Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

Construction and Building Materials 198 (2019) 172–181

Contents lists available at ScienceDirect

Construction and Building Materials


journal homepage: www.elsevier.com/locate/conbuildmat

Experimental investigations on the influence of cover depth and


concrete quality on time to cover cracking due to carbonation-induced
corrosion of steel in RC structures in an urban, inland environment
Mike Otieno ⇑, Jacob Ikotun, Yunus Ballim
School of Civil and Environmental Engineering, University of the Witwatersrand, Johannesburg, Private Bag 3, WITS, 2050, Braamfontein, Johannesburg, South Africa

h i g h l i g h t s

 Time to carbonation-induced steel corrosion cover cracking is investigated.


 Steel mass loss to initiate surface cracking increases with increase in cover depth.
 Blended cement concretes required higher steel mass loss to cause cover cracking.
 Higher w/b ratio concretes required longer times to cover cracking.
 An empirical relationship for time to cover cracking is proposed.

a r t i c l e i n f o a b s t r a c t

Article history: This paper reports on a study that was aimed at developing a coherent empirical relationship for the time
Received 30 July 2018 to cracking of the cover concrete in a reinforced concrete structure, subjected to reinforcing steel corro-
Received in revised form 5 November 2018 sion that is induced by carbonation in an inland, urban environment in South Africa. In particular, the
Accepted 23 November 2018
study considered the influence of cover depth and the quality of concrete in the cover zone on the extent
Available online 3 December 2018
of corrosion at the time of cover concrete cracking. The results of the study were then used to develop an
empirical relationship to estimate the time to cover cracking and the associated extent of reinforcing
Keywords:
steel corrosion. Test concretes were prepared using three binder types: plain Portland cement (PC –
Carbonation
Corrosion
CEM I 52.5N), 70/30 PC/FA (fly ash) and 50/50 PC/GGBS (ground granulated blastfurnace slag) at relatively
Steel radial loss high w/b ratios of 0.60 and 0.95. 13.2 mm crushed granite stone and granite crusher sand were used as
Cover depth aggregates. RC prism specimens were prepared using 20 mm diameter high yield deformed steel bars
Porosity placed at cover depths of 12 mm, 20 mm or 30 mm. Sufficient 100 mm companion concrete cube speci-
mens were prepared for permeability, porosity tests and carbonation depth measurement. The results
indicate that, as the cover depth increases, a higher extent of corrosion is required to initiate cracking
at the concrete surface. Also, at the same cover depth, the higher w/b ratio concrete, which was more por-
ous, more permeable and had lower tensile strengths, required more steel corrosion to initiate surface
cracking. Blended cement concretes required a higher extent of steel corrosion to initiate cover cracking
than the PC concretes.
Ó 2018 Elsevier Ltd. All rights reserved.

1. Introduction and chloride ingress, of which chloride-induced corrosion is known


to be more pernicious than carbonation-induced corrosion. In both
Corrosion of steel in concrete has been identified as a major cases, the service life of corrosion-affected RC structures is charac-
cause of deterioration in reinforced concrete (RC) structures, and terized by two main phases – initiation and propagation as shown
much resources are spent on maintaining, repairing and rehabili- in Fig. 1. In the past, the service life of both carbonation- and
tating corrosion-damaged RC structures. The two main causes of chloride-induced corrosion-affected RC structures has been quan-
steel corrosion in concrete have been identified as carbonation tified only by the initiation phase, ti, [1–5]. However, as the num-
ber of deteriorated RC structures continues to increase rapidly,
and in an effort to meet both durability and sustainability require-
⇑ Corresponding author. ments of corrosion-affected RC structures, service life prediction
E-mail address: Mike.Otieno@wits.ac.za (M. Otieno).

https://doi.org/10.1016/j.conbuildmat.2018.11.215
0950-0618/Ó 2018 Elsevier Ltd. All rights reserved.
M. Otieno et al. / Construction and Building Materials 198 (2019) 172–181 173

Time to structural failure A majority of steel RC structures in urban inland environments


