Sie sind auf Seite 1von 79

AP-T07

SERVICE LIFE PREDICTION OF REINFORCED CONCRETE STRUCTURES

AUSTROADS

Service Life Prediction of Reinforced Concrete Structures First Published 2000

Austroads Inc. 2000 This work is copyright. Apart from any use as permitted under the Copyright Act 1968, no part may be reproduced by any process without the prior written permission of Austroads.

National Library of Australia Cataloguing-in-Publication data: Service Life Prediction of Reinforced Concrete Structures ISBN 0 85588 576 9 Austroads Project No. N.T&E.9813 Austroads Publication No. APT07/00

Project Manager Rod McGee, Department of Infrastructure, Energy and Resources, Tasmania Prepared by Guangling Song, ARRB Transport Research Ahmad Shayan, ARRB Transport Research Published by Austroads Incorporated Level 9, Robell House 287 Elizabeth Street Sydney NSW 2000 Australia Phone: +61 2 9264 7088 Fax: +61 2 9264 1657 Email: austroads@austroads.com.au www.austroads.com.au

Austroads believes this publication to be correct at the time of printing and does not accept responsibility for any consequences arising from the use of information herein. Readers should rely on their own skill and judgement to apply information to particular issues.

SERVICE LIFE PREDICTION OF REINFORCED CONCRETE STRUCTURES

Sydney 2000

Austroads Incorporated Austroads is the association of Australian and New Zealand road transport and traffic authorities whose mission is to contribute to development and delivery of the Australasian transport vision by: !" supporting safe and effective management and use of the road system !" developing and promoting national practices !" providing professional advice to member organisations and national and international bodies. Within this ambit, Austroads aims to provide strategic direction for the integrated development, management and operation of the Australian and New Zealand road system through the promotion of national uniformity and harmony, elimination of unnecessary duplication, and the identification and application of world best practice. Austroads is governed by a council consisting of the chief executive (or an alternative senior executive officer) of each of its eleven member organisations. Member organisations

!" !" !" !" !" !" !" !" !" !" !"

Roads and Traffic Authority New South Wales Roads Corporation Victoria Department of Main Roads Queensland Main Roads Western Australia Transport South Australia Department of Infrastructure, Energy and Resources Tasmania Department of Transport and Works Northern Territory Department of Urban Services Australian Capital Territory Commonwealth Department of Transport and Regional Services Australian Local Government Association Transit New Zealand

Executive Summary
Service life prediction based on corrosion deterioration of reinforcement is critical to the safety of a reinforced concrete structure. Although considerable work in this area has been done, the prediction theory has not been well established. This report reviews and discusses the existing models, aiming at providing a background knowledge for future work in this area. The contents of this report can be summarised as follows: The deterioration processes of reinforced concrete structures due to corrosion damage are mostly caused by the ingress of chloride ions and carbonation of the cover concrete. The processes can generally be classified into three stages: initiation, stable propagation and unpredictable propagation. The service life prediction mainly refers to the first two stages. In the carbonation initiation stage, the diffusion mechanism plays a basic role in the carbonation process, based on which theoretical and empirical modifications and other more complicated formulation have been proposed and used when more complicated transport mechanisms other than diffusion are taken into account in the carbonation process. The pH threshold associated with the neutralisation of the pore solution by carbonation is discussed. The normal prediction procedures are also briefly addressed. The initiation of chloride induced corrosion is closely related to the ingress of chloride. The binding effect of the concrete matrix with free chloride ions makes a significant contribution to the chloride penetration process. Diffusion and binding are always the basic transport mechanism for chloride ions in concrete. However, in some cases when immigration and other transport mechanisms are also responsible for the ingress of chloride ions, multi-transport, more complicated and even empirical models have to be introduced to characterise the initiation process of the chloride induced corrosion. The chloride threshold is a critical parameter. Estimating relevant parameters from measured chloride profiles according to a selected model is usually involved in a prediction technique at this stage. Critical corrosion amount (corrosion depth, or thickness of corrosion product layer) is an important parameter in the stable propagation stage, which is addressed in most prediction models, including simple extrapolation, diffusion controlled corrosion modelling, laboratory simulation, and empirical formulation, etc. The prediction of this stage still requires the corrosion rate to be measured and the present corrosion state (amount) determined. Relatively less work has been done in association with this stage than the first stage. The final chapter of the report briefly summarises a methodology of service life prediction and identifies current problems with the existing prediction techniques.

Contents

1. 2.

INTRODUCTION.............................................................................................................1 DETERIORATION OF REINFORCED CONCRETE STRUCTURES....................2 2.1 2.2 2.3 Causes of corrosion damage.........................................................................................2 Deterioration due to corrosion of reinforcement.........................................................2 End of service life .........................................................................................................4

3. 4.

SERVICE LIFE PREDICTION APPROACHES.........................................................6 INITIATION OF CARBONATION INDUCED CORROSION ..................................8 4.1 Carbonation mechanisms and processes.....................................................................8 4.1.1 Diffusion of carbon-dioxide....................................................................................8 4.1.2 Carbonation reactions ...........................................................................................8 4.2 Mathematical models ...................................................................................................9 4.2.1 General diffusion model ........................................................................................9 4.2.2 Simple diffusion model ........................................................................................13 4.2.3 Empirically modified models...............................................................................14 4.2.4 Theoretically modified models ............................................................................14 4.2.5 Other complicated mechanisms ..........................................................................16 4.3 Carbonation situation ................................................................................................17 4.3.1 Carbonation depth ...............................................................................................17 4.3.2 pH threshold for corrosion...................................................................................17 4.4 Prediction procedures.................................................................................................20 4.4.1 Determination of extent of carbonation..............................................................20 4.4.2 Selection of prediction models.............................................................................21 4.4.3 Obtaining model parameters...............................................................................21

5.

INITIATION OF CHLORIDE INDUCED CORROSION........................................ 23 5.1 Ingress processes and mechanisms ...........................................................................23 5.1.1 Ingress and transport of chloride in concrete.....................................................23 5.1.2 Binding of chloride in concrete............................................................................24 5.2 Mathematical models .................................................................................................26 5.2.1 Simple diffusion model ........................................................................................26 5.2.2 Diffusion-binding model ......................................................................................29 5.2.3 Multi-transport-mechanism models ...................................................................31 5.2.4 Complicated models .............................................................................................32 5.2.5 Empirical models .................................................................................................37 5.3 Threshold concentration of chloride ..........................................................................38 5.4 Prediction techniques.................................................................................................40 5.4.1 Chloride profile ....................................................................................................40 5.4.2 Selection of models...............................................................................................40 5.4.3 Relevant parameters ...........................................................................................41

Contents (continued)

6.

CORROSION PROPAGATION OF REINFORCING STEEL.............................. 43 6.1 Corrosion processes and mechanisms .................................................................... 43 6.2 Critical corrosion amount ....................................................................................... 45 6.3 Prediction models .................................................................................................... 46 6.3.1 Simple extrapolation......................................................................................... 46 6.3.2 Diffusion controlled model................................................................................ 47 6.3.3 Laboratory simulation models ......................................................................... 48 6.3.4 Empirical prediction models............................................................................. 49 6.4 Prediction procedures.............................................................................................. 50 6.4.1 Reconfirmation of deterioration stage ............................................................. 50 6.4.2 Estimation of corrosion rate ............................................................................. 50 6.4.3 Determination of the present corrosion state.................................................. 51 6.4.4 Selection of prediction models.......................................................................... 51

7.

CONCLUDING REMARKS ....................................................................................... 52 7.1 7.2 Methodology of service life prediction .................................................................... 52 Problems to be solved .............................................................................................. 53

8.

REFERENCES ............................................................................................................ 55

Service life prediction of reinforced concrete structures

1. Introduction
Each structure has its own design service life according to certain codes, but such a parameter set for the sake of safety is often an estimate. Concrete is usually designed to be resistant to corrosion, and corrosion induced damage is not a main factor considered in the design. In some cases, the design service life could be too conservative; while, in some other occasions, it could be too risky. Therefore, reasonable and reliable prediction of service life of a structure based on corrosion induced premature damage is always an important issue to be addressed. A reliable prediction of service life based on corrosion state of the reinforcement is critical to the safety of reinforced concrete. It can also greatly affect the asset owners budget. A reasonable and accurate service life prediction can allow taking of necessary precautions in time, avoid unnecessary cost due to conservative design, and reduce risk of potential disaster due to the delay in taking precautionary measures. However, such service life prediction is always complicated and difficult due to two basic facts: A) service life is influenced by many unpredictable factors, such as the environmental parameters; B) the mechanisms dominating service life could vary from time to time. These difficulties are even more significant for reinforced concrete structures in the field because their deterioration could have various causes, such as a low grade of concrete used in severe environments, bad casting practices or design, chemical attacks from the environment, detrimental reactions in concrete (e.g. alkali aggregate reaction), corrosion of reinforcement, etc. Even though most of these deterioration processes have been widely investigated, the potential for deterioration to be induced by corrosion of reinforcement was not recognised for many years. This was because there was a misunderstanding that the reinforcement was non-corrodable due to the high alkalinity of the pore solution in the concrete and to the barrier provided by the cover concrete against aggressive agents from outside. Now it is generally accepted that corrosion of reinforcement can critically determine the service life of a reinforced concrete structure. Nevertheless, the history of investigations of corrosion of reinforcement is still relatively short, and the understanding of mechanisms involved is far from comprehensive. Therefore, service life predictions based on the corrosion mechanisms are still in their initial stages. The aim of this project is to provide an overview of available models for service life prediction of reinforced concrete structures.

Service life prediction of reinforced concrete structures

2. Deterioration of reinforced concrete structures


Signs of distress and deterioration can sometimes be found on concrete bridges in the form of cracking, delamination and spalling of concrete by rust, etc. However, more commonly, some reinforced concrete bridges may appear to be in good shape, but could actually be suffering from very serious corrosion attack on the reinforcement. The different manifestations of damage indicate different deterioration processes (mechanisms) the structures have undergone or are undergoing and different stages of deterioration.

2.1 Causes of corrosion damage


Many factors can induce deterioration in a reinforced concrete structure, such as earlyage defects of the concrete itself, freeze/thaw, abrasion, acid dissolution, alkaliaggregate reaction, sulphate attack, carbonation, chloride contamination, etc. Most of these processes are directly responsible for deterioration of the cover concrete, the discussion of which is outside the scope of this report. However, damage of the cover concrete leading to corrosion of reinforcement is the main issue to be addressed in this report. In addition to the factors that accelerate concrete deterioration accelerators, there are two processes which do not detrimentally affect the integrity and quality of the cover concrete to a great extent, but can directly trigger corrosion of reinforcement: 1) carbonation and 2) chloride contamination of concrete. It should be stressed that when corrosion induced damage becomes visible, deterioration is usually at a late stage and it may be too late to take any prevention or protection measures. Therefore, the service life prediction based on corrosion damage of reinforcement is particularly important. Carbonation induced corrosion most commonly occurs in relatively dry environments in which there is sufficient carbon oxide to diffuse into the cover concrete. However, in chloride containing environments, the ingress of chloride is usually faster than the carbonation process, and it is more likely to cause premature end of service of a structure.

2.2 Deterioration due to corrosion of reinforcement


Generally, deterioration of reinforced concrete starts with carbonation or ingress of chloride in the cover concrete. After the carbonation front or the chloride contamination reaches the vicinity of the reinforcement, the corrosion of reinforcement is initiated and damage commences.

Service life prediction of reinforced concrete structures

Fagerlund (1985) has schematically presented the different stages of the general deterioration process due to corrosion of reinforcement. Tuutti (1982) presented a curve of the development of corrosion damage with time. Purvis et al (1994) proposed an empirical equation to approximate this process:

100 1 $ a e # bt

(1)

where a and b are empirical constants; index S (in %) is a function determined by the percentages of delamination, spalling and bars with chloride above the corrosion threshold (Cl). Many investigators [Tuutti (1977, 1982), Cady et al (1984), Schiessl (1987), Broomfield (1995), Blankvoll (1997), Amey et al (1998)] have accepted the concept that deterioration experiences two stages: initiation (ti) and propagation (tp). In the first stage (ti), the main processes would be the ingress of chloride or CO2 into the cover concrete, and the transport of chloride ions or the progress of carbonation to the vicinity of steel reinforcement. In this stage, the corrosion rate of reinforcement is still very low, as the content of chloride is still far below the corrosion threshold and the pH value of the pore solution still high enough in the vicinity of the reinforcement to maintain the passivity of the reinforcement. The duration of the first stage (ti) could vary widely, depending on the aggressiveness of the environments and the quality of the cover concrete. Nevertheless, mechanisms which dominate this stage are the transportation of chloride or CO2 through the cover concrete. These mechanisms will not change significantly throughout this stage. In the second stage (tp), corrosion of steel in concrete is activated, and from then on the corrosion will develop further with time, which may finally lead to cracking and spalling of the cover concrete, if corrosion products are built up on the surface of steel. In this stage, the most important mechanisms are the corrosion reactions on the reinforcement. The corrosion rate of reinforcement involved in the second stage can change with time, particularly when cracking or spalling of the cover concrete occurs as a result of the corrosion of reinforcement. When the corrosion of reinforcement is activated, the deterioration of reinforced concrete (e.g. reduction of cross section, the propagation of pitting depth, or the increase in the thickness of the rust layer) will be governed by certain corrosion mechanisms, such as transportation of Fe2+ across the rust layer or O2, in the cover concrete.

Service life prediction of reinforced concrete structures

However, as corrosion damage of the reinforcement becomes more severe with time, the cover concrete would further deteriorate in respect of its corrosion protection effectiveness, through further cracking and spalling. As a result, the deterioration of the cover concrete can greatly worsen the corrosion conditions of the reinforcement by dramatically increasing the ease of ingress of chloride and CO2, the supply of corrosion depolarisation reagent O2, or removal of the rust layer which has a protection effect. Development of corrosion at this stage hasa different mechanism from that before the cover concrete is affected. After severe cracking or spalling of the cover concrete, the reinforcement may become directly exposed to the environmental agents, so that unpredictable changes of environmental parameters, such as wetting/drying, heating/cooling, etc. would directly impact on the corrosion of reinforcement with short response times. Therefore, the deterioration of reinforced concrete caused by the corrosion of reinforcement might better be classified into three stages: initiation (ti), stable propagation (tsp), unpredictable propagation (tup) (see Figure 1). They are dominated by different processes: transportation of chloride or CO2 in the concrete; corrosion of reinforcement under relatively stable conditions in uncracked cover concrete, and corrosion of reinforcement in cracked concrete, respectively.

deterioration degree (corrosion damage)

i ni t i at i on ( t i)

s t abl e pr opagat i on ) ( t s p

com pl et el y dam aged unpr edi ct abl e pr opagat i on ( t up)

af t ercr acki ng oft he cover concr et e

time

Figure 1. Deterioration stages of reinforced concrete due to corrosion of reinforcement

Service life prediction of reinforced concrete structures

2.3 End of service life


The above-mentioned stages (ti, tsp, and tup) are the basis for service life prediction. If all the stages, i.e. ti , tsp and tup , could be known, then the service life will be reasonably predicted. The service life of a structure may be different from the natural life or designed life. It seems that different researchers have different understandings of what should be defined as the service life for prediction purposes. Basically, service life can be defined as the period until repair becomes necessary [Geiker et al (1993)]. Visible changes such as large cracking, spalling and delamination are usually considered as severe damage. When the extent of these types of damage rise to a certain percentage of structural elements, repair or rehabilitation work may become necessary [Purvis et al (1994)]. For instance, Weyers, et al, (1994) found that bridge decks had reached the end of their functional service life when 5 to 14% of the deck area was spalled, delaminated or patched with temporary asphalt patches, and 4% cracking, spalling and delamination represented the end of functional service life of the substructures. Chamberlin et al (1993) reported an estimation of service life of an overlay on a concrete bridge deck by extrapolating its historical performance data. The overlay performance was also assessed by the percentage of deck area damaged by delamination or spalling (associated with either bond failures or reinforcement corrosion). The practical service life can also be defined as the duration before certain critical values of measurable properties are reached. This is a more widely adopted definition. In many cases, the definition of the service life is dependent on the deterioration mechanisms. For example, the time for the initiation of steel corrosion by ingress of chloride through cover concrete [Tuutti (1982)] or carbonation of cover concrete [Morinaga et al (1994)], has been taken as an important measure of concrete service life. In practice, there may be a risk to continue the service of a structure in its unpredictable propagation stage when damage, such as spalling and cracking, already exists. For the sake of safety, the service of such a structure should be limited or stopped as soon as the structure gets into its unpredictable propagation stage. On the other hand, it may also be too conservative to cease service at the onset of the stable corrosion stage, because the corrosion damage to the reinforcement may be still insignificant at that moment. It may be safe and economical to continue the service of a structure at this stage until certain amount of corrosion damage has occurred (reasonable prevention measures should be taken in this stage). Therefore, the service life (ts) could be defined as:

Service life prediction of reinforced concrete structures

ts & ti + tsp

(2)

where ti is the initiation time and tsp the stable propagation time. The rationale for such a definition is as follows: 1) At the end of the stable propagation stage, the reinforcement in concrete has been corroded to some extent, which has started to detrimentally affect the cover concrete. Therefore, it could be the right moment that some repair or protection actions should be taken. Or the service of the structure should be stopped for the sake of safety. The deterioration processes in the initiation and stable propagation stages dominated by certain predictable mechanisms, so the deterioration processes are relatively easy to predict compared with that in the unpredictable propagation stage which are affected by too many varying environmental factors.

2)

In this report, service life prediction will be based on the initiation and stable propagation stages.

Service life prediction of reinforced concrete structures

3. Service life prediction approaches


To predict service life, the following facts must be known: A) the mechanisms governing the initiation and stable propagation stages, B) the relevant parameters involved in the mechanisms, and C) the present state of the reinforced concrete. Currently, prediction approaches have been classified into the following different categories [Clifton (1993)]: 1) 2) 3) 4) 5) estimations based on experience, deductions from performance of similar materials, estimations based on accelerated testing, mathematical modeling based on the degradation and corrosion processes, and application of some other new concepts, such as reliability and stochastic methods, etc.

For the first approach, the prediction is usually based on in-situ inspection. The performance of a concrete structure evaluated at certain time intervals can be extrapolated to the future [Sayward (1984)]. In the prediction, the mechanisms, the relevant parameters, and the stage that the structure is at, all need to be estimated, so this approach is not very reliable. However, it is very simple compared with other methods. The second approach is also mainly based on experience. However, before the prediction is made, a sufficient database of the performance of similar materials must be available. In some cases, empirical expressions are proposed for the progress of deterioration with time [Purvis et al (1992)]. As the relevant information is provided by the database (which should be much more accurate than that estimated according to experience), this approach is more reliable than the first method. However, the mechanisms dominating the deterioration process, and the stage that the structure is at, are still not accurately determined. Prediction based on laboratory tests uses accelerated or non-accelerated techniques to simulate the deterioration process of reinforced concrete in the field. Due to the fact that the deterioration of concrete under laboratory conditions may be different from the natural deterioration process in the field, it is essential to examine the correlation between laboratory results and field damage. As the mechanisms dominating the deterioration processes can be investigated under laboratory conditions, and the relevant parameters involved in the mechanisms could also be obtained by laboratory tests, the third approach is more reliable than the first two. This kind of test has been commonly employed as a standard method of assessing reinforced concrete durability [Vesikari (1986), Kalousek et al (1972)].

