Você está na página 1de 11

448

Chloride-induced corrosion of steel


in concrete
G K Glass and N R Buenfeld
Imperial College, London, UK

Summary
The corrosion of steel reinforcement in concrete
structures induced by chloride ion contamination is
a major problem. Deterioration starts with the loss
of protection provided by the concrete cover as the
result of chloride ingress. This is followed by
corrosion initiation and then propagation. Recent
advances in models of chloride penetration into
concrete, based on mathematical models of the
physical transport processes and an analysis of

empirical data, have been made. Inhibitive and


aggressive properties of solid phases present at the
steel which affect corrosion initiation have been
identified. Improved methods have been developed
to monitor corrosion rates as well as chloride
ingress into the concrete cover. An improved
understanding of the deterioration processes has
led to new developments in repair techniques.

Prog. Struct. Engng. Mater. 2000; 2: 448d458

Introduction
Concrete normally provides a non-aggressive
environment to reinforcing steel. During cement
hydration the aqueous phase rapidly acquires a high pH
that is buffered to resist downward changes at pH values
below 12.5 by the sparingly soluble hydration products
of cement. When steel is in contact with such an alkaline
solution, it is normally passive. The thermodynamically
most stable corrosion products are insoluble oxides, and
a passive oxide film covers the surface, presenting
a barrier to further metal dissolution[1]. However,
conditions may arise that render the passive film
unstable, leading to corrosion of reinforcing steel. The
two most important causes of this are carbonation and
chloride contamination of the concrete[2].
Carbonation is a problem that mainly affects
buildings. A reduction in the pH of the concrete pore
solution, which renders the passive film unstable, is
induced by the dissolution of atmospheric carbon
dioxide in the pore solution. The moisture content of
the concrete has a marked effect on both the
carbonation rate and subsequent corrosion rate. Water
is necessary for the carbonation reaction, but if the
pores are completely filled, the ingress of carbon
dioxide is severely hindered[3]. Once the carbonation
depth has reached the steel, significant rates of
corrosion require the presence of moisture to lower the
electrical resistance of the environment. The problem
may be arrested by removing severely carbonated
concrete and applying an anti-carbonation coating to
the concrete surface[4].
Copyright ^ 2000 John Wiley & Sons, Ltd.

The subject of this review is chloride contamination.


This may result in local disruption of the passive
film[5]. The source of chloride ions may be internal or
external. Internal sources include contamination of the
mix materials and the use of calcium chloride as a set
accelerator in construction. Limitations are placed by
current codes of practice on the acceptable levels of
chloride contamination resulting from the use of
contaminated mix materials, while the use of
chloride-containing admixtures for reinforced
concrete is generally not permitted[6]. Thus,
internal chlorides tend to affect only older existing
structures. External sources of chlorides include
de-icing salts and sea salt in marine
environments[7].
Chloride-induced corrosion deterioration of
concrete is a major problem[8,9]. It has, for example,
been estimated that the remediation of concrete
bridges in the USA, undertaken as a direct result of
chloride-induced corrosion of the reinforcing steel,
would cost the US state highway departments $5
billion in the year 2000[10]. In addition to bridges, other
highway structures, buildings and marine facilities are
affected (Fig. 1). An estimate of the annual cost
incurred in the UK as the result of corrosion damage to
concrete is 750 million[11].
Deterioration starts with the loss of protection
provided by the concrete cover. No physical damage
occurs during this stage. This is followed by corrosion
initiation and then propagation. These processes are
reviewed in this order, and monitoring and
remediation techniques are then addressed.
Prog. Struct. Engng Mater. 2000; 2:448}458

STEEL CORROSION

Fig. 1 Expansive corrosion products on steel in chloridecontaminated concrete

Chloride penetration
CONCRETE PORE SYSTEM
Concrete consists of a graded mix of aggregate
particles in a cement paste matrix. The cement paste
consists of unhydrated cement, hydration products
and the residue of the water-filled space, which give
rise to capillary porosity. The capillary porosity
reduces as hydration proceeds. The hydration
products consist mainly of calcium silicate hydrate gel
which has a porosity of around 28% (the gel pores),
and calcium hydroxide. The cement paste envelops
the aggregate particles and has a far higher porosity.
Capillary pores are up to 1 m in diameter, whereas
gel pores are around 2 nm. Concrete may also contain
entrained air, entrapped air and other voids.
Intentionally entrained air voids are bubbles typically
0.1 mm in diameter and are distributed evenly
throughout the cement paste. Accidentally entrapped
air usually forms very much larger voids, often up to
several millimetres in diameter. This air will typically
account for 1.5% of the volume of the concrete. Other
air voids include internal cracks, voids under
aggregate particles created by bleed water in concrete,
and honeycomb voids[12].
All concrete is porous, and chloride ion transport
occurs through the pore solution contained in these
pores. At low porosities, entrained and entrapped air
voids are distinct cavities, isolated from one another,
and hence will have little influence on transport in
concrete. When they are not filled with solution, they
will act to block ion transport. Cracks and other large
voids may be oriented to produce a continuous
network through the concrete, thereby significantly
reducing resistance to transport. In their absence,
capillary pores will dominate transport if
interconnected and filled with pore solution[13].
The rate of any transport process will depend on the
volume fraction, tortuosity and connectivity of the
pores. This is determined by factors such as the
water/cement ratio (w/c), cement content, cement
Copyright ^ 2000 John Wiley & Sons, Ltd.

449

fineness, cement type, use of cement replacement


materials, for example, ground granulated
blastfurnace slag (GGBS); pulverized fuel ash (PFA);
and silica fume (SF), concrete compaction, and degree
of hydration[14]. Reducing w/c (which controls the
original spacing of the cement grains) and prolonged
hydration may, for example, result in the capillary
pores becoming blocked by gel so that they are
interconnected solely by gel pores.
In addition to chloride ions Cl\, other ions present
in the pore solution in concrete include sodium Na>,
potassium K>, calcium Ca>, hydroxyl (OH\) and
sulphate SO24\. In fresh cement paste the typical
concentrations of Na> plus K> and OH\ is greater
than 0.2 mol l\1 (M), SO24\ is between 0.01 and 0.05 M
and Ca2> less than 0.01 M[15]. This will depend on the
chemical composition of the cement, w/c, use of
cement replacement materials, degree of hydration
and moisture content. In aged concrete, the
concentration of OH\ may fall to less than 0.05 M[1,16].
Cations are generally less mobile than anions[17].

