.6


d = (3.2 x t xk +1.5 x t ) x Krel/C (g)

Figure 12 shows the lines of best fit for indoor and outdoor exposure derived from the data
in BR216. It also shows a line from the above equation using a relative humidity of 85% in
the Wierig equation for Krel. This gives a reasonable agreement with the outdoor BRE data.

Predicted outdoor carbonation depths based on


measured outdoor and indoor data
50
carbonation depth/mm

best fit for outdoor data


Predicted 20-year

40
best fit for indoor data
30

20 outdoor carbonation depth


using indoor data and
10 assumed RH of 85%

0
0 1 2 3
-16 2 0.5
(Oxygen Permeability/10 m )

Figure 12. Predicted outdoor carbonation depths based on measured outdoor and indoor
data

The graph suggests that the above expression could be used to predict the depth of
carbonation in different environments based on a measured value of oxygen permeability
and assuming that the average relative humidity of the environment is known. There are
only very limited data available at present allowing the model to be tested. However, data
52
are available in CR 216/94 . This report includes carbonation and permeability data for PC,
PC/ggbs, PC/limestone filler, PC/pfa, PC/pfa/ggbs and PC/ggbs/limestone filler concretes.
3
The buffering capacity has been normalised to the equivalent of a PC content of 300 kg/m
using the equation:

43
C = (PC content + 0.6 x ggbs content)/300 (h)

Pfa and limestone are assumed not to contribute to the buffering capacity. The outdoor
Licensed copy from CIS: qacentral, HBK Contracting Company W.L.L., 25/06/2016, Uncontrolled Copy.

carbonation depth (normalised to 20 years) is plotted against the oxygen permeability in


Figure 13.

Outdoor carbonation depth (normalised) against permeability for EU


cement types
35

30
Carbonation depth/mm

25

20

15

10

0
0 0.5 1 1.5 2 2.5 3
-16 2 0.5
(Oxygen permeability/10 m )

Figure 13 Carbonation depths for CR216/94 data plotted against O2 permeability together
with predicted curve

Although there is considerable scatter about the best straight line and the data are limited
the graph shows that there is good agreement between the line of best fit (the solid line) and
that predicted (shown by the long dashed line) using the equation derived using indoor data
from BR216:

d = (3.2 x t0.49 x k0.5 +1.5 x t0.6) x Krel/C (I)

The standard deviation for all data is 4.4 mm when normalised to 20 years. If this
agreement extends to other data sets then this should be a means of predicting carbonation
rates in environments with different relative humidities based on a measure of the oxygen
permeability and details of the cement content. However, this needs to be tested against
other data. The carbonation rate is dependent only on the buffering capacity, the
permeability and on environmental factors, particularly the relative humidity and local CO2
concentration.

7.1.3 Environmental concentration of CO2

There are only very limited data on the dependence of the carbonation rate on the
environmental concentration of CO2. The concentrations of CO2 have not been routinely
monitored in experimental programmes at BRE. However, the indoor CO2 concentration
53
has been monitored at regular intervals over the course of a day . It has been found to be
in the range 0.0312 - 0.0533% with a mean value of 0.0369% and a standard deviation of
0.0033%. The CO2 levels in a separate room was in the range 0.0367 - 0.044 %. The
average value was 0.0387% with a standard deviation estimated to be 0.0013%. A

44
separate series of measurements found the CO2 concentrations to be in the range 0.0337 –
0.0420% (mean value is not available). Based on these data the average value for the CO2
level for BRE samples has been assumed to be 0.037% (based on that for room 004). The
standard deviation has been taken as 0.0033%. However, the standard deviation reflects a
Licensed copy from CIS: qacentral, HBK Contracting Company W.L.L., 25/06/2016, Uncontrolled Copy.

variation in the CO2 level over time. The variability of CO2 levels with location is likely to be
a more valid factor to consider in modelling. In the absence of any other data this variability
has been assumed to be included in the scatter about the line of best fit in plots of
carbonation rate against oxygen permeability shown above.

Models of carbonation that specifically address CO2 levels generally take the rate of
carbonation to be proportional to the square root of the atmospheric concentration (or the
partial pressure difference) of CO2 following Fick’s law. The equation relating carbonation
rate to oxygen permeability can therefore be modified to reflect this - assuming an average
CO2 level of 0.037%.

No data are currently available for the outdoor CO2 levels at Garston. Neville says that the
concentration of CO2 in rural areas is about 0.03%. Parrott gives a value of 0.034% (for
Hawaii - although atmospheric levels are not thought to vary significantly globally - at least
away from urban environments). If these figures are similar to those on the external
exposure sites at Garston then the differences in observed outdoor carbonation rates are
likely to arise from differences in relative humidity on exposure. The value in unventilated
laboratories can reach 0.12% (Neville). Levels of up to 0.3% can occur in buildings. The
carbonation rate under such conditions, assuming similar concrete and relative humidity will
be about 3 times higher than that found for indoor exposure in the BRE programmes.

7.1.4 Temperature dependence of carbonation rate


If carbonation is taken to be a diffusion-controlled process and is expected to follow an
54
Arrhenius-type relationship :

D = Doexp(-E/KT) (m)

where D is the diffusion coefficient at temperature T, Do is a constant and E is the activation


energy. However, no data are available to allow this analysis to be carried any further at this
stage.

7.1.5 Carbonation initiation time model (based on BRE data only)


Based on the above analyses the following equation for the time to initiation for carbonation-
induced corrosion can be derived:

d = (3.17t0.49k0.5 x ([CO2]/0.037)1/2 +1.47t0.49 ) x Krel/C (j)

where

[CO2] is the atmospheric CO2 concentration expressed as a percentage.

C = (PC content + 0.6 x ggbs content)/300 (k)


and
Krel = -0.000508(RH)2 + 0.0556(RH) - 0.4472 (l)
0.5
The standard deviation for d (assuming a normal distribution) can be taken as about 0.98t
as a worst case based on all BRE data (i.e. indoor and outdoor).

45
7.1.6 Relating mix design parameters to 28-day oxygen permeability and
carbonation rate
The 2 key controlled parameters affecting the carbonation rate are the buffering capacity
Licensed copy from CIS: qacentral, HBK Contracting Company W.L.L., 25/06/2016, Uncontrolled Copy.

and the permeability of the concrete. Buffering capacity can be readily predicted from the
cement content and type (although it may also depend on the w/c ratio if hydration is not
complete). However, permeability is harder to predict.

Based on a preliminary analysis of oxygen permeabilities given in BR216 the values given in
Table 15 could be used as provisional inputs to the model to allow likely carbonation rates to
be predicted in the absence of measured values.

-17 2
Table 15. 28-day oxygen permeability data (/10 m ) for BR216 concretes

1-day cure 3-day cure 7-day cure


Concrete type and grade Average 95% limit* Average 95% limit* Average 95% limit*
C35 PC 49 70 27 40 18 32
C25 PC 65 85 36 50 25
C45 PC 44 60 24 35 20
C35 15% pfa 30 40 22 35 12 24
C25 15% pfa 62 70 32 50 17
C45 15% pfa 25 35 18 30 13
C35 30% pfa 33 50 18 25 11 20
C25 30% pfa 54 70 22 30 21
C45 30% pfa 18 35 16 25 9
C35 50% pfa 35 70 14 35 7 10
C25 50% pfa 41 75 11 30 10
C45 50% pfa 16 50 3 10 3

* Average values and standard deviations have been estimated from all available C35 data. Fewer
data are available for C25 and C45 concretes and the 95% confidence limits given are based on
the standard deviations for equivalent C35 data.

46
7.1.7 Relationship between compressive strength and carbonation rate
The correlation between indoor carbonation depth and 28-day compressive strength for
BR216 data is shown in Figure 14.
Licensed copy from CIS: qacentral, HBK Contracting Company W.L.L., 25/06/2016, Uncontrolled Copy.

20 year (normalised) carbonation depth against


compressive strength for BR216 concretes
60

50
Carbonation
depth/mm

40

30

20

10

0
10 20 30 40 50 60
28-day compressive strength/MPa

Figure 14 Carbonation depth against compressive strength for PC and PC/pfa concrete
from BR216

The Excel STEYX function (which gives the standard error of the predicted y-value for each
x in the regression) gives a value of 5.47 mm for this data (the value for the equivalent
analysis using air permeability was 4.29 mm). These data suggest that the carbonation rate
has a better correlation with the oxygen permeability than with the compressive strength. It
could, however, still be used as the basis for a prediction of the carbonation rate although
the level of additional cover that would be required to give the same confidence level will be
greater than if the oxygen permeability is used.