Service life prediction of reinforced concrete structures

Recently, with the development of computer techniques and the accumulation of data, mathematical simulation and modelling has been widely proposed for more reliable service life prediction. This kind of prediction requires knowledge of the dominating mechanisms, relevant parameters and current deterioration stage. One of the typical models currently widely used is the modelling of transportation of chloride in the cover concrete based on Ficks diffusion law. In this case, the transportation of chloride in concrete is believed to be mainly governed by the diffusion mechanism, which provides a detailed mathematical expression for the model. The relevant parameters critical to prediction, such as diffusion coefficient of chloride in concrete, surface concentration of chloride and chloride content depth profile in the concrete, etc., are usually experimentally determined. Based on the chloride profile, the corrosion stage of the structure can easily be identified. Therefore, this kind of approach is theoretically much more reliable than the previous three, and should have a promising future. However, some difficulties have been encountered in the application of this kind of approach, e.g. the contribution of mechanisms other than the diffusion, the variation of the relevant parameters with time and from part to part in the same structure, etc. These difficulties can greatly affect the reliability of this kind of prediction. In theory, if all the mechanisms affecting the deterioration processes and the corresponding parameters can be identified, then a prediction model can be accurately established and service life reasonably predicted. In practice, this is very difficult in many cases due to the complexity of the concretes and real-world environments, as well as a large number of unknown factors. To cope with these difficulties, some new concepts, such as reliability and stochastic methods [Clifton (1993)], have been proposed. However, these trials need further verification and more evidence from the field. In summary, the fourth approach is currently the most reasonable and reliable, and hence the most promising in the future. Therefore, in this report, only the fourth approach will be reviewed.

Service life prediction of reinforced concrete structures

4. Initiation of carbonation induced corrosion


Carbonation is an ongoing chemical reaction between carbon-dioxide in the atmosphere and some calceous components in the cementitious matrix, leading to the formation of calcium carbonate. It is one of the main causes of corrosion of reinforcement in concrete structures, and is a common phenomenon in old or badly built structures. In a concrete structure exposed to the atmosphere, a carbonation front develops and gradually advances inward from the surface of the concrete structure. After the front reaches the reinforcement, corrosion of the reinforcement is initiated, then the deterioration moves into the next stage, the corrosion propagation stage.

4.1 Carbonation mechanisms and processes


The carbonation phenomena have been widely investigated and well summarised [Aschan (1963), Hamada (1968), Smolczyk (1968), Kondo et al (1968), Wittmann et al (1986), Parrott (1987), Reardon et al (1990)]. Carbonation is regarded as being governed by a CO2-transport process accompanied by the chemical reactions that form calcium carbonate.

4.1.1 Diffusion of carbon-dioxide


Even though convection effects and the diffusion of other ions could contribute to the CO2 transport process, the diffusion of CO2 in concrete pores is generally believed to be the main mechanism responsible for the process, and can be theoretically expressed by Ficks diffusion law in a similar manner as that for chloride ingress [Li et al (1987)]. The difference in concentration of free carbon-dioxide within the cavities and pores is the driving force for the diffusion. The carbonation can progress throughout the concrete provided sufficient CO2 is available in the surrounding environment and the exposure time is long enough.

4.1.2 Carbonation reactions


After CO2 transports into concrete, it causes carbonation reactions in the concrete. Although the reactions involved in the carbonation process are very complicated [Richardsson (1988), Papadakis et al (1991a,1991b)], they are usually briefly expressed as:

Service life prediction of reinforced concrete structures

CO2 + H2O (pore solution)' H2CO3

(3)

H2CO3 + Ca(OH)2 ' CaCO3 + 2H2O


These reactions mainly occur in the pore solution.

(4)

The pore solution of plain concrete mainly consists of sodium, potassium, and much less calcium ions in equilibrium with hydroxide ions at a pH value between 12 to 14. The production of the CaCO3 precipitate during carbonation consumes the Ca(OH)2 in the concrete (reaction (4)) and lowers the pH value of the pore solution. In addition to the above principal carbonation reactions, CO2 in the concrete could still participate in some other reactions. For example, CO2 can directly react with the hydrated cement products, such as the calcium silicate hydrate and calcium aluminosulfates and cause their disintegration. Usually, the latter is significant only in badly deteriorated concrete. The lowered pH level of the pore solution in concrete provides an aggressive medium environment for the steel, because the passive film on the reinforcing steel, which protects the steel from corrosion, becomes unstable when pH is lower than 11 [RILEM (1976)], thereby steel can be readily corroded in carbonated concrete.

4.2 Mathematical models


It should be stressed that the transport of CO2 and carbonation reactions occur simultaneously in the concrete. The former is usually much slower than the latter [Klopfer(1978)], hence the overall carbonation process, in most cases, is dominated by CO2 transport mechanisms.

4.2.1 General diffusion model


If only the diffusion of CO2 in the pores and its consumption by carbonation of concrete are considered in the carbonation process, then this process can be mathematically expressed by the following equation, which has been proposed in similar format by several authors [Fukushima (1991), Papadakis et al (1991b), Mauda et al (1993), Xu (1995)]:

10

Service life prediction of reinforced concrete structures

(C CO 2 ( (C % ( DCO 2 CO 2 )# f CO 2 (t (x (x

(5)

where, CCO2 is concentration of free CO2 in concrete, x is the space coordinate, t is the time coordinate, DCO2 is the diffusion coefficient of the CO2 in the concrete, fCO2 is the overall CO2 consumption rate of the carbonation reactions, a concentration of free CO2 dependent function. This is a generalised model, which is based on a basic assumption: I) Diffusion and consumption of CO2 by carbonation reactions are the main mechanisms dominating the carbonation process.

In addition, three assumptions or approximations are also widely used to simplify the above equation: II) III) IV) The concrete matrix through which the CO2 diffusion occurs is uniform in microstructure and properties; The microstructure and properties of the concrete do not change significantly with time; and The microstructure and properties of the concrete do not change significantly with the concentration of the diffusing CO2.

Under these three assumptions, the diffusion coefficient DCO2, will be constant, independent of time (t), space (x), and CO2 concentration (CCO2). In this case, equation (5) can be further simplified into:

(C CO 2 ( 2C CO 2 % De ,CO 2 (t (x 2
where De,CO2 is effective diffusion coefficient:

(6)

De ,CO 2 %

DCO 2 (f 1$ CO 2 (C CO 2

(7)

which represents the transport rate of free CO2 under the carbonation effect. If V) The free CO2 consumption function fCO2 is simply assumed to be linearly dependent on the concentration of free CO2, i.e. kCO2 = (fCO2/(CCO2 = constant,

11

Service life prediction of reinforced concrete structures

Then, the effective diffusion coefficient would become a constant too:

De ,CO 2 %

DCO 2 1$ k CO 2

(8)

In this case, Equation (6) is a typical Ficks second diffusion equation, which has an analytical solution under semi-infinite boundary condition and zero initial condition: VI) The concentration (Cs,CO2) of CO2 on the concrete surface is constant, and the interior of the concrete far from the carbonation front has zero concentration of CO2, i.e.

CCO2|x=0 =Cs,CO2 =const;

CCO2|x=- = C-,CO2 = 0
(9)

VII) The concentration (Ci ) of CO2 in the concrete before carbonation is zero, i.e.

CCO2|t=0 = Ci,CO2, = 0

(10)

Conditions VI) and VII) are actually another two approximations, as for a field structure the concentration of carbondioxide in the surrounding air could change with time, and also the structure could be carbonated to some degree during its construction process, so semi-infinite and initial concentrations of CO2 could not be zero. Based on the above assumptions I) to VII), the following analytical solution to equation (6) can be obtained:

CCO2 = Cs,CO2 {1-erf [x/(De,CO2 t)1/2]}

(11))

which will give a distribution of concentration of CO2 from the concrete surface to the interior at a given time, as shown in Figure 2.
)

erf ( + )% 1# erfc( + )%

2 ,

e # z dz
2

12

Service life prediction of reinforced concrete structures

CCO2, concentration of CO at time t=t

Cs,CO2

0 0

xt

x, depth from concrete surface

60

Figure 2. Typical distribution of CO2 concentration in concrete Correspondingly, the change of the concentration of CO2 in a given position (depth) in the concrete can be schematically illustrated as Figure 3.

CCO2, concentration of CO2 at depth x=xt

Cs,CO2

0 0

tx

t, carbonation time 400000

Figure 3. Change of CO2 concentration with time at a given depth in concrete

13

Service life prediction of reinforced concrete structures

A very important conclusion from the above solution (11) is that any given concentration (C0,CO2) of CO2 at a given depth x0 in the concrete and at a given moment t0 will advance inward to the interior of the concrete with time in a parabolic manner: (12)

XCO2 = A0,CO2
1/2

(13)

A0,CO2= x0 /t0 ;

1/2

t>t0, XCO2>x0

where, XCO2 is the new position of the constant concentration (C0,CO2) of CO2 at carbonation time t (Figure 4).

18
concentration (C0,CO2 ) of CO2 XCO2, depth of a selected

x0

0 0

t0

t, carbonation time

1500000

Figure 4. Progress of CO2 of a given concentration with carbonation time in concrete Theoretically, CCO2 will not decrease to 0 unless at an infinite distance from the concrete surface. However, it is more practical to choose a depth Xc as carbonation depth or carbonation front, at which the concentration CC is low enough, and beyond which the concentration of CO2 and the carbonation process can be neglected. As Xc is actually a particular point of X, it also follows the above parabolic equation (12), so:

Xc = Ac t1/2 ; (Ac = constant)

(14)

implying that the carbonation depth also advances with time in a steadily reducing rate. It is obvious that the assumptions I) to VII) are not reasonable in any cases. Only when all these assumptions are satisfactorily met, the solutions and conclusions about the distribution, change, and progress of the concentration of CO2 in concrete can reliably reflect the realistic situation of field concrete structures.

14

Service life prediction of reinforced concrete structures

4.2.2 Simple diffusion model


The above general model can be further simplified by assuming that the flow of carbondioxide through the carbonated zone is subject only to the concentration of carbondioxide at the concrete surface and the distribution of CO2 concentration is linear along the depth of the concrete [Kondo et al (1968), Hamada (1968)]. This can be mathematically expressed by

(CCO2/(x = Cs,CO2 /Xc= const.

(15)

Based on this assumption, the general model equation (5) becomes much simpler, having a solution for the progress of carbonation depth with time:

Xc = (2De,CO2 Cs,CO2 K)1/2 t1/2 =Ac t1/2

(16)

where K and Ac are constants. This solution is similar to equation (14) which is based on the general model. However, the theoretical base of this solution (16) is not as solid. Moreover, the concentration distribution of CO2 does not follow a straight line along the depth of concrete, as assumed for the derivation of Eq(16).

4.2.3 Empirically modified models


In practice, the measured results are sometimes similar to the theoretical predictions made by general or simple diffusion models, but cannot be satisfactorily simulated by them. For simplicity, some empirical modifications could be made to equations (12) and (15) to simulate or curve-fit the measured results in the field. The general format of the modified equations of carbonation depth is:

Xc = A tn + B
where A and B are constants, and n could have a range of values around .

(17)

It has been noted that the linearity of Xc vs. t1/2 does not always start from time t=0, i.e. there appears to exist an initial depth (B) of carbonation in some cases [Nagataki et al (1986)]. The initial depth could be ascribed to the drying of the concrete surface that allows easy carbonation at early ages. In addition, a non-parabolic (square root) function of carbonation time was also proposed by Parrot et al (1991) for wet conditions. In that case, the power differs from and depends on the relative humidity of the environment.

15

Service life prediction of reinforced concrete structures

Generally, the empirically modified models do not have a solid theoretical base. Even though practical results could be better simulated by them, the reasonability of the predictions still need to be carefully analysed.

4.2.4 Theoretically modified models


As mentioned earlier, the conclusions based on the general and simple diffusion models are only applicable under certain conditions where those assumptions are completely met. However, in practice, all these assumptions do not always hold, because the carbonation process can be affected by many factors. Concrete parameters can significantly affect CO2 transport and carbonation processes. The carbonation rate can be dependent on porosity, pore size, cement content, w/c ratio, aggregate/cement ratio, etc. [Papadakis et al (1991a,1991b)]. This makes the prediction much more complicated. For example, in some cases, the influence of heterogeneity of concrete is very significant in the carbonation process, so space-dependent parameters would be involved in the model. In other cases, the porosity, pore size and other concrete parameters can still be modified by the deposition of the carbonation product, CaCO3, [Broomfield(1997)], which means that some parameters involved in the model could be time-dependent. Research work has been conducted on this issue to further improve understanding of the influence of concrete parameters on carbonation. In a theoretical and experimental study, Li et al (1987) found that pores with radii over 320 are most responsible for CO2 transport in concrete which is ascribed to the fact that this pore size is larger than the free-path of CO2 gas molecules for atmospheric carbon dioxide. Papadakis et al (1991A) calculated the possible amounts of Ca(OH)2 and other hydration products produced by each cement component, and their (empirical) contribution to reducing the effective path for CO2 diffusion. They proposed empirical formulae to estimate the effect of total porosity and relative humidity on the diffusion coefficient of CO2 in concretes measured in their tests. Hamada (1968) concluded that concrete with lesser proportions of Ca(OH)2 in the matrix would carbonate faster than those with higher proportions of this substance. However, Smolczyk (1968) believed that a high content of free Ca(OH)2 cannot protect any concrete against carbonation. Houst (1991) simultaneously determined the diffusion coefficients of oxygen and carbon-dioxide through hydrated cement paste as a function of relative humidity, and found that the coefficients increased with increasing w/c ratio and decreasing relative humidity. Similar trends for CO2 diffusion coefficients have also be reported by other authors [Masuda et al (1993), Cahydi et al (1993)]. In making predictions, the influences of the above factors are usually reflected in the parameters involved in equation (19). For example, some researchers have expressed A as a function of w/c [Hamada (1968), Zhang et al (1998)] whilst others [Smolczyk (1968), Nagataki et al (1986), Schubert (1987), Gebauer (1982), Scholz et al (1984), Costa et al (1992)] related A to the concrete strength. Parrot et al (1991) used a carbonation depth function which had concrete permeability (to gas) and relative humidity as relevant parameters.

16

Service life prediction of reinforced concrete structures

Environmental factors also have a significant influences on CO2 transport and carbonation processes. Temperature can increase the diffusion coefficient of CO2, hence the carbonation process is usually speeded up at high temperature. Atmospheric concentration of CO2 directly affects Cs,CO2, so the carbonation is normally more severe if a structure is exposed in an environment with a high atmospheric concentration of CO2. Rain and wind can also influence humidity, Cs,CO2, and transport mechanisms of CO2 in concrete, so the effects on the carbonation process can be complicated. It was reported [Klopfer (1978)] that the greatest carbonation rate occurs when concrete is exposed to relative humidities of between 50% and 70%. An atmosphere with 100% relative humidity can prevent carbonation of concrete, as the water in the pores close to the surface of the concrete impedes the penetration of carbon dioxide. If the relative humidity is lower than 30%, the carbonation of concrete would not be significant, since the carbonation reaction requires water. The higher the strength class of the concrete and the cement content, the more slowly the concrete is carbonated. A higher content of calcium hydroxide in the concrete, i.e. higher cement content could also lead to smaller depth of carbonation in the concrete. Unfortunately, all these factors including concrete parameters and environmental influences could vary with time (t), space (x), and concentration of CO2, etc. in the field. This means that the relevant parameters of those diffusion based models, such as De,CO2 and Cs,CO2, etc. (which are closely associated with concrete and environment parameters) are not constants. For example, in some cases, the Cs,CO2 could change seasonally, or De,CO2 may not longer be constant, but depend on time, space, CO2 concentration, etc. If these practical factors are taken into account in the model, then the equation describing the carbonation process would become very complicated and may have no analytical solutions. Some authors [e.g. Fukushima (1987), Fukushima (1991),] have made some considerable effort in clarifying these issues. Fukushima (1991) introduced a function that described the long term as well as periodical variations in the atmospheric carbon-dioxide concentration, and assumed that the CO2 consumption rate during the carbonation process is linearly dependent on the CO2 concentration. Therefore, the boundary condition Cs,CO2 is no long constant, and also the term describing CO2 consumption in equation (5) becomes a function of CCO2. In this case, they obtained a complicated solution to the modified equation, which may be hard to apply in practice. Masuda et al (1993) proposed a non-linear reaction term in replacement of the carbonation consumption term in equation (5). In their models, the carbonation consumption rate is not only determined by CO2 concentration, but also by concentration of Ca(OH)2 in the concrete. They gave an approximate solution to this equation. Similar attempts have been made by Cahyadi et al (1993) who considered the carbondioxide consumption term as a function of time, and took the internal transport of water within the pore system into account.

17

Service life prediction of reinforced concrete structures

All these modified models may be more reasonable than the general one, and could theoretically provide more reliable prediction of carbonation process, as some realistic factors influencing the carbonation process have been considered in these models. However, the complicated models usually would not have solutions as simple as the general or simple diffusion models, and could even have no analytical solutions in some cases. Therefore, the simple expression (equation (14) or (16)) for the progress of carbonation depth could not be obtained from these models. Obviously, they would make the prediction of carbonation depth difficult in practice.

4.2.5 Other complicated mechanisms


Some other mechanisms could also be responsible for the transport of CO2 in concrete, for example: 1) Convection effect: internal movements of the pore solution as a result of wetting and drying. This would not only speed up the transport of carbondioxide, but also eventually accelerate the movements of carbonate ions, alkali ions, etc., from place to place. Diffusion of ions in the pore solution: the ions involved in the carbonation reactions diffuse away or into the carbonating zones. This could also increase or decrease the carbonation process.

2)

If these mechanisms dominate the carbonation process, then the models mentioned earlier would be invalid, and significant error would be introduced in service life prediction if it was only based on the diffusion models.

4.3 Carbonation situation


Carbonation situation refers to the extent and intensity of carbonation in concrete.

4.3.1 Carbonation depth


As mentioned earlier, carbonation depth Xc is the distance from the concrete surface where carbonation originally starts to where carbonation starts to become insignificant. Within this zone, the properties of concrete change from its original uncarbonated condition, particular the pH value of the pore solution is decreased. Based on the simple or general diffusion model, the carbonation depth is a parabolic (square root) function of carbonation time, and can generally be expressed as:

Xc = Ac t1/2

(18)

18

Service life prediction of reinforced concrete structures

where constant Ac is closely related to the concrete properties, particularly the permeability, amounts of calcium hydroxide and moisture in the concrete, and the concentration of CO2 in the environment. Theoretically this function may be less accurate than those derived from other modified models which take more realistic factors into account. However, its simplicity and capability of fitting field results has made this equation overwhelmingly popular as a base for carbonation prediction. In research, in spite of the arguments on how to archive this simple equation, the general conclusion that there existed a square root relationship between the depth of carbonation and the carbonation time has been supported by various investigators [Smolczyk (1968), Parrott (1968), Goodgrake (1978), Tuutti (1982), Richardsson (1988), Papadakis et al (1991a,1991b)].