TRANSPORT PROCESSES
Chloride ingress into concrete can occur by a number
of mechanisms. These include diffusion due to
a concentration gradient, migration in an electric field,
and water flow[7]. The rate of diffusion is described by
the diffusion coefficient. This parameter gives the flux
of a species (quantity passing through a unit area per
unit time) per unit concentration gradient. Migration
is determined by the ionic mobility. This mobility
gives the average velocity per unit electric field. Water
flow may result from a pressure gradient, absorption
into partially dry concrete, wick action or electroosmosis (the movement of water under the influence
of an electric field) each of which may have its own
transport coefficients[12].
Chloride transport is affected by the pore structure
and the interaction between charged species. The
physical effect of the pore structure is generally
incorporated into the measured values of the transport
coefficients defining the various transport processes.
In addition, chemical interactions between ions in the
pore solution and with the pore walls affect the
movement of chloride ions through concrete. The
resistance to chloride transport is usually several
orders of magnitude greater in concrete than what
might be predicted from data obtained in simulated
pore solutions[13,14].
There are two main effects on chloride transport
arising from interactions between ions in the pore
solution, namely, the separation of charge is
constrained and the effective concentration may differ
from the actual concentration of a species.
The condition of electroneutrality implies that,
when no ion source or sink is present, the net quantity
of positive charge transported by cations must be
equal to the negative charge transported by anions,
Prog. Struct. Engng Mater. 2000; 2:448}458

450
after any initial transient effects have decayed. The
transport of an excess of ions of one particular type
(e.g. Cl\) in a given direction may occur only if other
ions of the same sign (e.g. OH\) are transported in the
opposite direction[18]. Thus, when the diffusion of
sodium chloride into concrete takes place, chloride ion
transport might be retarded by its association with
a slower, charge-balancing sodium ion if no significant
counter-diffusion of hydroxyl or sulphate ions occurs.
It may be noted that, when an external electric field is
applied, a source or sink for ions is provided by the
electrode reactions. This allows the positive and
negative ions to move in opposite directions, the
proportion of the current carried by each ion being
determined by their individual mobilities, charge
number and concentration[19].
The effective concentration of an ion resulting from
its non-ideal behaviour is termed its activity. The
theoretical description of the transport of a species
may, to a good approximation, be expressed in
terms of its concentration when the concentration is
low. When dealing with higher concentrations and
strong electrolytes the errors may be more
significant[20].
The two main interactions between ions and the
pore walls are chloride binding and membrane
effects[21,22]. Chloride binding in concrete may be
defined as the interaction between the porous matrix
and chloride ions which results in the ions effective
removal from the pore solution. All cements bind
a proportion of the chloride present. This action will
remove chloride from the transport process as well as
alter the pore solution concentration and therefore the
concentration gradient driving diffusion. While there
are many factors associated with the constituents of
the concrete, the composition of the pore solution and
the external environment that affect chloride binding,
the most important factor is the cement type. A change
in w/c has a significant effect on the free chloride
concentration, but this action results mainly from the
change in pore solution volume and its effect on the
quantity of bound chloride is much smaller[23].
Ion exchange membrane effects are produced by
a surface charge on the pore walls. This charge results
from the dissolution of a net quantity of ions with the
opposite charge. The charge on the pore walls is
balanced by an equal and opposite charge in the pore
solution. Indeed, for narrow pores or low
concentrations, only charge-balancing ions may be
present in the pore solution. An ion exchange
membrane therefore favours the transport of the
charge-balancing species through its pore system[24].
The sign and value of the surface charge on the pores
in cement paste will depend on the electrolyte
composition and pH as well as the binder type and
composition of the hydration products[12,22]. In the
absence of membrane effects, the difference in the
mobilities of the anions and cations will produce
a potential difference, termed a junction potential.
Copyright ^ 2000 John Wiley & Sons, Ltd.

CONCRETE CONSTRUCTION

Fig. 2 Chloride profile in concrete

When this is enhanced by membrane effects, it is


termed a membrane potential[25].
Models of chloride penetration into concrete may be
based on mathematical models of the physical
transport processes or an analysis of empirical data.

MODELLING PHYSICAL TRANSPORT PROCESSES


The flux J of an ion in an ideal solution is given by
a form of the NernstPlanck equation:
J"!Dp

*C
*
!C #Cu
*x
*x

(1)

where C, Dp, , , and u are the concentration,


diffusion coefficient, ionic mobility, electric potential
and average mass velocity, respectively[26]. The three
terms on the right-hand side of eq. 1 represent the flux
due to diffusion, migration and water movement
respectively. The spatial distribution of the species is
then given by a solution of the equation:
*C *J
" #r
*t *x

(2)

where r is the rate of production/consumption of the


species[12].
The model may be simplified if one transport
process dominates chloride ingress. If this process is
assumed to be diffusion, and the surface chloride
concentration C0 and diffusion coefficient are constant
(independent of time), and chloride binding may be
neglected (r"0), eqs 1 and 2 have an exact solution.
For diffusion into a semi-infinite medium, this
solution is given by:
C(x, t)"C0 erfc

2(Dpt

(3)

The pore system diffusion coefficient Dp characterizes


the velocity of ions under the influence of a
concentration gradient through the pore system[27]. If
the effects of linear chloride binding are included in
eqs 1 and 2, the pore system diffusion coefficient in
eq. 3 will be replaced by the apparent diffusion
coefficient Da (Fig. 2)[7].
Prog. Struct. Engng Mater. 2000; 2:448}458

STEEL CORROSION
However, the underlying assumptions render the
model represented by eq. 3 an inaccurate description
of chloride ingress into concrete for most cases. The
diffusion coefficient will change with time and depth,
depending on the variable moisture content of the
concrete and continued hydration of the cement[28].
The surface chloride content will also vary with time,
owing to variations in the exposure environment.
Chloride binding may be nonlinear and time
dependent. The principal transport process may not be
diffusion. For example, in cyclic wet/dry
environments, ingress may occur via absorption of
chloride-contaminated water[29].
As noted above, the condition of electroneutrality
imposes constraints on the transport process. Thus, for
example, the diffusion coefficients referred to above
should be determined under conditions that include
this effect. The diffusion coefficient describing ionic
diffusion in the absence of this constraint is referred to
as the self-diffusion coefficient D$. The mobility of
ions (the average velocity of the ions per unit electric
field) is related to the self-diffusion coefficient through
the NernstEinstein equation:
zFD$
"
RT

(4)

where z, F, R and T are the charge number, Faraday


constant, gas constant and absolute temperature,
respectively[17]. The effects that are included in the
value of the diffusion coefficient will be dependent on
how it is measured[27].
The various constraints and transport processes
may be considered separately and their effects
combined[30]. This combination assumes that the
individual processes are additive, as is the case in an
ideal solution. To consider the constraint imposed by
electroneutrality, a separate form of eq. 1 with the
appropriate self-diffusion coefficient is required to
model the transport of each type of ion. Numerical
methods are then necessary to solve the resulting
equations. The solutions will be complex, particularly
if the effects of time-dependent chloride binding,
charge separation and more than one transport
process are considered[31,32]. When an electric field is
imposed, oxidation and reduction reactions may
induce changes affecting the local pressure and
electric field[19,33]. If the model is too complex, it may
require parameters that are not readily measurable.
Nevertheless, some success has been achieved with
mathematical models of a combination of physical
transport processes[32,34,35].