7.1.8 Variation in carbonation depth for individual concrete prisms


The values used in the analysis of carbonation depth with time are averages of values from
2 prisms, with up to 12 measurements made on each prism. A more detailed analysis of the
data shows that the carbonation depth for an individual prism can vary appreciably. The
following graph shows the variation of carbonation depths for individual prisms of PC and
PC/pfa concretes stored indoors for 4 years. The data have been normalised to set the
average carbonation depth for each prism to 100. Assuming a normal distribution the
standard deviation is 21.1.

47
distribution of carbonation depths for individual 75x 75 x300 mm
prisms (all average depths normalised to 100)
Licensed copy from CIS: qacentral, HBK Contracting Company W.L.L., 25/06/2016, Uncontrolled Copy.

140
120
100
Frequency

80
60
40
20
0

More
5

15

25

35

45

55

65

75

85

95

105

115

125

135

145

155

165
Nornalised depth

Figure 15 Variation in carbonation depth for measurements on 75 x 75 x 300 mm prisms


after 4 years (analysis assumes that the mean depth for each cube is 100)

7.1.9 Summary of analysis of carbonation data


The use of available BRE data in assessing approaches to modelling carbonation has
shown the following:

• Carbonation can generally be modelled using a square root of time relationship.

• The mix design (especially cement content) is a key parameter in determining the
carbonation rate. However, curing conditions and duration, and exposure conditions are
also very important and are difficult to take into account based on the data that are
available at present. It is therefore difficult to reliably predict rates of carbonation for a
particular mix design.

• A model that relates rate of carbonation to the oxygen permeability measured at 28 days
has been developed using BRE data and gives reasonable results when tested against
other data. However, its applicability over a wider range of mixes and conditions cannot
be tested.

• The rates of carbonation predicted using the PII model and the BRE model are
reasonably consistent.

• More data are required for the effects of curing and of different exposure conditions on
the rate of carbonation.

48
7.2 Chloride ion diffusion coefficients

Chloride ingress can be modelled using the equation for diffusion given in Section 4. The
equation can be used to model future chloride ingress in existing structures based on
Licensed copy from CIS: qacentral, HBK Contracting Company W.L.L., 25/06/2016, Uncontrolled Copy.

measurements of changes in chloride ion profiles with depth within the structure. However,
for the model to be used in design apparent diffusion coefficients and surface chloride
concentrations need to be derived from laboratory samples and real structures for different
concretes, curing conditions and exposure conditions. The diffusion coefficient and surface
chloride level are also time dependent and this can have a considerable effect on the
predicted rate of chloride ingress as discussed in the following sections.
64
The ‘Ageddce’ model developed by Bamforth et al uses empirical relationships derived
from available data to predict diffusion coefficients and surface chloride levels. However, it
allows the predicted diffusion coefficient and surface chloride level to be overwritten with
specific values in modelling chloride ingress.

Diffusion coefficients and surface chloride levels derived from BRE data are summarised
and discussed below.

7.2.1 Portland cement concretes


50,55,56
Data are available for C35, C25 and C45 concretes after 1 - 10 years exposure.
Diffusion coefficients and surface chloride concentrations are given in Table 16.

Table 16 Diffusion coefficients and surface chloride levels for PC concretes taken from
reference 50 (unless otherwise stated).
-12 2
D/10 m /s Cs/% (by wt cement)
Description Av SD Av SD Comments
C25 11.5 Max: 15.3 3.78 Max: 4.16 5 samples available (3 at 2 years, 1 at 4 years
3
(250kg/m Min: 8.1 Min: 3.15 and 1 at 10 years). Surface chloride level
w/c = 0.76) based on all available data. Diffusion
coefficient based on 2 year data only.
C35 9.09 3.85 3.11 0.72 Data available for 6 different curing regimes (5
3 o
(300kg/m and 20 C and 1, 3 or 7 days damp cure) and at
w/c = 0.63) Max:14.8 Max: 5.05 test ages of 1 - 10 years (18 samples in all).
Min: 4.4 Min: 2.63 Values given in this table are based on all
available data, irrespective of curing or test
age.
3 57
300 kg/m , 6.46 1.48 3.35 1.04 2 year data
w/c = 0.60
57
6.29 0.97 4.43 0.97 4-year data
3 58
310 kg/m , 5.96 ±1 4.26 ±2.5 2 year data
w/c = 0.60
58
5.2 ±1.1 5.1 ±1.1 4 year data
C45 3.4 Max: 4.7 3.29 Max: 4.46 5 samples available (3 at 2 years, 1 at 4 years
3
(350kg/m Min: 2.0 Min: 2.75 and 1 at 10 years). Surface chloride level is
w/c = 0.54) based on all available data. Diffusion
coefficients based on 2 year data only.
16
C40 15.5 12.2 2.53 0.45 Bamforth
(w/c = 0.66) For 1m x 0.5m x 0.3m block
59
W/c = 0.45 1 0.4 Hobbs assumptions for CEM I concretes
based on field data
59
W/c = 0.4 0.7 0.4 Hobbs assumptions for CEM I concretes
based on field data

49
An analysis of these data shows the following:

1. Apparent chloride ion diffusion coefficients for PC concrete measured using 300 x 100 x
Licensed copy from CIS: qacentral, HBK Contracting Company W.L.L., 25/06/2016, Uncontrolled Copy.

100 mm prisms increase at later test ages as a result of sample geometry. Chloride
ions will ingress from each face and will therefore lead to higher levels at the centre of
the prism than expected from predictions made using data obtained at earlier ages.
These D values are not suitable for use in predicting the rate of chloride ingress into the
concrete although a more detailed analysis of the data could presumably be carried out
60
(this is, however, beyond the scope of the current study). Bentz, Evans and Thomas
have applied a 2-D model to the same data and have shown that the diffusion
coefficients for PC concretes are lower than those obtained from the 1-D model (as used
here). 2 year data have therefore been used in preference to 4 and 10 year data.

2. Based on data collected after 2 years exposure the diffusion coefficient falls as the
strength grade increases.

3. For C35 PC concretes the effects of curing regime and ongoing curing on exposure do
not appear to have a significant effect on the diffusion coefficient. The lack of time
dependence of the diffusion coefficient for PC concretes is consistent with Bamforth’s
observations (1999). However, there may be a small drop in the mean diffusion
coefficient between 1 and 2 years.

7.2.2 PC/pfa concretes


Mean and standard deviations for D and Cs for C35 pfa concretes are given in the following
tables. Data are taken from references 50,55 and 61 unless otherwise stated.

Table 17. Mean and standard deviations of D and Cs for 30% pfa C35 concretes. (PC
content = 242kg/m3, pfa content = 104kg/m3, w/c = 0.56)
-12 2
D/10 m /s Cs /wt % cement
Age Mean σ Mean σ
1 year 3.59 0.87 3.15 0.82
2 years 1.63 0.26 3.57 0.78
4 years 0.85 0.24 4.22 0.62
10 years 0.45 0.12 4.83 1.33

Table 18. Mean and standard deviations of D and Cs for 30% pfa C35 concretes from the
57
limestone filler cement series (w/c = 0.6, cement content = 300kg/m3).
-12 2
Series Age D/10 m /s Cs /wt % cement
FFA 2 years 1.91 3.22
5 years 0.8 3.51
FFB 2 years 2.49 4.63
5 years 1.22 4.08

50
Table 19. Mean D and Cs for 30% pfa C25 and C45 concretes taken from References
31,36 and 37.
-12 2
D/10 m /s Cs /wt % cement
Licensed copy from CIS: qacentral, HBK Contracting Company W.L.L., 25/06/2016, Uncontrolled Copy.