4.3.2 pH threshold for corrosion


The carbonation degree could be indicated by various parameters, including the ratio of CaCO3 (carbonation product) over calcium containing components of cement matrix; the ratio of bound CO2 over maximum capacity to bind CO2, etc [Power (1962),]. However, as far as corrosion of reinforcement is concerned, the parameter of principal interest would be the pH value of the pore solution, as this parameter indicates the corrosivity of the concrete and is closely related to the stability of reinforcement in the concrete. The higher the CO2 concentration in the concrete, the higher the consumption of OH-, and the lower pH value of the concrete. If carbonation is governed by the diffusion of CO2 in concrete, and the other carbonation reactions can be approximately regarded as steady or meta-steady processes, then an inverse proportional relationship could easily be obtained between the concentration of CO2 and concentration of OH- in concrete. In fact, experiments have demonstrated a clear decrease in the pH level as the carbonation degree (characterised by CaCO3/Ca(OH)2 ratio) increases (Figure 5) [Ohgishi et al (1983), Parrott (1987)].

19

Service life prediction of reinforced concrete structures

Figure 5. relationship between pH level and CaCO3/Ca(OH)2 ratio determined by X-ray in a concrete wall [Ohgishi et al (1983), Parrott (1987)] Since the concentration of CO2, as well as the carbonation degree, decreases with the depth of concrete (Figure 2), the typical pH profile in concrete would tend to increase from the concrete surface to the interior (Figure 6) [Forrester (1976), Broomfield (1997)].

Figure 6. Typical pH depth profile [Broomfield (1997)]

20

Service life prediction of reinforced concrete structures

As shown in Figure 6 by Broomfield (1997), the threshold of pH for initiation of corrosion of reinforcement is about 11. In fact, some degree of carbonation would not necessarily cause corrosion of reinforcement. Hence, the carbonation depth (Xc) described by equation (18) is not closely related to the depth where corrosion of reinforcement could occur. In this sense, the carbonation depth Xc characterising the real carbonation extent cannot be used for accurate prediction of carbonation induced corrosion of reinforcement. The reasonable depth corresponding to the initiation of corrosion of reinforcement should be the depth where the pH value of the pore solution is lower than the pH threshold. This depth, XpH, should have a corresponding concentration of CO2 (CpH), which is a particular point on the CO2 concentration depth profile (Figure 2), and can also be formulated by:

XpH = ApH t1/2

(19)

where ApH is still a constant. The depth XpH as a function of carbonation time corresponds to the position where corrosion of reinforcement could occur. In other words, if the depth of pH threshold (XpH) reaches the reinforcement in concrete, then the reinforcement could be corroded. Therefore, from equation (19), when the corrosion of reinforcement will take place can be predicted, provided that current depth of pH threshold and the constant ApH involved in equation (19) are known.

4.4 Prediction procedures


In order to predict the carbonation progress, the current carbonation situation should be determined, then a reasonable mathematical model should be selected, and finally the necessary parameters required should be measured, estimated or calculated.

4.4.1 Determination of extent of carbonation


Normally, concrete samples should be taken from structures for the purpose of the determination of the extent of carbonation (either in the field directly or in the laboratory). A number of methods are available for either qualitative or quantitative measurement of carbonation, such as X-ray diffraction [Richardsson (1988)], infra-red absorption spectroscopy [Slegers et al (1976)], thermal analysis [Kroone et al (1959)], mineralogical analysis [Meyer (1969)], and chemical analysis [Hamada (1968), Tuutti (1979), Litvan et al (1986), Kondo et al (1968), Conway (1957)].

21

Service life prediction of reinforced concrete structures

In practice, carbonation of a concrete specimen is determined simply by means of spraying it with phenolphthalein, a pH indicator which is pink at pH values greater than 9.0-9.5 and colourless at pH values below 9.0. This is the most popular method for field testing. As far as corrosion of reinforcement is concerned, it demonstrates the extent of the pH drop in concrete due to carbonation, which is directly responsible for the corrosion damage of reinforcement in concrete. The phenolphthalein indicator method of monitoring the depth of carbonation has been recommended by RILEM (1984). The indicator is prepared as a solution of 1% phenolphthalein in 70% ethyl alcohol. The samples should be broken rather than sawn to open fresh surfaces for application of the indicator. A solution of 3% phenolphalein and 95% denaturised ethyl alcohol is also suggested in the Swedish standard SS13 72 42. In fact, various researchers have used different types of acid/base indicators and different procedures of using phenolphthalein [Meyer (1969), Hobbs (1988)]. For example, the concentration of the phenolphthalein could vary from 1 % to 3% and that of the ethyl alcohol from 95% to 30%; other indicator types that have been tried include thymolphthalein and Alizarin R Yellow [Meyer (1969)]. The phenolphthalein indicator method of monitoring the depth of carbonation has proven convenient and yields reproducible results. The main limitation of the method is that it only indicates the depth at which the pH level is below 9.0-9.5. This does not necessarily correspond to either carbonation depth or pH threshold depth. In other words, the depth of colour change of the specimen (Xi) is different from the carbonation depth (Xc) and the pH threshold depth (XpH) (Figure 6). Therefore, it is generally concluded that the phenolphthalein indicator provides an approximate and incomplete picture of carbonation [Parrott (1987)]. If Xi is regarded as Xc or XpH, then a few mm error in the predicted carbonation depth and corrosion depth could occur, which could correspond to a few years of error in the predicted service life. Nagataki et al (1986) found that initiation of corrosion of reinforcement occurred earlier than the carbonation front, determined by the phenolphthalein indicator, reached the steel.

4.4.2 Selection of prediction models


In most cases, equation (18) or (19) is sufficiently reliable to model carbonation progress, based on which the initiation of carbonation induced corrosion of reinforcement can reasonably be predicted. However, in some cases, the measured results significantly differ from the prediction of equation (18) or (19), for which the general and simple diffusion models appear to be unsuitable for service life prediction. In these situations, the theoretically modified models and/or the empirically modified equations should be considered. It is relatively easy to use the empirically modified equation (17) to fit practical results. However, the meanings of the parameters involved need to be carefully analysed, as they lack theoretical foundation. On the other hand, the theoretically modified models could more realistically describe the actual carbonation process, but their complicated solutions (sometime even no analytical solutions), could be meaningless to practical service life prediction.
22

Service life prediction of reinforced concrete structures

4.4.3 Obtaining model parameters


To predict service life according to certain models, the parameters involved in the models should be known. There are several basic methods to obtain the parameters, including field testing and laboratory analysis, simple measurement and artificial acceleration. The selection of these methods depends on the models chosen for prediction. If equation (18) or (19) is used, then the simplest way to obtain parameter Ac or ApH should be the phenolphthalein spraying method. As mentioned earlier, even though Xi is unequal to Xc and XpH, they follow the same form of equation. For a simple and approximate prediction, phenolphthalein indicator can be used to estimate XpH with acceptable errors. For example, based on equation (19), the relevant parameter ApH can be obtained as follows: First, the carbonation time t is easily known at that moment, provided that carbonation started right after the construction of the structure. Then use the measured Xi as XpH and t to calculate ApH according to ApH = Xi/t1/2. The most attractive advantage of this method is its easy operation and suitability for field use. However, the accuracy of the parameter relies on one measurement, so the reliability of the calculated parameter could not be high, and consequently the prediction based on the parameter has low credibility. Measurements at different times are recommended, even though this would make the prediction more time consuming. Once ApH is known, determine how long it would take the carbonation depth to be equal to the depth of reinforcing steel in concrete. This would indicate the time for the initiation of corrosion. If empirically modified model (17) is used for prediction, then the simple phenolphthalein method can still be used in the field. However, more than three different times of measurements are essential, and the time intervals between the different measurements should be specially selected to simplify the calculation of parameters A, B and n. Therefore, it is more time consuming, but still very simple and flexible and is particularly suitable for field use. If the prediction is based on the general diffusion model (equation (11)), then parameters Cs,CO2 and De,CO2 need to be estimated. Cs,CO2 could be directly measured in the field, i.e., the atmospheric CO2 concentration, and De,CO2 can be measured by specially designed experiments in the laboratory. Another way to obtain these parameters is to measure the carbonation degree (or concentration of CO2) depth profile, through which the above parameters can also be calculated using equation (11). The methods that can be used to obtain the depth profile have been mentioned earlier in section 4.4.1. If theoretically modified models are selected, then more parameters need to be estimated. The commonly used methods for determining De,CO2, Cs,CO2 and CCO2 depth profile would still be similar to those used to obtain the parameters for equation (11). Other parameters involved may need special measurement. Usually, these models have more complicated equations and solutions. To obtain relevant parameters based on these models, numerical solutions based on computation techniques could be preferable.

23

Service life prediction of reinforced concrete structures

5. Initiation of chloride induced corrosion


The ingress and accumulation of chloride into the cover concrete is the first stage of chloride induced corrosion of reinforced concrete. Browne (1982) assumed that the corrosion-free life of a structure was the time for chloride ions to reach a critical concentration at the surface of reinforcement. Chloride ingress is more complicated than the carbonation process. It is less uniform across the cover concrete, and there is no such clear chloride front as in the case of carbonation. The chloride-induced corrosion is relatively fast and localised, and difficult to remedy after initiation [Blankvoll (1997)]. Corrosion of reinforcement is highly dependent on the process of ingress of chloride ions into concrete and transportation to the reinforcement steel. Understanding of these process is important in dealing with the corrosion problem.

5.1 Ingress processes and mechanisms


Chloride ions originate from different sources, such as 1) deliberate addition of chloride salts to concrete as accelerators; 2) use of water containing Cl-; 3) contaminated aggregates; 4) sea salt spray and direct sea water wetting; 5) deicing salts; and 6) use of chemicals. Etc. Normally, chloride ions exist in concrete in two forms: 1) dissolved in the pore solution as free chloride; and 2) adsorbed on cement gel or combined with hydrated cement phases and aggregates as bound chloride. Only free chloride can accelerate the corrosion of steel in concrete. The bound chloride is inert to steel before it is dissolved into solution and becomes free chloride. Typically about 40~50% of the total chloride in concrete is bound [Gaynor (1985)]. Dhir et al (1990) estimated that the ratio between soluble and insoluble chloride in concrete was roughly 3 to 1. Stoltzner et al (1997) found that the ratio of free to total chloride contents in concrete with low-alkali, sulfateresistant cement was about 50~70%, and for portland cement concrete 35~60%.

5.1.1 Ingress and transport of chloride in concrete


Ingress and transport always refers to the free chloride ions. After the free chloride ions are bound (adsorbed or chemically reacted), they would not travel in concrete until they become free (desorbed or dissolved) under certain conditions. Chloride can travel in concrete and reach the reinforcing steel through diffusion, capillary movement or electrical migration, etc. Chloride transport in concrete can be greatly affected by mechanisms other than diffusion. Mechanisms such as hydroxide leaching, osmotic pressure, hydraulic flow, suction and filtering [Volkwein (1993)], and other membrane effects may all affect the transport of chloride through the concrete pores.

24

Service life prediction of reinforced concrete structures

The penetration of chloride into a few centimeters of the outer cover concrete is initially dependent upon capillary suction, but the ingress of chloride ion into a greater depth would mainly be governed by long term diffusion [Bamforth et al (1990), Bamforth et al (1994)]. On a dry surface the uptake of chloride ion is mainly due to absorption; the chloride containing solution is absorbed into empty micro-cracks and pore spaces, and then penetrates further by capillary suction [Goto (1981a)]. If the surface is wet, the initial entry is likely to be by permeation or diffusion. In a field concrete structure the ingress of chloride is very complicated and might be governed by a combination of several mechanisms. Figure 7 is a schematic illustration of some possible transport processes in a sea wall [Concrete Society Discussion Document(1996)]. Because of the combined contributions of different transport mechanisms in the sea wall, the deepest penetration of Cl- was found at about 1~2 metres above sea level, and the highest Cl- concentration was detected at the surface layer in the wall about 8 metres above the sea level.

Figure 7. Schematic diagram of the transport process of chloride in a sea wall [Concrete Society Discussion Document (1996)]

25

Service life prediction of reinforced concrete structures

5.1.2 Binding of chloride in concrete


Binding of chloride refers to the adsorption of free chloride ions by cement gel and/or their chemical reactions with hydrated cement. When free chloride ions from the environment penetrate into concrete, some can be captured by the hydrated products through physical adsorption (physical binding) or chemical adsorption or reactions (chemical binding). The huge surface area of hydrate gel supplies plenty of active sites for the physical binding, while the chemical binding could be realised through reaction with the aluminate phases and formation of Friedels salt. Even though detailed mechanisms for the binding of chloride in concrete are still unclear, it is generally believed that there is an equilibrium between the adsorption of free chloride from pore solution, and the desorption of bound chloride from cement gel. Some researchers [Tuutti (1982a), Arya et al (1990a), Fishcher et al (1984)] have assumed that the equilibrium relationship was only a simple linear adsorption isotherm. However, it was also suggested [Dhir et al (1990)] that the relationship between bound chloride and free chloride was much more complicated, and could be non-linear [Blunk et al (1986), Byfors (1990), Sandberg et al (1993a), Akita et al (1995)]. Pereira et al (1984) suggested a Langmuir isotherm to describe the chloride binding behaviour. However, the Langmuir isotherm equation could only give a good expression for the chloride binding behaviour at low concentration. In order to fit the experimental data at intermediate concentration ranges, a Freundlich adsorption isotherm [Tang (1993)] and a modified BET [Xu (1990)] model were proposed. The relationship between free chloride and bound chloride is affected by binder type, degree of hydration, amount of pore solution, and other ions in the pore solution. Many researchers [Arya et al (1990), Tritthart (1989), Tritthart (1989a), Page et al (1986a)] have proposed a correlation between the tricalcium aluminate (C3A) content and the capacity of binding chloride ions through the formation of insoluble calcium chloroaluminates. Higher C3A contents bind more chloride ions, resulting in lower chloride ion level in the pore solution. Tang et al (1992) reported that the total amount of alumina and iron oxide in cement determined the chemical binding capacity; fly ash/cement binder contained higher amounts of alumina and iron oxide, so had a higher chemical binding capacity; slag cement formed finer hydrated products, and had higher physical binding capacity too. This is in addition to their beneficial effects on concrete microstructure and permeability. The hydroxide concentration in pore solution has a significant influence on chloride binding [Tuutti (1982), Tritthart (1989a), Byfors (1990), Page et al (1991)]. The higher the hydroxide concentration, the less chloride will be found in the pore solution, probably as a result of change balance. Temperature can also alter the chloride binding capacity. The amount of bound chloride decreased as temperature increased [Larsson (1995)]. Both physical and chemical desorption reaction tend to increase at higher temperature, decreasing the chloride binding capacity. This could also be because hydrated cement is more crystalline at higher temperature, with reduced adsorption capacity.

26

Service life prediction of reinforced concrete structures

Some other factors such as curing temperature, curing age, original alkalinity also affect the chloride binding capacity of concrete [Arya et al (1995), Byfors et al (1986)]. It has been suggested that the presence of superplasticiser in concrete could lower the chloride binding capacity [Haque et al (1995)]. Larsen (1997) found that the chloride uptake and the pore solution composition in concrete were affected by wetting/drying and temperature, but the effects were influenced by chloride concentration. Carbonation strongly decreased the chloride uptake and reduced the chloride concentration of concrete. Page et al (1991) pointed out that cement hydrates could bind a substantial portion of chloride in an insoluble form below a total chloride content of 1%, and that beyond this point the binding capacity is largely exhausted. The concentration of sulfate ions has also been reported to significantly influence the chloride binding capacity of a given binder [Yonezawa (1988), Tritthart (1989a), Byfors (1990), Sandberg et al (1993)]. Xu (1997) reported that calcium sulphate and sodium sulphate had different effects on chloride binding and pore solution chemistry; at the same sulphate content, cement pastes containing calcium sulphate had a higher chloride binding capacity than those containing sodium sulphate; this might be due to their different effects on OH- ion concentration in the pore solution; sodium sulphate increased the alkalinity of pore solution whereas calcium sulphate decreased it; so sodium sulphate had little effect on the ratio of Cl-/OH- while calcium sulphate significantly increased the ratio.

5.2 Mathematical models


A number of models have been put forward to describe the ingress and transport of chloride [Helffereich et al (1970), Frantz et al (1976), Crank (1979)] in concrete. However, the basis for those models is still the diffusion equation, based on which some other mechanisms and influencing factors are taken into consideration.

5.2.1 Simple diffusion model


Provided that the following assumptions hold: I) II) III) IV) V) VI) Diffusion of chloride is the principal mechanism dominating the chloride transport process in concrete, no other mechanisms are taken into account. The concrete matrix through which the diffusion occurs is uniform in microstructure or properties; The microstructure and properties of the concrete do not change significantly with time; and The microstructure and properties of the concrete do not change significantly with the concentration of the diffusing chloride. The concentration (Cs,Cl) of chloride on the concrete surface is constant; The initial concentration (CI,Cl) of chloride throughout the concrete is uniform,

27

Service life prediction of reinforced concrete structures

then the general expression for the diffusion of Chloride in concrete can be expressed by:

(C Cl ( 2C Cl % DCl (t (x 2

(20)

where CCl is concentration of free chloride in concrete, and DCl is the diffusion coefficient of free chloride. So the concentration of chloride at depth x and at time t can be expressed as:

CCl = Cs,Cl - (Cs,Cl Ci,Cl) erf [x/(4DCl t)1/2]

(21)

where Cs,Cl is concentration of the free chloride in the surface layer of cover concrete; Ci,Cl is the background concentration of the free chloride (which is equal to the concentration of free chloride in the concrete interior, unaffected by the ingress of chloride). The above solution has a typical concentration distribution of free chloride as illustrated in Figure 8. Correspondingly, the change of free chloride concentration with time at a given depth in concrete can be schematically presented by Figure 9. They are quite similar to those of the carbonation profile and carbon-dioxide change with time in concrete (Figure 2 and Figure 3).

1.20E+00 CCl , concentration of Cl at time t=tx Cs,Cl

Ci,Cl xt

0.00E+00 0

60 x, depth from concrete surface

Figure 8. Distribution of the concentration of free chloride ions in concrete at a given time

28

Service life prediction of reinforced concrete structures

1.2 CCl , concentration of Cl- at depth x=xt Cs,Cl

Ci,Cl 0 0

tx

t, chloride ingress time

3000000000

Figure 9. Change of the concentration of free chloride ions with time at a give depth in concrete Based on equation (21), the ingress of chloride ions in concrete with time can also be formulated by:

XCl = A0,Cl t1/2

(22)

where XCl is the depth of a selected concentration of chloride ions, A0,Cl is a constant which is equal to x0 /t01/2, and xo and t0 are the initial position of the selected chloride concentration and the corresponding initial time; as the chloride ions always ingress deeper into concrete with time, so t>t0, X>x0. Equation (22) is also similar to equation (12), and can give a similar progress diagram (Figure 10).

18
concentration (C0,Cl ) of ClXCl , depth of a selected

x0

0 0

t0

t, chloride ingress time

1500000

Figure 10. Ingress of a given chloride concentration in concrete with time

29

Service life prediction of reinforced concrete structures

Equations (20), (21) and (22) only describe the one dimensional diffusion process of chloride in concrete. The diffusion equation and the solution for the two-dimension model are much more complicated [Berke et al(1997)]. Theoretically, the above equations can only describe the diffusion of free chloride in concrete, rather than the distribution of bound chloride, as free chloride ions are travelling through the concrete [Sagues et al (1996)]. To deal with the total chloride distribution in concrete, the equation should be further modified and the meanings of some parameters involved could need redefining, leading to the following diffusionbinding model.