EMPIRICAL MODELS
Empirical models are based on an analysis of
empirical data. For example, many chloride profiles
follow the form given by eq. 3, even though the
underlying assumptions are incorrect. This provides
an empirical basis for the use of eq. 3 to describe the
Copyright ^ 2000 John Wiley & Sons, Ltd.

451

chloride profile. It is important to note that, even


though the chloride profile is characteristic of
a diffusion process, it is not necessary to assume that it
was caused solely by diffusion[36].
One area where eq. 3 has been used in an empirical
manner is in the development of probabilistic models
of chloride ingress into concrete. The ingress of
chloride, corrosion initiation and the subsequent
propagation of corrosion damage depend on a number
of factors that may be assessed only with substantial
uncertainty. The uncertainty arises not only from
uncertainties in the material characteristics and
environmental conditions, but also from uncertainties
in the mechanism of the deterioration process[37].
A probabilistic model attempts to address these
uncertainties[38].
The principal input parameters of a chloride ingress
model that have significant statistical scatter are the
chloride content at the concrete surface, the transport
coefficient, the cover depth and the chloride threshold
level[39]. The output is the predicted probability to
corrosion initiation. More complex models have
included the effects of the subsequent rate of
deterioration and change in load to predict structural
reliability[40]. Such models are gaining acceptance as
a method of assessing risk of failure[38].
Another important empirical approach to modelling
chloride ingress uses neural network analysis to
identify the relationships between a range of possible
parameters describing the material and exposure
conditions and the resulting chloride profile in
reported data[41]. A complex equation is represented
by a network of elements that process input data to
generate an output. The method is equivalent to
finding the best fit to data in multidimensional space,
and is able to separate the effects of factors that are
otherwise difficult to assess individually in
a laboratory environment. However, it may be
necessary to constrain an output to prevent nonphysical predictions[23]. Nevertheless the technique
has shown some promise in predicting chloride
profiles[41].

Chloride-induced corrosion initiation


MECHANISM OF CORROSION INITIATION
The oxides which make up the passive film on iron are
thermodynamically stable in the alkaline environment
in concrete even when chloride ions are present[1]. In
this situation, corrosion tends to be localized and
chloride-induced corrosion initiation in concrete
follows the model of pitting corrosion. It is a two-stage
process in which pit nucleation is followed by pit
growth[42]. The causes of pit nucleation are still subject
to much debate. However, it is widely recognized that
pit nucleation is usually followed by repassivation. To
prevent repassivation, pit nucleation must be
accompanied by a local fall in pH and increase in
Prog. Struct. Engng Mater. 2000; 2:448}458

452

CONCRETE CONSTRUCTION

Fig. 4 The resistance to a reduction in pH and soluble chloride


content determined on chloride-contaminated OPC concrete[42]
Fig. 3 Backscattered electron image of a polished cross-section
of steel in concrete[45]

chloride content at the pit nucleation site. The local fall


in pH renders the passive film locally unstable and the
presence of chloride ions promotes the dissolution of
iron and stabilizes the local fall in pH through the
formation of hydrochloric acid[43].
The effect of this mechanism is evident at voids at
the steelconcrete interface. It was noted that corrosion
tends to initiate at the location of such defects. This
was attributed to the absence of a lime layer on the
steel at these locations which would otherwise release
hydroxyl ions to prevent a local fall in pH, thereby
inhibiting corrosion initiation[44]. However, evidence
of any substantial increase in lime at the steel is limited
(Fig. 3), and inhibitive effects have also been attributed
to many other hydration products of cement that
dissolve to release hydroxyl ions at high pH
values[43,45].
Titrimetric methods have been developed to
determine the resistance to pH reduction that gives
rise to this inhibitive property of the solid phases. One
such method has been termed differential acid
neutralization analysis[46]. In this case, the resistance to
pH reduction is given by the acid added per unit pH
reduction of an aqueous suspension of ground solid.
An example of the data obtained on an aqueous
suspension of ground OPC concrete that had
previously been exposed to a neutral chloride solution
is given in Fig. 4. The dissolution of the solid phases
gives rise to peaks in the data at various pH values.
Chloride ions bound in the solid phase may act to
promote corrosion via the same mechanism that
operates when hydroxyl ions bound in the solid phase
act to inhibit corrosion[42,47]. The pH-dependent
dissolution characteristics of chloride are included in
Fig. 4. It was noted that, for this sample, virtually all
the chloride that was bound in the solid phases was
released before the pH reduced to 11. As the passive
film is still thermodynamically stable at this pH, it
might be assumed that all the chloride released will be
available to sustain local passive film breakdown.
Copyright ^ 2000 John Wiley & Sons, Ltd.

Evidence of the participation of bound chloride in


corrosion initiation comes from the absence of any
significant correlation between chloride binding and
chloride threshold level[5].