Age C25 C45 C25 C45


2 years 2.54 (σ = 1.0) 1.23 (σ = 0.18) 2.68 (σ =0.48) 3.64 (σ = 0.59)
4 years 1.48±0.24 0.73±0.13 3.08±0.14 3.5±1.0
10 years 0.71±0.34 0.27±0.07 3.72±0.5 5.56±0.63

Table 20. Mean and standard deviations of D and Cs for 15% pfa C35 concretes.
-12 2
D/10 m /s Cs /wt % cement
Age Mean σ Mean σ
1 year 4.51 0.739 3.69 0.79
2 years 2.30 0.551 4.36 0.66
4 years 1.30 0.22 4.08 0.40
10 years 0.88 0.29 4.79 0.16

Table 21. Mean D and Cs for 15% pfa C25 and C45 concretes.
-12 2
D/10 m /s Cs /wt % cement
Age C25 C45 C25 C45
2 years 2.93±0.54 1.76±0.07 3.81±0.26 3.72±0.31
4 years 2.38 1.3 2.87 3.32
10 years No data 0.44 No data 5.7

Table 22. Mean and standard deviations of D and C s for 50% pfa C35 concretes.
-12 2
D/10 m /s Cs /wt % cement
Age Mean σ Mean σ
1 year 3.27 0.81 2.36 0.47
2 years 1.69 0.34 2.40 0.38
4 years 0.97 0.09 2.50 0.31
10 years 0.42 0.16 2.70 0.60

Table 23. Mean D and Cs for 50% pfa C25 and C45 concretes.
-12 2
D/10 m /s Cs /wt % cement
Age C25 C45 C25 C45
2 years 2.12±0.4 1.48±0.1 1.72±0.22 2.43±0.13

61 3
Table 24. D and Cs for 40% pfa concretes . Cement content = 300kg/m
-12 2
Sample no. Age w/c D/10 m /s Cs /wt % cement
35 2 years 0.57 0.72 2.42
5 years 0.57 0.56 2.24
38 2 years 0.58 1.96 3.01
5 years 0.58 1.22 2.25

51
The tables show that both D and Cs are time dependent. The time dependence of D arises
from the ongoing hydration of the cementitious component of the concrete. As pfa hydrates
16
slowly in comparison with PC this time dependence is more marked than in PC concretes .
The time dependence of the surface chloride concentration is harder to explain. If all data
Licensed copy from CIS: qacentral, HBK Contracting Company W.L.L., 25/06/2016, Uncontrolled Copy.

are considered the mean and standard deviation for the surface chloride level for C35 pfa
concretes are 3.81 and 1.04 % respectively.
16
The dependence of D on time can be empirically modelled using a relationship of the form:

D = atn

Where a and n are coefficients derived from regression analyses.


16
For pfa concretes Bamforth gives values of -11.47 for log a and -0.7 for n. Values taken
from BRE samples are summarised in Table 25. Based on 2-10 year data for 30% pfa C45
-12 2 -12 2
concretes a = 2.36 x 10 m /s and n = -0.92. For C25 concretes a = 4.18 x 10 m /s and
n = -0.80. There are insufficient data points for standard deviations to be determined for D
and Cs for 30% pfa C25 and C45 concretes. For PC concretes it is difficult to interpret data
50
from BRE216 as the diffusion coefficient appears to increase at later ages due to sample
geometry. 1 and 2 year data suggest that D may fall with time. A value of n of -0.26 is
obtained for C35 concretes if 10-year data are discounted. The value of a for these
-12 2 16
concretes is 9.09 x 10 m /s. Bamforth uses a value of n of -0.264 for PC concretes.

Table 25. Time dependence of diffusion coefficient for C35 PC and pfa concretes
16
(Parameters refer to the equation D = atn after Bamforth )
-12 2
% pfa a/10 m /s n Equation for Cs
0 9.09 -0.26 No time dependence
15 4.04 -0.71 Cs = 0.0968t + 3.82
30 3.25 -0.92 Cs = 0.095t + 3.829
50 3.22 -0.88 Cs = 0.0376t + 2.33

7.2.3 Modelling the time to the onset of corrosion


A spreadsheet model that predicts the rate of chloride ingress and the time to corrosion
initiation has been developed based on BRE data. The chloride ingress model is very
similar to those developed elsewhere as the same basic equation (i.e. Fick’s law) is
commonly used. The model has been extended to include chloride threshold levels and
corrosion propagation (although there are only very limited data for the latter).

7.2.4 Summary of analysis of chloride ingress data


• Chloride ingress can generally be modelled using Fick’s 2 law of diffusion. This
nd

requires data for diffusion coefficients and surface chloride levels over a wide range of
mix designs, curing conditions, exposure conditions and sample ages. Data from real
structures are also required.
• An analysis of BRE data has allowed diffusion coefficients and surface chloride levels to
be estimated for a range of Portland cement and PC/pfa concretes. Where possible
mean values and standard deviations have been determined (assuming a normal
distribution is followed).

52
• Time dependencies for the diffusion coefficient and the surface chloride level have been
determined. These are particularly important for blended cement concretes due to the
slow hydration of the addition.

nd
A spreadsheet model using Fick’s 2 law has been developed. This allows the rate of
Licensed copy from CIS: qacentral, HBK Contracting Company W.L.L., 25/06/2016, Uncontrolled Copy.

chloride ingress and the time to corrosion initiation to be predicted.

7.3 BRE data on corrosion rates


h
There are only limited data available for assessing the rate of chloride-induced corrosion .
These data have been used to estimate the rate of corrosion in µm/year. However, it is
difficult to determine the nature of any time (and chloride) dependence of the corrosion rate.
64
The relationship between corrosion rate and chloride level has been reported to be
exponential although it is not possible to draw this conclusion from the available BRE data.
If the rate of corrosion is assumed to depend on the chloride level then the weight loss data
can be used to derive empirical relationships between the measured chloride level and the
corrosion rate.

A plot of corrosion rate (derived from the weight loss) against the measured chloride level at
the rebar after 1 year is given Figure 16 for C35 concretes (with and without pfa) taken from
50,55,56
BRE data .

Average corrosion rate against chloride level for C35


concretes after 1 year (from BR216)
25
Corrosion rate (micro

20
m/year)

15

10

0
0 0.5 1 1.5 2 2.5 3 3.5
Chloride level (weight %)

Figure 16 Average corrosion rate against chloride level for C35 concretes after 1 year.

The above graph shows an exponential line of best fit. The amount of corrosion after a
given time for C35 concrete could therefore be predicted using the following (empirical)
i
relationship :

h
The weight loss of reinforcement in concrete prisms, recorded after 1, 2, 4 and 10 years in the tidal
50
zone, was given in BR216 and associated data.
i
There are insufficient data from BRE studies to indicate the exact nature of the relationship
between chloride level and the amount of corrosion. It is possible, based on these data, that the
corrosion rate is independent of chloride level once the propagation stage has commenced and
that the correlation between amount of corrosion and chloride level may reflect the fact that both of
these quantities are time dependent. The rate may therefore depend on the availability of oxygen

53
amount of corrosion = t x 1.1e0.94C

where C is the measured chloride level at time t. The following could be used as an
Licensed copy from CIS: qacentral, HBK Contracting Company W.L.L., 25/06/2016, Uncontrolled Copy.

approximate worst case:

Amount of corrosion = t x 2.1e0.94C

Relationships for the corrosion rate have also been derived assuming a linear relationship
between corrosion rate and the chloride concentration. These are shown in Table 25.
Values derived using C45 concretes are also given.

Table 25. Estimated corrosion rates derived from BRE data

Corrosion rate (µm/year)


C35 C45
0.94C
best fit (exponential) 1.1e 0.9e0.55C
worst case (exponential) 2.1e0.94C e0.8C
Best fit (Linear) 3.16C-0.88 1.21C+0.484
Worst case (Linear) 4.12C-0.663 -

C is the chloride ion concentration at the reinforcement at time t.

There appears to be little or no dependence on the pfa content although the rate of
corrosion is dependent on the grade of the concrete. There was insufficient data to
estimate the corrosion rate for the C25 concrete although for a given chloride level the rate
of corrosion appeared to be greater than in otherwise equivalent C35 concrete. There are
64
no BRE data for corrosion rates under different exposure conditions. Bamforth et al have
j
used the following relationship for the rate of corrosion for concrete in the tidal zone:

CR = 0.46e1.84C

and moisture and the properties of the concrete rather than the chloride level. BRE data also
suggest that the corrosion rate falls with time, despite generally rising chloride levels. These trends
need to be analysed in more depth.
j
Note this rate is the instantaneous value rather than the rate averaged over time as given in the
relationship derived from BRE data. The model Taywood/PII uses the average value over small
time intervals to determine the amount of corrosion within that increment and hence the total
amount of corrosion. This approach is not possible with the BRE weight loss data.

54
8 Conclusions
Models for predicting time to failure for concrete are increasingly being used as tools for
service life design and prediction. The corrosion of reinforcement, induced by carbonation or
Licensed copy from CIS: qacentral, HBK Contracting Company W.L.L., 25/06/2016, Uncontrolled Copy.

by chloride ingress, is perhaps the major durability problem affecting concrete structures in
the UK. These two degradation processes, together with abrasion, are also the most
suitable for service life modelling. Other forms of concrete degradation are best avoided
through choice of materials or through the use of measures such as protective coatings.

Reinforced concrete deterioration is generally considered to follow a 2-stage process, with a


corrosion initiation stage, in which the reinforcement is protected by the cover concrete, and
a propagation stage, in which corrosion proceeds. Chloride ingress in saturated concrete is
nd
generally modelled using Fick’s 2 law of diffusion, although other processes such as wick
action and convection are important where the concrete is periodically dry. The inputs for
these models can either be estimated from measurements on actual structures or predicted
using empirical relationships derived from laboratory and field concretes. Carbonation leads
to a reduction in the pH of the concrete pore solution. Carbonation depth generally shows a
square root relationship with time. The rate depends on the physical properties of the
concrete and its pH buffering capacity, which is in turn dependent on the cement content.
Corrosion is initiated when the carbonation front reaches the reinforcement.