Level of corrosion-induced damage

in South Africa undergo premature deterioration due to


carbonation-induced corrosion of steel which leads to a variety
of inter-related serviceability effects including cover cracking,
delamination and spalling. Other more severe effects include loss
Time to concrete spalling of steel cross-section area which ultimately affects structural
integrity. However, for service life design, cover cracking is a ser-
viceability limit state [14] which can be used to indirectly infer
Time to cover cracking the ultimate limit state conditions through estimation of critical
steel radial loss which is directly related to the corrosion rate. This
study, which focused on carbonation-induced corrosion, was car-
ried out in an urban inland environment in Johannesburg which
Time is located in South Africa’s Gauteng Province. At an altitude
Initiation phase Propagation phase of 1700 m above sea level, Johannesburg is characterized by
wet summer (temperature: 15–26 °C, relative humidity: 70%)
(ti) (tp)
and dry winter (temperature: 4–20 °C, relative humidity: 50%)
Fig. 1. Service life of corrosion affected RC structures incorporating sub-phases
seasons [15].
[9–11].
1.1. Carbonation-induced corrosion of steel in concrete
and life-cycle management beyond the initiation phase is now
receiving more attention than in the past [6–8]. As a result, there In comparison to chloride-ingress in concrete, natural carbona-
have been trends towards extending the service life definition of tion takes place at a relatively slow rate mainly due to the low con-
RC structures prone to steel corrosion. In addition to the initiation centration of CO2 in the atmosphere (0.03 – 0.04%), but can be
phase, there have been attempts to quantitatively incorporate the faster in urban and industrial environments where the concentra-
corrosion propagation phase (See Fig. 2). tion of CO2 is relatively high (>0.04%) [16–18] and continues to
The impetus to include corrosion propagation in the design ser- rise. Nevertheless, it eventually leads to initiation of corrosion of
vice life of RC structures can be assigned to a combination of the the embedded steel and subsequently corrosion propagation,
following premises: which can result in corrosion-induced damage including delamina-
tion, cover cracking and spalling. It therefore deserves much atten-
(i) the rapid increase in the number of deteriorated RC struc- tion especially when considering inland RC structures. The
tures due to corrosion and the resulting need to develop chemical processes that sustain carbonation in concrete involve
pro-active maintenance and repair strategies for RC reactions between dissolved carbon dioxide (CO2) in the form of
structures; weak carbonic acid (H2CO3) and calcium hydroxide (portlandite,
(ii) at the end of the initiation phase, the RC structure may still Ca(OH)2) in the pore solution (or hydrated cement paste) resulting
have adequate residual serviceability and strength with or in the precipitation of mostly calcium carbonate (calcite, CaCO3) in
without remedial measures [12]; the concrete pores as shown in Eqs. (1) and (2) [19].
(iii) the propagation phase may in some circumstances be suffi-
ciently long to merit quantification and hence be part of the CO2 þ H2 O ! H2 CO3 ð1Þ
structure’s service life, t, e.g. in the presence of service load-
H2 CO3 þ H2 O ! CaCO3 þ H2 O ð2Þ
induced concrete cracking where active corrosion initiation
may be almost instantaneous in the presence of corrosion- Other calcium-bearing phases (e.g. C–S–H, C2S and C3S) are also
inducing species i.e. ti = 0 or t = tp [13]; susceptible to carbonation, and possible products include, among
(iv) for new RC structures exposed to corrosion-aggressive envi- others, aragonite and vaterite [20–23]. Carbonation affects both
ronments, the propagation phase may extend for a relatively the concrete microstructure and the durability of steel RC struc-
long period of time after corrosion initiation depending on tures. The following is a summary of some of the main conse-
the performance limit state indicator adopted (e.g. quences of carbonation of concrete:
corrosion-induced cover cracking) and the effectiveness of
the maintenance and repair actions undertaken. (i) Reduction in pH: Carbonation depletes the hydroxyl ions in
(v) for existing RC structures, there is always a need to predict the alkaline concrete pore solution and therefore leads to a
the propagation state of the structure (during the corrosion reduction in the pH (loss of alkalinity) of carbonated con-
process), for maintenance, repair and/or protection evalua- crete from 12.5 to as low as 8.3 [24]. This is detrimental
tion purposes. considering that the drop in pH leads to a destruction of

Epoxy coating Concrete prism 50 mm 50 mm


Reainforcing steel

12, 20 or 30 mm
100 mm

12, 20 or 30 mm
200 mm 100 mm

Fig. 2. Longitudinal and cross-section of a test RC prism.


