You are on page 1of 65

Draft Report on

Basic Understanding of Various Deterioration


Mechanisms in Concrete Structures: With
Emphasis on Rebar Corrosion

Prof. B. Bhattacharjee & Kamal Kant Jain


Department of Civil Engineering, IIT Delhi
Hauz Khas, New Delhi, India - 110016

Table of Contents
Table of Contents .........................................................................................................ii
List of Tables .............................................................................................................. iii
List of Figures............................................................................................................. iii
Chapter 1: Introduction ................................................................................................1
Chapter 2: Literature Review .......................................................................................6
2.1 Carbonation.........................................................................................................7
2.1.1 Effect of carbonation of concrete...............................................................8
2.1.2 Primary factors influencing carbonation..................................................9
2.2 Chloride Ingress................................................................................................11
2.2.1 Chloride penetration in concrete.............................................................11
2.2.2 Primary factors affecting the chloride ingress .......................................13
2.3 Corrosion...........................................................................................................16
2.3.1 Free Energy consideration .......................................................................18
2.3.2 Corrosion Kinetics ....................................................................................20
2.3.3 Primary factors affecting steel corrosion in concrete............................23
2.3.4 Corrosion diagnosis and measurement methods ...................................25
Open circuit potential (OCP) measurements: .................................................25
Concrete resistivity measurement:...................................................................26
Linear polarization resistance (LPR) measurement:.......................................28
Tafel Extrapolation:.........................................................................................30
Electrochemical impedance spectroscopy (EIS): ............................................31
2.3.5 Damages due to corrosion ........................................................................33
Common damages in RC Structures:...............................................................33
Stress corrosion cracking (SCC): ....................................................................34
Hydrogen embrittlement (HE): ........................................................................35
2.3.6 Corrosion control techniques...................................................................36
Patch Repair: ...................................................................................................36
Cathodic Protection:........................................................................................37
Electrochemical chloride extraction (ECE) and Realkalization: ....................40
Inhibitor Application: ......................................................................................41
Coatings:..........................................................................................................41
Design Considerations: ...................................................................................42

ii

2.4 Alkali aggregate reactions (AAR).....................................................................46


2.5 Acid Attack ........................................................................................................49
2.6 Sulphate attack..................................................................................................51
2.7 Freeze - Thaw damage......................................................................................53
2.8 Fatigue, Shrinkage and Creep..........................................................................57
2.8.1 Fatigue........................................................................................................57
2.8.2 Shrinkage ...................................................................................................58
2.8.3 Creep...........................................................................................................59

List of Tables
Table 2.2.1: Primary factors affecting chloride ingress...............................................14
Table 2.3.1: Electrode potentials for anodic and cathodic corrosion reactions ...........20
Table 2.3.2: Primary factors affecting steel corrosion in concrete ..............................25
Table 2.3.3: Corrosion condition related with HCP measurements [7].......................26
Table 2.3.4: Corrosion risk from resistivity.................................................................28
Table 2.3.5: Guidelines for rebar condition assessment using icorr ..............................31
Table 2.3.6: Relative volume ratios for various iron oxides in comparison to iron ....33

List of Figures
Fig. 2.3.1: Reinforced concrete performance against corrosion deterioration.............16
Fig. 2.3.2: Schematic presentation of corrosion process in reinforced concrete .........17
Fig. 2.3.3: Cathodic and anodic over potentials due to polarization [5]......................21
Fig. 2.3.4: Open circuit potential (OCP) measurement [6]..........................................26
Fig. 2.3.5: Circuit for electrical resistance measurements ...........................................27
Fig. 2.3.6: Guard ring test set-up .................................................................................29
Fig. 2.3.7: Schematic polarization curves showing Tafel extrapolation [14] ..............30
Fig. 2.3.8: Equivalent circuit diagram for simple corrosion system [18] ....................32
Fig. 2.3.9: Schematic Nyquist Plot ..............................................................................32
Fig. 2.3.10: Schematic application diagram for CP using impressed current..............38
Fig. 2.3.11: Schematic application diagram for CP using sacrificial anodes ..............40
Fig. 2.3.12: Schematic diagram for application of ECE ..............................................40
Fig. 2.4.1: Effect of AAR on concrete [2] ...................................................................47
Fig. 2.5.1: Effect of Acid Attack on concrete [3] ........................................................49

iii

Fig. 2.7.1: Types of freeze - thaw damage [5] .............................................................55

iv

Chapter 1: Introduction
Concrete is the most versatile material used in the construction industry worldwide.
Benefits such as mould ability, robustness, fire resistance, readily availability and
durability along with allowed variations in mix design to obtain desirable physical
properties suitable to various applications have made concrete ideal for use in bridge
construction. Concrete in bridges is used mainly in the form of reinforced or
prestressed concrete, mainly using steel for reinforcement purposes. This steelconcrete system in general is highly durable and is used extensively.
But because of severe environment, exposure to harmful chemicals, unplanned higher
loads, inadequate materials selection and poor workmanship etc., Concretes ability to
perform well above its serviceability limits over planned design period i.e. durability
of concrete is adversely affected.
Under action of one or more of above mentioned detrimental factors, various
damaging mechanisms may occur in concrete bridges, of which the commonly
detected are:
Carbonation
Chloride Ingress
Corrosion of reinforcement/prestressing steel
Alkali aggregate reactions
Acid Attack
Sulphate attack
Freeze - Thaw damage
Shrinkage, Fatigue and Creep
Carbonation of concrete occurs when carbon dioxide (CO2) from the atmosphere
reacts with calcium hydroxide (Ca(OH)2) in hydrated concrete in the presence of
moisture(H2O) and alkalis (Na+, K+ etc.) forming calcium carbonate (CaCO3). This
reaction causes reductions in the alkalinity of the concrete due to acidic nature of CO2
gas. Reduction in alkalinity impairs the natural passivating nature of concrete i.e.
capability of forming a protective oxide layer around steel reinforcement. Carbonation
starts on the surface and gradually penetrates the concrete, when carbonation front

reaches reinforcing or prestressing steel depassivation occurs and thereafter allows


corrosion to take place under presence of corroding substances.
Chloride ingress can be termed as a process in which chlorides from surroundings,
permeates through the porous concrete mainly by the means of diffusion and/or
absorption processes. The chloride ions attack the passive layer and therefore lead to
depassivation and corrosion initiation when present in sufficient concentration at rebar
level; this concentration level is commonly known as chloride threshold. Furthermore
chlorides available act as catalyst for corrosion process as they are not consumed in
the process and allow the corrosion process to proceed quickly. This is a great
concern in marine splash zones.
Corrosion in concrete occurs mainly because of chlorides and/or CO2 ingress, as they
depassivate the protective oxide layer formed around steel embedded in naturally
alkaline concrete. Corrosion can be defined concisely as loss of steel due to oxidation
of iron (electrochemical process) followed by reactions producing iron oxides,
resulting in production of high voluminous of rust.
Corrosion causes several problems and deteriorations in bridges which ultimately lead
to the premature failure with respect to designed service life. Most commonly
detected deteriorations are:
Reduction in effective diameter of reinforcement steel due to consumption of
iron, leading to reduction in load carrying capacity
Rust stains on concrete surface impairing the aesthetics
Production of high voluminous rust products which cause tension cracks or
spalling in surrounding concrete
In case of prestressed concrete bridges corrosion can lead to stress corrosion
cracking as strands are typically under high tension
Hydrogen embrittlement in prestressing steel may take place which further can
lead to brittle failure of system
Alkali- aggregate reaction (AAR) also called sometimes as alkali- silica reaction
(ASR), is chemical reaction that takes place between the alkali hydroxides in the
hydrated concrete and certain siliceous aggregates. The reaction product absorbs
2

water and in some situations expansion stresses sufficient to crack the concrete can be
developed in the concrete. The reaction is usually slow and it can take many years for
visual damage to develop under normal conditions.
Acid attack on concrete is promoted because of concretes alkaline nature and
reaction with acids may occur instantaneously at surface if concrete is exposed to acid
smoke, acid rain and aqueous solutions containing acids such as sulphuric, nitric,
hydrochloric and acetic acid, etc. This reaction mainly causes the dissolution of
Ca(OH)2 in presence of acids forming calcium salts. Therefore acid attack damages
concrete as components of the cement paste break down due to consumption of
Ca(OH)2 during contact with acids, leading to leaching and reduction in strength.
Presence of sulphuric acid is a highly damaging as it results in a combination of acid
attack as well as sulphate attack.
Sulphate attack on concrete can be defined as chemical reaction between the sulphate
(SO4-2) ions and hydrated calcium aluminates and/or the Ca(OH)2 components of
hardened concrete in the presence of water. The products resulting from these
reactions are calcium sulphoaluminate hydrate (Ettringite), and calcium sulphate
hydrate (Gypsum). These solids have a very high volume than the solid reactants and
as a consequence cause expansion, tensile stresses are produced that may result in
breakdown of the hardened paste and ultimately in breakdown of the concrete.
Freeze Thaw damage occurs in bridges when wet concrete is exposed to cycles of
freezing and thawing. The principle mechanism is a consequence of the expansiveness
of water as it freezes within the confined concrete pores. This is due to the significant
(around 9%) increase in volume as water turns to ice. This generates hydraulic
pressure in the pores, exposing the pore walls to stress that can cause severe
disruption to the concrete pore structure, ultimately leading to loss of durability due to
reduction in strength and development of cracks in concrete.
Cracking of concrete may also take place in bridges because of following factors:

Shrinkage (Volumetric decrease with time due to drying process of concrete)

Fatigue (Material deterioration due to repeated variations of stress or strain)


Creep (Time dependent deformation due to sustained load)
3

If concrete is subjected to one or more of the abovementioned damaging mechanisms;


their adverse affects make bridges more prone to development of various defects, of
which commonly detected are:
Change in microstructure
Strength reduction
Leaching/Staining
Debonding of reinforcement
Loss of prestress
Crack development
Spalling
So it can be commented and reasoned that synergetic effect of various damaging
mechanisms is responsible for occurrence of various defects as reported. And it can
therefore be reasoned that presence of the various deteriorating factors, will impair the
performance of bridge and so the ability to be functional without risk of failure till
end of the planned service life of the concrete bridge i.e. reduces age of the bridges.
But corrosion of steel reinforcement/prestressing is a single major concern that
primarily needs to be taken care of in concrete bridges as corrosion is almost
inevitable and has several severe consequences as mentioned earlier. Therefore to say
that even corrosion alone will extensively impair durability of concrete bridges would
not be an overstatement.
Damages caused due to corrosion call for unplanned repair works for further use of
the affected bridges for risk free functioning. This is a major economic issue
worldwide because of high repair costs. A study by FHWA/NACE in US estimated
direct cost of corrosion as $8.3 billion per year for bridges, over the next 10 years
from which $3.8 billion is for replacement purposes, $2.0 billion for maintenance of
existing decks, $2.0 billion for maintenance of existing substructure components and
$0.5 billion for painting bridges [1].
If the time at which the serviceability limit will be reached in a given RC structure
subjected to corrosion and other simultaneous damaging mechanisms, is reasonably
4

known then effective monitoring and maintenance systems can be used in time to
avoid sudden need for repair. It will help in cutting down the high costs for unplanned
repairs and will cause fewer disruptions to proper functioning of bridge over their
design life.
But to model durability and hence serviceability analytically in a deterministic
manner is a difficult task as durability is affected by various mechanisms discussed.
Therefore it will not only require good knowledge and understanding of various
streams of science and engineering such as: Concrete engineering, corrosion
engineering, fracture mechanics, solid mechanics etc.; but also tools capable to
combine these effects reasonably accurately. Likewise it will not be reasonable if
experimental results alone as obtained in laboratory or at some site are taken as
standards because of temporal and spatial variations in various parameters involved
that will alter the degradation mechanisms significantly.
Therefore it is much necessitated that, a model that includes knowledge of experts
from various fields having significant exposure to problem of durability of bridges
combined with experimental investigation, be used to provide a better insight about
the serviceability status of a given bridge structure. For this purpose, diagnosis of
extent of degradations can be made possible with judicious combination of fuzzy
mathematics which can transform experts knowledge from qualitative to quantitative
and models for reinforced/prestressed concrete behavior from experimental data
obtained using non destructive tests (NDT) and/or semi destructive tests (SDT). And
furthermore modeling prognosis will allow us to estimate the remaining service life of
a bridge structure.
Once the extent of various degradations and remaining service life of a bridge are
known appropriate measures can be implemented to improve or sustain the health of
the structure and a maintenance methodology can be developed.
References
1. Virmani Y. P., Payer J.H. and Koch G.H. et.al., Corrosion costs and preventive
strategies in the United States, FHWA Publication, No. FHWA-RD-01-156,
(2002).
5

Chapter 2: Literature Review


As mentioned in introduction, concrete structures are usually subjected to various
combinations of deteriorating mechanisms which lead to degradation in performance
of structure. These degradations are inevitable in concrete structures and therefore
must be accounted for during design phase for realistic assessment of service life of
structure.
But various limitations such as, incomplete understanding on spatial and temporal
variations of degradation mechanisms, synergetic effects of various degradations etc.,
make service life assessment task complex to undertake at design stage. Same
concerns needs to be taken care of even if the service life assessment is carried out
after the signs of damage in the structure have been observed.
For Bridges the structural service life thus assessed is mandatory for development of
management systems. These systems will allow the stakeholders to optimally utilize
the resources and funds available, to minimize the maintenance and repair costs.
So there is a need to develop the basic understanding of various damage mechanisms
involved, existing service life assessment techniques and bridge management systems
already in use, to arrive at a realistic bridge management system. This then can
logically be adopted for the abovementioned purpose.
Literature review reported in following sections will helps us to develop the
understanding and will provide essential information on:
Commonly observed degradation mechanisms in concrete
Service life assessment models in existence
Existing Management systems
And furthermore shall provide with the course for the actions to be taken for
development of an efficient management system.