There are a considerable amount of data, primarily from laboratory samples and field trials,
which can be used in calibrating and validating models. However, the large number of
variables involved means that the data are often very limited in important areas. The
reliability of models based on these data is consequently limited. There is therefore a need
for more data (from laboratory work and data from real structures) in determining:
• The effects of different cement compositions and curing regimes on the transport
properties of the concrete.
• The effects of different exposure conditions (including the effects of changes in relative
humidity, short-term application of deicing salts rather than constant marine exposure,
the combined effects of deterioration processes).
• The rate of the corrosion process and its dependence on measurable parameters such
as environmental conditions and properties of the concrete.
• The effects of different protection and repair strategies.
• The effects of ageing on the transport properties of the concrete.
• The application of models derived primarily from laboratory samples to real structures.

Available BRE data has been used in assessing approaches to modelling carbonation and
chloride ingress. A model that relates rate of carbonation to the oxygen permeability
measured at 28 days has been developed using BRE data and gives reasonable results
when tested against other data. However, its applicability over a wider range of mixes and
conditions cannot be tested. More data are required linking carbonation rates and oxygen
permeability to mix design, curing conditions and duration, and to the exposure conditions.

An analysis of BRE data has allowed apparent chloride diffusion coefficients and surface
chloride levels to be estimated for a range of concretes. Where possible mean values and
standard deviations have been determined (assuming a normal distribution is followed).
Time dependencies for the diffusion coefficient and the surface chloride level have also
been determined. These are particularly important for blended cement concretes due to the
slow hydration of the addition. However, the data are again spread thinly over a wide range
of variables. More data are required linking diffusion coefficients and surface chloride levels
to mix design, curing conditions and duration, and to the exposure conditions. A statistical
approach needs to be taken to the analysis of the available data.

55
There are only very limited data available for use in modelling the corrosion propagation
stage and further data are required if reliable predictions are to be made. BRE data are
limited to tidal exposure but are consistent with data from other sources.
Licensed copy from CIS: qacentral, HBK Contracting Company W.L.L., 25/06/2016, Uncontrolled Copy.

56
9 Further reading
There are a number of initiatives, reports and other publications that address service life and
whole life costing issues. These are summarised below. Other references are included at
Licensed copy from CIS: qacentral, HBK Contracting Company W.L.L., 25/06/2016, Uncontrolled Copy.

the end of this report.


62
BRE Digest 444 on the Corrosion of Steel in Concrete (Parts 1 to 3; Part 4 in
preparation) discusses the durability of steel in concrete, investigation and assessment
in existing structures, and protection and remedial work.
• BRE Digest 455, corrosion of steel in concrete – service life design and prediction
(2001).
• BRE Report BR367: Study on whole life costing

63
BRE Report BR316: A review of service life design of concrete structures’ .

64
Taywood PIT project. This work has provided a detailed review of field and also
provided models for chloride and carbonation-induced corrosion. The main report from
this project is due to be published by the Concrete Society in 2000.

65
Duracrete . This EU BRITE-EURAM project has investigated probabilistic service life
prediction including case studies.

66
Concrete Society report : Developments in durability design & performance-based
specification of concrete. Concrete Society Special Publication CS109.

67
Hetek . This programme of research projects, which addressed resistance of concrete
to chloride ingress and freeze-thaw, was funded by the Danish Road Directorate.
• Further aspects of the concepts of service life planning are given in the draft
68
International Standard on service life design .
• Highways Agency Document BD57 design for Durability

69
ENV1991-1 provides important general performance requirements for structures and
categories of design life.

6
RILEM 14: Durability design of concrete structures
• Architectural Institute of Japan, 'Principal guide for service life planning of buildings',
English Edition, 1993.
• Service life prediction – state of the art report, ACI 365.1R-00 (2000).

57
10 References
1 BRE Digest 455, Corrosion of steel in concrete –service life design and prediction
Licensed copy from CIS: qacentral, HBK Contracting Company W.L.L., 25/06/2016, Uncontrolled Copy.

(2001).

2 LW Masters and E Brandt, Systematic methodology for service life prediction of building
materials and components. Report of the TC 71-PSL, prediction of service life of
building materials and components, Materials and Structures, 22, 385 (1989).

3 pr EN1990 Eurocode: Basis of structural design.

4 ‘Developments in durability design & performance-based specification of concrete’,


Concrete Society Special Publication CS109 (1996).

5 RJ Currie and PC Robery, ‘Repair and Maintenance of reinforced concrete, BR254,


1994.

6 Durability design of concrete structures, RILEM Report 14, Ed. A Sarja & E Vesikari,
E&FN Spon, London (1996).

7 C Edvardsen and L Mohr, ‘Duracrete – a guideline for durability-based design of


concrete structures’, Procs FIB Symposium ‘Structural concrete – the bridge between
people, Prague, October 1999.

8 AJM Siemes and S Rostam, ‘Durability, safety and serviceability – a performance based
design’, IABSE colloquium: basis of design and actions on structures, Delft, March 27-
29 1996, TNO report No. 96-BT-R0437-001.

9 R Breitenbucher, C Gehlen, P Schiessl, J Van Den Hoonaard & T Siemes, ‘Service life
design for the Western Scheldt tunnel, Durability of building materials and components
8 (volume 1), Ed. MA Lacasse & DJ Vanier, NRC Research Press, Ottowa, Canada
(1999).

10 P Bamforth, Probabilistic performance-based durability design of concrete structures, in


Management of concrete structures for long-term serviceability’, p33, Thomas Telford,
London (1997).

11 K. Tuutti, ‘Corrosion of steel in concrete’. CBI – Forskning: Report Fo 4.82, Swedish


Cement and Concrete Association, Stockholm, 1982, pp 469.

12 Corrosion of steel in concrete, RILEM Report of the Technical Committee 60-CSC, Ed.
P Schiessl, Chapman & Hall, London (1988).

13 AV Saetta, RV Scotta & RV Vitaliani, ‘Analysis of chloride diffusion into partially


saturated concrete’, ACI Materials Journal, 90, 441-451 (1993).

14 L-O Nilsson, ‘On the uncertainty of service life models for reinforced marine concrete
structures’, in Life prediction and ageing management of concrete structures, Proc Int.
RILEM Workshop, Ed. D Naus, (2000)

15 Service life prediction – state of the art report, ACI 365.1R-00, American Concrete
Institute (2000).

58
16 P Bamforth, ‘The derivation of input data for modelling chloride ingress from eight-year
UK coastal exposure trials’, Mag. Conc. Res, 51, 87 (1999).
Licensed copy from CIS: qacentral, HBK Contracting Company W.L.L., 25/06/2016, Uncontrolled Copy.

17 L Meljbro, ‘The complete solution of Fick’s second law of diffusion with time-dependent
diffusion coefficient and surface concentration’, Durability of concrete in saline
environments, published by Almquist & Wiskell Tryckeri (Uppsala) for Cementa AB p
127 (1996).

18 CL Page, NR Short &A El Tarras, ‘Diffusion of chloride ions in hardened cement paste’,
Cem. Conc. Res., 11, 395 – 406 (1981).

19 A Neville, ‘Chloride attack of reinforced concrete: an overview’, Materials and


Structures, 28, 63 (1995)

20 G Sergi, SW Yu and CL Page, ‘Diffusion of chloride and hydroxyl ions in cementitious


materials exposed to a saline environment’, Mag. Conc. Res., 44, 63 (1992).

21 AM Butler, ‘Capillary absorption by concrete’, Concrete, 31 (4), 23 (1997).

22 PR Vassie & RJ Walker, Corrosion of reinforcement in concrete caused by wetting and


drying cycles in chloride-containing environments. Appendix 6, The natural wetting and
drying of the surface zone of a concrete block: monitoring in situ (2000).

23 M Alisa, Corrosion of reinforcement in concrete caused by wetting and drying cycles in


chloride-containing environments. Appendix 6, Modelling ingress of chloride in concrete
(2000).

24 RD Browne, ‘Design prediction of the life of reinforced concrete in marine and other
chloride environments’, Durability of Building Materials, Elsevier Science Publishing Co.
Vol. 3, 1982.

25 P Vassie, ‘Reinforcement corrosion and the durability of concrete bridges’, Proc. Instn
Civ. Engrs, Part 1, 76, 713 (1984).

26 K Tuutti (Repair philosophy for concrete structures, pp 159-169 in ‘Concrete durability


and repair technology’, Proc. Int. Conf. ‘Creating with Concrete’, Dundee, September
1999, Ed. RK Dhir & MJ McCarthy (1999).