174 M. Otieno et al. / Construction and Building Materials 198 (2019) 172–181

the thin protective film on the steel surface and therefore an empirical model to predict the service life of carbonation-
significantly increases the risk of steel corrosion initiation induced corrosion-affected RC structures in the South African
[25]. inland environment.
(ii) Changes in porosity: Carbonation leads to changes in poros-
ity of the concrete. The change can be a decrease (in the case
2. Experimental set-up
of Portland cement (PC) paste), or an increase (in the case of
blended cement paste [20]. In PC concretes, the formation of
A summary of the concrete mix proportions and selected con-
the highly insoluble products (calcite, aragonite, varetite)
crete properties is given in Table 1. Two binder types 70/30 PC/
with higher volumes (with varetite > aragonite > cal-
FA (plain Portland cement CEM I 52.5 N/fly ash) and 50/50 PC/
cite > portlandite) than original reactants have been
GGBS (ground granulated blastfurnace slag) and two w/b ratios
reported. These products precipitate [26] and block the con-
(0.60 and 0.95) were used. The high w/b ratio of 0.95 was used
crete pore network resulting in a reduction in the total
mainly to simulate extremely poor on-site practices including poor
porosity and permeability of the carbonated concrete [27].
compaction and inadequate concrete curing. Compared to the 0.60
From this viewpoint, carbonation has a beneficial effect with
w/b ratio concretes, the 0.95 w/b ratio concretes were expected to
respect to penetrability of especially PC concretes with high
have relatively more open pore structures which would facilitate
portlandite contents. However, coarsening of the pore struc-
higher steel corrosion rates in the concretes. A w/b ratio lower than
ture may be associated with the formation of additional sil-
0.60 was not be used in this study due to time constraints related
ica gel due to decomposition of C-S-H [28]. Considering that
to the very slow nature of natural carbonation-induced steel corro-
blended cement concretes have significantly lower quanti-
sion and corrosion-induced damage e.g. cover cracking [53,54].
ties of portlandite, carbonation of these concretes results
13.2 mm crushed granite stone and granite crusher sand were used
in a coarser pore structure than that of PC concretes.
as aggregates. 20 mm diameter high yield deformed steel bars
(iii) Microcracking: Carbonation may also lead to possible occur-
(260 mm in length) of average ultimate tensile strength
rence of microcracks (extensive in some cases e.g., [29])
(450 MPa) were embedded in the RC prism specimens. The steel
which makes it difficult to predict the change in transport
bars were cleaned by wire-brushing to remove mill scale before
properties of species through the carbonated concrete
taking initial mass measurements of each of the bars.
[27,29]. There have been reports of increased penetrability
Concrete was mixed in a horizontal drum mixer. For each con-
of carbonated concrete with respect to water absorption,
crete mix, two RC prism specimens (100  100  200 mm) were
oxygen diffusion and chloride diffusion [28,30]. The cracking
cast with two identical reinforcing steel bars (20 mm) embedded
is suggested to be caused by differential decalcification
in the concrete at the same top and bottom cover depths
shrinkage due to a gradient of Ca/Si ratios in the concrete.
(12 mm, 20 mm or 30 mm) as shown in Fig. 1. Five companion
It has also been reported that due to their low portlandite
unreinforced concrete prisms were also cast for each mix; two of
content, blended cement concretes are expected to be more
these were used to periodically monitor the depth of carbonation
susceptible to carbonation-induced cracking but this may to
front in the corresponding RC specimens while the remaining three