2.1 Carbonation
In concrete, under normal conditions, reinforcing steel does not corrode. As in the
highly alkaline environment of concrete, iron is oxidized to Fe2O3 / Fe3O4 / -FeOOH.
These oxides provide a tightly adhering passive oxide film (passive layer) on the
surface of the steel reinforcement, which limits the access of oxygen and moisture to
rebar and thus arrests corrosion. Drop in pH might cause depassivation of this film,
thus allowing corrosion to proceed [1, 2].
In moist environments, carbon dioxide CO2 present in the air, diffuses through pores
of concrete and forms an acid aqueous solution that can react with alkaline
constituents of concrete, mainly calcium hydroxide Ca(OH)2 formed during hydration
process of cement paste and tends to neutralize the alkalinity (Ca(OH)2 in the porewater of the concrete is the main reason for its alkalinity) of concrete, this
phenomenon commonly is termed as carbonation of concrete and is commonly
detected in concrete. The reactions in presence of alkalis and moisture, involved in
carbonation process in concrete can be written as [1, 3]:
CO2 + H2O
H2CO3 + Ca(OH)2

H2CO3

(2.1.1)

CaCO3 + 2H2O

(2.1.2)

This can be summed up as:


CO2 + Ca(OH) 2

H2O
NaOH

CaCO3 + H2O

(2.1.3)

In this process CO2 diffuses through the concrete pores and the rate of movement of
the carbonation front can be approximated reasonably using laws of diffusion, which
states that the rate of movement is inversely proportional to the distance from the
surface as presented in equation below [3]:

dx D0
=
dt
x

(2.1.4)

Where D0 = CO2 diffusion coefficient


x = Distance from surface of concrete
This gives variation of CO2 penetration depth called commonly as carbonation depth
with respect to time. And the relation which is obtained by integration of Eqn 2.1.4

under the assumption that diffusion coefficient is constant in space (homogenous


concrete) and time can be written as following:

d = Kt

(2.1.5)

Where d = Depth of carbonation (mm)


t = Time (yr)
K=Carbonation coefficient (mm/yr0.5)
K can be assumed as a measure of the rate of penetration of carbonation for given
concrete and environmental conditions. This relationship (Eqn 2.1.5) is widely
accepted for depth of CO2 penetration in concrete [1, 3-4].

2.1.1 Effect of carbonation of concrete


Carbonation does not cause any damage to the concrete itself, although it may cause
the concrete to shrink. Indeed, in the case of concrete obtained with Portland cement,
it may even reduce the porosity and lead to an increased strength [1]. However,
carbonation has important effects on corrosion of embedded steel. As main
consequence of carbonation is the drop in pH of the concrete pore solution to the pH
range of 8-9, hence approaching neutrality, which otherwise is alkaline normally
having the pH greater than 12.5 [1, 3]. This leads to depassivation of protective oxide
layer around steel reinforcement formed due to alkaline nature of concrete and thus
impairs the electrochemical protection of rebars against corrosion, and allow
reinforcement corrosion initiation, which is a big problem in bridges as will be
discussed in detail in following sections.
If chlorides are not present in concrete, the pore solution following carbonation is
composed of almost pure water. This means that the steel in humid carbonated
concrete corrodes as if it was in contact with water. Presence of chlorides in the
concrete, even in such small quantities that in themselves they would not give rise to
corrosion leads to a more damaging situation. As depassivation of reinforcement by
carbonation will allow the present chlorides to induce corrosion and thus damage the
structure. The presence of a small amount of chlorides in concrete may be due to the
use of raw materials (water, aggregates) containing these ions or to the penetration of
chlorides from the external environment (seawater, de-icing salts). The complete
corrosion process will be discussed in detail in following sections [1].

The carbonation reaction starts at the external surface and penetrates into the concrete
producing a low pH front. The measurement of the depth of carbonation is normally
carried out by spraying an alcoholic solution of phenolphthalein on a freshly broken
face. The areas where pH is greater than 9 take on a pinkish color typical of
phenolphthalein in basic environment, while the color in carbonated areas remains
unchanged [3-5].

2.1.2 Primary factors influencing carbonation


Relation obtained using Ficks law is an approximation as due to the carbonation
process, concrete pore structure is modified as carbonation proceeds. Deviations
prevail due to cracks, changes in concrete composition and moisture levels with
depth. Therefore rate of carbonation depends mainly on environmental factors and
properties of the concrete. Following factors have significant influence on rate at
which carbonation will take place in concrete and thus govern the depth of carbonated
concrete as presented in table 2.1.1:

Factor

Effect on carbonation of concrete

CO2 Concentration

As the CO2 content in the surrounding air increases, the

in surrounding air
Permeability of
concrete

carbonation rate increases


The diffusion process takes place through permeable pores
in the concrete, therefore higher permeability leads to high
rate of carbonation
The rate of diffusion of CO2 consequently decreases with
an increase in water content of the concrete until it becomes

Degree of concrete

zero in water-saturated concrete. As diffusion of CO2

pore saturation

within concrete is facilitated only through the aerated pores


The carbonation reaction occurs only in the presence of
water therefore it becomes negligible in dry concrete
Carbonation rate and depth of carbonation in concrete may

Wetting & drying


cycles

be variable even in different parts of a single structure. If


the structure is subjected to periodic cycles of wetting and
drying as it will affect the saturation level of concrete
significantly

Wetting of concrete occurs more rapidly than drying,


leading to water saturation of concrete, so more frequent
although shorter periods of wetting cause more reduction in
the rate and depth of CO2 penetration than less frequent and
longer periods. This means that the wetting time, as well as
the frequency and duration of wettingdrying cycles are
important parameters
Cracks in concrete

Cracks in concrete increase the carbonation rate manifolds


and it may provide direct access of CO2 at rebar level

Table 2.1.1: Primary factors influencing carbonation of concrete

References
1. Bertolini L., Elsener B., Pedeferri P. and Polder R., Corrosion of Steel in
Concrete: Prevention, Diagnosis, Repair, Wiley-VCH, Weinheim Publication

(2004).
2. Montemor M.F., Simoes A.M.P. and Ferreira M.G.S. Chloride-induced corrosion
on reinforcing steel: from the fundamentals to the monitoring techniques Cement

& Concrete Composites, Vol. 25, pp. 491502 (2003).


3. Broomfield John P., Corrosion of Steel in Concrete: Understanding, Investigation
and Repair, E&FN Spon. London (1997).
4. Pade C. and Guimaraes M. The CO2 uptake of concrete in a 100 year
perspective, Cement and Concrete Research, Vol. 37, pp. 13481356 (2007).
5. Jung W.Y., Yoon Y.S. and Sohn Y.M., Predicting the remaining service life of
land concrete by steel corrosion Cement and Concrete Research, Vol. 33, pp.

663677 (2003).
6. Apostolopoulos C.A. and Papadakis V.G., Consequences of steel corrosion on
the ductility properties of reinforcement bar Construction and Building

Materials, Article in press (2007).

10

2.2 Chloride Ingress


Chloride contamination of concrete is a frequent cause of corrosion of reinforcing
steel and leads to extensive corrosion damage after depassivation has taken place,
which can occur either due to carbonation and/or due to chlorides having sufficient
concentration at rebar level to cause depassivation. Chloride content in concrete that
alone might lead to depassivation of protective oxide layer formed around
reinforcement and further allowing corrosion to initiate, is commonly termed as
Threshold chloride content [1]. Depassivation in case of chloride ingress is assumed
to occur through formation of hydrochloric (HCl) acid which leads to reduction in
alkalinity of concrete. As chlorides draw the iron ions through the passive layer,
create rust, produce HCl acid and release the chloride ions again in the solution to
continue the attack and therefore chlorides act only as a catalyst for the corrosion
process to start [2,3].

2.2.1 Chloride penetration in concrete


The main reason for chloride contamination of concrete is penetration from the
environment. Chloride penetration in concrete from the environment is mainly a
diffusion process, which produces varying chloride content in the concrete
characterized by high chloride content near the external surface and decreasing
contents at greater depths. The studies worldwide have shown that, in general chloride
profile can be approximated using Ficks second law of diffusion under the following
assumptions [1, 4-6]:
Concentration of the Cl- ions at concrete surface, does not change with time
Diffusion coefficient is independent of time and concrete is homogeneous, so
that diffusion coefficient does not vary through the thickness of the concrete
Concrete does not initially contain chlorides i.e. (C = 0 for x > 0 at t = 0).
These assumptions are rarely met in real structures. As surface concentration might be
subjected to changes based on exposure environment and concrete is not
homogeneous. Also chlorides may already be present in structures due to [7-8]:
Use contaminated mixing water and aggregates (for instance by using water,
sand and gravel extracted from sea)
Addition of chloride based admixtures such as calcium chloride.
Application of chloride based coatings or salts
11

Only in concrete completely and permanently saturated with water can chloride ions
penetrate by pure diffusion. In most situations, other transport mechanisms such as
permeation and convection due to difference in hydraulic pressure (moisture content)
and migration due to difference in electrical potential also contribute to chloride
penetration. For instance, when dry concrete comes into contact with salt water,
initially capillary suction of the chloride solution occurs. Furthermore, binding due to
adsorption or chemical reaction with constituents of the cement paste alters the
concentration of chloride ions in the pore solution [1]. Nevertheless Ficks second law
can be used for a good approximation, considering the total content of chlorides
usually, hence including bound chlorides and can be written as [4-6]:
C
2C
= DApp. 2
t
x

(2.2.1)

This is an initial value boundary problem which on integration under assumed initial
value and boundary conditions gives us:

C ( x, t ) = Cs 1 erf

2 DApp.t

(2.2.2)

Where C(x,t) = Total chloride content at time t at depth x from surface of the concrete
DApp. = Apparent diffusion coefficient for chloride (m2/s)
Cs = Surface chloride content (% by mass of cement or concrete
erf = Error Function: erf ( x) =

e t dt

The apparent diffusion coefficient and the surface chloride content are calculated by
fitting the experimental data to Eqn. (2.2.2) and are often used to describe chloride
profile measured on real structures. DApp. is often used as the parameter that describes
the resistance of concrete to chloride penetration, when performances of different
materials are compared. The lower DApp is, the higher the resistance to chloride
penetration is. It should, however, be observed that, while the diffusion coefficient
obtained from pure (steady-state) diffusion tests can be considered as a property of the
concrete, the apparent diffusion coefficient obtained from real structures also depends
on other factors (such as the exposure conditions or the time of exposure). Therefore,
results obtained under particular conditions, especially during short-term laboratory
tests, may not be applicable to other exposure conditions [1].

12

In principle, if DApp and Cs are known and can be assumed to be constant in time, it is
possible to evaluate the evolution with time of the chloride profile in the concrete and
then to estimate the time t at which chloride threshold will be reached and corrosion
will initiate. Though in general DApp and Cs, cannot be assumed as constants in the
case of real structures where binding as well as processes other than diffusion take
place.