CHLORIDE THRESHOLD LEVEL


The chloride threshold level may be defined as the
quantity of chloride at the steel that is necessary to
sustain local passive film breakdown, and hence
initiate the corrosion process. It is usually presented as
a ratio of the total chloride to cement content of
concrete (expressed as a weight percentage). Typical
values range between 0.2 and 2.5% by weight of
cement[47]. While the chloride content is relatively easy
to determine, it should be noted that the cement
content is often only estimated, as laboratory
verification of this content is more difficult.
A ratio of the free chloride and hydroxyl
concentrations in the pore solution of the concrete has
also been promoted as an index to represent the
chloride threshold level[48,49]. This is based on the
hypothesis that only species in solution can affect
corrosion risk. However, as indicated above, this is not
the case. For example, there is no evidence of any
substantial correlation between chloride binding and
chloride threshold level[5]. Indeed, the standard
cement types that are known to bind very different
amounts of chloride appear to have very similar
threshold levels (Fig. 5)[47,50]. Furthermore, the use of
free concentration ratios is not supported by the
observation that this does not result in a narrow range
of threshold levels[51]. Indeed, it exacerbates this
problem. Other anomalies associated with the
moisture content of the concrete also arise when pore
solution concentrations are used[47].
More recently, a new index to rank the risk of
corrosion initiation has been proposed[52]. This index is
based on the soluble chloride present after the pH of
a ground suspension of concrete is reduced to the
endpoint of the phenolphthalein indicator, and the
acid that is required to reduce the pH to this level. This
representation includes the effects of the
Prog. Struct. Engng Mater. 2000; 2:448}458

STEEL CORROSION

453

binder (pH buffering) and the condition of the


steelconcrete interface (electrochemical potential and
millscale). Corrosion initiation may be prevented if the
concrete is too dry, while in very wet environments
the corrosion rate following initiation may be limited
by the restricted access to oxygen. One of the effects of
a dry environment may be to prevent moisture
entering the entrapped air voids at the steel.

Corrosion rate
Fig. 5 Chloride contents at which corrosion was observed for
various cement types[50]

Fig. 6 Risk of corrosion initiation as a function of chloride


content[53]

pH-dependent dissolution characteristics of the solid


phases and overcomes the measurement difficulties
that are associated with the determination of cement
content or pore solution concentrations. It may be
expressed as a mole ratio (moles Cl\ released/moles
H> consumed). A mole ratio of 0.01 would typically
equate to a total chloride content of 0.5% by weight of
cement.
A substantial amount of work has been undertaken
on chloride threshold levels, but this has not yet
resulted in a model refining the threshold level for
a particular set of circumstances. The most important
factor appears to be the presence of voids at the steel.
These voids will present a corrosion risk when they
are partially filled with pore solution. However, the
effect of this parameter has not yet been reported. It
therefore gives rise to a high level of apparently
random scatter. Thus, the chloride threshold is at
present best considered in terms of representing
a probability of corrosion initiation (Fig. 6)[53].
Numerous other factors may affect the chloride
threshold levels to a smaller extent[54]. These include
factors associated with the external environment
(moisture, temperature and chloride source), the
barrier properties of the concrete cover (w/c and cover
depth), the chemical properties of the cementitious
Copyright ^ 2000 John Wiley & Sons, Ltd.

Once passive film breakdown has occurred,


a corrosion cell is established. An ionic current flows
between the anodic areas, where metal dissolution is
occurring, and the cathodic areas, where oxygen is
being reduced. This results in an increase in chloride
content and the production of acid at the anodic areas.
Further disruption of the adjacent passive film then
occurs[55].
Corrosion rates may be expressed as a current
density, a rate of weight loss or a rate of section loss.
For steel, a corrosion current of 1 mA m\2
approximates a weight loss of 10 g m\2 yr\1, which in
turn approximates to a section loss of 1 m yr\1.
Reinforcing steel corrosion rates of 1 mA m\2 or less
are considered to be negligible and are not likely to
result in disruption of the concrete cover. Higher
corrosion rates are considered to be significant and, if
oxygen has relatively easy access to the steel (the pore
system is not saturated with water), the general
corrosion rate may exceed 100 mA m\2[55]. Local
corrosion rates of up to 1000 mA m\2 are possible.
The essential processes in a corrosion cell are an
anodic reaction (metal dissolution), a cathodic reaction
(typically oxygen reduction), the movement of
electrons through the metal between the anodic and
cathodic sites and the movement of ions through the
environment between these sites. The rate of corrosion
may be controlled by any one of these processes.
However, the resistance to electronic current flow
within a rebar is usually very small. Thus, in practice
the rate at which a metal is consumed can be
controlled by the kinetics of either the anodic or
cathodic reactions, or by the resistance to current flow
between the anodic and cathodic sites through the
concrete. The associated controlling mechanisms are
termed anodic control, cathodic control or resistive
control. Reinforced concrete is exposed to a wide
variety of environments, and any one of these three
mechanisms may dominate, depending on the
exposure conditions[2].
In submerged marine environments, chloride
contamination occurs relatively easily. However,
access to oxygen is severely restricted by its low
concentration in solution and its slow transport rate
through waterlogged concrete. The corrosion rate will
then depend on the rate at which oxygen can reach the
steel. It is said to be under cathodic control[55].
Prog. Struct. Engng Mater. 2000; 2:448}458

454

Fig. 7 Steel corrosion rate as a function of chloride content and


relative humidity in carbonated OPC mortar[56]

In above-ground structures, the presence of


a passive film often restricts the overall rate, the
corrosion rate effectively being under anodic control.
To increase the corrosion rate, further passive film
breakdown must occur. One of the factors affecting the
corrosion rate is therefore the level of chloride
contamination. The chloride content will also affect the
corrosion rate in dry or carbonated concrete via its
effect on the resistivity of the environment. Indeed,
chloride contamination may restrict the drying
process[56]. An example of this is given in Fig. 7.
Other factors such as temperature and relative
humidity will also affect the corrosion rate, and
advances have been made-in monitoring these
effects[57]. The quantity of corrosion has also been
related to the risk of corrosion-induced cracking of the
concrete cover[58]. These parameters may be used in
refining models of the risks associated with corrosioninduced deterioration after corrosion initiation has
occurred.

Monitoring corrosion
CURRENT CONDITION
Monitoring may be used to assess existing
deterioration and predict future performance. An
assessment of existing deterioration is a standard
requirement in the repair of corrosion-damaged
concrete[59]. The current condition of reinforced
concrete may be defined by the state of the
reinforcement and the deterioration and effective loss
of the cover concrete. If the steel is corroding, spalling
of the concrete cover can be a safety hazard, while the
loss of steel cross-section and bond strength may affect
load-carrying capacity.
Visual inspection may be used to assess a problem
in its advanced state. Traditional non-destructive,
in-situ techniques to monitor corrosion and its extent
include corrosion potential mapping, delamination
testing, corrosion rate measurement, and concrete
resistivity measurement. Corrosion potential and rate
measurements give an indication of the condition of
the passive film on the steel[57]. The history of the local
Copyright ^ 2000 John Wiley & Sons, Ltd.