some extent be offset by their low penetrability prior to car-
were used for flexural strength tests.
bonation [31].
For each concrete mix, two 100 mm companion concrete cube
specimens were cast for permeability and porosity tests. Immedi-
In South Africa, models have been developed to predict the pro-
ately after casting, the specimens were covered with polyethene
gress of carbonation front in concrete in order to predict the time
sheets for about 24 h. After this duration, the specimens to be used
to steel corrosion initiation [32,33]. However, the prediction of
for the permeability and tensile strength tests were cured in water
the propagation phase is still lacking, and this paper is a step
(23 ± 2 °C) for 28 days while the specimens to be used for carbon-
towards filling this gap. The approach taken in this paper is to pre-
ation and corrosion monitoring were cured in water for seven days,
dict a serviceability limit state in the propagation phase of steel
and air-dried in the laboratory for 30 days before exposing them to
corrosion. In this case, cover cracking has been adopted as a limit
accelerated carbonation in a chamber (details provided later).
state in the propagation phase of steel corrosion [34]. At the time
of cover cracking, the RC structure is deemed to be still structurally
sound and any maintenance and/or repair actions can be imple- 2.1. Oxygen permeability test
mented optimally at relatively low costs [35–38].
Oxygen permeability test [55] was carried out after 28 days cur-
1.2. Corrosion-induced cover cracking ing of the specimens. Two concrete cubes from each concrete mix
were cored to obtain four concrete disc specimens (two from a
A number of studies [34,39–42] have focused on predicting the 100 mm cube) of 70 ± 2 mm diameter and 30 ± 2 mm thickness.
time to steel corrosion-induced cover cracking, and a number of The discs were oven-dried at 50 ± 2 °C for 7 days ± 4 hrs. The disc
models have been proposed. The existing cover cracking models specimens were then cooled to room temperature (23 ± 2 °C) in a
can be categorized into two – empirical and analytical. Empirical desiccator for 3 h before commencement of the test. The oxygen
models are primarily based on regression analysis of experimental permeability index (OPI) is defined as –Log k where k (m/s) is
data and observations e.g., time-to-cracking. Examples include the coefficient of oxygen permeability determined, following
models by Alonso et al. [43], Rodriguez et al. [44], Morinaga [45] Darcy’s equation, from the rate of pressure decay through a sample
and Torres-Acosta and Sagues [46]. Analytical models are primarily when placed in a falling head permeameter (initial gas (oxygen)
based on the concepts of solid mechanics [47]. Examples include pressure of 100 ± 5 kPa); the rate of pressure decay can be
models by Bazant [48], Liu [49] and El Maadawy and Soudki [50]. expressed using the Darcy equation, allowing the coefficient of
Some of the main parameters used in these models, and which permeability (k, m/s) to be determined [56,57]. The coefficient of
have been identified to affect time to cover cracking, include steel permeability k is defined as (xVgd)  (ln(Po/P))  (RAht)1 where
corrosion rate (or mass loss or steel radial loss or amount of corro- x is the molecular mass of the permeating gas (kg/mol), V is the
sion products), cover depth to bar diameter ratio, bar spacing, and volume of the pressure cylinder (m3), g is the acceleration due to
material properties of concrete (concrete quality) [43,51,52]. The gravity (m/s2), d is the sample thickness (m), R is the universal
study reported in this paper focused on cover cracking to develop gas constant (Nm/Kmol), A is the cross sectional area of specimen
M. Otieno et al. / Construction and Building Materials 198 (2019) 172–181 175