2.2.2 Primary factors affecting the chloride ingress

As the chloride penetration in concrete is controlled by various properties of concrete


and environment, we can say that chloride induced corrosion in RC structures will
also be significantly influenced by these parameters. Most of the concrete parameters
that affect carbonation process will in general also have the similar affect on the
chloride ingress process as both are diffusion process through pores of concrete. The
parameters that have significant effect are presented in table 2.2.1
Factor

Effect on chloride ingress and induced corrosion


Lower threshold chloride content will lead to depassivation

and initiation of steel corrosion in comparatively shorter time


The chloride threshold depends primarily on:

Threshold chloride
content

pH of concrete, i. e. the concentration of OH- in the pores:

(Controls alkalinity of concrete)


Potential of the steel: (O2 availability on steel surface -

High availability lowers the threshold)


Presence of voids at the steel/concrete interface:

(Promotes local acidification, leading to pitting corrosion)


Permeability of
concrete
Surface chloride
concentration
Chloride binding
capacity of concrete

High permeability of concrete promotes chloride ingress by

providing chlorides an easy access to rebar


High surface chloride concentration will lead to high rates of

ingress leading to an early depassivation, this controls the


chloride ingress in splash zones and in marine environments
Presence of tricalcium-aluminate (C3A) governs the chloride

binding capacity of concrete, as chlorides form tricalciumchloro-aluminate (Fridels salt) on reaction with C3A

13

So only a fraction of the total chlorides are available for

corrosion initiation as bound chlorides can not participate


under normal conditions.
Thus chloride binding can delay the corrosion initiation due to

higher threshold chloride content, as it will have to take into


account the effect of bound chlorides
Cracks in concrete

Cracks in concrete will allow an easy access to chlorides at

rebar level and may lead to rapid depassivation


Ingress of chlorides in concrete takes place through chloride

bearing solution. When wetting takes place, chlorides enter


inside the concrete through pores in the solution phase. When
Wetting and drying

drying takes place, chlorides remain deposited in concrete

cycles

pores. On re-wetting, these deposited chlorides further move


inside the concrete
Therefore chloride ingress proceeds in steps, when the

structure is subjected to repeated cycles of wetting and drying.


Table 2.2.1: Primary factors affecting chloride ingress
References
1. Bertolini L., Elsener B., Pedeferri P. and Polder R., Corrosion of Steel in
Concrete: Prevention, Diagnosis, Repair, Wiley-VCH, Weinheim Publication
(2004).
2. West J. S., Larosche C. J., Koester B. D., Breen J. E. and Kreger M. E., State-ofthe-Art Report About Durability of Post-Tensioned Bridge Substructures,
Research Report 1405-1, Center for Transportation Research, Austin (1999).
3. Glass G. K., B. Reddy, and Clark L. A., Making reinforced concrete immune
from chloride corrosion, Proceedings of the Institution of Civil Engineers,
Construction Materials, Vol. 160(CM4), pp. 155164 (2007).
4. Chatterji, S., Bernhardsvej C., Frederiksberg C., On the applicability of Fick's
second law to chloride ion migration through portland cement concrete, Cement
and Concrete Research, Vol. 25(2), pp. 229-303 (1995).

14

5. Costa A. and Appleton J., Chloride penetration into concrete in marine


environment-Part I: Main parameters affecting chloride penetration, Materials
and Structures, Vol. 32, pp 252-259 (1999).
6. Costa A. and Appleton J., Chloride penetration into concrete in marine
environment- Part II: Prediction of long term chloride penetration, Materials and
Structures, Vol. 32, pp 354-359 (1999).
7. Broomfield John P., Corrosion of Steel in Concrete: Understanding, Investigation
and Repair, E&FN Spon. London (1997).
8. Glass G. K. and Buenfeld N. R., Chloride-induced corrosion of steel in
concrete, Prog. Struct. Engng. Mater., John Wiley & Sons, pp. 448 -458 (2000).
9. Siegwart M., Lyness J. F. and Cousins W., Advanced analysis of published data
on chloride binding in concrete, Magazine of Concrete Research, Vol. 55(1), pp.
4152 (2003).

15

2.3 Corrosion

Corrosion of reinforcement steel bars (rebars) in RC structure has drawn attention of


several researchers over the world because of the interesting inter disciplinary
mechanisms involved and magnitude of problems it causes.
Concrete is an alkaline material having its pH in range > 12.5 normally. Because of its
alkaline nature concrete provides as an excellent material for rebar protection against
corrosion through formation of an adhering passive oxide layer around rebar in
reinforced concrete. But under the action of depassivating agents mainly CO2 and/or
chloride ions, this protecting layer starts deteriorating. Consequent exposure to these
elements ultimately breaks the passive layer and allows corrosion initiation to take
place. Time taken for corrosion to initiate is called as initiation time [1, 2]. Once
initiated, corrosion propagates under supportive conditions i.e. availability of
chlorides, moisture and oxygen at rebar level and in process produces rust as end
product. Surrounding concrete not having enough pore space to occupy rust produced
is subjected to tensile stresses and this leads to crack development in concrete which
with time appear at surface, as concrete is a poor material in tension.
Crakes developed catalyze the corrosion process by providing easy access to
chlorides, CO2, O2 and moisture to rebar level. This process leads to spalling of
concrete and thus making structure susceptible to untimely failure earlier than its
design life as structure attains the serviceability limit earlier than planned. Time taken
after corrosion initiation to reach the defined serviceability limit is commonly termed

Deterioration

as propagation time [1].

Serviceability Limit

Initiation
Time

Failure

Propagation
Time

Time

Fig. 2.3.1: Reinforced concrete performance against corrosion deterioration

16

In simple words rebar corrosion can be defined as, An electrochemical process in


which iron (Fe) is oxidized to ferrous ions (Fe++), these ions react with hydroxyl ions
(OH-) available because of oxygen (O2) reduction in presence of water (H2O), to
produce ferrous hydroxide (Fe(OH)2). Thus formed Fe(OH)2 reacts with available O2
and H2O to produce ferric hydroxide (Fe(OH)3), which constitutes high voluminous,
granular hydrated red rust (FeO(OH)+H2O). Reactions involved are as followed [3]:
Fe Fe++ + 2 e -

(2.3.1)

O2 + 2 H2O + 4 e - 4 OH -

(2.3.2)

Fe++ + 2 OH - Fe(OH)2

(2.3.3)

4 Fe(OH)2 + O2 + 2 H2O 4 Fe(OH)3

(2.3.4)

In this process oxidation of metal [Eqn. 2.3.1] occurs at anode (Negative pole) while
reduction [Eqn. 2.3.2] takes place at cathode. Electrons flow from anode to cathode in
metal and thus current flows from cathode to anode in metal. Negative OH- ions flow
from cathode to anode in concrete pore water (electrolyte) and so current flows from
anodic area to cathodic area in concrete. Similarly positive metal ions diffuse from
metal to concrete at anode while electron flows from metal to concrete at cathode.
Concrete with pore Cl & CO2
water as electrolyte

Anode

O2 2 H2O

Cathode

4e

Passive layer

4 OH

Fe Fe++ + 2 e

Iron dissolution
after depassivation

High Voluminous
Red Rust - Fe(OH)3

Rebar

Fig. 2.3.2: Schematic presentation of corrosion process in reinforced concrete


Now as mentioned that corrosion is an electrochemical process represented by
reactions [2.3.12.3.4], and therefore the question whether corrosion will take place or
not, is governed by thermodynamics of the reactions involved as it does in case of any
other chemical reaction. Corrosion thermodynamics can be understood by considering
the change in free energy of the corrosion system [4].

17

2.3.1 Free Energy consideration

Free energy change (G) of any reaction is a measure of driving force for a reaction
and is represented by difference in free energies of formation of products and
reactants i.e.
G = Gproducts - Greactants

(2.3.5)

For electrochemical (cell) reactions change in free energy can be related with the
electrode (cell) potential in following manner [3-5]:
G = - nFE0

(2.3.6)

Where G =Change in free energy


n = No. of electrons involved in reaction
E0 = Standard electrode (cell) potential with reference to S.H.E.
If value of G is negative (or Positive E0) for a reaction then reaction will take place
spontaneously as written, otherwise may take place in reversible direction if G is
positive (or negative E0) and reaction equilibrium is represented by G or E0 equals
zero [4, 5]. For corrosion cell; electrode (cell) potential is the difference between
cathodic and anodic half cell potentials, which consequently is equal to the difference
in electrode potentials for oxygen reduction at cathode (Eqn. [2.3.2]) and iron
oxidation at anode (Eqn. [2.3.1]) i.e.
E0Corr = E0Cathode - E0Anode

(2.3.7)

Also we can say from discussion above on corrosion process that:


Esta < Estc

(Electrons move from anode to cathode in steel)

Econa > Econc

(OH- ions move from anode to cathode in concrete)

Esta - Econa < Estc - Econc

(Considering both above inequalities together)

Also we can write for anodic and cathodic half cells that:
Anode Half Cell Potential = Ea = Esta - Econa

(2.3.8)

Cathode Half Cell Potential = Ec = Estc - Econc

(2.3.9)

Where Esta = Potential of steel at anode


Estc = Potential of steel at cathode
Econa = Potential of concrete at anode
Econc = Potential of concrete at cathode

18

Values for these half cell potentials are available in literature and are equal to 1.229
volts and -0.440 volts respectively for cathodic and anodic cells at 25 oC with respect
to Standard Hydrogen Electrode (SHE) [3]. It explains that standard cell potential is
positive for corrosion cell. Therefore corrosion reactions will occur under conducive
environment i.e. after depassivation of protective oxide layer under standard
conditions, because corrosion cell potential E0 is positive and hence value of G is
negative, leading to a spontaneous reaction. Another tool for predicting the corrosion
reactions is the Pourbaix diagram, which deals with the thermodynamic equilibrium
of reactions involved and demonstrates a qualitative picture of what can happen at a
given pH and potential. The Pourbaix diagram illustrate the relations of the potential
established between a pure metal (electrode) and its ions in solution (electrolyte), and
the pH of the electrolyte; indicating which pH-potential conditions might lead to
corrosion. In reinforced concrete, water acts as electrolyte while iron rebars serve as
electrode [4].
Cell (Electrode) Potentials for various reactions involved in corrosion process can be
obtained using Nernst equation [3, 5]. It represents variation of cell potential from
standard potential values depending on absolute temperature and rate of electrode
reactions involved and can be written as:
E = E0 +

RT Qoxid
ln

zF Qred

(2.3.10)

Where E = Electrode potential


E0 = Standard electrode potential
R = Gas constant = 8.314 J/oK
T = Absolute temperature (oK)
z = No. of electrons involved in reaction at electrode
F = Faradays Const. = 96500 (Coulombs)
Qoxid / Qred = (Rate of oxidation reaction) / (Rate of reduction reaction)
Rate of electrode reactions i.e. of oxidation and reduction depends on activity (a) of
various species participating in reactions. But activity for solvent can be written as
mole fraction of solvent in the liquid which in corrosion reactions is water having
mole fraction equals unity. Activity for dissolved matters in diluted solutions can be

19

represented using molar concentration of dissolved species in mol/l and for solid
matter and pure liquids varies very little and can be taken as unity [5]. Using these
concepts electrode potentials for reactions [2.3.1 2.3.2] can be obtained as presented
in table 2.3.1.
Reaction

E0(Volts)

Nernst Equation for Elect. Poten. (E)

Fe Fe++ + 2e -

- 0.440

E0 + (RT/2F)*ln([Fe++] / [Fe])

O2 + 2H2O + 4e - 4OH -

1.229

E0 + (RT/4F)*ln([O2][H2O]2 / [OH-])

Table 2.3.1: Electrode potentials for anodic and cathodic corrosion reactions
Rebar corrosion in concrete once initiated is then propagated, in process deteriorating
the RC structure. Question on what controls the rate of corrosion reaction can be
answered with the help of theory of kinetics, which explains rate of reactions.

2.3.2 Corrosion Kinetics

For corrosion to occur, two electrochemical reactions, an oxidation and a reduction


reaction, must proceed at the same time and at the same rate to preserve the electroneutrality of the system, leading to equilibrium. But potential of half cell being
dependent on the standard redox potential, concentration of reactants and temperature;
will change as reactions take place on the electrode.