CONCRETE CONSTRUCTION
corrosion rate can give an indication of any local loss
in steel section. Resistivity measurements give an
indication of the aggressive nature of the cover
concrete[60], while cracks may be detected by
ultrasonic pulse velocity measurements[53].
Acoustic emission has recently been used to follow
the development of cracks in the concrete cover[61].
Novel non-destructive methods of imaging the
reinforcement in-situ are under development[6264].
These methods are based on an analysis of the effect of
the reinforcing steel on the local magnetic field, and an
analysis of radar images of the concrete.
Destructive tests include removing samples for
laboratory analysis to determine chloride profiles[65]. It
is generally advised that non-destructive techniques
should be calibrated with a limited amount of
destructive assessment if possible[53].

MONITORING TO PREDICT PERFORMANCE


Monitoring may be used to identify future problems to
facilitate planned, cost-effective maintenance and
provide data to improve future decisions. To predict
performance, the rate at which the corrosion
protection afforded by the cover is lost and the rate of
corrosion-induced deterioration following passive
film breakdown are required.
The rate of loss of protection provided by the cover
is determined by the source of chloride and its rate of
ingress. Approximate transport coefficients that
characterize diffusion, migration, the movement of
water and the capacity of the cement to immobilize
chloride may be extracted from samples removed for
laboratory analysis. However, factors such as
temperature and degree of saturation, which are likely
to be time dependent, will affect the transport
processes occurring and should be monitored on
site[66].
Some innovation is necessary to extract transport
coefficients by non-destructive techniques. Recent
developments have been made in impedance and
resistivity measurements[67]. These contain
information on the moisture content, pore structure
and degree of contamination of the concrete. Such
techniques have already been used to study wetting
and drying of concrete[29]. Other methods of
predicting deterioration have used galvanic corrosion
probes[68] and chloride ion sensors[69]. However, the
long-term stability of chloride ion sensors is unproven
and galvanic corrosion probes should be installed
prior to casting the concrete to accurately reflect the
changing local environmental conditions.
Once corrosion initiates, the corrosion rate will
determine the subsequent rate of deterioration. As
noted above, average data are required as this varies
with time. A popular method of obtaining the
corrosion rate is to measure the polarization
resistance. Recent advances in corrosion rate
measurements include the analysis of potential time
Prog. Struct. Engng Mater. 2000; 2:448}458

STEEL CORROSION

Fig. 8 Transformation of a coulostatically induced potentialdtime


transient (top) into the frequency domain (bottom) to give the
polarization resistance[71]

transients produced by galvanostatic or coulostatic


perturbations to extract parameters characterizing
a corroding steel interface (Fig. 8)[70,71].

Repair of chloride-contaminated concrete


The repair of concrete structures is becoming
increasingly important, owing to the long service lives
required and the high cost of building and
maintaining the infrastructure. The 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) 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[72]. 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[73].
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. Recent developments
include placing a small sacrificial anode covered with
an appropriate mortar in the patch repair material.
Zinc in a mortar, which prevents zinc passivation and
is able to accommodate the expansive corrosion
products, has been used[72].
Impressed current cathodic protection has been
extensively used as a repair technique for
chloride-contaminated concrete. A related technique is
the short-term process of electrochemical chloride
removal. It was believed that these techniques where
fundamentally different. The principal protective
effect of cathodic protection was the achievement of
a sustained negative potential shift, while the principal
effect of chloride removal was to extract chloride from
the concrete[55]. However, recent work suggests that
Copyright ^ 2000 John Wiley & Sons, Ltd.

455

the mechanism of protection operating in both cases is


very similar; namely, the current changes the
environment at the steel which promotes passive film
formation[74]. Indeed, the removal of chloride may not
be as important as the precipitation of solids
containing inhibitive hydroxyl ions at the location of
voids at the interface where the risk of corrosion
initiation is high. This will raise the chloride threshold
level[75].
Recent innovations in these techniques have
included their extension to areas where the current is
applied intermittently[33,74], improvements in the
assessment criteria[76], and improvements in the
anodes used[77]. Developments have also been made in
overcoming the problems associated with the use of
sacrificial anodes applied to the surface of concrete[78].
Other proposed methods of treating
chloride-contaminated concrete use surface
treatments. These may promote drying of the concrete,
and hence reduce the risk of corrosion initiation and
slow down chloride ingress. However they tend not to
be successful when corrosion damage is already
occurring[73]. Inhibitors have also been used. However,
inhibitors applied to the concrete surface as a repair
technique have no substantial track record, and
published data suggest they may even stimulate
corrosion-induced deterioration in some cases[79].

Conclusions
The corrosion of steel reinforcement in concrete
structures induced by chloride ion contamination is
a major problem. Deterioration starts with the loss of
protection provided by the concrete cover. Physical
damage does not occur during this stage. The high pH
of the concrete environment maintains a stable passive
oxide film on the reinforcing steel. This is followed by
corrosion initiation and then propagation.
Chloride ingress into concrete can occur via
diffusion, migration and water flow. It is affected by
the pore structure, interactions with the pore walls and
interactions with other ions in the solution. Two
important interactions with the pore walls are chloride
binding and membrane effects. Interactions with other
ions in the solution are associated with the constraint
imposed by charge separation and the activity
coefficient. Models of chloride penetration into
concrete, based on mathematical models of the
physical transport processes and an analysis of
empirical data, have been developed.
Chloride-induced corrosion initiation of the steel
requires a local fall in pH to follow a pit nucleation
event. The solid hydration products have an effective
buffering capacity that resists such a local pH
reduction. Thus, corrosion initiation is promoted by
the presence of voids at the steel. The local pH
reduction also results in the release of most of the
Prog. Struct. Engng Mater. 2000; 2:448}458

456

CONCRETE CONSTRUCTION

locally bound chloride which then acts to prevent


repassivation.
Once passive film breakdown has occurred, the
corrosion rate is controlled by either the kinetics of the
anodic or cathodic reactions, or by the resistance to
current flow between the anodic and cathodic sites
through the concrete. Reinforced concrete is exposed
to a wide variety of environments and any one of these
three mechanisms may dominate, depending on the
exposure conditions.
Monitoring may be used to assess existing
deterioration and predict the future performance.
Recent advances include developments in the area of
non-destructive imaging of the reinforcing steel and
the use of impedance and resistivity measurements to
extract information on the moisture content and
chloride contamination of the cover concrete.
The repair of chloride-induced corrosion damage
has always presented problems because of the
localized nature of the corrosion which provides
a form of electrochemical protection to the steel in the
adjacent chloride-contaminated concrete. Advances
have been made in the understanding of
electrochemical protection, and its area of application
has been extended to include intermittent cathodic
protection. Galvanic anodes, which may be embedded
in repair mortars and applied to the surface of the
concrete, have been developed.