Table 1
Summary of concrete mix proportions (kg/m3) and selected concrete properties.

Binder composition 100% PC 70/30 PC-FA 50/50 PC-GGBS


w/b ratio 0.95 0.60 0.95 0.60 0.95 0.60
Mix label PC-95 PC-60 FA-95 FA-60 SL-95 SL-60
PC (CEM I 42.5N) 255 350 179 245 128 175
Fly ash – – 77 105 – –
Ground granulated blastfurnace slag – – – – 128 175
Granite crusher sand 910 849 910 849 910 849
13.2 mm granite stone 1002 1015 1002 1015 1002 1015
Water 242 210 242 210 242 210
28-day compressive strength (MPa) 22.6 48.7 23.1 51.9 14.4 33.4
28-day flexural strength (MPa) 2.1 4.9 2.0 4.5 0.6 3.5

(m2), h = absolute temperature (K) and Po, P are the pressures at continued; the duration of the carbonation-induced corrosion
start of test and at time t (sec) respectively. The higher the index propagation experiments varied from specimen to specimen
value, the less permeable is the concrete. depending on the time to cover cracking. A summary of the time
to surface cracking of the concrete in the various specimens is pre-
sented later in Fig. 11. The actual surface crack widths, even though
2.2. Determination of concrete porosity
important for serviceability assessments [58,59], were not mea-
sured in this study due to lack of sensitive equipment. The cor-
The same concrete disc specimens used for the oxygen perme-
roded steel bars were then extracted from the cracked specimens
ability test were used to determine concrete porosity. Porosity
for determination of gravimetric mass loss.
tests were carried out immediately after the oxygen permeability
tests. The concrete disc specimens were weighed to determine
the initial oven-dry mass (Mso) and then vacuum-saturated 2.4. Gravimetric steel mass loss
(between –75 and –80 kPa) with saturated calcium hydroxide
solution. The negative pressure was maintained for 3 h ± 15 min, After extraction from the specimens, the corroded steel bars
after which saturated Ca(OH)2 was allowed to flow into the tank were cleaned of corrosion and loosely attached cement paste using
up to a head of 40 mm above the specimens. The negative pressure a wire brush. The bars were then dipped in a chemical solution pre-
was then restored and maintained for 1 h ± 15 min, after which the pared of 500 ml of hydrochloric acid and 3.5 g of hexamethylenete-
specimens were allowed to soak in the saturated Ca(OH)2 solution tramine in 500 ml of reagent water ASTM G1-03 [60]. This process
for a further 18 ± 1 h. The saturated surface dry masses (Msv) of the was repeated until the steel bars were visually observed to be free
vacuum-saturated specimens were then determined and the con- of both corrosion products and any hardened cement paste. The
crete specimen porosity (g, %) determined as 100  (Msv – Mso)/ mass of the reinforcing bars were then measured to determine
(Adqw) where Msv (g) is the vacuum-saturated mass of the speci- the final mass. The difference between the initial (before corrosion)
men, Mso (g) is the initial mass of the specimen, A (mm2) is the and final (after corrosion) mass of the reinforcing steel bars was
cross-sectional area of the specimen, while ‘d’ (mm) and qw evaluated to establish the mass loss, ml (grams). The average of
(103 g/mm3) are the specimen thickness and density of water four steel mass loss readings was calculated for two specimens of
respectively. For a particular concrete mix, four concrete disc spec- the same concrete mix.
imens were tested for porosity.
3. Results and discussion
2.3. Corrosion initiation and propagation
3.1. Oxygen permeability and porosity tests results
The corrosion specimens were pre-carbonated in a carbonation
chamber with 10% CO2, relative humidity of 55 ± 5%, and tempera- The oxygen permeability and porosity results are presented in
ture of 20 ± 2 °C. A 10% CO2 concentration was used in order to Fig. 4. It is clear that permeability and porosity increase with
accelerate the carbonation of concrete which is a very slow process increasing w/b ratio. This trend can be attributed to the capillary
in nature due to the low concentration of carbon dioxide in the porosity, volume and interconnectivity of the pores which increase
atmosphere [18]. The relative humidity in the chamber was estab- along with an increase in the w/b ratio [61,62]. Garboczi [62]
lished and maintained using saturated magnesium nitrate salt attributed the increased capillary porosity in higher w/b ratio con-
solution [14]. Carbonation testing was carried out weekly by spray- cretes to their lower binder content which results in inefficient
ing phenolphthalein solution on a freshly cut surface of the carbon- packing of the cement particles and the aggregates; it can also be
ation concrete specimens, and the carbonation front measured attributed to the less space-filling and less voluminous hydration
using a Vernier calipers to 0.05 mm accuracy. When the measured products [19,63]. For a given binder type, a clear trend in both
carbonation front was greater than or equal to the cover depth of the permeability and porosity results can be observed for concretes
the steel reinforcement, the RC specimens (shown in Fig. 1) were of different w/b ratios (see Fig. 4). The results also show that for
removed from the carbonation chamber and placed in an unshel- either porosity or permeability, and for a given w/b ratio, the
tered natural urban inland environment in Johannesburg – see blended cement concretes are less porous than, but show relatively
Fig. 3(a). comparable permeability values with, the PC concretes. This can be
While in the exposure site, the RC specimens were monitored attributed to the effect of the microstructure densification in the
daily throughout the experimental period until the first visible lon- blended cement concretes [64]. Similar results have been reported
gitudinal surface crack (parallel to the line of reinforcement) in the past [65–67]. The OPI (or k) results can also be used to deter-
appeared on the concrete cover (Fig. 3(b)) over any one of the steel mine the diffusion coefficients of the various concretes which is
bars. The time to corrosion-induced surface cracking of the con- useful in estimating the time to carbonation-induced corrosion ini-
crete for each specimen was noted visually and the monitoring dis- tiation [68].
176 M. Otieno et al. / Construction and Building Materials 198 (2019) 172–181

Fig. 3. RC prisms exposed to unsheltered inland environment (a), and cracked prisms (b).

10.5
Error bars: 95% confidence interval 16
OPI (–Log k, m/s)

10.0 14

Porosity (%)
9.5 12

9.0 10

8.5 8
0.60 0.95 0.60 0.95 0.60 0.95 0.60 0.95 0.60 0.95 0.60 0.95
100% 70/30 50/50 100% 70/30 50/50
PC PC/FA PC/GGBS PC PC/FA PC/GGBS

(a) (b)

Fig. 4. Oxygen permeability index (OPI) and porosity results.