So usually, the electrode

potential for half cell reactions shifts away from equilibrium, as a result of the net cell
reaction occurring, i.e. a net electric current flowing through the interface between
metal and electrolyte, and retards reactions at electrode [5].
The deviation from equilibrium is called polarization, and we say that the electrode is
polarized. A simple measure of polarization is the overvoltage (), i.e. the difference
between the equilibrium electrode potential and the cell potential. When corrosion
process takes place on a surface, the electrode potential adopts the cell potential value,
which is somewhere between the equilibrium potential of the cathodic and anodic
reactions leading to polarization, which slows down the reaction at electrode. The
polarization can briefly be categorized as following, based on the source of origin:

20

Activation polarization: Polarization due to kinetic hindrance of the rate

controlling step of electrode reaction


Concentration polarization: Polarization due to change in concentration of

electrolyte in vicinity of electrode


Ohmic (Resistance) polarization: Polarization due to resistance of the

electrolyte and of any films/material surrounding the electrode surface


Ec
Cathode Over potential
(c) < 0
Ecell
Ea

Anode Over potential


(a)> 0

Fig. 2.3.3: Cathodic and anodic over potentials due to polarization [5]
There exists an energy barrier that the species involved in a reaction (atoms/ ions)
have to overcome to be transferred to a new state. The rate-determining step may be
ion or electron transfer across the interface or some kind of conversion of a species
involved in the reaction. Activation polarization is caused by the resistance against
these reactions itself at the metalelectrolyte interface. For activation polarization the
relationship between current density, i and overvoltage, is given by the Tafel
equation, which states that in absence of concentration and ohmic polarization, the
activation polarization is proportional to logarithm of current density [5]. Tafels
equation for activation over potential is given as:
i
RT

* ln
zF
icorr

= 2.303*

(2.3.11)

Where, = Over voltage


= 2.303/z = Transfer coefficient
icor = Corrosion current density
This can further be classified as:
a = a*ln(ia/icor) = Ea Ecell
c = c*ln(ic/icor) = Ecell - Ec

21

In cases, where activation polarization dominates the mass or charge transfer across
the interface is rate determining. On the other hand, the mass transport within the
solution may be rate determining, due to dominance of concentration polarization.
This implies either there is a shortage of reactants at the electrode surface or that an
accumulation of reaction products occurs. Effect on over potential due to
concentration polarization can be obtained using the Nernst equation, which takes in
to account the effect of concentration of participating species in potential calculation.
This can be explained using the effect of oxygen availability during corrosion process,
which is a major rate determining factor as will be discussed later. Where cathode
over potential due to change in concentration of oxygen (conc.) can be calculated as
mentioned below [5]:
RT
zF

conc. =

Cf

*ln

Ci

(2.3.12)

Where Cf - Ci = Change in oxygen concentration


Also the resistance of system (R) to current flow due to corrosion process (I);
significantly affects the over potential in some cases. This resistance may be due to
surface layers on metals (passive layer) and materials surrounding (concrete) and it
leads to change in over potential. This phenomenon is termed as resistance
polarization and the change in over potential due to resistance (res.) is given by:
res. = RI

(2.3.13)

Now rate of any reaction can be defined as rate of consumption of involved reactant
species or production rate of end products. Which for electrochemical (ionic)
reactions involved in corrosion process can be obtained using Faradays law which
gives us:
r = I / (z*F)

(2.3.14)

Where: r = rate at which reactants are consumed [mol/s]


z = Valancy of participant in reaction
I = Current
F = Faradays Const. = 96500 C/moles
So we can say the rate of corrosion process will depend on the equilibrium cell
potential (corrosion potential Ecorr) and corresponding current flowing in the system
(corrosion current Icorr) in case of steel corrosion in concrete. This rate i.e. corrosion

22

current is a direct and simple measure for two main deteriorating effects of corrosion
i.e.; Amount of iron consumed in the corrosion reaction and corresponding amount of
rust produced. Commonly the rate is expressed in corrosion current per unit area i.e.
current density. Higher the rate is more damaging it will be to reinforced concrete
structures as it will cause more damage to reinforcement by consuming more iron in
even smaller time. But due to various controlling parameters the overall process rate
is not as simple to model as it looks. Some parameters that have significant influence
on corrosion rate are briefly discussed in following section along with their influence
on overall corrosion process.

2.3.3 Primary factors affecting steel corrosion in concrete

It is well accepted that if mixed properly, concrete provides as an excellent material


for construction and also protects steel from corrosion attack as long as passivation is
not destroyed due to action of harmful agents such as CO2 and chlorides. But poor
construction practices, exposure to severe environments and harsh working conditions
cause several damages in concrete. Steel corrosion is the most common and
extensively studied problem that occurs in reinforced/prestressed concrete structures
including bridges, due to the scale of problems it causes. Steel corrosion in concrete is
controlled by various environmental and concrete parameters that will alter the rate of
corrosion reactions significantly. Table 2.3.2 presents primary factors that control
corrosion process in concrete.
Factor

Effect on Corrosion in Concrete


Increase in cover thickness improves performance of

Cover Thickness

concrete against corrosion as longer time is required for


Chlorides / CO2 to reach at rebar level

Permeability of concrete

Less permeable concrete reduces rate of diffusion of Cl-/

CO2 in concrete, thus delaying the corrosion initiation time


High resistivity of concrete impedes the flow of charge

Resistivity of concrete

through concrete, therefore causing significant reduction in


corrosion current and hence rate of corrosion is controlled

Oxygen availability

As can be observed that O2 is required both at anode and

cathode for corrosion reactions to take place and production

23

of rust [Eqn 2.3.2-2.3.4]. Therefore oxygen availability may


significantly alter the rate and end products in corrosion
reactions and therefore the affect on the structure will vary
Black rust (Fe3O4) forms in case of restricted O2 supply

while red rust (Fe(OH)3) is commonly produced otherwise


Concrete pore saturation plays an important role as it

controls various factors affecting corrosion such as:


Degree of pore
saturation in concrete

Diffusion rates of Cl- and CO2 ingress


Resistance of concrete: Increase in saturation leads to

decrease in resistivity
Corrosion can occur only in partially saturated

conditions as H2O and O2 both are must for corrosion


Chloride ion content depends on concrete constituents,

Chloride ion content

exposure conditions and controls the corrosion initiation


time, as depassivation will occur when concrete Cl- content
exceeds the threshold chloride content.
pH of concrete pore water is critical as the depassivation

pH of concrete pore
water

can only take place when pH falls in range of 9. Therefore


it will require more time and more amount of CO2 and/or
Cl- to depassivate the oxide layer, in a concrete having high
pH pore water, thus delaying the initiation of corrosion
Affects pH of the pore water significantly following the

Carbonation

acidic nature of CO2, which neutralizes alkalinity of


concrete and thus plays an important role in depassivation
and corrosion initiation time.
More the cracks more the corrosion and vice versa, as

cracks will allow an easy access to CO2 and/or chlorides


Cracks in concrete

which consequently will accelerate the corrosion process


and therefore rust production, which further will cause
formation of more cracks due to high volume of rust.
Stray current often arising in railway structures where

Stray currents

electric lines are used, can induces pitting corrosion on


buried metal structures, leading to severe localized damage

24

Wetting and drying


cycles

Corrosion is dependent on moisture content (pore

saturation) in concrete and so is significantly influenced by


the duration and frequency of wetting and drying cycles

Table 2.3.2: Primary factors affecting steel corrosion in concrete

2.3.4 Corrosion diagnosis and measurement methods

The detection and measurement of corrosion in concrete structures is essential.


Although there are several methods for the diagnosis, detection and measurement of
corrosion in reinforcing steel, there is no consensus regarding which method assesses
corrosion levels in reinforced concrete structures most accurately. Various techniques
used for detection and/or measurement of corrosion provide important information on
the causes, detection or rate of reinforcement corrosion in concrete. These techniques
are mainly based on the electrochemical aspects of corrosion process. Following are
the commonly used techniques for corrosion detection or measurement.
Open circuit potential (OCP) measurements:
The main method of detection of corrosion is the half-cell potential (HCP)
measurements and it can easily be obtained using open circuit potential (OCP)
measurement techniques. This is the most typical procedure to the routine inspection
of reinforced concrete structures.
The tendency of any metal to react with an environment is indicated by the potential it
develops in contact with the environment. In reinforced concrete structures, concrete
pore water acts as an electrolyte (environment) and the reinforcement will develop a
potential depending on the rate at which iron dissolves (controls corrosion rate) and it
may vary from place to place. This potential can be obtained using OCP measurement
technique, the schematic diagram for which is shown in Fig. 2.3.4 [6]. The principle
involved in this technique is essentially measurement of corrosion potential of rebar
with respect to a standard reference electrode, such as saturated calomel electrode
(SCE), copper/copper sulfate electrode (CSE) etc.. As per ASTM standard C876 [7],
the probability of reinforcement corrosion is as follows in Table 2.3.3.

25

Fig. 2.3.4: Open circuit potential (OCP) measurement [6]


Open circuit potential (OCP) values (mV)

Probability

of

Reinforcement

w.r.t. SCE

w.r.t. CSE

Corrosion

OCP < - 276

OCP < - 350

High (> 90%) risk of corrosion

-275 < OCP < - 126

-350 < OCP < - 200

Uncertain (May be taken as 50%)

OCP > -125

OCP > -200

Low (<10% ) risk of corrosion

Table 2.3.3: Corrosion condition related with HCP measurements [7]


Advantages of OCP measurement:
It is a non-destructive test that can collect enormous quantity of data from a large

structural area without many disruptions to the functioning of the structure, which
can be used to establish structure potential map
OCP measurement is useful in finding out the anodic and cathodic sites in

reinforced concrete structures


Limitations of OCP measurements:
Potential readings cant be taken as a sufficient as criterion, since they are affected

by various factors such as: Polarization due to limited oxygen availability,


Concrete porosity and the system resistance
OCP values only provide information for corrosion probability and does not

indicate the rate of corrosion


Concrete resistivity measurement:
The electrical resistivity of concrete is an important parameter concerning with the
intensity of corrosion process. Because corrosion process involves charge transfer,
therefore in concrete with high electrical resistivity the corrosion process will be slow
26

compared to concrete with low resistivity in which the current can easily pass through
concrete between anodic and cathodic areas.
Two different techniques, namely AC and DC measurements are used for
determination of electrical resistivity. In these measurements both surface and
embedded probes are applied. Applying a constant electric field between the two
embedded electrodes and measuring the resulting current as a voltage drop over a
small resistance accomplish the DC measurements.
The AC measurements can be conducted both by means of two and four-pin methods.
The most common surface mounted probe is known as the Wenner array, which uses
four probes. An alternating current (AC) is passed between the outer electrodes and
the potential between the inner electrodes is measured. Concrete resistivity is usually
measured by using the Wenner four probe method [6]. Schematic setup for the
resistivity measurement is shown in Fig 2.3.5.

R = V/I
= 2aR

Probes
Sponges
Concrete
Rebar

Fig. 2.3.5: Circuit for electrical resistance measurements


A known current `I' is impressed on the outer probes and the resulting potential drop
`V' between the inner probes is measured and resistance `R' is given by V/I. and then
resistivity of concrete () can be calculated as following:
= 2aR

(2.3.15)

Where a = inner electrode distance in cm


R = measured resistance in ohm
The resistivity measurement is a useful additional measurement to aid in identifying
problem areas or confirming concerns about poor quality concrete in context of

27

corrosion and is being increasingly used to evaluate concretes corrosion


characteristics. Measurements can only be considered along side other measurements
as reinforcing bars will interfere with resistivity measurements. It has been established
that the likelihood of corrosion in concrete is inversely proportional to the electrical
resistivity of concrete [8]. It has also been shown that for concrete, corrosion
performance and resistivity can be related as follows in Table 2.3.4 [6, 9, 10].
Resistivity (.cm) Corrosion risk
> 20,000

Negligible

10,000 - 20,000

Low

5,000 - 10,000

High

0 < 5,000

Very high

Table 2.3.4: Corrosion risk from resistivity


Linear polarization resistance (LPR) measurement:
LPR monitoring has been developed to address the need of a non-destructive
technique, which enables accurate assessment of the condition of corroded reinforced
concrete structures and has become a well-established method of determining the
instantaneous corrosion rate measurement of reinforcing steel in concrete. The
technique is rapid and non-intrusive, requiring only localized damage to the concrete
cover to enable an electrical connection to be made to the reinforcing steel. In LPR
measurements the reinforcing steel is perturbed by a small amount from its
equilibrium potential. This can be accomplished potentiostatically by changing the
potential of the reinforcing steel by a fixed amount E, and monitoring the current
decay I, after a fixed time. Alternatively it can be done galvanostatically by applying
a small fixed current I, to the reinforcing steel and monitoring the potential change
E, after a fixed time period. In each case the conditions are selected such that the
change in potential E, falls within the linear SternGeary range of 1030 mV [6].
The polarization resistance Rp, of the steel is then calculated using following equation:
Rp = E / I

(2.3.16)

From which the corrosion rate Icorr, can then be calculated using
Icorr = B / Rp

(2.3.17)

28

Where, B is the SternGeary constant. For which a value of 26 mV is adopted in case


of active steel and 52 mV in case of passive steel [11-13]. Also B can be obtained
using Tafels slopes for anodic and cathodic polarization curves i.e. a and c
respectively using the following equation:
B=

a c
2.303( a + c )

(2.3.18)

In order to determine the corrosion current density, icorr, the surface area A, of
polarized steel needs to be accurately known and then:
icorr =Icorr / A

(2.3.19)

In a conventional LPR test the perturbation is applied from an auxiliary counter


electrode on the concrete surface. The surface area of steel assumed to be polarized is
that lying directly beneath the auxiliary electrode. However the current flowing from
the auxiliary electrode is unconfined and can spread laterally over an unknown, larger
area of steel. This can lead to inaccurate knowledge of the surface area of steel
polarized and will therefore result in an error in the calculation of the corrosion
current density. In order to overcome the problem of confining the current to a
predetermined area, a secondary auxiliary guard ring electrode surrounding the inner
auxiliary electrode is commonly used. The principle is that the confinement current is
maintained by the outer guard ring electrode during LPR measurement. This
confinement current prevents the perturbation current from central auxiliary electrode
spreading beyond a known area [6]. Fig 2.3.6 shows the schematic set-up arrangement
of LPR measurement using guard ring technique.
Reference electrode
Auxiliary electrode

Sensor electrodes

Guard ring
electrode
Working
electrode

Fig. 2.3.6: Guard ring test set-up

29

Tafel Extrapolation:
The Tafel extrapolation technique (TP) is another electrochemical method for
calculating corrosion rate based on the intensity of the corrosion current (Icorr) and the
Tafel slopes. Tafel slopes also could be used to calculate corrosion rate with LPR as
mentioned earlier. This technique is also based upon application of either steady fixed
levels of current, followed by monitoring of the potential (galvanostatic) or
application of specific potential followed by monitoring of the current (potentiostatic).
The main difference between these two methods is that the change in potential must
be kept within Stern- Geary range i.e. less than 30 mV for the LPR technique, while
the change of potential can go up to 250 mV for the Tafel extrapolation technique.