5.2.2 Diffusion-binding model


In the above simple diffusion model, the ingress and transport of chloride in concrete is ascribed to the diffusion of free chloride alone. However, in reality the binding of chloride in concrete is unavoidable, which would consume the free chloride ions during its diffusion. So, the diffusion equation for the free chloride should be expressed as:

(C Cl ( 2C Cl % DCl # f Cl (t (x 2

(23)

where fCl is a function dependent on the concentration of free chloride ions dependent function, representing the rate for the binding and consumption of free chloride ions. This equation can be converted into:
(CCl ( 2CCl % De,Cl (t (x 2

(24)

where, De,Cl is the effective diffusion coefficient of the free chloride ions in concrete, which has combined the binding effects into the diffusion process:

De ,Cl %

DCl (f 1$ Cl (C Cl

(25)

In the case of the binding rate being assumed to be linearly dependent on the concentration of free chloride ions, i.e., kCl=constant=(fCl/(CCl, the effective diffusion coefficient would become a constant:

De ,Cl %

DCl % consta 1$ k Cl

(26)

30

Service life prediction of reinforced concrete structures

Under this condition, the diffusion-binding model (24) has the same diffusion equation format as the simple diffusion model, except for the different physical meanings of their diffusion coefficients. In some cases, if a linear relationship is assumed between the concentrations of bound (Cb) and free chloride ions:

Cb,Cl = . CCl + /;

. and / are constants

(27)

then the concentration (Ct,Cl) of total chloride ions in concrete would be:

Ct,Cl = (1+.) CCl +/

(28)

and equation (24) could be transformed into:

(C t ,Cl (t

% De ,Cl

( 2C t ,Cl (x 2
(29)

This means that the total chloride distribution in concrete could also obey the diffusion equation, provided that linear equilibrium is established between the bonded and free chloride ions [Sagues et al (1996), Fishcher et al (1984)]. Therefore, the diffusion equation (with an effective diffusion coefficient) sometimes can be directly used to describe the total chloride profile in concrete, as demonstrated by Nilsson (1992), Tang et al (1995a), and Xu et al (1995a). As it is more convenient to measure a total chloride profile, rather than a free chloride profile, Eq (29) is very useful in practice. Therefore, equation (29) is more practical, as long as its precondition of linear relation between bound and free chloride can hold. It should be stressed that equation (24) has the same solutions as equations (21), and (22) except that the DCl is replaced by De,Cl. However, due to the difference between the values of De,Cl and DCl, the depth profiles of free chloride and total chloride could be different. Also, the biding effect can change the chloride profile in concrete. For example, if the binding capacity is high, i.e. De,Cl is small, then the profile would be steep.

31

Service life prediction of reinforced concrete structures

5.2.3 Multi-transport-mechanism models


The above models about ingress and transport of chloride are limited to a stationary pore solution system. In reality, concrete structures may be exposed in different environments with a gradient of water pressure or vapour pressure. Therefore, the ingress and transport of chloride in concrete could occur by mechanisms other than diffusion, such as hydraulic flow of chloride solution due to a gradient of water pressure [Edwardsen (1995)], capillary suction of chloride-bearing solution in an unsaturated pore system due to the surface tension of pore walls [Collepardi et al (1989), Akita et al (1995) ], convection of chloride solution due to wick action [Buenfeld et al (1995)], moisture flow and evaporation due to a gradient of vapour pressure [Tuutti (1982), Ohama et al (1995)], etc. However, all these mechanisms operating in the field are too complicated to be formulated by mathematical equations. It is difficult or even impossible to quantify the influences of these combined processes on chloride ingress and transport [Nilsson et al (1996)]. In addition to the mechanisms, migration under electric field may be another important process for free chloride ions that carry negative charges in the pore solution. The migration effect is particularly significant for the reinforced concrete structures that are subjected to stray current interference, galvanic corrosion effect, or are under cathodic protection in the field. Furthermore, the non-uniform distribution of ions in concrete due to the heterogeneity of concrete could build up statistic charge, which generates electric fields in the concrete, and affects the transport of negatively charged chloride ions. If the migration is combined in the transport model of free chloride ions, then the model can be mathematically expressed as:

( 2C (CCl FE (CCl ) % De,Cl ( 2Cl # (t RT (x (x

(30)

where F is Faraday constant, E electric field strength in concrete, R gas constant, and T absolute temperature. Equation (30) has an analytical solution [Tang et al (1992a, 1993, 1996a)]:

C Cl %

C s ,Cl

x $ 0De ,Cl t x # 0De ,Cl t [ )$ erfc( )] e 0x erfc( 2 2 De,Cl t 2 De ,Cl t

(31)

where 0=FE/(RT). This solution gives a steeper chloride profile than the simple diffusion and diffusion-binding models. It should be stressed that the above solution is theoretically obtained under the condition that De,Cl is a constant independent of t, x, and CCl. In practice, the microstructure of concrete, its chloride binding capacity, the influence of other ions on chloride mobility, etc, can all affect the chloride profile. Therefore, the measured chloride profile may not be as sharp as that predicted by equation (31) [Tang et al (1995)].

32

Service life prediction of reinforced concrete structures

5.2.4 Complicated models


In the laboratory, the above diffusion based models for the ingress and transport of chloride may be considered valid, as diffusion, permeability, and absorption of chloride can be measured under controlled conditions. However, even under strictly controlled conditions, the measured data could still deviate significantly from theoretical predictions due to some unexpected factors not incorporated into the predictive models. The application of equation (21) in analysing chloride profiles in concretes sampled from structures has been shown to be only partly successful. Swamy et al (1994) summarised the results obtained from more than 1000 sets of data on chloride content in concretes exposed to various marine conditions, including submerge, tidal, splash, atmosphere, etc. They showed that the estimated CS,Cl and De,Cl changed with time, and varied with depth and exposure conditions. Specifically, the Cl- diffusion coefficient of inner concrete (larger than 30 - 50 mm) was found to be quite different from that of the surface concrete. Weyers et al (1994) showed that the increase in chloride content in the concrete surface (0-0.5 in) approached that estimated by the square root of time function, however it was nearly constant after about 5 years. The above indicates that the assumptions based on which the diffusion based models are established are not reasonable in some cases. For example, in practice, there are at least two basic facts that violate the above assumptions: 1) 2) the surface concentration of chloride is not constant and varies with season and environmental factors (temperature and wetness, etc); and the diffusion coefficient is not constant and varies with age and depth in concrete.

The variable surface concentration of chloride is the first main source of error in the prediction model. The change of environmental temperature and the wetting/drying of cover concrete can accelerate chloride uptake or wash away the accumulated chloride on the concrete surface, making the surface concentration change with time. Sometimes due to some reported condensation effects, the surface concentration of chloride in a saturated concrete can reach a level which is two to four times higher than that of the surrounding saline solution, even though diffusion is still the principal transport process in this case [Nagataki et al (1993)]. For example, the surface chloride concentration may increase in concrete in the tidal and splash zones of bridges and areas where de-icing salt is used in winter and where drying of the surface takes place, whereas rain may cause it to leach out [Saetta et al (1993), Fluge (1997), Andersen et al (1997)]. The variation in CS can also be caused by the chemical reaction of cement paste with the environment, e.g. carbonation and leaching [Larsen (1997)] can reduce the capacity of cement paste to adsorb chloride ions. The diffusion coefficient of chloride can be greatly enhanced by sulfate attack [Jones et al (1994)]. Because of these influences, in most practical cases, the chloride concentration profiles are anomalous in the surface layer [West et al (1985a)].

33

Service life prediction of reinforced concrete structures

The most common practice to deal with the variation of surface concentration is to discard the first few millimeters of the concrete surface and take the concentration of chloride at about 10 mm depth as a constant pseudo surface concentration [Broomfield (1997)]. It is expected that the concentration of chloride at this depth would be constant. Obviously this is an approximation, and could introduce significant errors. Establishment of a mathematical relationship between the surface concentration of chloride and time is another attempt to cope with the difficulty arising from the nonconstant surface concentration. Uji et al (1990) reported that the surface concentration of chloride was a function of the square root of time. Such an expression was regarded as the best estimate for surface concentration of chloride in practical application [Uji et al (1990), Purvis et al (1994), Amey et al (1996)]. Lin (1990) proposed another expression for the surface concentration of chloride:

Cs,t = u [1 - exp(-v t)]

(32)

where u and v are constants. With this exponential equation to describe the time dependent surface concentration of chloride, the diffusion equation can be resolved and an analytical solution obtained. It should be borne in mind that, the practical change of surface concentration of chloride can not follow one given function of time [Swamy et al (1994)]. Even though the above mathematical treatments can result in a better correlation between measured and analysed curves, particular caution should still be taken when drawing conclusions based on such analyses, because the physical meaning of the time dependent function, for the change of surface concentration of chloride with time, is not clear. The second source of error in the use of the diffusion based equations is the variable diffusion coefficient. This can change with temperature, time, the depth of concrete and from site to site. Therefore, the ingress of chloride can be strongly influenced by many factors like environmental conditions, the microstructure and defects of concrete, reaction of concrete constituents with chloride, heterogeneity of concrete, and surface charge on the hydrated cement phases; all of which depend on time, temperature, and thickness of concrete. For example, it was reported that the diffusion coefficient apparently decreased with time [Mangat et al (1994), Johansen et al (1995)]. In addition, coarse aggregate could effectively block chloride diffusion [Castro et al (1993)]; and the reaction of chloride with the concrete constituents could delay the penetration of chloride into deep concrete [Collepardi et al (1972), Goto et al (1981), McCarter et al (1992)]. Therefore, it was found [Bentz et al (1996)] that solution (21) could result in significant error in the calculation of the diffusion coefficient. In order to deal with the difficulty caused by the variable diffusion coefficient, various attempts have been made by many researchers using more flexible diffusion equations [Collepardi et al (1972), Uji et al (1990), Purvis et al (1992) and Poulsen (1997)]. Maage et al (1993,1996,1997) developed a time-dependent diffusion coefficient Dt:

34

Service life prediction of reinforced concrete structures

t 1 Dt % D0,t ( 0 ) t

(33)

where 1 is a parameter varying from 0.4 to 0.8 (determined on laboratory concrete specimens depending on w/c, curing time and the interactions between concrete and the exposure environment, t0 is the age at which the exposure to chloride starts, and D0.t and Dt are the apparent diffusion coefficients at the start and at time t respectively. Lin (1990) also proposed a time dependent diffusion coefficient, but the coefficient was a function of cement type and w/c ratio. The influence of w/c ratio on the diffusion coefficient, which is highly time dependent, is also represented by the model proposed by Maage et al (1996, 1997):

d 1#1 ) )2 /( ti % tC ( 2 t C Dt

(34)

2%

erfc( ) C S ,Cl # C i ,Cl

2 C th # C i ,Cl

(35)

where tC is the time when the structure (not yet damaged) is inspected, Cth is the chloride threshold level for initiation of corrosion of reinforcement, and d is the thickness of concrete cover, and the Dt has the same meaning as equation (33). Considering that both surface chloride content CS.Cl and diffusion coefficient De,Cl change with time, Mejlbro (1996) and Poulsen, (1997) proposed a four parameter model to estimate the distribution of chloride in concrete: x 3 p( C(x, t )% Ci $ ( CS # Ci ) ) 2 4 t ex Daex where (36)

CS % Ci $ S p4 p

(37)

D 5 t ex p S p % C1[ ( 1 ) ] D100 1# t ex

(38)

35

Service life prediction of reinforced concrete structures

3 p (z )%

n 2n 2n $1 7 (p $ 1) 2z ) 2z ) p (n )( (p # 05 .) ( # 7 (p $ 05 2n ) 2n $ 1) ( ! .)n % 0 ( ! n%0

$-

$-

(39)

p(n)=p(p-1)(p-2) 8(p-n+1)
(40)

p%

C100 /C1) l og( 100 # t ex D100 l og( ) 1 # t ex D1

(41)

t 0 t 1#0 4 %( ) # ( ex ) t ex t

(42)

5%

1 1 l og( ) 2 t ex

(43)

t 0 Da % Daex ( ex ) t

(44)

where subscript ex denotes the age (in years) at first exposure of the concrete to chloride, and subscripts 1 and 100 denote 1 and 100-year respectively. Eq(44) was derived on the basis of 2-year laboratory experiments, which might not be applicable for long-term predictions. In addition, some other mathematical expressions have also been used [Tang (1996)] to describe the decrease of diffusivity with age of concrete. It should be noted that the surface chloride concentration and the diffusion of chloride ions in the concrete are correlated with each other. For the same concentration distribution or the same amount of chloride penetrating into the concrete, a larger CS,Cl corresponds to a smaller De,Cl, and vice-versa. This fact throws doubt on the conclusions that the chloride content in the surface concrete increases with time while the diffusion coefficient decreases. In fact, in a thin surface layer of concrete, the absorption of chloride can reach its equilibrium state very quickly [Tang et al (1992b), Steen (1997)], and it is unlikely that it could keep increasing significantly for decades, provided that the chloride concentration in the surrounding environment has a stable average level in the long term.

36

Service life prediction of reinforced concrete structures

Furthermore, the diffusion coefficient is theoretically temperature dependent. Page et al (1981) used a temperature dependent diffusion coefficient to study the influence of temperature on the ingress of chloride by considering the contribution of activation energy to chloride transport. Amey et al (1998) proposed to use a semi-empirical formula to estimate the effect for concrete:

T g (T1 # T2 ) DT 2 % DT 1 ( 2 ) e T1

(45)

where T1 and T2 refer to different temperature, and DT1 and DT2 are the diffusion coefficients at temperature T1 and T2. g is a constant depending on w/c. When an analytical solution is hard to obtain after introducing time, space, concentration dependent parameters into the diffusion based models, a numerical calculation can be considered. The simplest method in this case might be to numerically solve the finite difference equation instead of differential equation [Tang (1996)]:

dC Cl C Cl (x $ 9x)# C Cl (x) C Cl ,i $1 # C Cl ,i : % dx 9x 9x

(46)

dC Cl C Cl ,j $1 # C Cl ,j : dt 9t C Cl ,i ,j $1 # C Cl ,i ,j 9t C Cl ,i $1,j $ C Cl ,i #1,j # 2C Cl ,i ,j 9x 2

(47)

% DCl ,i ,j

(48)

where CCl,i,j represents concentration of free chloride ions at depth xi and time tj; the subscripts i and i+1 indicate two adjacent positions, and 9x is a small increment in distance; the subscripts j+1 and j denote two adjacent times. Tang (1996) demonstrated the application of the finite difference equation to numerically solve the problem of chloride diffusion into concrete, with the following special mathematical treatments: !" Diffusion coefficient is assumed to be a function depending on temperature, time and space :

DCl ,i ,j % DCl ,0

T0

; ft ( T j ); f t ( t j ); f x (x i )

(49)

!" Concentration (CC;,0,j) of chloride in water at the concrete surface (x=0) is assumed to periodically (yearly) fluctuates with time:

37

Service life prediction of reinforced concrete structures

C Cl ,0,j %

C Cl ,0 max $ C Cl ,0 min 2

C Cl ,0 max # C Cl ,0 min 2

sin( 2,

c t j # t0

365

(50)

!" Temperature variation is treated in the same way. !" Bound chloride is assumed to be dependent on various factors:
A > AE > OH ] [ ? < fb 5 i,j # 1 ? b ( 1 # 1 )<W % # Qb, Ci,j 0 exp ( 1 ) exp ? < Cl , i,j OH i j , ? < [ ] 1000 OH R T T ? ? initial ,j # 1 < j 0 < @ = @ =

(51)

where 0OH=0.59 is a constant for the effect of OH- on chloride binding, W is the content of hydration products in concrete, Eb is the activation energy for binding, fb is a binding constant, and 5 is an empirical coefficient of Freundlich adsorption isotherm. The prediction of chloride distribution in concrete by this approach was claimed to be very successful [Tang (1996)]. Simulation by finite-difference method can also be made by using an implicit method proposed by [Crank (1975)]: Ci,j $1 # Ci,j 9t % D B Ci#1,j # 2Ci,j $ Ci #1,j Ci $1,j $1 # 2Ci,j $1 $ Ci #1,j $1 E $ C F 2 2 2D 9x ) 9x ) ( ( G

(52)

which has been applied in simulating chloride distribution in bridge decks by Hoffman et al (1994).

5.2.5 Empirical models


As presented earlier, the analytical solutions to the chloride ingress and transport models are usually very complicated, and sometimes the analytical solutions could not be obtained in mathematics, if all the realistic influences are combined. Therefore, such models are sometimes impractical in the prediction of the initiation time of chloride induced corrosion. In these cases, empirical models may be able to simplify the prediction equations and be more practical in use. Some empirical models have been used to predict service life, based on laboratory and field experience [Maage et al (1997), Steen (1997)]. In recent work [Takewake et al (1988), Maage et al (1993, 1995), Uji et al (1990), Swamy et al (1994), Poulsen (1996), Frederiksen et al (1997), Mejlbro (1996)] concerning models of chloride transport in concrete, besides particular attention being paid to the time-dependent surface concentration and diffusion coefficient, more variables have also been introduced into the models to make them more flexible [Poulsen (1997), Steen (1997)].

38

Service life prediction of reinforced concrete structures

Among the empirical predictions, modification of diffusion-based models could be the simplest method to fit measured chloride penetration data. For example, the depth of the ingress of chloride (XCl) can also be determined in another form proposed by Cady et al (1983):

XCl = K1 t1/2 + K2
where K1 and K2 are constants.

(53)

Besides the diffusion-based expressions, various empirical formulae have been proposed to estimate the remaining life of reinforced concrete structures [Beaton et al (1963), Clear (1976), Cady et al (1983)]. For example, capillary effect usually has complicated influences on the transport and ingress of chloride, especially for the concrete exposed to wetting - drying conditions where this effect can be significant. In this case, pure theoretical modelling would be difficult, and some reasonable empirical modification based on well known phenomena could be acceptable and useful. It has been found that chloride diffusion lagged behind the advancing absorbed water front because of the binding of chloride by cement paste. Lunk et al (1998) used a square root function of time with empirical modifications to describe such a process:

X Cl %

H 1. 25 2 2 1 rt $( r02 # r3k ) ( 1# e #7t ) ' 2Ir0 7

(54)

where XCl is the depth of chloride penetration, H is the surface tension and I is the viscosity of water, 3k is the residual water capacity (unfilled volume of pores at the start of the exposure), r0 is the effective pore radius at time t = 0 (start of exposure), ris the effective pore radius at time t = -, 7 is the deviation coefficient (to express the deviation of the penetration from the Jt - Law), and t is the elapsed time. This model was used in an estimation of the minimum cover thickness required for a retaining wall after rehabilitation, which was to be exposed to de-icing salt at its lower part in winter, and to chloride-bearing spry and mist at the upper part. In the calculation, the service life was taken as the cumulative exposure time for which the structure was subjected to occasional chloride attack; and the necessary thickness was estimated as twice the depth of the chloride front at the assumed service time.