3.2. Steel mass loss required for cover cracking to the increase in the volume of pores in the concrete cover
which act as reservoirs for the steel corrosion products, and
Fig. 5 presents the results for average steel mass loss at surface- partly to the greater depth of the material to resist tensile
cracking for the different concretes (binder type and w/b ratio) and stresses due to corrosion. Cracking will only occur when
cover depths. The results show that: the porous zone between the steel bar and the surrounding
concrete has been filled up with the steel corrosion products
(i) increasing the cover depth (from 12 to 20 to 30 mm) led to and/or sufficient tensile stresses build up that exceed the
an increase in steel mass loss required for surface cracking tensile strength of the concrete [14,34,69–71].
of the concrete (see Fig. 6). This trend is attributed partly

6 Note: "12","20", "30": Cover depth to steel (mm)


(1.59)

*: Theoretical uniform corrosion rate ( µA/cm2)


Gravimetric steel mass loss (g)

(1.28)

5
(2.03)
(2.33)
(0.95)

(0.96)

4
(1.45)
(1.45)
(0.93 µA/cm2)*

(1.29)
(0.79)

(1.74)
(1.40)

(1.50)
(0.94)

3
(0.60)

(1.11)
(0.68)

0
12 20 30 12 20 30 12 20 30 12 20 30 12 20 30 12 20 30
0.60 w/b 0.95 w/b 0.60 w/b 0.95 w/b 0.60 w/b 0.95 w/b
100% PC 70/30 PC/FA 50/50 PC/GGBS
Fig. 5. Average steel mass loss (and corresponding uniform corrosion rates) at cover cracking.
M. Otieno et al. / Construction and Building Materials 198 (2019) 172–181 177

6.0 6.0

Mass loss (g) for 30 mm cover depth


Mass loss (g) for 20 mm cover depth
5.0 5.0

4.0 4.0

3.0 3.0

PC-60 PC-95 PC-60 PC-95


2.0 2.0 FA-60 FA-95
FA-60 FA-95
SL-60 SL-95 SL-60 SL-95
1.0 1.0
1.0 2.0 3.0 4.0 5.0 6.0 1.0 2.0 3.0 4.0 5.0 6.0

Mass loss (g) for 12 mm cover depth Mass loss (g) for 20 mm cover depth
(a) (b)
Fig. 6. Effect of cover depth on steel mass loss at cracking.

(ii) lower quality (0.95 w/b ratio) concretes, which were more tures, the determination of steel mass loss is based on indirect non-
porous, more permeable (see Fig. 4) and had lower tensile destructive methods such as the measurement of steel corrosion
strengths (see Table 1), require more steel mass loss for sur- rate. Ultimately, the estimation of the loss in steel cross-section
face cracking than relatively higher quality (0.60 w/b ratio) due to corrosion is vital. The steel mass loss results presented ear-
concretes (see Fig. 7). The higher porosity and permeability lier can be converted to steel radial loss (dr) using the equation
in these concretes explains the higher steel mass loss dr = mloss/2prq where q (g/cm3) is the density of the reinforcing
required for cracking [43,69]. It can also be argued that in steel bar and mloss (g) is the measured steel mass loss. The corro-
these concretes, the rate of production of steel corrosion sion in the steel bars obtained from the prisms after cracking
products (i.e. the corrosion rate) governs the time-to- was observed to be uniform along the length of the steel (see
cracking more that the amount of corrosion products; the Fig. 8). The calculated radial loss for the steel bars is presented in
rate of production of steel corrosion products must exceed Fig. 9 which, as expected, follows the same trend as that presented
that of diffusion of the products into the concrete pores for earlier for mass loss (Fig. 5).
sufficient tensile stresses to build up in the concrete to cause
cracking.
(iii) blended cement concretes require higher steel mass losses 3.3. Time to steel corrosion-induced cover cracking
to cause cover cracking than PC concretes (see Figs. 5 and
7). Results presented earlier show that these concretes are From the results presented and the ongoing discussion, it is
less porous and permeable, but were pre-carbonated, a pro- apparent that the time to steel corrosion-induced cover cracking
cess known to increase permeability due to internal micro- is a function of steel mass loss (corrosion rate), cover depth and
cracking [72–75]. This facilitates the migration or diffusion concrete properties (porosity and permeability) related to its qual-
of the corrosion products away from the corrosion sites ity; this is summarized in Figs. 10 and 11. These results follow the
which prolongs the time to cracking [43,69]. trends presented in earlier sections; it is clear that the time to
cracking increases with increase in cover depth and decreases with
Even though gravimetric steel mass loss with time is a good increase in w/b ratio. For a given w/b ratio and cover depth, the
experimental measure of actual corrosion rate, in existing RC struc- time to cracking increases in the order PC > PC/FA > PC/GGBS; the
blended cement concretes exhibit shorter times to cover cracking
compared to the PC concretes. This is despite the higher mass
6.0 losses (see previous section) in these concretes. This clearly indi-
cates that corrosion rates in the carbonated blended cement con-
cretes were higher than those in the carbonated plain PC
5.0
Mass loss (g) for 0.95 w/b