Fig. 2.3.7: Schematic polarization curves showing Tafel extrapolation [14]


In TP, corrosion rate can be calculated using straightforward substitution of Tafel
slope values (a and c) to get the corrosion current then, by calculating corrosion rate
using Eq.2.3.20 as following [6, 14]:

I = I corr exp {S1 ( E Ecorr )} exp { S2 ( E Ecorr )}


Where S1 = 2.303 / a ; S2 = 2.303 / c
E = Potential at timet
I = Current at timet
Icorr = Corrosion current (A)
Ecorr = Corrosion potential (V)

30

(2.3.20)

The corrosion current density can then be obtained by using the exposed surface area
(A) of steel, using Eq. 2.3.19 as mentioned earlier. Main disadvantage of using Tafels
extrapolation technique is that it is an electrochemically destructive technique, as
application of external activation (voltage or current) may alter the electrochemical
behavior of the system significantly.
Rebar condition can be interpreted on the basis of corrosion current density, and
following guidelines have been suggested and are widely accepted for the rebar
condition assessment [15, 16].
Corrosion Current Density (icorr)

Rebar Corrosion Condition

A/cm2
icorr < 0.1

Negligible

0.1 < icorr < 0.5

Low to moderate corrosion

0.5 < icorr < 1.0

Moderate to high corrosion

icorr > 1.0

High corrosion rate

Table 2.3.5: Guidelines for rebar condition assessment using icorr


Electrochemical impedance spectroscopy (EIS):
Electrochemical impedance spectroscopy (EIS) is a technique that works in the
frequency domain. The basic concept involved in EIS is that an electrochemical
interface can be viewed as a combination of passive electrical circuit elements, i.e.,
resistance, capacitance and inductance [17]. This technique is commonly known by
the name AC impedance technique and is a useful non-destructive technique for
quantifying corrosion of steel rebars embedded in concrete. In this technique an
alternating voltage of about 10-20 mV is applied to the rebar and the resultant current
and phase angle are measured for various frequencies. The response to an A.C. input
is a complex impedance Z, which is the ratio of A.C. voltage to A.C current and has
both real (resistive) and imaginary (capacitive or inductive) components Z and Z
respectively. For a simple corrosion system equivalent circuit can be drawn as
presented in Fig. 2.3.8 and then equivalent complex impedance of the system can be
written as:

31

Z eq. = Rs +

Rp

(2.3.21)

1 + CR p

Where Rs = Ohmic resistance of the cover concrete


Rp = Polarization resistance of system
C = Capacitance of the system
= Operating frequency
1/C = Impedance of capacitor
C
Rs
Rp
Fig. 2.3.8: Equivalent circuit diagram for simple corrosion system [18]
Eq. 2.3.21 on simplification gives real and imaginary parts of impedance i.e.
Z = Rs +

Z"=

Rp

1 + (CR p )

(2.3.22)

CR p 2

1 + (CR p )

(2.3.23)

On simplification (elimination of ) above relations give us equation of a circle.


Plotting the imaginary impedance against the real impedance gives a semicircle, with
a diameter equal to Rp with center at Rs + 0.5*Rp, these plots are commonly known as
Nyquist plots and can be presented as shown in Fig.2.3.9.
Z
max

Rs

Rp
Simple Corrosion System

Fig. 2.3.9: Schematic Nyquist Plot

32

At the highest point on the semicircle the frequency can be found by interpolation
and then double-layer capacitance value is then given by Eq. 2.3.24 as following:
C=

(2.3.24)

max R p

In practice, an AC Impedance response will often be a combination of several


different semicircles, due to different RC parallel components, which could arise from
film resistance, diffusional impedance etc.[17]. The A.C. impedance technique has the
advantage that it can give more information than DC LPR measurements, but it can be
very time-consuming due to wide range of frequencies it needs to be performed upon
to get the Nyquist plot. Therefore its use has been generally confined to the laboratory
rather than on structures in the field [6].

2.3.5 Damages due to corrosion

Corrosion in RC structure is most commonly observed damaging mechanism. As in


case of corrosion rebar is oxidized to produce high voluminous rust. The volume of
rust produced depends on the level of oxidation (thus on corrosion products formed)
and can be 2-6 times more than that of steel depending on the constituents. Typical
oxides formed during oxidation have the following volume ratios relative to the
volume of the iron have been reported by various researchers and are presented in
table 2.3.6 [19-22].
Iron Oxide

VOxide / VIron

FeO

1.7 - 1.8

Fe3O4

2.0 - 2.2

Fe2O3

2.1 - 2.2

Fe(OH)2

3.6 - 3.8

Fe(OH)3

4.0 - 4.2

Fe(OH)33H2O

6.2 - 6.4

Table 2.3.6: Relative volume ratios for various iron oxides in comparison to iron
Common damages in RC Structures:
The rust produced initially occupies the pore space surrounding the rebar, once all the
pore space available is used then; concrete is subjected to tensile stresses due to

33

excess volume of remaining rust. This ultimately causes cracks in concrete which is a
poor material in tension. This will lead to spalling of cover concrete in area
surrounding the corroding rebar and hence is a major concern as it will affect the
serviceability of the structure significantly by making it accident prone. In general
following deteriorating consequences of corrosion are common in RC structures:
Rust stains on concrete surface due to leaching of corrosion products, this

impairs the aesthetics


Reduction in effective cross sectional area of rebar as iron is consumed in the

process, leading to significant reduction in load carrying capacity of structure.


Impairs the bonding of rebar with surrounding concrete, and thus adversely

affects load transfer between concrete and rebar


Cracking and spalling of the cover concrete due to tensile forces arising from

excessive volume of corrosion products, reducing structural serviceability


Rebar becomes more accessible to the aggressive agents due to cracking,

leading to further corrosion at an accelerated rate


Pitting (localized) corrosion of the rebar may also occur and is more

dangerous than uniform corrosion because it reduces the cross-section area of


rebar to a point, leading to a catastrophic failure of the structure.
In concrete bridge structures use of prestressed steel strands is common, and in this
case corrosion is of more concern. As corrosion can lead to catastrophic brittle failure
prestressed steel strands due to any of the following mechanism:
Stress corrosion cracking (SCC):
SCC can be defined as highly localized corrosion that produces cracking in the
simultaneous presence of corrosion conditions and externally applied or residual
tensile stresses. SCC induces sudden, brittle, unexpected failure as SCC-initiated
cracks propagate across the components load-carrying section i.e. in planes
perpendicular to tensile stresses. In simple words it can be termed as anodic
corrosion under high sustained tensile stresses in case of prestressing steel in
concrete [5, 15, 23-25]. Most significantly, the process progresses at a far faster rate
than the more commonly recognized corrosion mechanisms, and shows almost no
signs of initiating and progress on the visible surfaces of the structure, such as
concrete cracking, spalling, and debonding. Existence of surface flaws including
34

longitudinal wire splits, wire seams created by overused/faulty drawing dies, and pits
in wire from atmospheric corrosion created during storage and prior to grouting, are
known to have provided crack initiation sites. Other locations susceptible to the
phenomenon include locations where moisture penetrates to the steel in areas where
steel is not passivated, e.g., with carbonated concrete, or steel not in direct contact
with concrete or grout.
Hydrogen embrittlement (HE):
Hydrogen embrittlement is a related phenomenon, which causes liberated atomic
hydrogen molecules to penetrate (through absorption) the steel lattice and
consequently leads to reduction in ductility of strands. This process can be termed as
Cathodic corrosion under tensile stress and will lead to brittle cracking of
prestressing strands due to embrittlement of hydrogen. As hydrogen can be created by
the electrochemical cathodic corrosion process itself if not available otherwise, for
example at a pit, or corrosion at manufacturing and wire drawing flaws, and from
cathodic protection systems which produce excessive electrical currents. Rupture is
sudden and brittle [5, 23-25]. Hydrogen embrittlement can also take place even in
absence of applied tensile stresses [25].
In case of SCC or HE the failure process develops such that initially, one or several
micro cracks are generated at the surface of the steel and then, these cracks grow and
propagate very quickly until they reach a certain depth, leading to the unexpected
brittle failure of the strand wire.
Main concern with SCC or HE is that, they cant be noticed by means of measuring
the corrosion rate, as the loss of metal is negligible and the corrosion potential
measured at the concrete surface may not indicate the development of the process.
Therefore, this type of corrosion cannot be electrochemically measured during its
occurrence. Only the risk of its appearance may be approached by the traditional
electrochemical techniques.
Even after the failure the identification of the nature of the failure is not an easy task.
It can only be detected by microscopic observation of the fractured surface when it is
fresh, and not corroded or contaminated. SHE or HE shall not to be mistaken with the
35

failure induced by chlorides in prestressed wires. The ions inducing localized


corrosion may also aim into a failure, but of simple reduction (localized attack) of the
cross section produced by the electrochemical process.

2.3.6 Corrosion control techniques

Corrosion is the most prominent among the factors that govern the performance of a
concrete structure during the service life of the structure. Therefore it is mandatory to
control the corrosion process in existing as well as in new structures if required, to
maintain the minimum required performance levels over the service life period of the
structure. In present scenario corrosion process can be controlled by various
techniques, which have established their application in field over the period of time.
The commonly utilized techniques and materials for corrosion control are:
Patch Repair
Cathodic protection (CP) using impressed current
Cathodic protection using sacrificial anode
Electrochemical chloride extraction (ECE)
Realkalization
Inhibitor application
Coatings
Use of corrosion resistant rebars
Design considerations

Patch Repair:
Patch repair is a traditional approach used for repair of corroded concrete structures;
where spalling has taken place or cracks appear on the concrete surface or structure
show signs of rebar corrosion over an area. And therefore areas having rebar
corrosion is usually identified by means of visual inspection, though sometimes
electrochemical methods may also be used. In this method basically the contaminated
concrete is replaced with fresh mortar.
In this method the concrete placed over the corroded rebars is removed completely by
breaking out the concrete behind the steel reinforcement. At all locations, the steel
must then be at very least wire brushed to remove all loose deposits or preferably grit
blasted or water jetted back to bright metal. Once the area to be repaired has been cut
36

out and cleaned thoroughly of all debris and dust, the repair can begin. Usually
designed repair mortars usually contain one or more of acrylic or other polymers,
shrinkage compensating additives, silica fume etc. to improve workability,
permeability, or ease of application. Mortar for repairs is then usually hand placed
packing the mortar on in layers. The repair must then be cured in properly. Once
finished and cured, there may be surface irregularities and will almost certainly be
color variation between the patch and the surrounding parent concrete. It is usual
therefore to apply a thin skim coat of a proprietary material, usually known as a
"fairing coat" to fill in any surface blemishes and to mask the patches themselves.
This may then be followed by the application of a final coating of anti-carbonation
paint which has the dual purpose of stopping further carbonation in unrepaired areas
and providing a pleasing and even color finish [10].
Patch repair is very handful in case of global corrosion where corrosion takes place
over a large area, that is in case of carbonated induced corrosion. But the patch repair
of chloride-induced corrosion damage presents problems because of the localized
nature of the corrosion. Active corrosion sites provide a form of electrochemical
protection (Cathodic protection due to incipient anode effect) for adjacent passive
steel that may also be exposed to chloride-contaminated concrete. The traditional
patch repair of corroding areas therefore triggers corrosion initiation in these passive
areas as the local protection is removed. Thus, when conventional patch repairs are
used to inhibit chloride-induced corrosion deterioration, all contaminated concrete
should be removed if further deterioration is to be avoided. One method of
overcoming this problem is to provide a sacrificial anode in the area of the patch
repair to continue the provision of local electrochemical protection. Placing small
sacrificial anodes such as Zinc covered with an appropriate mortar in the patch are
used commonly [10, 26-27].
Cathodic Protection:
Cathodic protection (CP) is a widely used and effective method of corrosion control.
Cathodic protection suppresses the corrosion current that causes damage in a
corrosion cell and forces the current to flow to the metal that is to be protected. The
basic fundamental is that cathodic areas in an electrochemical cell do not corrode.
Therefore if all the anode sites on rebar were forced to function as cathodes, then the
37

rebar would be a cathode and corrosion would be eliminated. To achieve this potential
of the rebar is artificially shifted so that it acts as cathode and hence becomes either
immune or passive. This phenomenon is appreciably used by converting active anodic
area of rebar in concrete to the cathodic ones, by the means of either an impressed
direct current (DC), or by connecting it to a sacrificial or galvanic anode [26-29].
Cathodic protection using impressed current:
In this technique the steel is connected to the negative terminal of an external
electrical power supply forcing it to undergo cathodic reaction as it receives current.
The anode, connected to the positive terminal of the power supply, is usually chosen
to be a relatively non-reactive conductor such as carbon or titanium so that its
corrosion rate is low. Fig. 2.3.10 provides the schematic application diagram for CP
using impressed current.
External
power source
Less reactive anode

__

Protection currents

Concrete & Rebar

Fig. 2.3.10: Schematic application diagram for CP using impressed current


This technique for corrosion control using CP of concrete structures requires the
following basic components:
DC power supply (Rectifier): A rectifier is used to convert alternating current

(AC) to direct current and provides required current to the CP system and
must be capable of providing either constant current or constant voltage to the
anode system.
Relatively inert anode material: It is used to distribute protective current to the

reinforcing steel and provides locations for anodic reactions to take place in

38

lieu of the reinforcing steel. By using relatively inert materials, such as


catalyzed titanium, anode consumption is minimized.
Instrumentation: such as embedded reference electrodes etc. for monitoring

purposes.
Though used extensively this technique has certain limitations that need to be taken
care of for an efficient protection system. Some of the limitations of this technique are
as followed:
Requires skilled labor and routine monitoring and maintenance of voltage,

current and anode etc.