27 L.-O. Nilsson, P. Sandberg, E. Poulsen, Tang L., A. Andersen & J.M. Frederiksen from
HETEK, A system for estimation of chloride ingress into concrete, Theoretical
background,

28 DA Haussmann, ‘Steel corrosion in concrete – how does it occur?’, Material Protection,


Vol. 6, Nov. pp 19-23 (1967).

29 L Tang and L-O Nilsson, ‘Service life prediction for concrete structures under seawater
by a numerical approach’, in ‘Durability of Building Materials and Components 7’, Proc.
7th International Conf. on Durability of building materials and components, Stockholm,
pp 97 – 106 (1996).

59
30 K Petersson, ‘Chloride threshold value and corrosion rate in reinforced concrete’,
Concrete 2000, Ed. RK Dhir & MR Jones, E&F Spon, London, Volume 2, pp461-471
(1993).
Licensed copy from CIS: qacentral, HBK Contracting Company W.L.L., 25/06/2016, Uncontrolled Copy.

31 RK Dhir, MR Jones &MJ McCarthy, ‘Pfa concrete: chloride-induced reinforcement


corrosion’, Mag. Conc. Res, 46, 269-277 (1995).

32 RK Dhir, MR Jones &MJ McCarthy, ‘Pfa concrete: chloride-induced reinforcement


corrosion’, Mag. Conc. Res, 46, 269-277 (1995).

33 R Currie & P Robery, Repair and maintenance of reinforced concrete, BRE 254.

34 L Parrott, ‘Carbonation and corrosion’, Presented at SCI meeting: ‘Preventing chloride


and carbonation induced corrosion of reinforced concrete’, (1998).

35 RAM Watkins & AP Pitt Jones, Carbonation: a durability model related to site data,
Proc. Instn Civ. Engrs Structs & Bldgs, 99, 155 (1993)

36 C Andrade, C Alonso & FJ Molina, ‘Cover cracking as a function of bar corrosion: part 1
– experimental test’, ‘Materials & Structures’, 26, 453-464 (1993).

37 L Parrott, ‘Carbonation and corrosion’, Presented at SCI meeting: ‘Preventing chloride


and carbonation induced corrosion of reinforced concrete’, (1998).

38 C Andrade and MC Alonso, ‘Values of corrosion rate of steel in concrete to predict


service life of concrete structures’, in ‘Application of accelerated corrosion tests to
service life prediction of materials’, ASTM STP 1194 (Ed. G Cragnolino & N Sridhar),
American Society for Testing and Materials, Philadelphia, pp 282-295 (1994).

39 JA Gonzalez, C Andrade, C Alonso & S Feliu, ‘Comparison of rates of general corrosion


and maximum pitting penetration of concrete embedded steel reinforcement’, Cem.
Conc. Res., 25, p257-264 (1995).

40 BRITE project 4062 ‘The residual life of reinforced concrete structures’.

41 JP Broomfield, ‘Corrosion of steel in concrete, understanding investigation and repair’, p


70, E&FN Spon, London (1997).

42 C Edvardsen & L Mohr, ‘Duracrete – a guideline for durability-based design of concrete


structures’, Procs. FIB Symposium ‘Structural Concrete – the Bridge Between People’,
Prague, Oct 1999.

43 A.M. Neville, ‘Properties of Concrete’, 4th Edition, Longman (1995). Page 308.

44 CL Page, NR Short &A El Tarras, ‘Diffusion of chloride ions in hardened cement paste’,
Cem. Conc. Res., 11, 395 – 406 (1981).

45 A Neville, ‘Chloride attack of reinforced concrete: an overview’, Materials and


Structures, 28, 63 (1995).

60
46 SE Benjamin and JM Sykes, ‘Chloride induced pitting corrosion of Swedish iron in
ordinary Portland cement mortars and alkaline solutions: the effect of temperature. 3rd
Int Symp. on Corrosion of Reinforcement in Concrete Construction, Ed. Page,
Licensed copy from CIS: qacentral, HBK Contracting Company W.L.L., 25/06/2016, Uncontrolled Copy.

K Treadaway & P Bamforth, (1990).

47 A Atkinson & J.A. Hearne, ‘Mechanistic model for the durability of concrete barriers
exposed to sulfate-bearing groundwaters’, Materials Research Society Proceedings,
176, 149 (1990).

48 Modelling of degradation, Duracrete Document BE95-1347/R4-5 (1998).

49 RB Polder & A Hug, ‘Penetration of chloride from deicing salt into concrete from a 30
year old bridge’, Heron, 45, 109 (2000).

50 MDA Thomas & JD Matthews, ‘Durability of pfa concrete’, BR216 (1994).

51 H-J Wierig, ‘Long time studies on the carbonation of concrete under normal outdoor
exposure’, Proc. RILEM seminar, Hannover (1984).

52 AP Barker & JD Matthews, ‘Concrete durability specification by water/cement ratio or


compressive strength for European cement types – a 2-year report’, BRE CR216/94
(1994).

53 Unpublished BRE data

54 Roy, Beng & Northwood, The carbonation of concrete structures in the tropical
environment of Singapore and a comparison with published data for temperate
climates, Mag. Conc. Res., 48, 293 (1996)

55 MDA Thomas & JD Matthews, ‘Chloride penetration and reinforcement corrosion in fly
ash concrete exposed to a marine environment’, Proc. 3rd CANMET/ACI Int Conf. on
Performance of Concrete in a Marine Environment, ACI SP-163 pp317 – 338 (1996).

56 JD Matthews, Personal Communication

57 JD Matthews, ‘Performance of limestone filled cements – update of 2 year durability


data’, BRE Internal note: N76/91 (1991).

58 JD Matthews, personal communication.

59 Minimum requirements for durable concrete, Ed. DW Hobbs, British Cement


Association, Crowthorne (1998).

60 EC Bentz, CM Evans and MDA Thomas ‘Chloride diffusion modelling for marine
exposed concretes’, in Corrosion of reinforcement in Concrete Construction, Ed. CL
Page, PB Bamforth & JW Figg, Royal Society of Chemistry, Cambridge, pp136-145
(1996).

61 JD Matthews, ‘Performance of pfa concrete in aggressive conditions: 3 marine


conditions’, BRE Laboratory Report, BR295 (1995).

62 Corrosion of steel in concrete (in 3 parts), BRE Digest 444 (1999).

61
63 B Marsh, ‘A review of service life design of concrete structures’, BR316 (1996).
Licensed copy from CIS: qacentral, HBK Contracting Company W.L.L., 25/06/2016, Uncontrolled Copy.

64 P Bamforth et al, ‘Guidance on the selection of measures for enhancing reinforced


concrete durability’, Taywood Engineering Ltd Report No 1304/98/10349, Taywood
Engineering Ltd , Southall, Middlesex, UK, March 1999 (in press).

65 C Edvardsen & L Mohr, ‘Duracrete – a guideline for durability-based design of concrete


structures’, Procs. FIB Symposium ‘Structural Concrete – the Bridge Between People’,
Prague, Oct 1999.

66 Developments in durability design & performance-based specification of concrete,


Concrete Society Special Publication CS109 (1996).

67 L.-O. Nilsson, P. Sandberg, E. Poulsen, Tang L., A. Andersen & J.M. Frederiksen from
HETEK, A system for estimation of chloride ingress into concrete, Theoretical
background, Danish Road Directorate, Report No. 54 (1997).

68 BS ISO 15686-1: 2000, Buildings and constructed assets. Service life planning. General
principles.

69 ENV 1991-1:1994, 'Eurocode 1 : Basis of design and actions on structures - Part 1:


Basis of design', CEN (European Committee for Standardisation), Brussels, 1994.

62
Annexe A. Other models for carbonation
Models of the initiation stage for carbonation-induced corrosion have been discussed
in the main text. Other models are summarised in this Annexe.
Licensed copy from CIS: qacentral, HBK Contracting Company W.L.L., 25/06/2016, Uncontrolled Copy.

A1 CEB TG V/1+2 “Service life design of concrete structures”.


The following model has been derived from Fick’s 1st law of diffusion in CEB TG
V/1+2 “Service life design of concrete structures”.

n
2k1 k 2 Deff C s t 
d= t o 
a  t 

where

a is the binding capacity for CO2 given by:

M CO2
a = 0.75.C.c.DH .
M CaO

Deff is the effective diffusion coefficient at defined compaction, curing and


environmental conditions.
Cs is the concentration of the passivator at the concrete surface
t is the time in service
t0 is a reference period
k1 and k2 are constant parameters for the influence of execution and environment on
Deff.
n is a constant parameter for the influence of meso climatic conditions
c is the cement content
C is the CaO content of the cement
DH is the degree of hydration
MCO2 and MCaO are respective molar masses of CO2 and CaO.