Slag concretes

4.0
PC (12 mm cover)
PC (20 mm cover)
3.0 PCconcretes PC (30 mm cover)
PC/FA (12 mm cover)
FAconcretes PC/FA (20 mm cover)
PC/FA (30 mm cover)
2.0
PC/GGBS (12 mm cover)
PC/GGBS (20 mm cover)
PC/GGBS (30 mm cover)
1.0
1.0 2.0 3.0 4.0 5.0 6.0

Mass loss (g) for 0.60 w/b


Fig. 8. Photograph of some of the specimens split open to retrieve the corroded
Fig. 7. Effect of w/b ratio and binder type on steel mass loss at cracking. steel bars.
178 M. Otieno et al. / Construction and Building Materials 198 (2019) 172–181

60
Note: "12","20", "30": Cover depth to steel

Steel radial loss, mm (×10 –3)


50

40

30

20

10

0
12 20 30 12 20 30 12 20 30 12 20 30 12 20 30 12 20 30
0.60 w/b 0.95 w/b 0.60 w/b 0.95 w/b 0.60 w/b 0.95 w/b
100% PC 70/30 PC/FA 50/50 PC/GGBS
Fig. 9. Average steel radius loss required to crack the cover concrete.

3.0
Note: "12","20", "30": Cover depth to steel (mm)
Time to cracking (years)

2.5

2.0

1.5

1.0

0.5

0.0
12 20 30 12 20 30 12 20 30 12 20 30 12 20 30 12 20 30
0.60 w/b 0.95 w/b 0.60 w/b 0.95 w/b 0.60 w/b 0.95 w/b
100% PC 70/30 PC/FA 50/50 PC/GGBS
Fig. 10. Average time to steel corrosion-induced cover cracking.

concretes; this is supported by literature [76]. This can be attribu- cretes used in the experimental work of this study. However, the
ted to the carbonation-induced microcracking of the concrete objective was to obtain an empirical relationship that is suitable
which allows easy access of corrosion-sustaining agents (O2, and for South Africa’s urban inland environment using commonly-
H2O) to the steel level to support the corrosion propagation. At used concretes ranging from good (0.60 w/b ratio) to very poor
high corrosion rates, the porous zone between the steel and con- (0.95 w/b ratio); this objective was met. Other parameters such
crete, and the internal microcracks due to carbonation are filled as temperature and relative humidity could not be incorporated
faster with corrosion products leading to shorter times for build- into the empirical relationship but will need to be considered in
up of tensile stresses that cause cracking. future studies to improve on its accuracy. It is also important to
Based on a regression analysis of the experimental data, the acknowledge that parts of the experimental work involved acceler-
time to cracking (tcr, years) and steel radial loss (dr,  103 mm) ated carbonation which may not be fully representative of natural
at cracking can be estimated as: carbonation. The empirical relationships will therefore need to be
validated using data from natural carbonation in in-service RC
t cr ¼ 0:12c0:72 e0:14f t ð3Þ
structures.

dr ¼ 1893q1:92 c0:232 ð4Þ


where c is the cover depth (mm), ft is the tensile strength (MPa) of 4. Conclusions
the concrete and q is the porosity of the concrete (%). These empir-
ical relationships will need to be validated in future using indepen- The objective of this study was to investigate the influence of
dent data from in-service RC structures to ascertain its accuracy. cover depth and concrete quality on steel carbonation-induced
The empirical relationship for time to cover cracking has obvi- cover cracking in RC structures in urban inland environments,
ous limitations in respect of its applicability which need to be specifically in Johannesburg, South Africa. Experiments were car-
acknowledged. It is limited in its scope with respect to the con- ried out and the results revealed the following findings:
M. Otieno et al. / Construction and Building Materials 198 (2019) 172–181 179

PC-60
2.50 PC-95
FA-60 SL-60
FA-95
2.25 R² = 0.98 R² = 0.92 SL-95
R² = 1.00 30 mm cover

2.00 R² = 0.93
Time to cracking (yrs)

R² = 0.74
R² = 0.99
R² = 1.00
1.75

1.50
R² = 0.88

20 mm cover
1.25

R² = 0.86
1.00 12 mm cover

0.75
1.5 2.0 2.5 3.0 3.5 4.0 4.5 5.0 5.5
Steel mass loss at cracking (g)

Fig. 11. Correlation grid for time-to-cracking, steel mass loss, cover depth and concrete quality.