It is difficult to assure a uniform current density in the high resistivity concrete

pore water environment


Cathodic protection is not necessarily effective in protecting steel exposed at

concrete cracks and at spalls in atmospheric applications


While using with prestressed system, one must be very careful as in presence

of cathodic reaction at steel atomic hydrogen may be produced and that may
lead to brittle failure due to hydrogen embrittlement
Cathodic protection using sacrificial (galvanic) anode:
Sacrificial anode systems for CP are based on the principle of dissimilar metal
corrosion and the relative position of different metals in the galvanic series. In this
procedure a sacrificial anode is connected with the rebars. and hence a galvanic cell is
set up by connecting the steel to a more reactive metal, usually zinc. The zinc then
undergoes the anodic reaction and corrodes whilst the steel is rendered entirely
inactive because the whole surface undergoes the cathodic reaction and the iron no
longer dissolves. This may also be thought of as the anodic sites on the steel being
shifted to the sacrificial anode material.The direct current is generated by the potential
difference between the sacrificial anode and reinforcing steel when connected and
thus doesnt require external power source. The sacrificial anode will corrode during
the process and is consumed. Current will flow from the anode, through the concrete,
to the corroding rebar. Because current is to travel through concrete therefore this
method is not very effective in high resistance concretes, which impedes the flow of
current [26-29]. Fig. 2.3.11 represents schematic application diagram for CP using
sacrificial anodes.
39

Sacrificial galvanic metal (Anode)

Connection

Concrete & Rebar


Fig. 2.3.11: Schematic application diagram for CP using sacrificial anodes
Electrochemical chloride extraction (ECE) and Realkalization:
Chlorides (Cl-) are negatively charged, so we can use an electrochemical process to
repel the chloride ion from the steel surface and move it towards an external anode,
So that the passivity of rebar system is restored as alkalinity surrounding the rebar wil
be restored. For removal of chlorides a process similar to cathodic protection is used
but with both external power source as well as an external sacrificial anode (used
temporarily) [10].
In this process the external power source supplies current to rebars, forcing them to
act as cathodes and the used external anode material attracts chlorides present, away
from the rebars. Selected anode material must be able to stand the procedure till
desired reduction in chloride reduction at rebar level has been achieved. Once the
desired Cl- concentration has been achieved at rebar level, the external anode material
may be removed. Fig. 2.3.12 presents schematic diagram for application of ECE.
External
power source
Anode material

Cl-

Cl-

Cl-

Protection currents

Concrete & Rebar

Fig. 2.3.12: Schematic diagram for application of ECE

40

__

Passing large amounts of electricity through concrete can have effects upon its
chemistry and therefore its physical condition, therefore an expert system is required.
ECE on the other hand may also accelerate alkali aggregate reactivity due to the
cathodic reaction that generates excess alkali at the steel surface. The main benefit of
this techniques are that structures is repassivated and the instrumentation need not be
installed permanently like CP and can be removed as soon as the goals are achieved.
Realkalization is a similar process in which instead of any metal anode an alkali
material such as Lithium (Li) is used, which undergoes the anodic reaction and
produces alkali ions (Li+) , which being positively charged are attracted by cathodic
steel thus increasing the alkalinity in surrounding of rebar.
Inhibitor Application:
Some chemical species inhibit corrosion. Chromates, silicates, and organic amines are
common inhibitors. The mechanisms of inhibition can be quite complex. In the case
of the organic amines, the inhibitor is adsorbed on anodic and cathodic sites and
impairs the corrosion current. Other inhibitors specifically affect either the anodic or
cathodic process. Still others promote the formation of protective films on the metal
surface. The use of inhibitors is favored in closed systems where the necessary
concentration of inhibitor is more readily maintained. Inhibitors can be incorporated
in a protective coating or in a primer for the coating. At a defect in the coating, the
inhibitor leaches from the coating and controls the corrosion [29].
Coatings:
Coatings can be applied to concrete structure at whole or to the rebars with the intent
that the corrosive substances are isolated from the rebar. And mainly are of following
three types [29]:
Metallic Coatings: The concept of applying a more noble metal coating on an active

metal takes advantage of the greater corrosive resistance of the noble metal. An
example of this application is tin-plated steel. Alternatively, a more active metal can
be applied, and in this case the coating corrodes preferentially, or sacrificially, to the
substrate. An example of this system is galvanized steel, where the sacrificial zinc

41

coating corrodes preferentially and protects the steel, and is commonly called as
anode coating.
Organic Coatings: The primary function of organic coatings in corrosion protection

is to isolate the metal from the corrosive environment. In addition to forming a barrier
layer to prohibit further corrosion, these coatings can contain corrosion inhibitors.
Inorganic coatings: Include porcelain enamels, chemical-setting silicate cement

linings, glass coatings and linings, and other corrosion resistant ceramics. These
coatings serve as barrier coatings.
Design Considerations:
The application of rational design principles can eliminate many corrosion problems
significantly. Some guidelines that shall be followed are:
Eliminate the dead spaces or crevices as corrosion often occurs there3
Storage of water shall be avoided at joints
Accurate concrete cover depth shall be provided
Limit the inherent chlorides in structures by using appropriate aggregates
If possible limit the crack width at tensile face of concrete surface, which is

close to the rebars


Where stress-corrosion cracking is possible, the components can be designed

to operate at stress levels below the threshold stress for cracking


Areas which are expected to get corroded shall be provided with easy access

for maintenance and repair.


As we can see that there are various techniques available for controlling the corrosion
in new as well existing structures. The one which is most appropriate shall be selected
for repair purposes and that will depends on several factors such as:
Cost effectiveness
Resource availability
Purpose and remaining design service life of structure
Current condition of structure
Corrosive substances present in the structure and surrounding environment

42

References
1. Indian Road Congress (IRC), An approach document for assessment of
remaining life of concrete structures, IRC: SP 60, New Delhi, India (2002).
2. ACI 222R Corrosion of Metals in Concrete, ACI Committee 222 Report,
American Concrete Institute, (1985).
3. Bazant Z. P., Physical model for steel corrosion in concrete sea structures Theory, Journal of Structural Division, ASCE, Vol. 105(6), pp. 1137-1153
(1979).
4. O'Brien Thomas F., Bommaraju Tilak V. and Hine Fumio Corrosion: Handbook
of chlor-alkali technology, Volume - V , Chemistry and material sciences,
Springer US (2005).
5. Einar Bardal, Corrosion and Protection, Engineering Materials and Processes,
Springer London (2004).
6. Ha-Won Song and Velu Saraswathy, Corrosion Monitoring of Reinforced
Concrete Structures A Review, International Journal of Electrochemical
Science, Vol. 2, pp. 1- 28 (2007).
7. ASTM: C876, Standard test method for half-cell potentials of uncoated
reinforcing steel in concrete, ASTM Standards, Vol. 03(02), (1991).
8. Gannon E.J., Cady P.D., Condition evaluation of concrete bridges relative to
reinforcement corrosion, Volume 1: state of the art of existing methods Strategic
Highway Research Program, Report: SHRP-S/FR-92-103, National Research
Council, Washington, DC, (1992).
9. Mehta P.K. and Monteiro P. J. M. Concrete: Microstructure, Properties, and
Materials Third edition. McGraw-Hill, pp. 415. (2006).
10. Grantham M. G. and Gray M. J., Diagnosis, Inspection, Testing and Repair of
Reinforced Concrete Structures M.G. Associates, Herts, (2000).
11. Ahmad S. and Bhattacharjee B., "A simple arrangement and procedure for in-situ
measurement of corrosion rate of rebar embedded in concrete, Corrosion
Science, Vol. 37(5), pp.781-791, (1995).
12. Daly A. F, Modelling of deterioration in bridges, Transport Research
Laboratory (TRL), BRIME Report: PL97-2220, (1999).
13. Broomfield John P., Corrosion of Steel in Concrete: Understanding, Investigation
and Repair, E&FN Spon. London (1997).

43

14. Perez Nestor, Electrochemistry and Corrosion Science, Kluwer Academic


Publishers, Boston, (2004).
15. Nuclear Energy Agency, Electrochemical Techniques to Detect Corrosion in
Concrete Structures in Nuclear Installations: Technical Note, Committee Report
on: The Safety of Nuclear Installations, NEA/CSNI/R(2002)21, (2002).
16. Liu T. and Weyers R.W., Modeling the Dynamic Corrosion Process in Chloride
Contaminated Concrete Structures, Cement and Concrete Research, Vol. 28(3),
pp. 365379, (1998).
17. Montemor M.F., Simoes A.M.P. and Ferreira M.G.S. Chloride-induced corrosion
on reinforcing steel: from the fundamentals to the monitoring techniques Cement
& Concrete Composites, Vol. 25, pp. 491502 (2003).
18. Marcus P. and Mansfeld F. Analytical Methods in Corrosion Science and
Engineering, Published by: Taylor & Francis Group, (2006).
19. Bhargava K., Ghosh A.K., Mori Y. and Ramanujam S., Model for cover cracking
due to rebar corrosion in RC structures, Engineering Structures, Vol. 28, pp.
10931109, (2006).
20. Pantazopoulou1 S. J. and Papoulia K. D., Modeling Cover-Cracking Due to
Reinforcement Corrosion in RC Structures, Journal of Engineering Mechanics,
Vol. 127(4), pp. 342351, (2001).
21. Bhargava K., Ghosh A.K., Mori Y. and Ramanujam S., Modeling of time to
corrosion-induced cover cracking in reinforced concrete structures, Cement and
Concrete Research, Vol. 35, pp. 2203 2218, (2005).
22. Lundgren K., Modelling the effect of corrosion on bond in reinforced concrete,
Magazine of Concrete Research, 54(3), pp. 165173, (2002).
23. Adrian C.T., Corrosion and Prestressed Concrete Bridges, In Proceedings: The
2005 Structures Congress and the Forensic Engineering Symposium, New York,
April 2024, (2005).
24. 12-CRC COMMITTEE, Corrosion of reinforcement and prestressing tendons: A
state of the art report, RILEM Committee report on Corrosion of Reinforcement
in Concrete, Materials and Structures, Vol. 9(3), pp. 187-206, (1976).
25. ACI 222.2R, Corrosion of Prestressing Steels, ACI Committee 222 Report,
American Concrete Institute, (2001).
26. Glass G. K. and Buenfeld N. R., Chloride-induced corrosion of steel in
concrete, Prog. Struct. Engng. Mater., John Wiley & Sons, pp. 448 -458 (2000).
44

27. Akbar A. S., Manual on Service Life of Corrosion-Damaged Reinforced


Concrete Bridge Superstructure Elements, National Cooperative Highway
Research Program, NCHRP Report: 558, Transportation Research Board
Washington, D.C., (2006).
28. Steven D. F., Understanding Corrosion and Cathodic Protection of Reinforced
Concrete Structures Corrpro Companies, Incorporated, Concrete Services
Group, Ohio. Website: http://www.corrpro.com/, Accessed: (2008).
29. Davis J.R., Corrosion: Understanding the Basics, Chapter 1: The Effects and
Economic Impact of Corrosion, ASM International, pp. 1-20, (2000).