This model simplifies to the following relationship if DH is considered to be constant

d = Kt (1 / 2 − n )

where K is a constant.
A2 Schiessl model

This model gives the following equation for time, t, to reach carbonation depth, d.
Licensed copy from CIS: qacentral, HBK Contracting Company W.L.L., 25/06/2016, Uncontrolled Copy.

a   d 
t = − . d + d ∞ . ln1 −  

b   d∞ 

where
a is the binding capacity of CO2,
b is a retardation factor (which depends on environmental factors, the concrete mix
and the dimensions of the structural element) given by:

b = f . D CO 2 , surface .C s + b

The value of b ranges from 0.8 – 1.5 x 10-9 kgm-2s-1 for concrete not affected by
rain and from 0.8–2.8 x 10-9 kgm-2s-1 for unprotected concretes in the open. It
increases with the number and intensity of wetting and drying cycles, with
increasing duration of wet periods and with increasing relative humidity. It also
increases with decreasing w/c ratio, with increasing cement content, duration of
moist curing and with the thickness of the structural element.

b is the amount of CO2 needed for the carbonation of counter-diffused substances


Cs is the surface concentration of CO2.
f is a factor to consider the decrease in diffusion coefficient due to an increase in
the concrete density caused by a higher moisture content and degree of
hydration in the inner section of the structure.

d∞ is the ultimate carbonation depth given by:

DCO2 , surface .C s
d∞ =
b

DCO2 is the CO2 diffusion coefficient of the concrete at depth d given by:

DCO2 (d ) = DCO2 , surface .(1 − fd )

The shape of the carbonation depth against time curve for this model approximates
to a d = ktx relationship, with x approaching 0.5 as d∞ increases. However, it does
not give values of x above 0.5 and so would not model all the available data. It does,
however, allow factors such as the changes in diffusion coefficient with time and the
counter-diffusion of Ca(OH)2 to be taken into account.
It may be difficult to apply this model in practice but it does help to identify factors
affecting carbonation rate and the final depth of carbonation.
Licensed copy from CIS: qacentral, HBK Contracting Company W.L.L., 25/06/2016, Uncontrolled Copy.

A3 Parrott

Parrott has derived the following equation for the depth of carbonation with time:

a.k 0.4 .t in
d=
c 0.5

where

k is the air permeability of cover concrete (this can be estimated from the relative
humidity in the cover concrete and the permeability of a specimen dried at 60% RH).
c is the CaO content in the hydrated cement matrix, n is an exponent determined
from known data and dependent on the relative humidity. a is a coefficient and can
be assigned the value of 64 based on published data. The dependence of n on the
relative humidity, RH, taken from data given by Parrott, is given in Figure A1.

Figure A1 n against relative humidity, RH, for concrete


taken from Parrott

0.6
0.5
0.4
0.3
n

0.2
0.1
0
0 20 40 60 80 100 120
RH/%

Parrott1 has given a separate set of data for n against r%. These data are derived
from experiments on cement pastes in which the weight gain on carbonation, and the
weight loss during thermogravimetric analysis were used to determine the extent of
carbonation with time after exposure at different relative humidities. Carbonation
against time curves were fitted and the value of n estimated. The values are shown
in Figure A2. The average value of n in these experiments was 0.5. Parrott says
that changes in the value of n may arise from changes in the rate controlling process
from gaseous diffusion through open pores to a relatively slow process in which lime
dissolves and diffuses through the liquid phase to the reaction front. The formation of
local low permeability layers may reduce the value of n at low humidities and low
reaction rates arising from the presence of thin films of sorbed water.
Figure A2. N against relative humidity, RH, for cement
paste, Taken from Parrott
Licensed copy from CIS: qacentral, HBK Contracting Company W.L.L., 25/06/2016, Uncontrolled Copy.

0.8
0.7
0.6
0.5
0.4
n

0.3
0.2
0.1
0
0 20 40 60 80 100 120
RH/%

Parrott also suggests that the relative humidity at which the maximum rate of
carbonation is observed may be greater the higher the concrete porosity. More
porous mixes are likely to carbonate rapidly at higher relative humidities. He has
shown that the degree of carbonation is mainly controlled by the empty porosity.
Parrott has also questioned the accuracy of phenolphthalein testing of carbonation,
particularly at low relative humidities. He says that, in these cases, carbonation does
not proceed to completion and that residual Ca(OH)2 can give misleading
phenolphthalein test results.

Parrott2 has pointed out that other measurable parameters can be used to
characterise the transport properties of cover concrete. These include 28-day
strength, 6-month carbonation depth, w/c ratio etc. He gives the following table
based on experimental data for the scatter of carbonation depth data plotted against
parameters that can be measured at early ages.

Options for control and assessment of carbonation (taken from Parrott, 1998)

Parameter Scatter of carbonation depths


(mm)
w/c ratio ±6
28 day strength ±5
Strength 8 days after cure ±3
Initial 4-day weight loss ± 4.5
Initial 4-day weight loss allowing for cement ±2
Carbonation depth at 6 months ± 1.5
Air permeability of cover concrete at 18 months ± 3.5

Parrott’s model is empirical in that the value of n is taken from experimental data.
The other parameters can be estimated or measured.
Parrott3 has shown that the carbonating zone can be quite wide (although evidence
on this is conflicting). In the example that he gives reductions in the level of Ca(OH)2
and increased levels of CaCO3 are observed in a zone 30-40 mm thick. A pH profile
taken from a separate source shows that the pH changes from about 8 in the
Licensed copy from CIS: qacentral, HBK Contracting Company W.L.L., 25/06/2016, Uncontrolled Copy.

carbonated zone to over 11 in the uncarbonated zone with the variation occurring
within 6 mm. However, the pH would be expected to be nearer 12.5 in uncarbonated
concrete and the zone over which the pH changes may therefore be greater than 6
mm.

A4 Taywood model

This model assumes that the chemical buffering capacity of the cement is the main
parameter affecting the carbonation rate. The C3A content of the cement, the period
of wet curing and the mean relative humidity of the environment are the other
important parameters. Empirical relationships between these parameters and the
carbonation rate have been determined using long-term test data. These data have
been used to determine k in the equation:

D = a +kt1/2

The dependence of k on the C3A content, the period of wet curing and the relative
humidity has been determined and used to derive a normalised curve for buffering
capacity against carbonation rate. The equivalent buffering capacity can be
determined from:

b = b1c1 + b2c2

where b = buffering capacity


b1 = buffering capacity of cement 1
b2 = buffering capacity of cement 2
c1 = mass of cement 1
c2 = mass of cement 2

The resulting curve of predicted k against measured k shows a reasonable


agreement, with the predicted rate of carbonation generally being slightly higher than
that measured. The approach is therefore considered to be adequate for design
purposes.

A5. Bob4

In this model the depth of carbonation, x, can be predicted using the equation:

150ckd
x= t
fc

where c is parameter representing the binding capacity (equal to 1 for a 42.5N PC


concrete).

k is a parameter for environmental conditions (mainly RH)


d is the surface concentration of CO 2
fc is the compressive strength at time t.
A6. Papadakis, Fardis & Vayenas5
Licensed copy from CIS: qacentral, HBK Contracting Company W.L.L., 25/06/2016, Uncontrolled Copy.

1/ 2
 2[CO2 ]0 De ,CO 2 
xc =  t
 [CH ] + 3[CSH ] 
 

where xc is the depth of carbonation at time t, De,CO2 is the effective diffusivity of CO2
in the carbonated concrete or mortar and [CH], [CSH] are molar concentrations. This
model is considered to be valid for constant diffusivity - ie if the concrete is in a state
of equilibrium with the environment (both in space and time). It can be applied for
indoor exposure and external sheltered exposure if an average value of D is utilised.
The model also assumes that the hydration reactions are virtually complete and that
the diffusion coefficient is not changing. The application of the model for predictive
purposes requires knowledge of the compositional parameters of the concrete (which
control the effective diffusivity) and the relative humidity (which determines the
degree of pore saturation - assumed to be constant).

Measured values of De,CO2 have been fitted to the empirical equation:

De,CO2 (m2s-1) = (1.64 x 10-6)εp1.8[1-(RH/100)]2.2

Where εp is the porosity of hardened cement paste calculated from the total porosity
of the concrete or mortar.