5.3 Threshold concentration of chloride


As discussed earlier, chloride ions generally penetrate into concrete interiors following a square root function of time (equations (22) and (53)), and at a certain depth, the concentration of chloride gradually increases with time. Therefore, chloride ions will sooner or later arrive at the vicinity of reinforcement in concrete, and the concentration of chloride at the surface of reinforcement will gradually increase to a certain threshold level, at which corrosion will be initiated. Such a level of chloride ions in concrete is termed the chloride threshold, above which severe corrosion of steel in concrete could be induced.

39

Service life prediction of reinforced concrete structures

A very low concentration of chloride ions in the pore solution cannot break down the passive film on the reinforcement and initiate corrosion of the reinforcement. However, the corrosion risk increases with the increasing content of chloride in the vicinity of the reinforcement, and after a certain level of chloride is reached at the surface of the reinforcement, the corrosion of reinforcement would become possible. In other words, the chloride threshold correlates the chloride contamination in concrete with the corrosion of reinforcement, which lays a practical base for prediction of the initiation of chloride-induced corrosion, using chloride profiles. Table 1. Chloride content thresholds reported in the literature
chloride concentration 0.1% total chloride by mass of dry concrete 0.15% water soluble chloride ion by mass of cement 0.17~2.5% of chloride by mass of cement 0.2% acid-soluble chloride by weight of cement 0.4% total chloride by weight of cement 0.4~0.8% total chloride by weight of cement Cl-/OH- ratio 0.29~0.3 0.6 reference Stoltzner et al (1997) Holden et al (1983), Popovics et al (1983) Glass et al (1995) Clear et al (1973) Broomfield (1997), Building Research Establishment (1982) Locke et al (1980) reference Diamond (1986) Hausmann (1967)

Table 2. Chloride limits recommended by different codes (% by cement mass, otherwise specified)
Prestressed concrete 0.06 % water soluble Cl- by cement mass 0.06 % water soluble Cl- by cement mass 0.08 % Acid soluble Cl- by cement mass 0.8 kg Acid soluble Cl-/m3 concrete 0.1 % Acid soluble Cl- by cement mass 0.1 % Acid soluble Cl- by cement mass Reinforced concrete Exposed to Cl 0.15% water soluble Cl- by cement mass 0.1% water soluble Cl- by cement mass Reinforced Dry or protected from moisture 1.0% water soluble Cl- by cement mass Reinforced Other construction 0.3% water soluble Cl- by cement mass 0.15% water soluble Cl- by cement mass Code

ACI 318-89

ACI 201 Committee ACI 222 Committee

0.2% Acid soluble Cl- by cement mass 0.8 kg Acid soluble Cl-/m3 concrete 0.4 Acid soluble Cl- by cement mass 0.4% Acid soluble Cl- by cement mass

3.0 kg Acid soluble Cl-/m3 concrete

Australia AS 3600-88 Norway NS 3420 UK BS 8110-85

40

Service life prediction of reinforced concrete structures

Table 1. lists some values of the chloride threshold reported in the literature. Various thresholds are also defined in a number of codes (see Table 2). The commonly accepted threshold value is usually in the range of 0.2~0.4% chloride ion by mass of cement. The above values of chloride threshold cannot absolutely be fixed, because the pH value, binding capacity of cement, moisture and oxygen contents, etc. can affect the threshold [Broomfield (1997), Tritthart (1989), Blankvoll (1997), Pettersson (1993), Thomas (1996), Hussain et al (1995), Dhir et al (1994)]. As far as corrosion of reinforcement is concerned, the position or depth (Xth) of the threshold level of chloride in concrete, is the main concern. Since Xth is a specific depth corresponding to the chloride threshold on the chloride profile curve, it also follows a square root function similar to equations (22) and (53). The accuracy of the threshold level determination can significantly affect the accuracy of the predicted service life. A higher value of threshold would lead to a longer predicted service life; and vice-versa, provided the other conditions are the same. Therefore, the threshold value should be carefully selected.

5.4 Prediction techniques


The prediction of initiation time of chloride-induced corrosion normally consists of three procedures: 1) obtaining of the chloride profile; 2) selection of prediction models; and 3) estimation of relevant parameters involved in the selected models.

5.4.1 Chloride profile


Chloride profile is the foundation for the prediction of service life. It is generally obtained through sampling of concrete at different depths by cutting successive slices from drilled cores or drilling successively from the concrete surface, then analysing the amount of the chloride ions in the specimens. There are many techniques used for the chloride analysis in concrete specimens, including neutron-gamma ray [Rhodes (1977), Rhodes(1980), Livingston (1986)]; X-ray fluorescence [Provebio et al (1997)]; valhard titration, quantab chloride titrator strips, and argentometric digital titration [Building Research Station (1977)]; specific ion probes [James Instruments Inc.(1990)], potentiometric titration [AASHTO T260-84 (1984)], expression and extraction techniques [Arya et al (1990a), Byfors et al (1986), Duchesne et al (1994), AASHTO T260-84 (1984), ASTM D1411 (1982), Larsen (1997), Xu (1997), Hussain et al (1994)], leaching (or water extraction) [Arya et al (1990a)]. etc. Detailed discussion on each approach is beyond the scope of this report. However, it is very important to understand that the profile obtained would be either a free chloride or total chloride profile, and the choice would determine the selection of prediction models.

41

Service life prediction of reinforced concrete structures

5.4.2 Selection of models


The simplest method [Broomfield (1997)] to calculate the initiation time for t chlorideinduced corrosion is directly based on the chloride profile which is simply regarded as a parabolic curve in this case. By fitting the profile to a parabolic curve, the chloride threshold depth can be determined. The advance rate of the chloride threshold through the concrete and the time for the threshold to reach the surface of steel reinforcement can be predicted using:

Xd %

x 0,th t 0,th

(55)

where Xd is the depth of steel reinforcement in concrete, x0,th is current threshold depth, t0,th is present time reference to the beginning of chloride ingress, t is predicted time when current chloride threshold will reach the steel reinforcement. In other cases when the chloride profile does not follow the above equation, the selection of prediction models is very difficult, because too many causes can lead to deviation of the chloride profile from the theoretical prediction based on Ficks second diffusion law. In the case of a non-typical profile, the analysis of the causes would be critical to the model selection. For example, if the concrete is submerged, the environmental factors are relatively stable except the seasonally changing temperature, so the complicated prediction model with a diffusion coefficient varying with time could be the base for prediction. If a structure is subjected to sea water splash, then not only the diffusion, but also other transport mechanisms, like hydraulic flow or convection etc., should be considered. In these cases, both the diffusion coefficient and the surface chloride concentration may need to be expressed as variables, rather than as constants. If the structure is under cathodic protection or under significant stray current influence, the modified diffusion model incorporating the electric field parameters (30) needs to be preferentially considered.

5.4.3 Relevant parameters


The relevant parameters involved in diffusion based prediction models basically refer to diffusion coefficient (De,Cl), surface concentration (Cs,Cl) and background concentration (Ci,Cl) of chloride ions. As to the complicated models, the relevant parameters still include those characterising the concrete and the environment. Basically, there are two ways to estimate these parameters: curve-fitting the chloride profile in the field concretes; and 2) measuring through specially designed laboratory experiments.

42

Service life prediction of reinforced concrete structures

There are several methods to estimate diffusion coefficient from measured chloride profile. For example, it can be conducted [West(1985)] as follows: first calculate a concentration ratio K from a chloride profile:

K = (Cs,Cl CCl ) / (Cs,Cl-Ci,Cl ) = erf [x (4De,Cl t )-1/2]

(56)

then look up the error function table, from which x (4De,Cl t)-1/2 can be determined, hence De,Cl calculated; finally based on the calculated De,Cl and measured chloride profile CCl ~ x, the rate of advance of the chloride can be predicted, i.e. the time taken for the chloride concentration to exceed the threshold value at the reinforcement surface can be determined. In fact, under certain conditions, equation (21) can [Poulsen (1990)] into: also be linearly approximated (57)

CCl = Ci,Cl + (Cs,Cl Ci,Cl) [1-x/(12De,Cl t)1/2]2

therefore, the [(CCl-Ci,Cl)/(Cs,Cl-Ci,Cl)]1/2 is linearly dependent on the depth x where CCl is measured. By plotting such a straight line, De,Cl can be obtained from its slope. Bentz et al (1996) have developed a numerical model to determine chloride diffusion coefficient using a direct finite difference implementation of Ficks second law. In fact, with the development of computerisation techniques, curve-fitting is becoming easier, simpler and more accurate, and it will become the main technique of obtaining the relevant parameters from available chloride profiles. Typically, chloride profiles are relevant long-term exposure of field structures, so methods based on chloride profile are invalid for a newly constructed structure. To predict the initiation time of chloride induced corrosion in new concrete, specially designed testing is required. In earlier research work the diffusion coefficient was determined by diffusion cell experiments [Ushiyama et al (1974), Page et al (1981)], which basically involves the following procedures: 1) a thin slice of specimen is used to separate two cells, one of which contains chloride solution and the other does not; 2) the increase in chloride ions in the originally chloride free cell is monitored; 3) the diffusion coefficient is calculated according to Ficks first law when a steady flow of the chloride ions is obtained. Similar cells have also been used by various researchers in their studies [Roy et al (1986), Hansson et al (1985)]. Based on the diffusion cell, quicker, non-steady state methods have also been employed to estimate the relevant parameters, such as coulomb test [Whiting (1981), AASHTO method (Designation T 277-83) (1983)], potential diffusion index [Dhir et al (1990a)], and embedded electrodes[Rechberger (1985), Hansen et al (1986)]. However, the validity of their results is still uncertain.

43

Service life prediction of reinforced concrete structures

Generally speaking, the chloride diffusion coefficient De,Cl for a concrete is about 2~3 x 10-8 cm2/s [Berke et al (1994, 1997)], and the typical surface chloride concentration Cs,Cl is about 18 kg/m3 in the splash/tidal zone in a severe marine environment [Browne (1982), Bamforth (1996), Berke et al (1994), Hartt (1990)]. However, various diffusion coefficients and surface concentrations have been widely reported and summarised [Bamforth (1996), Stoltzner et al (1997), Sagues et al (1997)]. Different measurements are likely to lead to different results. By means of curve fitting, the maximum CS,Cl was reported to be 0.4% by mass of concrete for samples obtained at a testing site on the Swedish west coast [Poulsen (1997)], 0.4-0.6% for concrete structures on the south coast of England [Bamforth et al (1993)], 4% by mass of cement (approximately 0.4-0.6% by mass of concrete) on the Japanese coast [Swamy et al (1994)], 0.47% by mass of concrete for 30 years old jetties in Tokyo Bay [Fukute et al (1998)]. The value of De,Cl is around 0.1 - 1 L 10-11 m2/s [Dhir et al, (1991), Swamy et al (1994), Fukute et al (1998), Maage et al (1997)] for concretes with water to cement ratio 0.50-0.60; and 0.2-1 L 10-12 m2/s for lightweight concretes incorporating silica fume. For results from diffusion cells, the reported diffusion coefficients are typically 0.1 - 10 L 10-12 m2/s for pastes of w/c 0.3-0.6 [Page et al (1981), Li et al (1986)].

44

Service life prediction of reinforced concrete structures

6. Corrosion propagation of reinforcing steel


The ingress of chloride and carbonation of concrete can both induce corrosion of reinforcement, and finally lead to damage to reinforced concrete structures. After corrosion of reinforcement is initiated due to the ingress of chloride ions or carbonation of concrete, in the initiation stage, the deterioration will get into the second stage, the stable propagation stage, at which the corrosion development of reinforcement becomes the main issue. Compared with the first stage, relatively little research has been conducted on the second stage to deal with the service life of concrete structures. The reason might possibly be that, once a reinforced concrete structure gets into the second stage, the corrosion of steel is expected to proceed at a high rate. This means that the risk ofdamage to the structure is much higher than that in the first stage. Therefore, repairing the damage and eliminating the potential risk have become more important and urgent than the prediction of service life. Another reason might be the uncertainty of the second stage. In most cases, it is hard to tell whether the corrosion rate of reinforcement in concrete would always increase, decrease or keep constant with time, as it depends on many factors. So, the prediction of the second stage is even more difficult than that of the first stage. However, from the corrosion point of view, the second stage may be more critical, as it is directly responsible for the final damage of a reinforced concrete structure.

6.1 Corrosion processes and mechanisms


Normally, the pore solution is rich in oxygen with a high pH value. So the passive film on the steel surface is stable, which consequently retards corrosion reactions of reinforcement in concrete. In this case, the steel can be well protected in concrete, and there will be no detectable corrosion damage. However, under the conditions that chloride content is above the chloride threshold level due to the ingress of chloride ions or the pH value of the pore solution is below the pH threshold due to carbonation, the protective passive film on reinforcement would not be stable and would break down. In other words, as soon as the concrete structure gets into its second stage of deterioration, corrosion of reinforcement could be initiated and severe corrosion damage could become likely. After the corrosion of reinforcement is initiated, there could be two basic forms of corrosion deterioration: 1) The immediate corrosion product (dissolved Fe2+ ) could move away from the steel surface. Hence, the cross-section of reinforcement could keep decreasing and finally the breakdown of reinforcement could result, which will further result in damage of the whole reinforced concrete structure.

45

Service life prediction of reinforced concrete structures

2)

The corrosion of steel could proceed at a high rate, but the final corrosion products could not move out of the corrosion site, but accumulate on the steel surface as a rust layer. This could generate expansion force and cause delamination, cracking and spalling of the cover concrete. Consequently, final damage of the whole reinforced concrete structure could eventuate.

The first form of deterioration can be facilitated when the concrete is saturated with water. Sufficient water in concrete is conducive to the anodic dissolution of reinforcement but restrictive to access of oxygen which is needed for further reaction with the dissolved iron to form less soluble products. Thus, the removal of the immediate corrosion product away from the steel surface would be relatively easy, and there would be a low possibility for rust to form and accumulate at the interface between the reinforcement and the concrete. In this case, dissolved corrosion product could move out of the concrete and form rust stains, on the surface of the concrete structure. Such a corrosion deterioration has no significant impact on the surrounding cover concrete, and the attack can easily be detected by the visible rust stain on the structure surface. The most adverse case is the second deterioration form. It mostly occurs when the pore solution is rich in oxygen and aggressive species and the cover concrete is not very wet. In this case, some of the dissolved corrosion product could be oxidised and transformed into rust, deposited at the interface between steel and the surrounding concrete. The volume of such rust is usually 2~6 times greater than the volume of the metal consumed by corrosion [Broomfield (1997)]. This means that an expansion stress will be generated at the interface between the steel and the surrounding concrete. The more the cross section of the steel is reduced by corrosion, the higher the resulting expansion stress. When the expansion stress exceeds the tensile strength of the cover concrete, the latter will crack. The cracks in the cover concrete will in turn act as shortcuts for the ingress of detrimental species, so the corrosion process will be further accelerated by the cracking of cover concrete. Therefore a self-catalytic deterioration process is triggered: corrosion ' expansion ' cracks in concrete ' ingress of detrimental species ' more serious corrosion ' more expansion ' more cracks ' more detrimental species. Finally spalling of the cover concrete takes place, further exposing the interior of the concrete to corrosion. This is the main course of damage of reinforced concrete associated with corrosion. It should be noted that the chloride- and carbonation-induced corrosion processes are different in their initiation mechanisms, and thereby they are also somewhat different in their deterioration morphologies. Chloride ions are likely to cause localised breakdown of the passive film, so the corrosion damage of reinforcement is relatively localised and deep. This could result in severe reduction of the cross-section of reinforcement without a significant amount of steel being corroded at some sites. On the other hand, the passive film of reinforcement would be thermodynamically unstable in carbonated concrete as long as the pH value of the pore solution is below the threshold level, i.e., the reinforcement could have an overall corrosion in the areas in contact with the carbonated concrete (pH lower than the threshold level). Therefore, the carbonation induced corrosion is usually more widespread but shallower on the reinforcement.

46

Service life prediction of reinforced concrete structures

After the corrosion of reinforcement is triggered (i.e., at the second stage), its further development depends on many parameters and environmental conditions that change with time. In practical cases, it is difficult to tell whether the corrosion rate is slowing down or increasing with time. Even under the condition that the concrete and environmental factors do not change dramatically with time, the corrosion rate could either be decreasing gradually due to the formation of corrosion products on the steel surface or increasing dramatically because of the continuous ingress of more chloride ions (or a greater degree of carbonation with time). Nevertheless, there is reason to believe that the corrosion of reinforcement in the second stage is relatively stable, because the chloride content or the pH value in the vicinity of the reinforcement can only increase or decrease gradually, and there are no dramatic changes of the factors that can affect the corrosion of reinforcement until the third stage of deterioration starts, i.e., spalling of concrete. Finally, the cover concrete might spall off due to accumulation of corrosion products at the interface between reinforcement and concrete. At this stage, the reinforcement would be subject to unpredictable and rapid environmental influences. In this report, this is termed the unpredictable corrosion stage.

6.2 Critical corrosion amount


As mentioned above, corrosion of reinforcement could eventually cause spalling or cracking of the cover concrete, which is easily visualised and means that repair is required to restore the structure, i.e., the termination of the service life of the concrete structure. Spalling or cracking of the cover concrete is closely related to the total amount of corroded reinforcement or the total amount of corrosion products being accumulated at the interface between the reinforcement and cover concrete. The relationship between spalling or cracking of concrete and the corrosion amount of the reinforcement is of interest and significance for the prediction of the second stage of deterioration. Generally, the total amount of corrosion of reinforcement that can lead to cracking or spalling of the cover concrete is defined as the critical corrosion amount (Qcr). In some sense, the visible spalling and cracking could be regarded as the criterion for the termination of service life of a reinforced concrete structure, and the critical corrosion amount is usually used as a decisive parameter in the service life prediction models. Various efforts have been made to estimate the critical amount of corroded rebar that can cause spalling of the cover concrete. Clemena et al (1992a,1992) found that the corrosion rate of rebar had a reasonable correlation with the metal loss of rebar, and that the threshold of metal loss of rebar that initiated delamination in concrete was about 3~6% by mass of metal. Rodriguez et al (1994), and Gonzalez et al (1997) found in laboratory tests that the critical section loss of metal to cause cracking and spalling was 15~40 Mm. A higher corrosion penetration depth of around 100Mm had also previously been proposed to initiate cracking [Grimes et al (1979), Andrade et al (1993)]. The cover concrete could also be induced by less than 100 Mm of steel section loss [Broomfield (1997)].

47

Service life prediction of reinforced concrete structures

Certainly, the distribution pattern of corrosion products, the ability of concrete to accommodate stress, the geometry of rebar distribution in concrete and rebar diameter all can affect the above results. Also, the risks of damage to rebars affected by uniform and localised corrosion are also quite different, even if their final mass losses might be the same. For example, the critical corrosion amount has been found to be related to the cover thickness and diameter of the reinforcing steel. Morinaga et al (1988, 1994) proposed the following empirical equation to correlate the critical corrosion amount per unit area (Qcr) on reinforcement, the thickness of the cover concrete (d), and the diameter of reinforcement ():

Qcr = 0.590 (1 + 2d/ )0.85

(58)

Andrade et al (1998) used corrosion currents, ranging from 3 to 100 A/cm2, to accelerate corrosion of steel in concrete specimens. Their results showed that if concrete crack width (Wt) was less than 1.0 mm, then it could be well related to the corrosion penetration (steel bar radius decrease) Px:

Wt = 0.05 + M (Px - Px0 )

(59)

where M is a constant about 0.01, Px0 is the initial penetration of corrosion into steel at crack initiation, which can be further expressed by:

Px0 = 83.8 + 7.4 d/ - 22.6 fc,

sp

(60)

where d/ is the ratio of concrete cover thickness to reinforcing bar diameter, and fc, sp is the concrete splitting tensile strength (in MPa). These authors also observed the phenomenon that higher applied currents caused smaller crack widths, and that cracks appeared more rapidly for the concrete with higher strength. The latter was attributed to the easier diffusion of corrosion products through a porous weaker concrete which has also higher capability of accommodating the corrosion products.