45

2.4 Alkali aggregate reactions (AAR)

Alkali Aggregate Reactions (AAR) also known as Alkali Silica Reactions (ASR) are
chemical reactions where alkali sodium and potassium ions in pore water solution
react with certain types of concrete aggregates, reaction is generally agreed to be a
dissolution reaction. This occurs as a result of the increased solubility of amorphous,
disordered or poorly-crystallized forms of silica minerals in high pH solutions. The
reaction forms a hygroscopic alkali-silica gel which absorbs water and swells. The
swelling forces generated may be sufficient to disrupt aggregate particles and the
surrounding concrete, causing expansion, cracking, and associated deterioration [1-4].
Aside from the environmental conditions (such as temperature and humidity) and the
alkali content etc., the development of AAR depends on the type of reactive rock
types in the aggregate and the aggregate grading (size). Some rock types are rapidly
reactive such as chert, and some rock types containing microcrystalline quartz are
slowly reactive as for instance the Norwegian rhyolite. The rapid reactions can
develop within a few years, while the slow reactions may develop for as much as 20
years before visual signs occur. Slow reaction could be the case if the content of
reactive rock types in the aggregate is very low, or the concrete contains no free
Ca(OH)2. Therefore if designed as par guidelines i.e. choosing aggregate constituents
on basis of their typical reactivity in AAR can ensure safety and durability of concrete
against AAR. The most reactive forms of aggregate are strained quartz, amorphous
silica, cryptocrystalline quartz, chalcedony, and chert. In general, aggregates
containing crystalline silica are stable and those with amorphous or very fine-grained
silica are reactive.
But if this is avoided, then AAR can cause significant damages in concrete. As it will
lead to cracking or unacceptable expansion of the concrete, which then disrupt the
microstructure of concrete. In field structures in many cases the cracking does not
lead to severe deterioration of the structure by itself, but it opens up for dangerous
side effects such as corrosion from CO2 and chloride intrusion, freeze thaw
mechanism etc.. This may lead to spalling of the concrete cover. This is an area of
concern which adversely affects the durability of concrete and therefore of bridges.

46

Two different theories are used as explanation for the mechanism causing the
expansion and ultimately the cracking of the concrete:

The silicious gel, which is a reaction product of the AAR, causes the
expansion of the concrete.

The primary expansion is caused by a solid expansion of the aggregate.


Secondary the gel can cause extra expansion when filling up closed cracks.

An expansion of several adjacent aggregate particles close to each other will lead to
tensile forces and will generate tensile stresses in surrounding concrete and if the
tensile stresses exceed the tensile strength of the concrete, cracking occurs. It has also
been observed that AAR causes a reduction in bond strength and a significant
reduction in the fatigue life. Prestress developed by the AAR expansion can enhance
the shear strength and stiffness of beams. Fig. 2.4.1 shows effect of aggressive AAR
reaction on concrete.

Fig. 2.4.1: Effect of AAR on concrete [2]


Visual detection of alkali-silica reactions is difficult in the early stages due to the
fineness of the cracks and may go unrecognized for years. If the concrete member is
unrestrained, visible concrete damage starts with small surface cracks exhibiting an
irregular pattern (or map cracking). When the expansive forces are restrained (e.g., by
reinforcement), the cracking pattern will be modified as the expansion will develop in
the direction of least constraint. Pop-outs and glassy appearing seepage of varying
composition can appear as a result of alkali-silica reactions [2].

47

References
1. Wigum B.J., Pedersen L.T., Grelk B. and Lindgard J. State-of-the art report: Key
parameters influencing the alkali aggregate reaction Report No. SBF52A06018,
SINTEF Building and Infrastructure Publication, (2006).
2. Naus D.J., Primer on Durability of Nuclear Power Plant Reinforced Concrete
Structures - A Review of Pertinent Factors, Report: NUREG/CR-6927, U.S.
Nuclear Regulatory Commission, Washington, D.C., (2007).
3. Basheer P. A. M., Chidiac S. E., and Long A. E., Predictive models for
deterioration of concrete structures, Construction and Building Materials, Vol.
10(1), pp. 27-37, (1996).
4. Grantham M. G. and Gray M. J., Diagnosis, Inspection, Testing and Repair of
Reinforced Concrete Structures M.G. Associates, Herts, (2000).

48

2.5 Acid Attack

Concrete having alkaline nature does not exhibit good resistance to acids in general
and has a natural tendency to neutralize the acids. In damp conditions, sulphur dioxide
(SO2) and carbon dioxide (CO2), as well as some other fumes present in the
atmosphere, form acids which attack concrete and combine with the calcium
compounds in the hydrated cement paste (i.e., calcium hydroxide, calcium silicate
hydrate, and calcium aluminate hydrate) to form soluble materials due to dissolution
of Ca(OH)2 and C-S-H, that then are readily leached away by aqueous solution from
the concrete[1-3].
Acid attack therefore cause dissolution and removal of hydrated cement paste in
concrete and leave a soft and very weak mass with high porosity and permeability.
Therefore it will allow other harmful species such as CO2 and Cl- to reach the rebar
through pores in the concrete, which can lead to rebar corrosion which will in turn
cause cracking and spalling of the concrete.
Acid Attack can therefore be typified by leaching, exposed aggregates and
disintegration in extreme cases. Therefore concrete undergoing acid attack will show
disintegration of the concrete in the form of loss of cement paste and aggregate from
the matrix. Fig. 2.5.1 presents effect of acid attack on concrete. It can sometimes be
difficult to diagnose since the acid is neutralized by the cement paste and may be
washed away. The common sources are spillage from acid tanks, acidic groundwater
and oxidation of sewage effluents, industrial plants, acid rains etc. [2 -3].

Fig. 2.5.1: Effect of Acid Attack on concrete [3]

49

The extent of acid attack is dependent on various factors such as:


Quantity of acid neutralizing material, such as the calcium hydroxide, C-S-H

gel, and limestone aggregates


Acidity of attacking acids
Solubility of the resulting calcium salts
Ability of hydrogen ions to be diffused through the cement gel (C-S-H) after

Ca(OH)2 has been dissolved and leached out


Nitric, hydrochloric, and sulfuric acids are very aggressive as their calcium salts are
readily soluble and removed from the acid front. Organic acids such as formic, acetic,
tannic and lactic are also harmful. Other acids such as phosphoric, carbonic, and
humic are less harmful as their calcium salts have low solubility and inhibit the attack
by blocking the pathways within the concrete. Oxalic acids have negligible affect on
concrete. Due to the large buffering capacity of concrete and the relatively small
amount of acid contained in rain, acid rains are not very harmful. Though carbonic,
humic, and sulfuric acids are the acids most commonly encountered by concrete since
they are found in natural ground waters and surrounding environment [2-3].
Acid Attack can be prevented or reduced by fixing it. This is achieved by treatment
with diluted water glass (sodium silicate) to form calcium silicates in the pores.
Surface treatment with coal-tar pitch, rubber or bituminous paints, epoxy resins, and
other agents can also be used. Surface coatings and a dense concrete with a low watercement ratio can also provide improved resistance to acid environments [2-3].
References
1. Basheer P. A. M., Chidiac S. E., and Long A. E., Predictive models for
deterioration of concrete structures, Construction and Building Materials, Vol.
10(1), pp. 27-37, (1996).
2. Grantham M. G. and Gray M. J., Diagnosis, Inspection, Testing and Repair of
Reinforced Concrete Structures M.G. Associates, Herts, (2000).
3. Naus D.J., Primer on Durability of Nuclear Power Plant Reinforced Concrete
Structures - A Review of Pertinent Factors, Report: NUREG/CR-6927, U.S.
Nuclear Regulatory Commission, Washington, D.C., (2007).

50

2.6 Sulphate attack

Attack by soluble sulphates is a well-appreciated problem for concrete structures. The


concrete is exposed to sulphates usually due to environmental actions, though they
can also be present in the concrete from binders or aggregates or from cement clinker.
Solid salts do not attack concrete but when present in solutions they can react with
hydrated cement paste. The most common solutions are sulphates of sodium (Na),
potassium (K), magnesium (Mg) and calcium (Ca).
These sulphates react with the hardened concrete and produce significant chemical
and physical effects. It causes chemical breakdown of cement paste as sulphate ions
attack alkali components of paste such as Ca(OH)2 and C-S-H and form hydrated
calcium sulphate (gypsum) and ettringites. These reactions lead to increase in volume
of the cement paste, thus changing the microstructure of concrete and producing
internal expansive stresses that cause deformation, cracking and eventually loss of
cohesion. The concrete subsequently disintegrates, with significant reductions in
compressive strength, tensile strength and elastic modulus. Damage usually starting at
the edges and corners and followed by cracking and spalling of the concrete. This
also enhances the permeability (continuity in the capillary pore structure) and thus
increases the migration of aggressive solutions into the interior and accelerates the
process of deterioration [1-4]. Concrete attacked by sulphates has a characteristic
whitish appearance because of formation of gypsum and etterngites.
Magnesium sulphate has a more damaging effect than other sulphates because it leads
to the decomposition of the hydrated calcium silicates as well as of Ca(OH)2 and of
hydrated C3A; hydrated magnesium silicate is eventually formed and it has no binding
properties.
The thaumasite form of sulphate attack (TSA) has recently been recognized as a
serious problem. Thaumasite is a calcium silicate sulphate carbonate hydrate that
turns concrete into a white pulpy mass. Unlike conventional sulphate attack, TSA
requires cold wet conditions and a source of carbonate, which can be the aggregate
used in the concrete [2].

51

The extent and form of sulphate attack depends on its concentration, the permeability
of the concrete, i.e. on the ease with which sulphate can travel through the pore
system and surrounding environment.
It is interesting to notice that, gypsum is added to the cement clinker in order to
prevent rapid set by the hydration of the C3A. Gypsum quickly reacts with C3A to
produce ettringite which is harmless because, at this stage, the concrete is still in a
semi-plastic state so that expansion can be accommodated. A similar reaction takes
place when hardened concrete is exposed to sulphates from external sources.
The effects of sulphate attacks in concrete can be controlled by following ways:
Reducing the C3A in the binders used in concrete
Reducing the Ca(OH)2 in concrete by addition of pozzolana
Reducing the permeability of the concrete; requires a low w/c ratio with good

compaction and curing


References
1. Neville A. M., Properties of Concrete, 4th edition, Wiley & Sons, New York
(1997).
2. Daly A. F, Modelling of deterioration in bridges, Transport Research
Laboratory (TRL), BRIME Report: PL97-2220, (1999).
3. Basheer P. A. M., Chidiac S. E., and Long A. E., Predictive models for
deterioration of concrete structures, Construction and Building Materials, Vol.
10(1), pp. 27-37, (1996).
4. Mangat P.S. and McLeish A. (editor), Underwater Concreting and Repair,
Chapter 6: Durability of concrete underwater, Elsevier Publication, (1994).

52

2.7 Freeze - Thaw damage

Concrete when in a saturated or near saturated condition if subjected to a wet


environment and freezing can be disrupted by a frost attack. Water enclosed in the
pores of the wet concrete will expand on freezing (approximately 9 % increase in
volume) and the high internal stresses so created can disrupt the surface. The effects
are intensified by subsequent freeze/thaw action as cracks develop which, in turn,
become filled with water.
Because as the temperature of concrete drops, freezing occurs gradually so that the
still unfrozen water in the capillary pores is subjected to hydraulic pressure by the
expanding volume of ice. Such pressure, if not relieved, can result in internal tensile
stresses in the concrete which on exceeding the tensile strength of the concrete the
pores will dilate and rupture may occur causing local failure. This would occur in
porous, saturated concrete containing no empty voids as liquid water cant move. On
subsequent thawing, the expansion caused by ice is maintained so that there is now
new space for additional water which may be subsequently imbibed. On re-freezing
further expansion occurs. Thus repeated cycles of freezing and thawing have a
cumulative effect, and it is the repeated freezing and thawing, rather than a single
occurrence of frost, which causes damage. Damage can occur in both the cement
paste and aggregate phases of the concrete. The damage is observed on exposed
surfaces of affected structures. [1-6].
Concrete deteriorates under the action of freezing and thawing due to one or more of
the following mechanisms:
Generation of hydraulic pressure due to freezing in capillaries
Diffusion of gel water into capillaries followed by freezing
Osmotic pressure resulting from partial freezing in capillaries of solutions with

a local salt concentration


Hydraulic and osmotic pressure develop in the pores when water freezes and expands.
Since there is a thermodynamic imbalance between the gel water and the ice,
diffusion of gel water into capillaries leads to a growth in the ice body and thus to an
increase of hydraulic pressure. Also, if salts are present in the pore solution the
osmotic pressure is increased since moisture tends to move toward zones of higher
53