[9] Smith JL & Virmani YP. Materials and methods for corrosion control of
reinforced and prestressed concrete structures in new construction. Federal Highway
Administration Report, FHWA-RD-00-081, Washington DC. June 2000.
[10] Federal Highway Administration. Building More Durable Bridges.
FOCUS Federal Highway Administration, Washington DC, September 1999. p. 6.
[11] BRE Centre for Concrete Construction. Guide to the maintenance,
repair and monitoring of reinforced concrete structures. DME Report No. 4.
Watford, UK: Building Research Establishment. May 2001.
* [12] Buenfeld NR, Glass GK, Hassanein AM & Zhang J-Z. Chloride
transport in concrete subjected to an electric field. Journal of Materials in Civil
Engineering 1998: 10(4): 220d228.

References and recommended reading


Papers of particular interest have been marked:
* Special interest
** Exceptional interest
** [1] Page CL & Treadaway KWJ. Aspects of the electrochemistry of steel
in concrete. Nature 1982: 297(5862): 109d115.
A far-sighted overview of the basic mechanism of corrosion of steel in concrete
and its implications.
[2] Glass GK & Buenfeld NR. Reinforced concrete d the principles of its
deterioration and repair. In: Macdonald S (ed) Modern Matters d Principles and
Practice in Conserving Recent Architecture. Shaftesbury, UK: Donhead Publishing.
1996. 101d112.
[3] Buenfeld NR, Hassanein NM & Jones AJ. An artificial neural network
for predicting carbonation depth in concrete structures. 2nd ASCE Monograph on
Artificial Neural Networks in Civil Engineering, Reston, VA: American Society of Civil
Engineers, 1997. 77d117.
[4] Seneviratne AMG, Sergi G & Page CL. Performance characteristics of
surface coatings applied to concrete for control of reinforcement corrosion.
Construction and Building Materials. 2000: 14(1): 55d59.
[5] Glass GK, Reddy B & Buenfeld NR. The inhibitive properties of
concrete in a chloride containing environment. 14th International Corrosion
Congress, Cape Town, September 1999: Paper No. 47.1.
[6] British Standards Institution. BS 1881, Part 124, The testing of
hardened concrete. London: British Standards Institution, 1988.
* [7] Bamforth PB, Price WF & Emerson M. International review of
chloride ingress into structural concrete. TRL Contractor Report 359, Transport
Research Laboratory, Edinburgh, Scotland, 1997.
A review of many aspects of chloride-induced corrosion that collates a wide
variety of data from many sources.
* [8] Creegan PJ, Graham JR, Tatro SR, Herreryherrera AE, Kaden
RA, McDonald JE & Schrader EK. Compendium of case-histories on repair of
erosion-damaged concrete structures. ACI Materials Journal 1994: 91(4): 408d409.
An overview of a number of case histories of catastrophic corrosion induced
failure in reinforced concrete.

Copyright ^ 2000 John Wiley & Sons, Ltd.

A review of basic transport processes affecting chloride ions in porous media.


[13] Ngala VT, Page CL, Parrott LJ & Yu SW. Diffusion in cementitious
materials 2. Further investigations of chloride and oxygen-diffusion in well-cured OPC
and OPC / 30% PFA Pastes. Cement and Concrete Research 1995: 25(4): 819d826.
[14] Atkinson A & Nickerson AK. The diffusion of ions through
water-saturated cement. Journal of Materials Science 1984: 19(9): 3068d3078.
[15] Page CL & Vennesland O. Pore solution composition and chloride
binding capacity of silica-fume cement pastes. Materials and Structures 1983: 16(91):
19d25.
[16] Hassanein AM, Glass GK & Buenfeld NR. Effect of intermittent
cathodic protection on chloride and hydroxyl concentration profiles in reinforced
concrete. British Corrosion Journal 1999: 34(4): 254d261.
[17] Atkins PW. Physical Chemistry. 5th edn. Oxford, UK: Oxford University
Press. 1994.
** [18] Yu SW, Sergi G & Page CL. Ionic diffusion across an interface between
chloride-free and chloride-containing cementitious materials. Magazine of Concrete
Research 1993: 45(165): 257d261.
Experimental and theoretical analysis of diffusion subject to the constraints
imposed by charge conservation and chloride binding.
[19] Li LY & Page CL. Finite element modelling of chloride removal from
concrete by an electrochemical method. Corrosion Science 2000: 42(12):
2145d2165.
[20] Li LY & Page CL. Modelling of electrochemical chloride extraction from
concrete: Influence of ionic activity coefficients. Computational Materials Science
1998: 9(3d4): 303d308.
[21] Glass GK & Buenfeld NR. The influence of chloride binding on the
chloride induced corrosion risk in reinforced concrete. Corrosion Science 2000:
42(2): 329d344.
[22] Zhang J-Z & Buenfeld NR. Membrane potential and its influence on
chloride transport in cementitious materials. In: Andrade C (ed) 2nd International
RILEM Workshop on Testing and Modelling the Chloride Ingress into Concrete. Paris,
France: Rilem Publications. 2000.
** [23] Glass GK, Hassanein NM & Buenfeld NR. Neural network modelling
of chloride binding. Magazine of Concrete Research 1997: 49(181): 323d335.
A detailed examination of factors affecting chloride binding that includes the
development of a model to predict binding isotherms.
[24] Kyrota J. Ions electrodes and membranes. 2nd edn. Chichester, UK: John
Wiley & Sons. 1992.
* [25] Zhang J-Z & Buenfeld NR. Measuring the membrane potential across
cement-based materials. Materials and Structures 2000: 33(232): 492d498.
The measurement and implication of the membrane potential that develops as
ionic contamination of concrete occurs.
[26] Weber WJ, McGinley PM & Katz LE. The nature and effects of
sorption processes in subsurface systems. In: Bear J & Corapcioglu MY (eds)
Transport processes in porous media. NATO ASI Series E: Applied Sciences. Vol. 202.
Netherlands: Kluwer. 1991. 543d582.
* [27] Glass GK & Buenfeld NR. Theoretical assessment of the steady state
diffusion cell test. Journal of Materials Science 1998: 33(21): 5111d5118.
An examination of the various types of diffusion coefficients that also introduces
a novel technique to determine chloride binding isotherms from chloride profiles.
[28] Climent MA, De Vera G, Lopez J, Garcia D & Andrade C.
Transport of chlorides through non-saturated concrete after an initial limited chloride
supply. In: Andrade C (ed) 2nd International RILEM Workshop on Testing and
Modelling the Chloride Ingress into Concrete. Paris, France: Rilem Publications. 2000.
[29] McCarter WJ, Chrisp TM & Ezirim HC. Discretized conductivity
measurements to study wetting and drying of cover zone concrete. Advances in
Cement Research 1998: 10(4): 195d202.