6.3 Prediction models

6.3.1 Simple extrapolation


The simplest way to predict the corrosion of reinforcement is to assume the corrosion rate of reinforcement to be a constant, so that the duration of the second stage can be simply regarded as the time for reinforcement to be corroded to the critical corrosion amount Qcr at such a corrosion rate.

48

Service life prediction of reinforced concrete structures

Clear (1989) found a very simple, broad relationship between the corrosion rate and the time required for corrosion development (Table 3). Table 3. Measured corrosion rate and time for corrosion development [Clear (1989)] corrosion rate (MA/cm2) 0.2~1 1~10 >10 time for corrosion development (year) 10~15 2~10 <2

It clearly shows that the larger the corrosion rate, the shorter the period for corrosion development. Obviously, this kind of prediction ignores the fact that the corrosion rate of reinforcement can hardly be constant in the long term.

6.3.2 Diffusion controlled model


Weyers (1998) stated that Fe2+ from reinforcing steel must diffuse through the rust layer before it can turn into corrosion products. This can also be understood as the diffusion of oxygen through the rust to the steel surface. Therefore the expansion is a diffusion controlled process. Based on this and the electrochemical balance during the corrosion reaction, Liu (1996) proposed a time-to-cracking mechanism that can be modelled by the following expression [Weyers (1998)]: t CR %
2 Wcrit 2k p

(61)

where tCR = is the time for the occurrence of cracking in the cover concrete kp = F (1/N),OIcorr , N = the mean of electrochemical equivalent volume of corrosion products O = diameter of the reinforcing bar Icorr = annual mean corrosion current density Wcrit = critical weight of corrosion products per unit length of reinforcement required to crack the cover concrete, which is expressed by:

V W S Wcrit % P rust T,O ( d s $ d o )$ st Q P st R U

(62)

where Prust = density of rust = thickness of pore band around the steel/concrete interface do Wst /Pst = volume of the corroded steel = thickness of corrosion products needed to generate tensile stress, ds which is a function of concrete tensile strength, modulus of elasticity, cover thickness, Poissons ratio, creep coefficient, and the size of reinforcing bar.

49

Service life prediction of reinforced concrete structures

The above equation was obtained based on the consideration that, the critical corrosion amount should be so much that they can fill the space between the steel and the surrounding concrete (which includes pore band and the cavity created by the corrosion) and produce enough stress to crack the cover concrete. Liu (1996) demonstrated that Eq (61) can give a good estimation of tcr. By analysing nearly 3000 measurements using a corrosion rate measuring device without the guard ring on a chloride contaminated concrete which was exposed to outdoor conditions for 5 years, Liu (1996) and Weyers (1998) found the corrosion current density to be a function of several parameters, such as the chloride content, temperature, the electric resistivity of the cover concrete, and the time after the initiation of corrosion, etc. This diffusion-based model has incorporated the retarding effect of corrosion products during the second stage, and in this sense, it is more reasonable than the simple extrapolation method. However, the concurrent enhancement effect due to the continuously increasing chloride level or carbonation degree does not appear to have been addressed. Therefore, in theory, it is still not a comprehensive model.

6.3.3 Laboratory simulation models


Service life prediction can sometimes be carried out in the laboratory under simulated conditions. Morinaga et al (1988, 1994) showed that

tp = Qcr / q

(63)

where tp is the time between the initiation of reinforcement corrosion to the occurrence of significant damage to concrete, Qcr is the critical amount of corrosion per unit area, which can be empirically determined by (58), and q is the rate of corrosion per unit area, and can be determined according to the ratio of corrosion rates of the steel in concrete determined under reference (testing) condition and that prevailing under the environmental condition:

q % q1

q2 q' 2

(64)

where q1 and q2 are the rates of corrosion of steel in experimental concrete specimens containing Cl- and stored under the reference condition (controlled temperature, humidity, and oxygen concentration) and the environmental conditions under investigation, respectively; and q2 is the corrosion rate of the structural member fabricated in the laboratory and determined under the reference condition. q2 and q2 are expressed as empirical functions of testing or environmental conditions, while q1 is further related to the w/c.

50

Service life prediction of reinforced concrete structures

The application of this model requires testing the corrosion rate of steel in concretes containing various contents of Cl- representing the increase in Cl- at the vicinity of the steel with time, and summing up the cumulative amount of corrosion by:
n

Qn year %

6 qi t '
i %1

(65)

where qi is the corrosion rate at the ith year, t = 1 year, and n is the number of years for which the corrosion rate is evaluated. The service life of concrete is judged by comparing Qn year with Qcr, the amount causing concrete cracking. When Qn year/Qcr > 1, n is regarded as the service life. In an evaluation of a 30 year old structure in Tokyo Bay, these researchers demonstrated that cracking due to corrosion is likely to occur when the ratio Q30 year/Qcr exceeds 1. Theoretically, the laboratory simulation model relies on the validity of the simulation conditions. However, it is impossible to completely repeat field situations in the laboratory. Therefore, these methods obviously have several limitations, such as being labour-intensive and time-consuming, not fully representative of exposure conditions, etc.

6.3.4 Empirical prediction models


Due to the complexity of the corrosion system for reinforced concrete structures, the simple extrapolation based on the assumption that the measured corrosion rate at a certain time is the average rate throughout the service life of a structure is obviously inaccurate [Broomfield et al (1993)]. However, there is still no rigorous mathematical equation to describe the developmental process of the corrosion of reinforcement, particularly for field concrete structures. In this case, empirical prediction would be an option. In the 1970s, an empirical equation was used by the Florida Department of Transportation for estimation of time-to-cracking (tcr), known as the Stratfall formula, and was later modified by Clear (1976):
. 129d 122 0. 42 Cl ] [ (w /c )

t cr %

(66)

where d is the cover concrete thickness, [Cl] is the chloride content of the water in contact with concrete, and w/c is the water-to-cement ratio [Purvis et al (1992)]. This type of empirical formula can have quite different empirical constants, for example, Clear (1976) proposed that for atmospherically exposed structures:

tcr= [(0.052 d1.22 t0.21 )/(CS,Cl0.24 (w/c) ]0.83

(67)

51

Service life prediction of reinforced concrete structures

where t is the age at which chloride content in concrete surface (CS,Cl) is measured. Several variations of this equation also exist [Purvis et al, (1994)]. Other similar expressions have also been proposed [Purvis (1992)]. Bazant (1979,1979a) developed a physical-mathematical model to determine the time required for the initiation of cracking in concrete due to chloride-induced corrosion, based on a steady state corrosion rate. He also found that the time to the start of cracking was a function of corrosion rate, cover depth, and mechanical properties of concrete. Among these, corrosion rate was the most significant parameter in determining the incubation time of cracking of cover concrete. As the corrosion rate is influenced by temperature and moisture conditions in the concrete, a single corrosion rate measurement may not give reliable predictions or an overall behaviour of concrete structure. Therefore, Clifton et al (1994) suggested that the effect of ambient moisture condition on the rate of steel corrosion in concrete could be roughly estimated by taking the corrosion rate in ambient relative humidity (RH) 70% as 1, and the corrosion rate at a humidity lower than 70% and at rain season as 0.1 and 10, respectively, and so to evaluate the average corrosion rate over a year according to the weather condition. However, this type of relationship should be used carefully because the moisture absorption ability of concrete varies from concrete to concrete and with age and temperature, and also because the corrosion rate changes with temperature. Overall, the empirical models could apparently have good fit with existing experimental data. However, there is less confidence in predicting the future corrosion behaviour of reinforced concrete, since empiricism means less theoretical guidance. Furthermore, there is still lack of comparison and evidence as to which model is more appropriate and whether the empirically based models can be applied to all structures.

6.4 Prediction procedures


The prediction of the duration of this stage (the stable propagation stage) requires several different steps, including reconfirmation of the deterioration stage, estimation of corrosion rate, and determination of the current corrosion stage of reinforcement.

6.4.1 Reconfirmation of deterioration stage


It is important to verify whether the structure being studied is in the second stage of corrosion deterioration.

52

Service life prediction of reinforced concrete structures

The primary verification of the deterioration stage can be achieved through visual inspection, which can basically reveal whether or not the structure has got into the third stage of deterioration. After the possibility of being in the unpredictable deterioration stage is excluded, the half-cell potential measurement (ASTM C876) can be used as the simplest way to reconfirm whether the structure is at the initiation stage or has got into the corrosion propagation stage. However, the half-cell-potential measurement may give misleading indication in cases of carbonation induced corrosion, or the structure being water saturated or coated, etc. The most reliable reconfirmation is to break out at some typical sites for close inspection of the corrosion situation of reinforcement.

6.4.2 Estimation of corrosion rate


Corrosion rate of reinforcement is also a critical parameter for prediction, and is involved in most prediction models. The accuracy of the estimated corrosion rate would decisively affect the predicted results. There are several techniques that can be used to estimate the corrosion rate of reinforcement, which are presented in another report by the authors [Song et al (1999)], and will not be repeated in this report. In spite of the methods for estimating corrosion rate, the basic concept of using corrosion rate in prediction models is to estimate when the corrosion amount reaches its critical level to crack the cover concrete.

6.4.3 Determination of the present corrosion state


Even though the current corrosion state is not specified in the above prediction models, the current corrosion amount is the basis for estimating the time needed for it to grow to the critical level. It should be borne in mind that the past corrosion of a structure should be subtracted from the whole service life in determining the remaining service life. The most reliable determination of the corrosion situation is still the breakout at some typical sites.

6.4.4 Selection of prediction models


The selection of prediction models for the second stage of deterioration is relatively empirical, depending on the predictors knowledge and understanding of the corrosion situation of the structure. At present, it cannot be expected that the prediction will have a high reliability, particularly for a long term prediction, because of the immaturity of all the existing models. Therefore, the selection of a good model is to some extent a trial-and-error process. In some cases, it may even be more realistic to establish a new model following an innovative idea for a corroding structure.

53

Service life prediction of reinforced concrete structures

7. Concluding Remarks
7.1 Methodology of service life prediction
In the face of a corroding reinforced concrete, to predict its remaining service life based on its corrosion situation, these basic steps could be followed: 1) Obtaining structural and environmental parameters: Such background information is critical to the identification of causes of corrosion damage. For example, if a structure is exposed to a marine environment, chloride induced corrosion could be the principal cause of corrosion deterioration; if the structure is in a relatively dry atmosphere far away from coastal environments, and chloride salts have not been found in the concrete, then carbonation should be considered as the possible cause of corrosion damage. The background parameters include concrete quality, thickness of the cover concrete, resistivity of the concrete, etc. which would be useful in understanding the corrosion deterioration process of the structure. 2) Field and laboratory testing to confirm causes of corrosion deterioration: This is essential information for the prediction of the first stage of deterioration. The selection of prediction models in the second deterioration stage also depends on the corrosion causes and mechanisms. The causes can normally be found by analysing the environmental conditions and the samples taken from the field structure. 3) Determination of current deterioration stage: Current deterioration stage also affects the selection of prediction models. It is strongly suggested to break out the cover concrete at some typical sites to confirm the current stage of corrosion deterioration, and to experimentally detect the carbonation depth and chloride penetration depth in the concrete at those typical sites. If the reinforcement is intact, the carbonation front is still far away from the reinforcement, and chloride concentration in the vicinity of the reinforcement is still far below the chloride threshold, then the structure is still in the first stage of deterioration. Otherwise, the structure should be regarded as being in the second stage. 4) Measurement of primary data: This includes obtaining chloride profile, carbonation depth profile, corrosion rate, current corrosion amount, etc. They are the essential data based on which the prediction models will be selected or established and verified. 5) Selection of prediction models: This is the critical procedure in the prediction. An unreasonable model means unreliable predicted results, even though the existing data can be fitted very well by this model. The selection of models is determined by the environmental conditions, corrosion causes, current deterioration stage, chloride and carbonation depth profiles, etc. Basically, the selected prediction model should at least be able to satisfactorily fit the primary data first. Normally, it is required to be able to give reasonable conclusions under typical conditions and have acceptable inferences in some assumed extreme cases. Otherwise, theoretical or empirical modification of the model would be necessary.

54

Service life prediction of reinforced concrete structures

6) Estimation of relevant parameters and prediction: The relevant parameters involved in the model could be directly estimated by curve-fitting of the primary data with the selected model. In some cases, the relevant parameters may need to be theoretically calculated, to be found from literature or to be measured by other experiments. After all the relevant parameters are known, the prediction model would become a determinate deterioration-time function, such that service life could be predicted by assuming a certain degree of deterioration as the criterion to terminate the service of the structure. The above methodology is not fixed, and could vary with the real situations.

7.2 Problems to be solved


As demonstrated earlier in this report, the service life prediction is far from successful, and there are a number of unsolved problems in the prediction: 1) The mechanisms of chloride ingress and carbonation are still unclear, particularly the influences of the transport mechanisms other than diffusion, which are usually involved in the field. Due to the insufficient knowledge of the transport mechanisms, it is impossible to correctly formulate their influences into mathematical equations. Therefore, the prediction models dealing with the first deterioration stage have difficulty in comprehensively characterising the real processes in field structures. 2) There is a lack of understanding of the corrosion process in the second stage of deterioration. As a very special corrosion system, a steel/concrete electrode system is much more complicated than the other typical steel/solution corrosion systems. At present, even the latter which have been intensively investigated have no reliable model to predict future corrosion trends. It is even harder to have a determinate equation to describe the more complicated corrosion process in the reinforced concrete system. 3) Difficulty in mathematically formulating environmental influences on the deterioration process, due to the unpredictable variations of the environmental factors with time, is another problem. This will make the models hard to solve due to the non-constant parameters involved. 4) It is still difficult to accurately determine some important parameters involved in the prediction model for a particular structure. For example, chloride threshold is a critical parameter determining the onset of the second stage of deterioration, and is involved in some models. However, such a parameter depends on many factors and can not be regarded as a fixed value for various structures under different conditions. The critical corrosion amount is also in a similar situation. Another example is the pH threshold associated with the carbonation process. The pH threshold is a thermodynamic concept which only indicates the possibility of a reaction taking place. For example, the pH value of the pore solution being below the threshold due to carbonation, only means that the corrosion of reinforcement has become possible, but it may not occur immediately or years later.

55

Service life prediction of reinforced concrete structures

Unfortunately, in the existing models, these issues have not been satisfactorily resolved. 5) There is limited data to verify the existing or newly proposed prediction models, particularly data from field structures. It is obvious that these problems limit the development of universal acceptable service life prediction techniques. The solution of these problems would undoubtedly lead to improvements in the prediction models.

56

Service life prediction of reinforced concrete structures

8. References
Akita, H., and Fujiwara, T. (1995), Water And Salt Movement Within Mortar Partially Submerged In Salty Water, Proceedings Of The International Conference On Concrete Under Severe Conditions. AASHTO Designation T 277-83 (1983), Standard method of test for rapid determination of the chloride permeability of concrete, America Assoc. of State Highway and Transportation Officials, Washington D.C. AASHTO T260-84. (1984), Standard Method Of Sampling And Testing For Total Chloride Ion In Concrete And Concrete Raw Materials, (American Association of State Highway Transportation Officers, Washington, DC). Akita H., and Fujiwara T. (1995), Water and salt movement within mortar partially submerged in salty water, Proc. Of the international conference on concrete under severe conditions, Japan, E&FN Spon, London. Amey, S.L., Johnson, M.A. Miltenberger, and Farzammehr, H. (1996), A Methodology For Prediction Service Life Of Concrete Structures Exposed To A Marine Environment, ACI, SP-163, (Detroit, USA). Amey, S.L., Johnson, D.A., Miltenberger, M.A., and Farzam, H., (1998), Prediction of service life of concrete marine structures: An environmental methodology, ACI Structural Journal, Vol.95, No.2, p.205-214. Andrade, C., Alonso, C. and Molina, F.J. (1993), Mater.Struct. 26:pp453. Andersen, A. and Paulsson, J. (1997), Measurements on seasonal and diurnal variations of environmental conditions surrounding a heavily trafficked bridge structure, Proceedings of International Conference: Repair of Concrete Structures from Theory to Practice in a Marine Environment, Svolvr, Norway. p.143-152. Andrade, C., Alonso, C., Rodriguez, J. and Ortega, L.M. (1998), On-site corrosion rate monitoring and its use to assess structural condition, Rehabilitation of Structures, Proceedings of the 2nd International RILEM/CSIRO/ACRA Conference, Melbourne. p.144-153. Arya, C. et al, (1990), Cement and Concrete Research, 20:pp.291. Arya, C., Newman, J.B. (1990a), Materials And Structures, 23:pp.319. Arya, C., Xu, Y. (1995), Cement and Concrete Research, 25(4):pp.893. Aschan, N. (1963), Thermogravimetric Inverstigation of carbonation phenomenon in concrete, Nordisk Betong, 3:pp275-283. ASTM D1411, (1982). Standard test methods for water soluble chlorides present as admixes in graded aggregate road mixes, (American Society for Testing and Materials, Philadelphia,PA ).

57

Service life prediction of reinforced concrete structures

Bamforth, P.B. and Pocock, D. (1990), Proc. 3th Symp Corrosion of Reinforcement In Concrete Construction, pp.19 (Elsevier Applied Science). Bamforth, P.B. and Price, W.F. (1993), Factors influencing chloride ingress into marine structures, High Performance Concrete in Severe Environments, ACI SP-140, P. Zia, editor, Vol.2, p.1105-1118. Bamforth, P.B. and Chapman-Andrews, B. (1994),Corrosion And Corrosion Protection Of Steel In Concrete, 1:pp.139 (Sheffield Academic Press). Bamforth, P.B.(1996), Definition of exposure classes and concrete mix requirements for chloride contaminated environments, in: C.L.Page, P.B.Bamforth and J.W.Figg (ed.) Corrosion Of Reinforcement In Concrete Construction, pp.176 (The Royal Society of Chemistry, Thomas Gram House, Science Park, Cambridge). Bazant, Z.P.(1979), Physical model for steel corrosion in sea structures-theory, Journal of the Structural Division, ASCE, ST6,pp.1137. Bazant, Z.P.(1979a), Physical model for steel corrosion in sea structures-applications, Journal of the Structural Division, ASCE, ST6, pp.1155. Beaton, J.L. and Stratfull, R.F. (1963), Environmental Influence of corrosion of reinforcing in concrete bridge substructures, Concrete bridge decks and Pavement Surface, HRR 14 HRB 73(Washington, DC). Bentz, E.C. Evans C.M. and Thomas, M.D.A. (1996), Chloride diffusion modelling for marine exposed concretes, in: C.L.Page, P.B.Bamforth and J.W.Figg, (ed.) Corrosion Of Reinforcement In Concrete Construction, pp.136(The Royal Society of Chemistry, Thomas Gram House, Science Park, Cambridge). Berke, N.S. and Hicks, M.C. .(1994), Corrosion, 50(3): pp.234. Berke, N.S., Hicks M.C. and Tourney, P.G. (1997), 100 year service lives in marine environments using belts and braces approach with calcium nitrite, In: A.Blankvoll (ed.) Proceedings Of International Conference-Repair Of Concrete Structures, From Theory To Practice In A Marine Environment pp. 163 (,Norway). Blankvoll, A.(1997). History of the gimsoystraumen bridge repair project, In A.Blankvoll (ed.) Proceedings of International Conference-Repair Of Concrete Structures, From Theory To Practice In A Marine Environment, pp.35 (Norway). Blunk, G., Gunkel P. and Smolczyk, H.G. (1986), On the distribution of chloride of chloride between the hardening cement paste and its pore solution, In: Proceedings Of The 8th International Congress On The Chemistry Of Cement, Vol.V, 85 (Brazil). Broomfield, J.P., Rodriguez, J. Ortega, L.M.and Garcia, A.M. (1993), Corrosion rate measurement and life prediction for reinforced concrete structures, In: Proceedings Of Structural Faults And Repair-93, vol.2:pp.155, (Engineering Technical Press, University of Edinburgh). Broomfield, J.P.(1995), Bulletin of Electrochemistry 11(3):pp.121.