A7. Saetta, Schrefler & Vitaliani

The authors point out that the process of carbonation itself will change the
permeability of the concrete and will therefore, over time, change the form of the
equations modelling the process. The rate of carbonation is given by:

∂c
= α 1 . f 1 (h). f 2 ( g ). f 3 (c). f 4 (T )
∂t

where:

α1 is a material parameter and depends on the characteristics of the concrete, f1(h) is


a parameter taking the relative humidity of the concrete into account. For RH<0.5 it
is set to 0. For RH’s between 0.9 and 1 it is set to 1. In the RH range 0.5 – 0.9 it is
given by:

f 1 (h) = 52 (h − 0.5)

f2(g) is a measure of the extent to which carbon dioxide has penetrated the concrete.
 g 
f 2 ( g ) =  
 g max 
g is 0 where CO2 penetration has not occurred and g = gmax = 1 where the CO2
concentration is at its maximum.

f3(c) describes the influence of the degree of carbonation on the carbonation rate:
Licensed copy from CIS: qacentral, HBK Contracting Company W.L.L., 25/06/2016, Uncontrolled Copy.

m
 c 
f 3 (c) = 1 −  
 c max 

where cmax is the maximum amount of calcium carbonate that can form and is usually
set to 1.

f4(T) gives an Arrhenius-type dependence on temperature.

α1 is a material parameter and depends on the characteristics of the concrete.

A8. Xu and Rodhe

Xu and Rodhe6 give the following analysis for the diffusion of CO2 into the concrete
and reacting with the pore walls:

∂C ∂ 2C ∂A
ε = εD − (1 − ε )
∂t ∂x 2
∂t

where ε is the porosity, C is the concentration of CO2, A is the ‘reacting part’, which is
considered, as a first approximation, to be equal to the dissolution of CO2 in pore
solution and obeys Henry’s law such that A =RC where R depends on the Henry’s
law constant.

∂C D ∂ 2C
=
∂t (1 − ε ) R ∂x 2
1+
ε

The solution of this equation is

x
C ( x, t ) = C 0 erfc( )
2 Deff t

where

D
Deff =
1− ε
1+ R
ε

The depth of carbonation d can be shown to be:


2C 0 Deff
d= t
1− ε π
C0 + Q
ε
Licensed copy from CIS: qacentral, HBK Contracting Company W.L.L., 25/06/2016, Uncontrolled Copy.

where Q is the capacity of the concrete in binding CO2. This suggests that d should
be proportional to the square root of time.

The following empirical relationships between carbonation depth and strength grade
have been suggested:

 1 
d ∝ . t
 f 
 w 

where fw = strength after water curing7.

d ∝ (β − f 28 .) t

where β is a constant and f28 is the compressive strength at 28 days8.

References

1 L Parrott, ‘Carbonation, moisture and empty pores’, Adv. Cem. Res., 4, 111
(1991/92)
2 L Parrott, SCI conference, Nov ’98
3 L Parrott, C&CA review
4 C Bob, Durability of concrete structures and specification, pp 311-318 in
‘Concrete durability and repair technology’, Proc. Int. Conf. ‘Creating with
Concrete’, Dundee, September 1999, Ed. RK Dhir & MJ McCarthy (1999)
5 Papadakis, Fardis & Vayenas Effect of composition, environmental factors and
cement-lime mortar coating on concrete carbonation, Materials and structures, 25,
293 (1995)
6 Xu and Rodhe Influence of alkali on carbonation of concrete, Durability of
building materials and components 7 (vol 1), Ed. C. Sjostrom, Spon (1996)
7 HG Smolczyk, Discussions to M Hamada’s paper: neutralisation (carbonation) of
concrete and corrosion of reinforcing steel, 5th ISCC, Tokyo, Vol. III, pp369-383
(1968).
8 Kokubu M & Nagataki S, Carbonation of concrete with fly ash and corrosion of
reinforcements in 20-year tests, ACI SP-114, pp315-329 (1989).
Annexe B. Additional analyses of BRE carbonation and
permeability data
Licensed copy from CIS: qacentral, HBK Contracting Company W.L.L., 25/06/2016, Uncontrolled Copy.

B1. Dependence of oxygen permeability on mix and curing parameters

There are only limited data available for 28-day oxygen permeability measurements. The
range of values is illustrated in the following graph (for PC and PC/pfa concretes).

Oxygen permeability at 28 days for BRE PC & PC/pfa data


35
30
25
Frequency

20
15
10
5
0

e
12
16
20
24
28
32
36
40
44
48
52
56
60
64
68
72
76
0
4
8

or
M
-17 2
Oxygen permeability/10 m

Figure B1. Oxygen permeability at 28 days for PC and PC/pfa concrete

The permeability falls with the duration of damp cure and the amount of pfa present as
shown in the following graph:

2 8 -d a y o x yg e n p e rm e a b ility a g a in s t d u ra tio n o f d a m p
c u re fo r P C a n d P C /p fa c o n c re te s
(ta k e n fro m B R 2 1 6 , S e rie s A )
50
2
Permeability/10 m

40
-17

0 % p fa
30 1 5 % p fa
3 0 % p fa
20
5 0 % p fa
10

0
0 10 20 30
D u ra tio n o f d a m p c u re /d a ys

Figure B2. Oxygen permeability against duration of damp cure


Figure B2 shows that the duration of damp cure has a substantial effect on the 28-day
oxygen permeability. The relative humidity of the environment to which the concrete is
exposed following damp cure will also have a significant effect. The pfa content has a
smaller, but still significant effect. There is a degree of variability between measurements
Licensed copy from CIS: qacentral, HBK Contracting Company W.L.L., 25/06/2016, Uncontrolled Copy.

made on identical samples as illustrated by the following graph of samples with similar mix
o
designs cured under water at 20 C for 28 days as shown in Figure B3.

Oxygen permeability against pfa content after 28 day


water cure for identical mix designs at 20oC (BR216 Series
6 A - D)

5
2
m
-17

A
Oxygen permeability/10

4
B
3
C
2 D
1

0
0 10 20 30 40 50 60
pfa content (%)

Figure B3. Permeability against pfa content for otherwise identical concrete showing
sample variability

The significant influence of the curing conditions on permeability, and consequently on the
rate of carbonation, makes it difficult to use mix design parameters to predict the rate of
carbonation (and is the main reason for using a measurable parameter such as the oxygen
permeability in making the prediction). In the absence of detailed information about the
curing time and conditions the mix design can therefore only be used to give a preliminary
indication of the likely rate of carbonation in a given environment. Table B1 gives minimum
curing periods from BS8110: part 1: 1997.
Table B1 Minimum curing and protection periods from BS8110: part 1: 1997

Cement type Ambient conditions Minimum curing period


after casting
Licensed copy from CIS: qacentral, HBK Contracting Company W.L.L., 25/06/2016, Uncontrolled Copy.

Average surface temperature of concrete


o o o o
5 C - 10 C T C (between 10 C
o
and 25 C
PC 42.5 or PC 52.5 Average 4 days 60/(t+10) days
to BS12, SRPC 42.5
to BS4027
Poor 6 days 80/(t+10) days
Other cements in Average 6 days 80/(t+10) days
BS5328: Pt 1: 1997
Poor 10 140/(t+10) days
All Good No special requirements

Where:
Good = damp and protected (RH>80%, protected from sun and wind)
Average = intermediate between good and poor
Poor = dry or unprotected (RH<50%, not protected from sun and wind)

The curing conditions used in BRE work could therefore be categorised as follows:

Under water for 28 days good


Damp for 1,3 or 7 days then in air at 90% RH good
Damp for 1,3 or 7 days then in air at 80% RH average
Damp for 1,3 or 7 days then in air at 65% RH average
Damp for 1,3 or 7 days then in air at 40% RH poor
o
At 20 C 2 days damp curing would be required at relative humidities of 65% and 80% if the
minimum curing period were to be achieved for PC concrete and 3 days for pfa concrete. In
these cases concretes damp cured for 1 day would not therefore have been subjected to the
minimum curing period. All concretes stored at 90% RH prior to exposure can be
considered to have received the minimum acceptable cure. Of the concretes stored at 40%
RH prior to exposure only those that have been damp cured for 3 days (for PC concretes)
o
and 7 days will have received an acceptable minimum cure. At 5 C only concretes cured for
7 days will have received an acceptable minimum cure. The following graph shows oxygen
permeability against % pfa for C35 concretes receiving an ‘acceptable’ minimum cure at
o o
20 C or 5 C taken from BR216.
O x y g e n p e r m e a b ilit y a g a in s t % p f a f o r
c o n c r e t e s r e c e iv in g ' a c c e p t a b le ' m in im u m
c u re
Licensed copy from CIS: qacentral, HBK Contracting Company W.L.L., 25/06/2016, Uncontrolled Copy.

35
2
Oxygen poermeability/10 m
-17

30

25

20

15
10

5
0
0 20 40 60
% p fa

Figure B4. Oxygen permeability against pfa content for concretes receiving acceptable
minimum cure

The graph shows that the average permeability falls as the pfa content is increased.
However, there is still a very wide range of measured permeabilities for otherwise identical
concretes cured under different conditions. It is possible, based on available laboratory
data, to determine the mean and standard deviations of the range of values and to predict a
‘worst case’ (i.e. maximum likely) value for a concrete of a given mix design cured under
minimum acceptable conditions.