salt concentrations. Osmotic pressure may also develop due to freezing alone. As
when the temperature drops, there is an initial period of super-cooling after which ice
crystals start to form in the large capillaries leaving alkalis in remaining unfrozen
solution. The result is an increase in the alkalinity of the unfrozen water in these areas
causing local increase in solute concentration. This creates an osmotic potential that
causes unfrozen water in neighboring capillaries to diffuse towards the site of
freezing. This in turn dilutes the high alkali water allowing more ice crystals to form
[5-6].
Damage to concrete due to freeze-thaw attack can take several forms such as [1-5]; a
schematic diagram of various damages is presented in Fig 2.7.1:
Internal damage: These damages are confined primarily to the mortar and are

associated with freeze-thaw damage of young concrete or cement paste that


doesnt have the pore structure which can resist stresses developed due to freezethaw attack. Internal damage mainly occurs due to freezing of water inside
concrete. Random cracks originate from the concrete but cracks also form parallel
to the exposed surface. These are of concern as they are also close and parallel to
the rebar.
Scaling: It is progressive delamination with each freeze-thaw cycle, often arising

due to application of de-icing salts or salts from sea which causes the temperature
of the concrete surface to change rapidly leading to thermal stresses. The surface
layer may become detached if the stresses occurring exceed the tensile or bond
strength holding the layer to the substrate causing eventual loss of sand and
aggregate particles and the concrete surface is gradually eroded. This affects the
strength of the member as well as the appearance and durability. Surface scaling is
a progressive problem. Scaling can be assumed to be linear with time so that the
loss of concrete cover can be estimated.
Pop- out: A pop-out is a small volume of concrete that has separated from the

body of the concrete to leave a roughly conical depression. The most common
cause of pop-outs is stress resulting from freeze-thaw action within the coarse
aggregate particle that causes cracking of the particle and simultaneous fracture of
the concrete between the particle and nearest concrete face.
D- cracks: Systems of vertical cracks visible on the surface and closely spaced at

joints and edges are called D-cracking and are commonly observed in pavements
54

since the concrete is most likely to be water saturated. When the surface shows Dcracking, the underlying concrete may likely be severely deteriorated. D-cracks
generally initiate within the non-durable aggregate [2].

Fig. 2.7.1: Types of freeze - thaw damage [5]


Factors controlling freeze-thaw action on concrete include:
Air entrainment: Size and spacing of air bubbles control the length of the flow

path to the nearest place for the water to escape, properly air-entrained concrete
with sufficient strength demonstrates good resistance to freeze-thaw
Water-cement ratio & curing: Controls permeability and strength, low w/c cement

ratio improves frost resistance


Presence of salts: Presence of salt from seawater or deicing salt, promotes scaling

damages in general
Degree of capillary saturation: Only highly saturated (80-90 % saturation level)

concrete is subjected to frost damage and this saturation level is called critical
saturation point
Aggregate materials: Durable and less porous aggregates reduce the damage
Temperature: Minimum temperature and rate of change affects freeze-thaw cycles

& ability of concrete to resist induced strains


Pore structure of concrete: Only capillary pores allow freezing of water to take

place and gel pores doesnt even at very low temperatures as they are very small
Number of freeze-thaw cycles: It is an progressive damage mechanism in general
Chemical constituents of concrete: Resistance can be increased by lowering C3A

and alkali content in concrete


55

Physical properties of concrete: Fracture strength and tensile strength of concrete

will decide the onset of cracks in concrete


Frost damage may be assessed by visual inspection. Practically assessment can also
be made after a number of cycles of freezing and thawing by measuring the loss in
mass of the specimen, the increase in its length, decrease in strength or decrease in the
dynamic modulus of elasticity, the latter being the most common.
References
1. Grantham M. G. and Gray M. J., Diagnosis, Inspection, Testing and Repair of
Reinforced Concrete Structures M.G. Associates, Herts, (2000).
2. Transportation Research Circular, E-C107 Control of Cracking in Concrete: State
of the Art, Transportation Research Board, Washington, D.C., (2006).
3. Mangat P.S. and McLeish A. (editor), Underwater Concreting and Repair,
Chapter 6: Durability of concrete underwater, Elsevier Publication, (1994).
4. Naus D.J., Primer on Durability of Nuclear Power Plant Reinforced Concrete
Structures - A Review of Pertinent Factors, Report: NUREG/CR-6927, U.S.
Nuclear Regulatory Commission, Washington, D.C., (2007).
5. Daly A. F, Modelling of deterioration in bridges, Transport Research
Laboratory (TRL), BRIME Report: PL97-2220, (1999).
6. Basheer P. A. M., Chidiac S. E., and Long A. E., Predictive models for
deterioration of concrete structures, Construction and Building Materials, Vol.
10(1), pp. 27-37, (1996).

56

2.8 Fatigue, Shrinkage and Creep

Fatigue, shrinkage and creep are damage mechanisms which arise under the action of
loads and dont involve chemical reactions with external agents. And stresses may be
generated due to applied loads, hydration process or environmental changes (thermal)
etc. and will cause progressive damages to concrete structures.
2.8.1 Fatigue

Fatigue is the result of applied repeated mechanical loadings, repeated reversal and
variation of stresses and strains or induced vibrations, and generally causes loss of
bond between the constituents of concrete, causing localized damage. For extreme
conditions, the resistance of the reinforced systems may be reduced or failures may
occur at applied stress levels less than yield [1-3].
In general concrete structures exhibit good resistance to fatigue and are designed for
influence of repetitive loads as par the design codes, so fatigue failure is unusual.
However, as structures age, there may be instances of local fatigue damage (due to
stress intensification, material flaws etc.) at locations where reciprocating equipments
cause repetitive loading, or at supports for pipes that exhibit flow induced vibrations.
Fluctuations in loading, temperature, or moisture content (that are not large enough to
cause failure in a single static application) lead to damage by fatigue in concrete
structures.
Fatigue damage initiates as microcracks in the cement paste, proximate to the large
aggregate particles and rebar. On continued or reversed load application, microcracks
may propagate to form structurally significant cracks that can expose the concrete and
reinforcing steel to severe environments or produce increased deflections. Ultimate
failure of a concrete structure in fatigue will occur as a result of excessive cracking,
excessive deflections, or brittle fracture. Therefore fatigue failure may be attributed to
the progressive deterioration of the bond between the coarse aggregate and the matrix
(typically starting along the interfacial zone between an aggregate and the paste). Also
fatigue failure may arise due to the coalescence of pre-existing micro-cracks in the
matrix, resulting in a single localized macro- crack under action of repetitive loads.
As cyclic loading proceeds, stresses are redistributed and the macro-crack width
decreases but never closes completely. Also Repeated cyclic loading causes
57

incremental bond slip and formation of sub microscopic cracks, which lead to stress
concentration due to notch effect [1-3].
The fatigue behavior of concrete structures is in general a function of the magnitude
of applied loads, strength and fracture energy of concrete. Fatigue resistance is also
affected by stress range, loading frequency and load history. In reinforced or
prestressed concrete, properties of the steel tend to control structural performance
since the steel carries the tensile loads [1-4].
2.8.2 Shrinkage

Shrinkage can be defined as reduction in volume of concrete as it hardens as water


escapes from matrix and evaporates further from the surface. Shrinkage can take place
in fresh concrete within first few hours after placement as well as in hardened
concrete after concrete has achieved its target strength. Shrinkage can result in
cracking and may affect the strength development in concrete.
Plastic shrinkage can occur at the surface of fresh concrete within the first few hours
after placement when concrete is still in plastic stage. If rate of evaporation of water is
faster than the rate of water bleeding from the matrix, the surface will begin to dry
resulting in high capillary stress development near the surface and cracks may appear.
This can be attributed typically to high temperatures, low ambient humidity, high
winds, and mixture ingredients and proportions [2, 5].
The potential for early-age cracking may increase with lower w/c concrete. This
phenomenon is termed as autogeneous shrinkage and refers to the loss of moisture
from the paste to allow hydration in a mix with low w/c. This suggests that in addition
to the evaporative effects, autogeneous effects may substantially add to the shrinkage
at early ages (in low w/c mixtures) thereby adding to the potential for plastic
shrinkage cracking.
Some, mostly igneous, aggregates can contain inclusions of weathered material in the
form of clay minerals. These minerals, in common with the clays encountered in the
ground, swell in the presence of moisture and shrink as they dry out in hardened

58

concrete paste. They phenomenon is termed as drying shrinkage of the concrete and
can cause a random crack pattern [5].
Cracks caused by plastic shrinkage can be quite wide (2-3 mm) on the surface of
concrete. It is also possible that the subsequent events including drying shrinkage and
loading can cause the plastic shrinkage cracks to propagate inside the concrete. These
cracks will make concrete more prone to damages as they will allow rapid access to
harmful material such as acids, chlorides and CO2 etc. and will also promote frost
attack, which might further damage the structures leading to the loss of durability.
The amount of shrinkage that occurs is directly related to the loss of water from the
concrete matrix, evaporation rate etc., greater evaporation leads to greater shrinkage.
Shrinkage cracking is also dependent on various factors such as the shrinkage rate and
magnitude, time dependent material property development, stress relaxation (creep),
strength, structural geometry, and the degree of structural restraint.
Shrinkage is a major problem for concrete structures specially for large flat structures
such as such as bridge decks and pavements, in which the exposed surface area is high
relative to the volume of the placed concrete [2, 5].
2.8.3 Creep

The phenomenon of a gradual increase in strain with time under a given level of
sustained stress is called creep. When a hydrated cement paste is subjected to a
sustained stress, depending on the magnitude and duration of applied stress, the C-SH will lose a large amount of the physically adsorbed water and may cause
rearrangement of bonds between the surfaces of the individual crystallites, and the
paste will show a creep strain. This can be of the same order of magnitude as drying
shrinkage and wmay cause significant deflections [6-7].
If a sustained load is removed, the strain decreases immediately by an amount equal
to the elastic strain at the given age; this is generally lower than the elastic strain on
loading since the elastic modulus has increased in the intervening period. This
instantaneous recovery is followed by a gradual decrease in strain, called creep

59

recovery. This recovery is not complete because creep is not simply a reversible
phenomenon.
Creep behavior of concrete is influenced significantly by various factors such that:
Concrete that exhibits high shrinkage generally also shows a high creep
When hydrated cement is completely dried, little or no creep occurs
Concrete with the lower relative humidity, shows higher creep
Strength of concrete has a considerable influence on creep and within a wide

range creep is inversely proportional to the strength of concrete at the time of


application of load. Therefore it follows that creep is closely related to the watercement ratio
Modulus of elasticity of aggregate controls the amount of creep that can be

realized and therefore concretes made with different aggregates exhibit creep of
varying magnitudes.
The rate of creep decreases continuously, and it is generally assumed that creep tends
to a limiting value. It has been estimated that 75 per cent of 20-year creep occurs
during the first year.
Creep of plain concrete does not by itself affect strength, although under very high
stresses creep hastens the approach of the limiting strain at which failure takes place.
The influence of creep on the ultimate strength of a simply supported, reinforced
concrete beam subjected to a sustained load is insignificant, but deflection increases
considerably and may in many cases be a critical consideration in design.
Another instance of the adverse effects of creep is its influence on the stability of the
structure through increase in deformation and consequent transfer of load to other
components. Thus, even when creep does not affect the ultimate strength of the
component in which it takes place, its effect may be extremely serious as far as the
performance of the structure as a whole is concerned. The loss of prestress due to
creep is well known and accounted for the failure of all early attempts at prestressing.
The effects of creep may thus be harmful. On the whole, however, creep is beneficial
in relieving stress concentrations and has contributed to the success of concrete as a
structural material.
60

Reference
1. Naus D.J., Primer on Durability of Nuclear Power Plant Reinforced Concrete
Structures - A Review of Pertinent Factors, Report: NUREG/CR-6927, U.S.
Nuclear Regulatory Commission, Washington, D.C., (2007).
2. Transportation Research Circular, E-C107 Control of Cracking in Concrete:
State of the Art, Transportation Research Board, Washington, D.C., (2006).
3. Gupta R.K., Residual Life Assessment of Steel Bridges, In: Proceedings of
National Conference on Advances in Bridge Engineering, Department of Civil
Engineering Indian Institute of Technology, Roorkee, (2006).
4. Sain T. and Chandra Kishen J. M., Damage and residual life assessment of
structures using fracture mechanics, 16th ASCE Engineering Mechanics
Conference, University of Washington, Seattle, (2003).
5. Grantham M. G. and Gray M. J., Diagnosis, Inspection, Testing and Repair of
Reinforced Concrete Structures M.G. Associates, Herts, (2000).
6. Mehta P.K. and Monteiro P. J. M. Concrete: Microstructure, Properties, and
Materials Third edition. McGraw-Hill, pp. 415. (2006).
7. Feldman R.F., Volume Change and Creep of Concrete, Canadian Building
Digest, CBD -119, (1969).

61

You might also like