58

Service life prediction of reinforced concrete structures

Broomfield, J.P., (1997), Corrosion of steel in concrete, Understanding, investigation, and repair, (Chapman and Hall). Browne R.D.(1982), Design prediction of the life for reinforced concrete in marine and other chloride environments, Durability of Building Materials, 1:pp122, (Amsterdam). Buenfeld N.R., Shurafa-Daoudi, M.T., and McLoughlin I.M. (1995), Chloride transport due to wick action in concrete, Proc. Of the RILEM International workshop on chloride penetration into concrete, France. Building Research Establishment (1982) Digest 264, The durability of steel in concrete: Part 2 Diagnosis and assessment of corrosion -cracked concrete, Building Research Establishment, (Garston, Watford, UK). Building Research Station (1977), Determination of chloride and cement content in hardened portland cement concrete, Building Research Establishment (Information Sheet, IS 13/77, England: Building Research Station). Byfors K., Hansson C.M. and Tritthart J., (1986), Cement and Concrete Research, 16:pp.760. Byfors, K. (1990), Chloride-initiated reinforcement corrosion, chloride binding, CBI report 1:pp.90, (Swedish Cement and Concrete Research Institute). Cady, P.D. and Weyers, R.E. (1984), Deterioration rates of concrete bridge decks, Journal of Transportation Engineering, Vol.110, No.1 , p.35-44. Cady P.D. and Weyers, R.E. (1983), Cement, Concrete and Aggregate, 5:pp.81. Cahyadi J.H., and Uomoto T., (1993) influence of environmental relative humidity on carbonation of concrete mathematical modelling, proceedings of the 6th international conference on the durability of building materials and components, Omiya, Japan, vol 2, pp1142-1151. Castro P., Maldonado l. and Coss R.de (1993), Corrosion Science, 35:1557. Chamberlin, W.P., and Weyers R.E. (1993), Field performance of latex-modified and low-slump dense concrete bridge deck overlays in the United States, Concrete Bridges in Aggressive Environments, edited by R.E. Weyers, ACI SP 151. p.1-16. Clear, K.C. and Hay R.E. (1973), Time-to-corrosion of reinforcing steel in concrete slabs, vol.1, Effect of mix design and construction parameters, Report No. FHWA-RD-73-32, (Federal Highway Administration, Washington, D.C.). Clear, K.C. (1976), Time to corrosion of reinforcing steel in concrete slabs, FHWA-RD76-70, (Washington, DC). Clear, K.C. (1989), Measuring the rate of corrosion of steel in field concrete structures, Transportation Research Record 1211, (Transportation Research Board, national Research Council, Washington, DC).

59

Service life prediction of reinforced concrete structures

Clemena G.G.(1992), Benefits of measuring half-cell potentials and rebar corrosion rates in condition surveys of concrete bridge decks(final report),(Virginia Transportation Research Council). Clemena G.G., Jackson D.R., and Crawford G.C., (1992a) Inclusion of rebar corrosion rate measurements in condition surveys of concrete bridge decks, Transportation Research Record 1347: pp37. Clifton, J.R. (1993), Predicting the service life of concrete, ACI Materials Journal, Vol.90, p.611-617. Clifton, J.R. and Pommersheim, J.M. (1994), Predicting remaining service life of concrete, Proceedings of the International Conference on Corrosion and Corrosion Protection of Steel in Concrete, Sheffield, UK, 1994. Edited by R. N. Swamy, Sheffield Academic Press. Vol.1, p.619-637. Collepardi M., Marcialis,A., and Turriziani R. (1970), The kinetics of penetration of chloride ions into the concrete, II Cimento, Vol.4, pp157-164. Collepardi M., Marcialis, A., and Turreziani, R. (1972), Penetration of chloride ions into cement pastes and concrete,J. Am. Cer. Soc., 55(10):pp.534-535. Collepardi M. and Biagini S. (1989), Effect of water/cement ratio, pozzolanic addition and curing time on chloride penetration into concrete, ERMCO 89. Concrete Society Discussion Document(1996), Development in durability design and performance-based specification of concrete, Concrete Society Special Publication CS109. Conway E.J.(1957), Microdiffusion analysis and volumetric error, 4th ed. Crosby Lockwood and Son Ltd,pp201-205. Costa, U., Facoetti, M., Massazza, F. (1992), Permeability and diffusion of gases in concrete, Proceedings of the 9th International Congress on the Chemistry of Cement, New Delhi, Vol.V, p.107-114. Crank, J. (1975), The Mathematics of Diffusion, 2nd edition, Clarendon Press, Oxford. Crank J.(1979), The Mathematics Of Diffusion, pp.414, (Oxford, Clarendon Press). Dhir, R.K., Jones M.R., and Ahmed H.E.H. (1990), Determination of total and soluble chloride in concrete, Cement and Concrete Research, 20 (4):pp.579. Dhir R.K., Jones, M.R., Ahmed H.E.H., and Seneviratne A.M.G. (1990), Rapid estimation of chloride diffusion coefficient in concrete, Magazine of Concrete Research, 42(152),pp.177-185. Dhir, R.K., Jones, M.R. and Ahmed, H.E.H. (1991), Concrete durability: estimation of chloride concentration during design life, Magazine of Concrete Research, Vol.43, No.154, p.37-47. Dhir, R.K., Jones M.R, and Mccarthy M.J. (1994), Mag.Concr.Res., 46:pp.269.

60

Service life prediction of reinforced concrete structures

Diamond, S.(1986), Chloride concentrations in concrete pore solutions resulting from calcium and sodium chloride admixtures, Cement, Concrete, and Aggregates, 8(2):pp.97. Duchesne, J., and Berube M.A. (1994), Cement and Concrete Research, 24(3):pp.456. Edwardsen C.(1995), Choride penetration into cracked concrete, Proceedings of the RILEM international workshop on chloride penetration into concrete, Oct., France. Fagerlund, G. (1985), Essential data for serve life prediction, NATO Advanced Workshop on Problems in Service life Prediction of Building and Construction Materials, Paris, 1984. Martinus Nijhoff Publishers, Dortrcht/Boston/Lancaster. Fishcher K.P., Bryhn O., and Aagaard P., (1984), Corrosion 40(7):pp.358. Fluge, F. (1997), Environmental loads on costal bridges, Proceedings of International Conference: Repair of Concrete Structures from Theory to Practice in a Marine Environment, Svolvr, Norway. p.89-98. Forrester J.(1976), Measurement of carbonation, Carbonation of concrete 1976 RILEM Symposium, paper 2.1, pp.6. Frantz J.D., and Mao H.K. (1976), Am.Jour.Sci., 276:pp.817. Frederiksen J.M., Nilsson L-O., Sandberg P., et al, (1997), A system for estimation of chloride ingress into concrete- theoretical background, report No. 83, (The Danish Road Directorate, Copenhagen, Denmark). Fukushima T.(1987), Theoretical investigation on the influence of various factors on carbonation of concrete, 4th international conference on durability of building materials and components, Singapore, pp662-670. Fukushima, T.(1991), Predictive methods on the progress of neutralization of concrete by unsteady state dynamic analysis considering the influence of tendency of ingress in concentration of atmospheric carbondioxide, 2nd canment/ACI international conference on durability of concrete., Montreal Canada, Supp. Paper, pp545-564. Fukute, T., Moriwake, A., Seki, H. and Kawada, H. (1998), Repair work strategy for concrete jetties deteriorated by salt attack, Rehabilitation of Structures, Proceedings of the 2nd International RILEM/CSIRO/ACRA Conference, Melbourne. p.279-290. Gaynor, R.(1985), September:pp.35. Understanding chloride percentage, Concrete International,

Gebauer, J. (1982), Some observations on the carbonation of fly ash concrete, Silicate Industrials, Vol.47, No.6, p.155-159. Geiker, M.R., Henriksen, C. and Thaulow, N. (1993), Design for durability: a case story, Concrete 2000, Dhir, R.K. and Jones, M.R. editors. EFN Spon. Vol.1, p.64-70. Glass G.K. and Buenfeld N.R. (1995), Chloride threshold levels for corrosion induced deterioration of steel in concrete, RILEM Whorkshop on Penetration of chloride into concrete, October.
61

Service life prediction of reinforced concrete structures

Goodgrake, C.J. (1978), Reaction of Beta Dicalcium Silicate and Tricalcium Silicate with Carbon Dioxide and Water, PhD. thesis, University of Illinois at UrbanaChampaign. Goto, S., and Roy D.M. (1981), Cement and Concrete Research, 11:pp.751. Goto, S. and Roy D.M., (1981A), Cement and Concrete Research, 11:pp575. Gonzalez J.A. Feliu S., and Rodriguez P., (1997), Corrosion, 53(1):pp.65. Grimes W.D., Hartt W.H., Turner D.H., (1979), Corrosion 35:pp.309. Hamada, M. (1968), Neutralisation of concrete and corrosion of reinforcement steel, 5ICCC, Tokyo, V3,pp343-368. Hansen T.C., Jensen H. and Johannesson T. (1986), Chloride diffusion and corrosino initiation of steel reinforcement in fly-ash concretes, Cements and concrete Research, 16(5):pp782-784. Hansson C.M., Strunge H., Markussen J.B., and Frolund T. (1985), The effect of cement type on the diffusion of chloride, Nordic Concrete Research, Publication No. 4, pp70-80. Haque M.N. and Kayyali O.A., (1995), vol.CE37 (2):pp.141. Australian Civil Engineering Transactions,

Hartt W.H.(1990), A concrete durability enhancement strategy for proposed Tampa Bay Bridge Construction, Consultants report, W.R.Grace & Co., CPD, Cambridge,MA, March pp.30. Hausmann D.A.(1967), Steel corrosion in concrete: how does it occur?, Materials Protection, 6:pp.19. Heleffereich F. , Katchalsky A. .(1970), Jour.Phys.Chemistry, 75:pp.308. Hobbs D.(1988), Carbonation of concrete containing slag, Magazine of Concrete research, 40(143):pp.69-78. Hoffman P.C. and Weyers R.E., (1994), Computer simulation of apparent diffusion in decks with concrete overlays, In: Richard E. Weyers (ed.), Phillip D.Cady International SymposiumConcrete Bridges in Aggressive Environments, SP 151-11, pp.197-219 (American Concrete Institute, Detroit, Michigan, USA). Holden W.R., Page C.L., Short M.R. (1983), The influence of chlorides and sulphates on durability of reinforced concrete, in : A.P.Crane (ed.), Corrosion Of Reinforcement In Concrete Construction, pp.143 (Ellis Horwood, Hichester). Houst Y.F. (1991), Influence of microstructure and water on the diffusion of CO2 and O2 through cement paste, 2nd Canmet/ACI International conference on Durability of concrete, Montreal Canada, Supp. Paper. Pp.141-159. Hussain S.E., Rasheeduzzafar, Al-Musallam A., and Al-Gahtani A.S. (1995), Factors affecting threshold chloride for reinforcement corrosion in concrete, Cem. Concr. Res. 25(7):pp.1543-1555.
62

Service life prediction of reinforced concrete structures

Hussain S.E. and Rasheeduzzafar, (1994) ACI Materials Journal, May-June: pp.264. James Instruments, Inc. (1990) , Cl test model Cl 500 Instruction Manual, Chicago, IL: pp.11 (James Instruments, Inc., Chicago, IL,). Johansen V.(1995), et al, Chloride transport in concrete, Concrete International, American Concrete Institute, July: pp.43. Jones, M.R., McCarthy, M.J. and Dhir, R.K. (1994), Chloride ingress and reinforcement corrosion in carbonated and sulphated concrete, Proceedings of the International Conference on Corrosion and Corrosion Protection of Steel in Concrete, Sheffield, UK, 1994. Edited by R. N. Swamy, Sheffield Academic Press. Vol.1, p.365376. Kalousek. G.L., Porter, L.C. and Benton, E.J. (1972), Concrete for long-term service in sulfate environment, Cement and Concrete Research, Vol.2, p.79-90. Klopfer H.(1978), Bautenschutz und Bausanierung, 1(3):pp.86 English by P.Aukland). (translated into

Kondo, R., Daimon, M., and Akiba, T., (1968), Mechanisms and kinetics of carbonation of hardened cement, 5ICCC, Tokyo, V3, pp402-409. Kranc S.C. and Sagues A.A. (1994), Corrosion, 50(1):pp.50. Kroone, B., and Blakely F.A.(1959), Reaction between carbondioxide gas and mortar, Journal of American concrete institute, pp.497-510. Larsen C.K.(1997), Effect of temperature, carbonation and drying and wetting on chloride uptake in concrete, In Aage Blankvoll (ed.) Proceedings Of International Conference-Repair Of Concrete Structures, From Theory To Practice In A Marine Environment . pp.153 (Norway). Larsen C.K.(1997), Effect of temperature, carbonation and drying and wetting on chloride uptake in concrete, In Aage Blankvoll (ed.) Proceedings Of International Conference-Repair Of Concrete Structures, From Theory To Practice In A Marine Environment . pp.153 (Norway). Larsson J.(1995), The enrichment of chlorides in expressed concrete pore solution submerged in saline solution , Proceedings Of The Nordic Seminar On Field Studies Of Chloride Initiated Reinforcement Corrosion In Concrete, Lund University of Technology, Report TVBM-3064, pp.171. Li, S. and Roy, D.M. (1986), Investigation of relations between porosity, pore structure, and Cl- diffusion of fly ash and blended cement pastes, Cement and Concrete Research, Vol.16, No.5, p.749-759. Li, Y. and Wu, Q. (1987), The mechanism of carbonation of mortars and the dependence of carbonation on pore structure, Concrete Durability-Katharine and Bryant Mather international conference, Atlanta, ACI SP 100, Ed. Scanlon J.M., pp1915-1943. Lin S.H.(1990), Corrosion, 46:pp.964.
63

Service life prediction of reinforced concrete structures

Litvan G. and Meyer A.(1986), carbonation of granulated blast furnace slage cement concrete during twenty years of field exposure, ACI SP-91, Madrid Proc. Vol 2,pp14451462. Liu Y., and Weyers R.E., (1996), Time to cracking for chloride-induced corrosion in reinforced concrete, in: C.L.Page, P.B.Bamforth and J.W.Figg (ed.), Corrosion Of Reinforcement In Concrete Construction, pp.88 (The Retail Society of Chemistry, Thomas Gram House, Science Park, Cambridge). Liu T., and Weyers R.W. (1997), Cement and Concrete Research, 28(3):pp.365. Livingston L.G.(1986), Diagnosis of building condition by neutron gamma ray technique, ASTM, 165, ASTM STP 901 Philadelphia, PA. Locke C.E., Siman A. (1980), ASTM STP 713, pp.3. Lunk, P. and Wittmann, F.H. (1998), Rehabilitation of concrete structures in chloridecontaining environment, Rehabilitation of Structures, Proceedings of the 2nd International RILEM/CSIRO/ACRA Conference, Melbourne. p.29-37. Maage M., Helland S., and Carlsen J.E. (1993), Chloride penetration in high performance concrete exposed to marine environment, Symposium on Utilisation of High Strength Concrete. (Lillehammer, Norway). Maage M., Poulsen E., VenneslanD O., Carlsen J.E. (1995), Service life model for concrete structures exposed to marine environment-initiation period, LIGHTCON report No.2.4, STF70 A94082, SINTEFF, (Trondheim, Norway). Maage, M., Helland, S., Puolsen, E., Vennesland, . and Carlsen, J.E. (1996), Service life prediction of existing concrete structures exposed to marine environment, ACI Materials Journal, Vol.93, No.6, p.602-612. Maage M., Helland S. et al (1997), Service life prediction of concrete in marine environment, In: Aage Blankvoll (ed.) Proceedings Of International Conference-Repair Of Concrete Structures, From Theory To Practice In A Marine Environment pp.177187 (Norway). Mangat P.S. and Molloy B.T., (1994), Magazine of Concrete Research, 46(169):pp.279. Masuda, Y. and Tanono, H., (1993), Prediction model for grecess of concrete carboationn, Proceedings of the 6th internaional conference on the durability of building materials and compaonents, Omiya, Japan, vol.2,pp1152-1161. Mccarter W.J., Ezerim H., and Emerson M. (1992), Magazine of Concrete Research, 44 (158):pp.31. Mejlbro L. (1996), The complete solution of Ficks second law of diffusion with timedependent diffusion coefficient and surface concentration, In: Proc. Of Cementas Workshop On Durability Of Concrete In Saline Environment , (Danderyd, Sweden). Meyer A. (1969),Inverstigations on the carbonation of concrete, 5th Conf on Chem of cement, vol3. Pp394-401.