B2. Relationships between the rate of carbonation and measurable parameters

The following graph shows the dependence of the normalised 20-year carbonation depth
(corrected for Portland cement content) on the pfa content for concretes that have received
an acceptable minimum cure. The graph suggests that the carbonation rate (when
corrected for cement content) is virtually independent of the pfa content and supports the
use of a model in which the PC content only is used in predicting the rate of carbonation as
discussed in the main text.
2 0 y e a r c a r b o n a t io n d e p t h ( c o r r e c t e d f o r
c e m e n t c o n t e n t ) a g a in s t % p f a f o r
c o n c r e t e s r e c e iv in g 'a c c e p t a b le ' m in im u m
Licensed copy from CIS: qacentral, HBK Contracting Company W.L.L., 25/06/2016, Uncontrolled Copy.

c u re

4 0 y = - 0 .0 7 7 1 x + 2 5 .0 8 4
3 5
3 0
Carbonation depth/mm

2 5
2 0
1 5
1 0
5
0
0 2 0 4 0 6 0
% p fa

Figure B5. Normalised 20-year carbonation depth (corrected for PC content) against pfa
content

The following graph shows the variation in the carbonation rate (relative to that for the same
concrete following a 7-day cure) for PC and PC/pfa concretes taken from BR216.

Effect of curing time on carbonation rate (graph shows


relative carbonation rates for 1 and 3 days relative to that
for 7 days for all BR216 data
Relative carbonation

3
2.5
2
rate

1.5
1
0.5
0
0 2 4 6 8

Damp cure time/days

Figure B6. Effect of damp curing time on carbonation rate

The wide scatter for 1 and 3 day data shows that there is not a simple relationship between
the duration of the damp cure and the relative rate of carbonation. The relative humidity of
the environment immediately following the damp cure has a considerable influence on the
continued hydration of the concrete and consequently on the rate of hydration.
The relationship between permeability (plotted on a log scale) for concrete and its w/c ratio
i
has been reported to be linear provided the wet concrete was workable. The following
graph shows 28-day oxygen permeability against w/c ratio for PC and PC/pfa concretes that
were damp cured for 3 days and stored to 28 days at different relative humidities.. The
Licensed copy from CIS: qacentral, HBK Contracting Company W.L.L., 25/06/2016, Uncontrolled Copy.

graph is consistent with there being a linear relationship. similar results are found for
concretes cured in water for 28 days.

40
m2

35
y = 74.609x - 21.365
-17

30
O2 permeability/10

2
R = 0.6773
25
20
15
10
5
0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8
w/c ratio

Figure B7. O2 permeability against w/c ratio for PC and pfa concretes damp cured for 3
days.

The dependence of indoor carbonation rate (normalised 20-year carbonation depth,


corrected for cement content) on the w/c ratio is illustrated in the following graph. Only data
for concretes cured for 28 days in water prior to exposure are included to eliminate any
variability arising from differences in curing conditions. Data for all cement types are
included. There are insufficient data points for a meaningful analysis based on PC
concretes only to be carried out.
N o rm alised 20 year carb o n atio n d ep th (co rrected fo r
cem en t co n ten t) ag ain st w /c ratio - 28 d ay w ater cu red
sam p les o n ly
Licensed copy from CIS: qacentral, HBK Contracting Company W.L.L., 25/06/2016, Uncontrolled Copy.

30
y = 1 8 .2 8 4 x + 5 .8 0 2 9
25 R 2 = 0 .3 8 8 8
carbonation depth/mm

20

15

10

0
0 .3 0 .4 0 .5 0 .6 0 .7 0 .8
w /c ratio

Figure B8. Dependence of 20-year normalised carbonation depth on the w/c ratio for
samples cured in water for 28 days.

The following graph shows the results for all data in BR216 (with no correction for cement
content):

Normalised carbonation depth (20 years) not corrected for cement content against w/c
2
ratio for all concrete (k = 0.2 - giving best fit from R )

70

60
Carbonation depth (mm)

y = 69.913x - 19.818
50
R2 = 0.3252
40

30

20

10

0
0.30 0.40 0.50 0.60 0.70 0.80 0.90 1.00 1.10
w/c ratio

Figure B9. 20-year carbonation depth against w/c ratio for all data included in BR216.

A similar graph taken from the same database is shown below for concretes that have
received an acceptable minimum cure (and ignoring 28-day data).
Normalised carbonation depth (20 years) corrected for cement
content against w/c ratio for concretes receiving 'acceptable'
minimum cure
Licensed copy from CIS: qacentral, HBK Contracting Company W.L.L., 25/06/2016, Uncontrolled Copy.

45
40
Carbonation depth

y = 39.252x + 2.9717
35
30
(mm)

25
20
15
10
5
0
0.3 0.4 0.5 0.6 0.7 0.8
w/c ratio

Figure B10. Normalised carbonation depth corrected for cement content against w/c ratio
for concretes receiving acceptable minimum cure

The Excel STEYX function gives a value of 7.49 for the scatter about the line of best fit.
The following graph (Figure 11) shows the correlation between the w/(c+kp) ratio (where p is
the pfa content and k is a factor representing its contribution to the properties of the
concrete) and indoor carbonation depth (normalised to 20 years but NOT corrected for
cement content).

Normalised carbonation depth (20 years) NOT corrected for cement content
against w/c ratio for concretes receiving 'acceptable' minimum cure (k = 0.2 -
giving best fit from R2)
60

50 y = 110.63x - 43.972
Carbonation depth (mm)

2
R = 0.7302
40

30

20

10

0
0.30 0.40 0.50 0.60 0.70 0.80 0.90
w/(c+kp) ratio

Figure 11. Normalised carbonation depth (not corrected for cement content) against
w/(c+kp) for concretes receiving an acceptable minimum cure
The same data but with the carbonation depth corrected for cement content are shown in
the following graph. The best correlation is obtained using a k value of 0.56 when the
correction for cement content is included. The standard deviation for these data is 4.5 mm –
this is comparable to that obtained for oxygen permeability v’s w/c ratio data, suggesting
Licensed copy from CIS: qacentral, HBK Contracting Company W.L.L., 25/06/2016, Uncontrolled Copy.

that this may be a possible approach to modelling the carbonation rate.

Normalised carbonation depth (20 years) corrected for cement content against w/c ratio for
2
concretes receiving 'acceptable' minimum cure (k = 0.57 - giving best fit for R ))k = 0.2 -
2
giving best fit from R )
45
40
Carbonation depth (mm)

35 y = 57.392x - 10.292
30 R2 = 0.5112
25
20
15
10
5
0
0.30 0.35 0.40 0.45 0.50 0.55 0.60 0.65 0.70 0.75 0.80
w/(c+kp) ratio

Figure B12. Normalised carbonation depth corrected for cement content against w/(c+kp)
for concretes receiving acceptable minimum cure.

B3. Accelerated carbonation

BR 216 includes a study of the rate of carbonation in concretes exposed to 10% CO2 levels.
The correlation between accelerated and normal carbonation rates is limited, however.
There is a sharp drop in the rate of accelerated carbonation after 1 - 3 days. The following
graph shows the correlation between accelerated and indoor carbonation (both are
1/2
normalised to 20 years assuming that carbonation follows an a + bt relationship). The
graph shows that there is a considerable degree of scatter, however, the gradient of the
graph (about 0.067) is comparable to that expected from the ratios of the assumed CO2
levels in each case. If the indoor CO2 level is 0.035% the ratio between accelerated and
indoor rates should be about 0.06.
0.5
i.e. dindoor/dacc = ([CO2]indoor/[CO2]acc)
Indoor carbonation (normalised to 20 years) against accelerated
carbonation (normalised to 20 years) for BR216 concretes
Licensed copy from CIS: qacentral, HBK Contracting Company W.L.L., 25/06/2016, Uncontrolled Copy.

80 y = 0.0673x + 23.022
Indoor carbonation
depth (mm)

60
40
20
0
0 100 200 300 400 500 600
Accelerated carbonation depth (mm)

Figure 13. Normal (indoor) and accelerated carbonation depths for BRE PC and PC/pfa
concretes

However, the scatter about the line of best fit is considerable and this approach, at least
based on the current data and analysis, does not appear to be a more suitable way of
predicting carbonation rates than that based on oxygen permeability data.

A closer analysis of both sets of data (i.e. indoor natural and accelerated) shows that much
1/2
of the scatter probably arises from the scatter in individual depth v’s t plots and the
consequent extrapolated 20 year values. There are only 4 - 5 data points for each sample
and there are problems in accurately determining the carbonation depth, especially at the
extremes of the range. It would be possible to go through both sets of data and remove
‘odd’ values and improve the correlation.
i
Glanville WH - the permeability of Portland cement concrete, BRS technical paper no. 3, 1931

You might also like