Prog. Struct. Engng Mater. 2000; 2:448}458

STEEL CORROSION
[30] Chatterji S. Transportation of ions through cement-based materials
d Part 1: Fundamental equations and basic measurement techniques. Cement and
Concrete Research 1994: 24(5): 907d912.
* [31] Truc O, Ollivier J-P & Nilsson L-O. Numerical simulation of multispecies transport through saturated concrete during a migration test d MsDiff code.
Cement and Concrete Research 2000: 30(10): 1581d1592.
An analysis of the transport of several species under the influence of an electric
field and the use of a migration test to determine the diffusion coefficient.
[32] Yu SW & Page CL. Computer simulation of ionic migration during
electrochemical chloride extraction from hardened concrete. British Corrosion
Journal 1996: 31(1): 73d75.
[33] Hassanein AM, Glass GK & Buenfeld NR. Chloride removal by
intermittent cathodic protection applied to reinforced concrete in the tidal zone.
Corrosion 1999: 55(9): 840d850.
** [34] Buenfeld NR, Shurafa-Daoudi MT & McLoughlin IM. Chloride
transport due to wick action in concrete. In: Nilsson L-O & Ollivier JP (eds) Chloride
Penetration into Concrete, Proceedings of the International RILEM Workshop. Paris,
France: RILEM Publications. 1997. 302d324.
An experimental and theoretical investigation of the effects of wick action,
diffusion and chloride binding on the chloride profile.
[35] Puyate YT & Lawrence CJ. Steady state solutions for chloride
distribution due to wick action in concrete. Chemical Engineering Science 2000:
55(16): 3329d3334.
** [36] Engelund S & Sorensen JD. A probabilistic model for chloride-ingress
and initiation of corrosion in reinforced concrete structures. Structural Safety 1998:
20(1): 69d89.
The development of a probabilistic model in which the statistical variation in
chloride content at the steel and its impact on repair and maintenance strategies
is assessed.
* [37] Faber MH. Reliability based assessment of existing structures. Progress in
Structural Engineering and Materials 2000: 2(2): 247}253.
An extension of the probabilistic model to include changes in load to predict
structural consequences.
[38] Faber MH, Berge HE, Tiller I & Hall ME. Reliability based assessment
of offshore concrete structures. Proceedings of the 18th OMAE Conference on
Offshore Mechanics and Arctic Engineering. St Johns, Canada, 12d15 July, 1999.
[39] Enright MP & Frangopol DM. Probabilistic analysis of resistance
degradation of reinforced concrete bridge beams under corrosion. Engineering
Structures 1998: 20(11): 960d971.
[40] Vu KAT & Stewart MG. Structural reliability of concrete bridges
including improved chloride-induced corrosion models. Structural Safety 2000: 22(4):
313d333.
** [41] Buenfeld NR & Hassanein NM. Predicting the life of concrete
structures using neural networks. Proceedings of the Institution of Civil Engineers
d Structures and Buildings 1998: 128(1): 38d48.
An overview, with examples of chloride-induced corrosion, of the use of neural
networks in the analysis of empirical data to predict service life.
** [42] Glass GK, Reddy B & Buenfeld NR. The participation of bound
chloride in passive film breakdown on steel in concrete. Corrosion Science 2000:
42(11): 2013d2021.
An extension of the basic mechanism of corrosion initiation which includes
important effects of many of the solid phases in cement.
[43] Glass GK, Reddy B & Buenfeld NR. Corrosion inhibition in concrete
arising from its acid neutralisation capacity. Corrosion Science 2000: 42(9):
1587d1598.
[44] Page CL. Mechanism of corrosion protection in reinforced concrete
marine structures. Nature 1975: 258(5535): 514d515.
[45] Glass GK, Yang R, Dickhaus T & Buenfeld NR. Backscattered
electron imaging of the steel-concrete interface. Corrosion Science, 2001: 43(4):
605d610.
[46] Glass GK & Buenfeld NR. Differential acid neutralisation analysis.
Cement and Concrete Research 1999: 29(10): 1681d1684.
* [47] Glass GK & Buenfeld NR. The presentation of the chloride threshold
level for corrosion of steel in concrete. Corrosion Science 1997: 39(5): 1001d1013.
A review of chloride threshold level data focusing on the indices used to present
this parameter.

Copyright ^ 2000 John Wiley & Sons, Ltd.