64

Service life prediction of reinforced concrete structures

Morinaga, S. (1988), Prediction of service lives of reinforced concrete buildings based on rate of corrosion of reinforcing steel, Special Report of Institute of Technology, Shimizu Corporation, No.23. June 1988. Morinaga, S., Irino, K., Ohta, T. and Arai, H. (1994), Life prediction of existing reinforced concrete structures determined by corrosion, Proceedings of the International Conference on Corrosion and Corrosion Protection of Steel in Concrete, Sheffield, UK, 1994. Edited by R. N. Swamy, Sheffield Academic Press. Vol.1, p.603618. Nagataki, S., Ohga, H., Kim, EK. (1986), Effect of curing conditions on the carbonation of concrete with fly ash and the corrosion of reinforcement in long-term test, Proceedings of 2nd CANMET/ACI International Conference on Fly Ash, Silica Fume, Slag, and Natural Pozzolans in Concrete, Madrid,. Ed. by Malhotra V.. ACI SP 91. Vol.I, pp.521-540. Nagataki S., Otsuki N., Wee T.H. and Nakashita K., (1993), Condensation of chloride ions in hardened cement matrix materials and on embedded steel bars, ACI Materials Journal, 90(4):3pp.23. Nilsson, L.O. (1992), A theoretical study on the effect of non-linear chloride binding on chloride diffusion measurements in concrete, P-92:13, Division of Building Materials, Chalmers University of Technology, Gteborg, Sweden. Nilsson L.O., Poulsen E., Sandberg P. and Sorensen H.E. (1996), Chloride penetration into concrete, state of the art, HETEK Report, AEC-Chalmers-Cementa. Ohama Y., Demura K. and Satoh J. (1995), Behaviour of chloride ions in unmodified and polymer-modified mortars, Proc. Of international conference on concrete under severe conditions, Japan, E&FE Spon. London. Ohgishi S. and Ono H., (1983), Study to estimate depth of neutralisation of concrete members, Cement Assoc. Japan Rev., pp.168-170. Page C.L., Short N.R., and El-Terras A. (1981), Diffusion of chloride ions in hardened cement pastes, Cement and Concrete Research, 11(3):pp.359-406. Page C.L., Buenfeld N.R., and Newman J.B., (1986a), Cem.Conc.Res., 16:pp.79. Page C.L., Lamber P., and Vassie P.R.W., (1991), Materials and Structures, 24(142):pp.243. Papadakis, V., Vayenas,C., and Fardis M.,(1991a), Physical and chemical characteristics affecting the durability of concrete, ACI Materials Journal, 88(2):186196. Papadakis, V., Vayenas,C., and Fardis M.,(1991b), Fundamental modeling and experimental investigation of concrete carbonation, ACI Materials Journal, 88(4):363373. Parrott, L.J., (1987),a review of carbonation in reinforced concrete, Cement and Concrete Association, Building Research Establishment, pp67.

65

Service life prediction of reinforced concrete structures

Parrott, L. and Chen, Z. (1991), Some factors influencing air permeation measurement in cover concrete, Materials and Structures, Vol.24, p.403-408. Pereira C.J., and Hegedus L.L., (1984), Siffusion and reaction of chloride ions in porous concrete, In: Proceedings Of The 8th International Symposium On Chemical Reaction Engineering, pp427, Publication Series No.87, (Edinburgh, Scotland). Pettersson K.(1993), The chloride threshold value and the corrosion rate in reinforced mortar specimens, in: Chloride Penetration Into Concrete Structure, (Chalmer University of Technology, Sweden, January). Popovics S., Simeonov Y., Boshinov G., Barovsky N. (1983), Durability of reinforced concrete in sea water, in: A.P.Crane (ed.), Corrosion Of Reinforcement In Concrete Construction, pp.19, (Ellis Horwood, Chichester). Poulsen E.(1990), The chloride diffusion characteristics of concrete: Approximate determination by linear regression analysis, Nordic Concrete Research No.1, Nordic Concrete Federation. Poulsen E. (1996), Estimation of chloride ingress into concrete and prediction of service lifetime with reference to marine RC structures, In: Proceedings Of Cements Workshop On Durability Of Concrete In Saline Environment, (Danderyd, Sweden). Poulsen E. (1997), Four-parametric description of marine exposure and concretes response to its chloride intensity, In: Aage Blankvoll (ed.) Proceedings Of International Conference-Repair Of Concrete Structures, From Theory To Practice In A Marine Environment pp.189-199 (Norway). Power t., (1962), a hypothesis on cabonation shrikage, J PCAR & D Lab., May pp.4050. Provebio E. and Carassiti F. (1997), Cement and Concrete Research, 27 (8):pp.1213. Purvis, R.L., Graber, D.R., Clear, K.C., and Markow, M.J. (1992), A Literature Review of Time-Deterioration Prediction Techniques, SHRP-C/UFR-92-613, National Research Council, Washington, D.C., U.S.A. Purvis, R.L., Babaei, K., Clear, K.C. and Markow, M.J. (1994), Life - cycle cost analysis for protection and rehabilitation of concrete bridges relative to reinforcement corrosion, SHRP-S-377, National Research Council, Washington, D.C., U.S.A. Reardon, E.J. and Dewaele, P., (1990), Chemical Model for the carbonation, Journal of the American Ceramic Society, vol.73 (6): pp1681-1690. Rechberger P.(1985), Electrochemical determination of whole chloride diffusion coefficients, Zement-Kalk-Gips, 38: pp679-684. Rhodes J.R. (1977), In Situ Determination Of The Chloride Content Of Portland Cement Concrete In Bridge Decks-Feasibility Study, FHWA-RD-77-26, pp.114, (Washington, D.C. FHWA, U.S.Department of Transportation).

66

Service life prediction of reinforced concrete structures

Rhodes J.R.(1980), In Situ Determination Of The Chloride Content Of Portland Cement Concrete In Bridge Decks, FHWA-RD-80-030, pp.59, (Washington, D.C. FHWA, U.S. Department of Transportation). Richardsson, M.G. (1988), Carbonation of reinforced concrete: Its causes and management, CITIS Ltd., Dublin, pp205. RILEM technical committee 12-CRC, corrosion of reinforcemtn and prestressing tendons, A, State of the art report, Meteriaux et Constructions, 9 (51):pp187-206. RILEM concrete permanent committee 18-CRC, measuremetn o fhardened concrete carbonation depth, Meteriaux et Constructions, 17(102):pp435-440. Rodriguez P., Ramirez E., Gonzalez J.A., (1994a), Method for studying corrosion in reinforced concrete, Magazine of Concrete Research, 46:pp.81. Roy, D.M., Kumar A. and Rhodes J.P. (1986), Diffusion of chloride and cesium ions in portaland cement pastes and mortars containing blast furnace slage and fly ash, in: Use of fly ash, silica fume, slag and natural pozzolans in concrete, Proc. 2nd Inter. Conf., Madrid, ACI SP-91,pp1423-1444. Saetta, A.V., Scotta, R.V. and Vitaliani, R.V. (1993), Analysis of chloride diffusion nto partially saturated concrete, ACI Materials Journal, Vol.90, No.5, p.441-451. Sagues A.A. and Kranc S.C. (1996), Effect of structural shape and chloride binding on time to corrosion of steel in concrete in marine service, in: C.L.Page, P.B.Bamforth and J.W.Figg (ed.) , Corrosion Of Reinforcement In Concrete Construction, pp.105 (The Royal Society of Chemistry, Thomas Gram House, Science Park, Cambridge). Sagues A. A. and Power R.G. (1997), Corrosion and Corrosion control of concrete structures in Florida-what can be learned, In: Aage Blankvoll (ed.), Proceedings Of International Conference-Repair Of Concrete Structures, From Theory To Practice In A Marine Environment:, pp. 49 (Norway). Sandberg P., and Larsson J., (1993), Chloride binding in cement pastes in equilibrium with synthetic pore solutions, In: Proceedings Nordic Seminar On Chloride Initiated Reinforcement Corrosion In Concrete, pp.13-14, (Gothenburg, Sweden, January). Sandberg P., and Larsson J. (1993a), Chloride binding in cement pastes in equilibrium with synthetic pore solutions as a function of [Cl] and [OH], in: L.O.Nilsson (ed.), Chloride Penetration Into Concrete Structures-Nordic Miniseminar, Publication P93:1, 98, ( Division of Building Materials, Chalmer University of Technology) Sanberg P.(1993a), and Larsson J., Chloride binding in cement pastes in equilibrium with synthetic pore solutions as a function of [Cl] and [OH], in: L.O.Nilsson (ed.), Chloride Penetration Into Concrete Structures-Nordic Miniseminar, Publication P-93:1, 98, (Division of Building Materials, Chalmer University of Technology). Sayward, J.M. (1984), Salt Action on Concrete, Special Report 84-25. US Army Corps of Engineers, Cold Regions Research & Engineering Laboratory. Schiessl P. (1987), Influence Of The Composition Of Concrete On The Corrosion Protection Of The Reinforcement, American Concrete Institute Special Publication SP100, In:Concrete Durability, Vol.2:pp.1633.
67

Service life prediction of reinforced concrete structures

Scholz, E. and Wierig, H-J. (1984), Carbonation of fly ash concrete, Proceedings of RILEM Seminar on the Durability of Concrete Structures under Normal Outdoor Exposure, Hannover, p.258-274. Schubert, P. (1987), Carbonation behavior of mortars and concretes made with fly ash, Proceedings of the International Conference on Concrete Durability, Atlanta, ACI SP. 100. Vol.2, pp.1945-1962. Slegers P., and Rouxhet P.(1976),Carbonation of the hydration products of tricalcium silicate, Cement and concrete research, 6(3):pp381-388. Smolczyk, H., (1968) discussion of principal paper on carbonation of concrete, 5ICCC, Tokoy, V.3,pp369-386. Song G., and Shayan A. (1999), Monitoring of steel corrosion in concrete, NSRP Report (for Austroads), (to be published). Steen P.E. (1997), Systematic bridge inspection, condition assessment and service life prediction, In: Aage Blankvoll (ed.), Proceedings Of International Conference-Repair Of Concrete Structures, From Theory To Practice In A Marine Environment, pp.201212, (Norway). Stoltzner E., Knudsen A., and Buhr B. (1997), Durability of marine structures in Denmark, In: Aage Blankvoll (ed.), Proceedings Of International Conference-Repair Of Concrete Structures, From Theory To Practice In A Marine Environment , pp.59 (Norway). Swamy R.N., Hamada H., Laiw J.C. (1994), A critical evaluation of chloride penetration into concrete in marine environment, in: Swamy R.N. (ed.) Proceedings Of The Conference On Corrosion And Corrosion Protection Of Steel In Concrete, Vol.1, pp.404419, (University of Sheffield, Sheffield). Takewake K., Mastumoto S. (1988), Quality And Cover Thickness Of Concrete Based On The Estimation Of Chloride Penetration In Marine Environment, ACI, SP-109, Detroit, USA. Tang L., and Nilsson L.O., (1992), Effect of conditions on chloride diffusivity in silica fume high strength concrete, In: Proceedings Of The 9th International Congress On The Chemistry Of Cement, Vol.V, pp.100 New Delhi. Tang L., and Nilsson, (1992a), Rapid determinatin of chloride diffusivity of concrete by applying an electric field, ACI Materials Journal, 49(1):49-53). Tang, L., and Nilsson, L.O. (1992b), Chloride binding capacity and binding isotherms of OPC pastes and mortars, Cement and Concrete Research, Vol.23, pp.247-253. Tang L., and Nilsson L.O.(1993), A rapid method for measuring chloride diffusivity by using an electrical field, in Nilson L.O. ed. Chloride penetration into concrete structures-Nordic Miniseminar, Division of Building Materials, Chalmers University of Technology, Publication P-93:1,pp26-35.

68

Service life prediction of reinforced concrete structures

Tang L., and Nilsson L.O, (1995), A discussion of the paper Calculation of chloride diffusivity in concrete from migration experiments in non-steady-state condition by Andrade C. et al, Cement and Concrete Research, 25(5):pp1133-1137. Tang, L., and Nilsson, L.O. (1995a), Transport of ions, in Performance Criteria for Concrete Durability State of-the-art report prepared by RILEM Technical Committee TC 116-PCD, edited by J. Kropp and H.K. Hilsdorf, RILEM Report 12, E&FN SPON, London. Tang L.(1996), Chloride Transport In ConcreteMeasurement And Prediction, PhD.Thesis, Dept. of Building Materials, Chalmers University of Technology, Sweden, Publication P-96:6. Tang L.(1996a), Electrically accelerated methods for determining chloride diffusivity in concrete, Magazine of concrete research, 48(176):pp173-179. Thomas M. (1996), Chloride threshold in marine concrete, Cem. Concr. Research, 26 (4): pp.513-519. Tritthart J.(1989), Cement and Concrete Research, 19: pp.586. Tritthart J.(1989a), Cement and Concrete Research, 19: pp.683. Tuutti, K. (1977), Stlets krrosionsfrlopp i ospruken betong en hypotes, CBI Forskning Fo 4:77. Tuutti K.,(1979), Service life of concrete structures-corrosion test methods, Swedish Cement and concrete reserch institute, pp.227-247. Tuutti, K. (1982), Corrosion of Steel in Concrete, Swedish Cement and Concrete Research institute (CBI) Research Report Fo 4:82. Tuutti K.(1982a), Corrosion Of Steel In Concrete, Swedish Cement and Concrete Research Institute, Stockholm, report F04, 468. Uji K., Matsuoka Y., Maruya T. (1990), Formulation of an equation for surface chloride content due to permeation of chloride, In: Proceedings Of The Third International Symposium On Corrosion Of Reinforcement In Concrete Construction, (Elsevier Applied Science, London, UK), pp.258-267. Ushiyama H., and Goto S. (1974), Diffusion of various ions in hardened portand cement paste, in: Proc. 6th congress on the chemistry of cement, Moscow, Vol.II-1,pp331-337. Vesikari, E. (1986), Service life design of concrete structures with regard to frost resistance of concrete, Nordic Concrete Research, No.5, 1986, p.215-228. Volkwein A.(1993), Convection of chlorides into concrete due to hydration suction and capillary suction, In: Proceedings 6th int. Conf. On durability of building materials and components, pp.279 (Omiya, Japan). West R.E. and Hime W.G., (1985a), Materials Performance, July: pp.29.

69

Service life prediction of reinforced concrete structures

Weyers, R.E., Fitch, M.G., Larsen, E.P., Al-Qadi, I.L., Chamberlin, W.P. and Hofman, P.C. (1994), Concrete bridge protection and rehabilitation: Chemical and physical techniques Service life estimates, Strategic Research Program, SHRP-S-668, National Research Council, Washington, D.C. Weyers, P.E. (1998), Servicelife model for concrete structures in chloride laden environments, ACI Materials Journal, Vol.95, No.4, p.445-453. Whiting D.(1981), Rapid determination of the chloride permeability of concrete, Report No. FHWA/RD-81/119, August, NTIS DB No. 82140724. Wittmann, F.,a nd Brieger, L. (1986), Simulation of carbonation of concrete, Esslingen International Colloquium, pp635-640. Xu A.(1990), The Structure And Some Physical Properties Of Cement Mortar With Fly Ash, Thesis, Dept. of Building Materials, Chalmers University of Technology, Sweden, Publication 90:9. Xu, A. (1995), Influence of alkali on air permeability and carbonation of concretes containing fly ash, Supplementary Papers of the 5h CANMET/ACI International Conference on Fly Ash, Silica Fume, Slag and Natural Pozzolans in Concrete, Milwaukee, U.S.A. Xu, A., and Sarkar, S.L., (1995a), Durability of concrete, in Cement Research Progress 1993, edited by L.J. Struble, American Ceramic Society, 1995, p.146-184. Xu Y.(1997), Cement and Concrete Research, 27 (12):pp.1841. Yonezawa T. (1988) , Pore Solution Composition And Chloride Induced Corrosion Of Steel In Concrete, British Ph S Thesis, Victoria University of Manchester, Corrosion and protection centre. Zhang, R., Edamoto, H., Nakazawa, T., Imai, F. and Shinnishi, N, (1998), Experimental evaluation of material deterioration of aged RC bridges, Rehabilitation of Structures, Proceedings of the 2nd International RILEM/CSIRO/ACRA Conference, Melbourne. p.246-257.

70

INFORMATION RETRIEVAL

Austroads (2000), Service Life Prediction of Reinforced Concrete Structures, Sydney, A4, 79pp, AP-T07/00

KEYWORDS: Concrete, Durability, Bridge, Service Life, Prediction, Modelling, Chloride, Carbonation, Steel, Reinforcement, Corrosion ABSTRACT: Concrete is the dominant material of construction in the more than 12,000 bridges and major culverts managed by road authorities in Australia and New Zealand. With an ageing asset, concrete durability is becoming an increasingly significant issue for bridge owners. Corrosion damage to affected structures mainly results from chloride ingress and carbonation. While considerable work has been done in the field of service life prediction of reinforced concrete structures, the prediction theory has not been well established. The report reviews and discusses existing models for service life prediction, and provides a basis for future work in the area. A proposed strategy to address problems with existing techniques is presented.

AUSTROADS PUBLICATIONS
Austroads publishes a large number of guides and reports. Some of its publications are: AP-1/89 Rural Road Design AP-8/87 Visual Assessment of Pavement Condition Guide to Traffic Engineering Practice AP-11.1/88 Traffic Flow AP-11.2/88 Roadway Capacity AP-11.3/88 Traffic Studies AP-11.4/88 Road Crashes AP-11.5/88 Intersections at Grade AP-11.6/93 Roundabouts AP-11.7/88 Traffic Signals AP-11.8/88 Traffic Control Devices AP-12/91 AP-13/91 AP-14/91 AP-15/96 AP-17/92 AP-18/96 AP-22/95 AP-23/94 AP-26/94 AP-29/98 AP-30/94 AP-34/95 AP-36/95 AP-38/95 AP-40/95 AP-41/96 AP-42/96 AP-43/98 AP-44/97 AP-45/96 AP-46/97 AP-47/97 AP-48/97 AP-49/97 AP-50/97 AP-51/98 AP-52/97 AP-53/97 AP-54/97 AP-55/98 AP-56/98 AP-57 & 58/98 AP-59/98 AP-60/98 AP-61/99 AP-62/99 AP-63/00

AP-11.9/88 AP-11.10/88 AP-11.11/88 AP-11.12/88 AP-11.13/95 AP-11.14/99 AP-11.15/99

Arterial Road Traffic Management Local Area Traffic Management Parking Roadway Lighting Pedestrians Bicycles Motorcycle Safety

Road Maintenance Practice Bridge Management Practice Guide to Bridge Construction Practice Australian Bridge Design Code Pavement Design RoadFacts 96 Strategy for Pavement Research and Development Waterway Design, A Guide to the Hydraulic Design of Bridges, Culverts & Floodways Strategy for Structures Research and Development Austroads Strategic Plan 19982001 Road Safety Audit Design Vehicles and Turning Path Templates Adaptions and Innovations in Road & Pavement Engineering Guide to Field Surveillance of Quality Assurance Contracts Strategy for Ecological Sustainable Development Bitumen Sealing Safety Guide Benefit Cost Analysis Manual National Performance Indicators Asphalt Recycling Guide Strategy for Productivity Improvements for the Road Transport Industry Strategy for Concrete Research and Development Strategy for Road User Cost Australia at the Crossroads, Roads in the Community A Summary Roads in the Community Part 1: Are they doing their job? Roads in the Community Part 2: Towards better practice Electronic Toll Collection Standards Study Strategy for Traffic Management Research and Development Strategy for Improving Asset Management Practice Austroads 1997 Bridge Conference Proceedings Bridging the Millennia Principles for Strategic Planning Assessing Fitness to Drive Cities for Tomorrow Better Practice Guide & Resource Document Cities for Tomorrow CD Guide to Stabilisation in Roadworks Australia Cycling 1999-2004 The National Strategy e-transport The National Strategy for Intelligent Transport Systems Guide to the Selection of Road Surfacings

These and other Austroads publications may be obtained from: ARRB Transport Research Ltd 500 Burwood Highway VERMONT SOUTH VIC 3131 Australia Telephone: Fax: Email: Website: +61 3 9881 1547 +61 3 9887 8144 donm@arrb.org.au www.arrb.org.au

or from road authorities, or their agent in all States and Territories; Standards New Zealand; Standards Australia & Bicycle New South Wales.

Das könnte Ihnen auch gefallen