457

[48] Alonso C, Andrade C, Castellote M & Castro P. Chloride threshold


values to depassivate reinforcing bars embedded in a standardized OPC mortar.
Cement and Concrete Research 2000: 30(7): 1047d1055.
[49] Kayyali OA & Haque MN. The Cl\/OH\ ratio in chloridecontaminated concrete d A most important criterion. Magazine of Concrete Research
1995: 47(172): 235d242.
[50] Breit W & Schiessl P. Concrete Durability. ACI SP 170. Detroit:
American Concrete Institute. 1998. 363.
[51] Yonezawa T, Ashworth V & Procter RPM. Pore solution
composition and chloride effects on the corrosion of steel in concrete. Corrosion
1988: 44(7): 489d499.
[52] Sergi G & Glass GK. A method of ranking the aggressive nature of
chloride contaminated concrete. Corrosion Science 2000: 42(12): 2043d2049.
[53] BRE Centre for Concrete Construction. Corrosion of Steel in
Concrete, Part 2, Investigation and assessment. BRE Digest 444. Watford, UK:
Building Research Establishment. February, 2000.
[54] Glass GK & Buenfeld NR. Chloride threshold levels for corrosion
induced deterioration of steel in concrete. In: Nilsson L-O & Ollivier JP (eds) Chloride
Penetration into Concrete, Proceedings of the International RILEM Workshop. Paris,
France: RILEM Publications. 1997. 429d440.
[55] Broomfield JP. Corrosion of steel in concrete, understanding,
investigation and repair. London, UK: E & FN Spon. 1997.
* [56] Glass GK, Page CL & Short NR. Factors affecting steel corrosion in
carbonated mortars. Corrosion Science 1991: 32(12): 1283d1294.
An examination of the processes controlling the corrosion rate of steel in
concrete and their influence on corrosion rate determination.
* [57] Gowers KR & Millard SG. Electrochemical techniques for corrosion
assessment of reinforced concrete structures. Proceedings of the Institution of Civil
Engineers d Structures and Buildings 1999: 134(2): 129d137.
A summary of the various methods of monitoring the corrosion rate of steel in
concrete.
[58] Schiessl P & Raupach M. Laboratory studies and calculations on the
influence of crack width on chloride-induced corrosion of steel in concrete.
ACI Materials Journal 1997: 94(1): 56d62.
[59] Concrete Bridge Development Group. Testing and monitoring the
durability of concrete structures. Technical Guide No 2. Crowthorne, UK: Transport
Research Laboratory. 2000.
[60] Gowers KR & Millard SG. Measurement of concrete resistivity for
assessment of corrosion severity of steel using Wenner technique. ACI Materials
Journal 1999: 96(5): 536d541.
[61] Shigeishi M, Colombo S, Broughton KJ, Rutledge H, Batchelor AJ
& Forde MC. Acoustic emission to assess and monitor the integrity of bridges.
Construction and Building Materials 2001: 15(1): 35d49.
[62] John G, Gaydecki P, Fernandes B & Silva I. Development of inductive
imaging of corrosion damaged reinforced and prestressed concrete structures.
Corrosion 2000. Paper 293. Houston, Texas: NACE. March 2000.
[63] Millard SG, Shaw MR, Giannopoulos A & Soutsos MN. Modeling of
subsurface pulsed radar for nondestructive testing of structures. Journal of Materials
in Civil Engineering (ASCE) 1998: 10(3): 188d196.
[64] Borchelt R, Edwards M, Ghorbanpoor A & Salam EA. Magneticbased NDE of prestressed and post-tensioned members d The MFL system. Federal
Highway Administration Report, FHWA-RD-00-026, Washington DC, May 2000.
[65] ComiteH Euro-International du BeH ton. Strategies for testing and
assessment of concrete structures. Guidance report. Bulletin DInformation No. 243.
Lausanne, Switzerland: CEB. 1992.
* [66] McCarter WJ, Chrisp TM & Basheer PAM. Sensors for condition
profiling of cover-zone concrete. In: Basheer PAM & Sloan TD (eds) COST521:
Corrosion of Steel in reinforced concrete structures. Annual Progress Reports.
Belfast, UK: Queens University. August 2000.
A comparison of site and laboratory data where resistivity is the main technique
used to assess the condition of cover-zone concrete.
** [67] McCarter WJ & Chrisp M. Monitoring water and ionic penetration into
cover-zone concrete. ACI Materials Journal 2000: 97(6): 668d674.
A comparison of embedded and surface probes that may be used to obtain
resistivity and other data to separate the effects of moisture and temperature
variations from continued hydration and ionic ingress.
[68] Raupach M & Schiessl P. Monitoring system for the penetration of
chlorides, carbonation and the corrosion risk for the reinforcement. Construction
and Building Materials 1997: 11(4): 207d214.

Prog. Struct. Engng Mater. 2000; 2:448}458

458
[69] Atkins CP, Scantlebury JD, Nedwell PJ & Blatch SP. Monitoring
chloride concentrations in hardened cement pastes using ion selective electrodes.
Cement and Concrete Research 1996: 26(2): 319d324.
[70] Law DW, Millard SG & Bungey JH. Galvanostatic pulse measurements
of passive and active reinforcing steel in concrete. Corrosion 2000: 56(1):
48d56.
* [71] Glass GK, Hassanein AM & Buenfeld NR. Obtaining impedance
information on the steel-concrete interface. Corrosion 1998: 54(11): 887d897.
A recent advance in obtaining corrosion rate information on steel in concrete.
* [72] Page CL & Sergi G. Developments in cathodic protection applied to
reinforced concrete. Journal of Materials in Civil Engineering (ASCE) 2000: 12(1):
8d15.
An overview of cathodic protection highlighting the impact of local protection on
the life of patch repairs in chloride-contaminated concrete.
[73] BRE Centre for Concrete Construction. Corrosion of Steel in
Concrete, Part 3, Protection and remediation. BRE Digest 444. Watford, UK: Building
Research Establishment, February 2000.
** [74] Glass GK, Hassanein AM & Buenfeld NR. Cathodic protection
afforded by an intermittent current applied to reinforced concrete. Corrosion
Science 2001: 43(6): 1111d1131.

CONCRETE CONSTRUCTION
The presentation of recent advances in the mechanism of cathodic protection and
their impact on designing an intermittent cathodic protection system.
[75] Glass GK & Buenfeld NR. The inhibitive effects of electrochemical
treatment applied to steel in concrete. Corrosion Science 2000: 42(6): 923d927.
* [76] Glass GK, Hassanein AM & Buenfeld NR. Cathodic protection
criteria for reinforced concrete in marine exposure zones. Journal of Materials in Civil
Engineering (ASCE) 2000: 12(2): 164d171.
A presentation of recent advances in understanding the protection criteria for
steel in concrete.
[77] Cramer SD, Covino BS, Holcomb GR, Bullard SJ, Collins WK,
Govier RD, Wilson RD & Laylor HM. Thermal sprayed titanium anode for
cathodic protection of reinforced concrete bridges. Journal of Thermal Spray
Techniques 1999: 8(1): 133d145.
[78] Covino BS, Bullard SJ, Holcomb GR, Russell JH, Cramer SD,
Bennett JE & Laylor HM. Chemically modified thermal-spray zinc anodes for
galvanic cathodic protection. Materials Performance 1999: 38(12): 28d32.
* [79] Page CL, Ngala VT & Page MM. Corrosion inhibitors in concrete
repair systems. Magazine of Concrete Research 2000: 52(1): 25d37.
An independent examination of the effectiveness of corrosion inhibitors as
a repair technique for corrosion-damaged concrete.

G K Glass
Department of Civil and Environmental Engineering,
Imperial College, London, SW7 2BU, UK
N R Buenfeld
Department of Civil and Environmental Engineering,
Imperial College, London, SW7 2BU, UK

Copyright ^ 2000 John Wiley & Sons, Ltd.

Prog. Struct. Engng Mater. 2000; 2:448}458

Você também pode gostar