(i) lower quality (higher w/b ratio) concretes required longer South Africa, Bradlow Foundation as well as Postgraduate Merit
times to cover cracking than higher quality (lower w/b ratio) Award of the University of the Witwatersrand, Johannesburg.
concretes. This should not be misinterpreted to mean that Gratitude is also extended to PPC Pty Ltd, AfriSam and Sika (SA)
higher w/b ratio concretes are better in prolonging the ser- Pty Ltd. for their support.
vice life of RC structures than lower w/b ratio concretes. In
fact, the opposite is the case as has been shown in previous
5. Compliance with ethical standards and conflict of interest
durability-related studies. The longer time to cover cracking
in the higher w/b ratio concretes increases the risk of struc-
The study presented in this paper has no ethics-related issues,
tural failure of RC structures due to higher steel radial loss.
and has no conflict of interest issues to disclose.
This is because it can be expected that post-cracking, the
corrosion rate in the lower quality concretes will be higher
than that in the higher quality concretes due to the dominat- References
ing effect of concrete quality on corrosion rate even in [1] V. Sirivivantnanon, Effect of cracking on service life of concrete, in: Proceedings
cracked concrete [7,77]. of the international conference by the Concrete Institute of Australia. 11th-
(ii) following the finding in (i), it was also found that the 14th September, Perth, Western Australia, 2001. pp. 273-279.
[2] LIFE-365,, Computer program for predicting the service life and life-cycle costs
blended cement concretes required higher steel mass losses
of reinforced concrete exposed to chlorides, ACI Committee 365 (2005).
to cause cover cracking than the plain PC concretes. This [3] DuraCrete, Probabilistic performance-based durability design: modelling of
finding means that even though blended cement concretes degradation, 1998. Document No. BE95-1347/R4-5, The Netherlands.
are less penetrable compared to plain PC concretes and are [4] L. Tang, Engineering expression of the ClinConc model for prediction of free
and total ingress in submerged marine concrete, Cement and Concrete
usually preferred in the construction of durable RC struc- Research 38 (8–9) (2008) 1092–1097.
tures, it should be borne in mind that carbonation can delay [5] J.R. Mackechnie, Predictions of reinforced concrete durability in the marine
the visible signs of steel corrosion on the concrete surface environment PhD Thesis, Department of Civil Engineering, University of Cape
Town, 1996.
(due to internal micro-cracking) which may lead to higher [6] M.G. Stewart, X. Wang, M.N. Nguyen, Climate change adaptation for corrosion
steel mass/radial losses It is however important to note that control of concrete infrastructure, Structural Safety Vol. 35(3) (2012) 29–39.
the impenetrability of blended cement concretes signifi- [7] A.N. Scott, M.G. Alexander, The influence of binder type, cracking and cover on
corrosion rates of steel in chloride-contaminated concrete, Magazine of
cantly slows down both the carbonation process and conse- Concrete Research 59 (7) (2007) 495–505.
quently the corrosion rate in the post-cracking phase of steel [8] M.B. Otieno, H.D. Beushausen, M.G. Alexander, Chloride-induced corrosion of
corrosion. steel in cracked concrete – Part II: Corrosion rate prediction models, Cement
and Concrete Research 79 (2016) 386–394.
[9] C.Q. Li, Reliability based service life prediction of corrosion affected concrete
Empirical relationships have been presented based on the structures, ASCE Journal of Structural Engineering 130 (10) (2004) 1570–1577.
experimental results obtained in this study. These relationships [10] K. Tuutti, Corrosion of steel in concrete. Swedish Cement and Concrete
Research Institute, S-100 44 Stockholm, Report No. CBI Research 4:82, ISSN
relate to (i) time to cover cracking, and (ii) steel radial loss at cover
0346-6906, 1982, p. 468.
cracking. These are important parameters for both the design and [11] C.Q. Li, Y. Yang, R.E. Melchers, Prediction of reinforcement corrosion in
maintenance of corrosion-affected RC structures. These relation- concrete and its effects on concrete cracking and strength reduction, ACI
ships will need to be validated as in-service data becomes Materials Journal 105 (1) (2008) 3–10.
[12] W.Y. Jung, Y.S. Yoon, Y.M. Sohn, Predicting the remaining service life of land
available. concrete by steel corrosion, Cement and Concrete Research 33 (5) (2003) 663–
677.
[13] M.B. Otieno, M.G. Alexander, H.-D. Beushausen, Corrosion in cracked and
Acknowledgements uncracked concrete - influence of crack width, concrete quality and crack re-
opening, Magazine of Concrete Research 62 (6) (2010) 393–404.
[14] Z. Cui, A. Alipour, Concrete cover cracking and service life prediction of
The authors acknowledge the financial support of The Concrete reinforced concrete structures in corrosive environments, Construction and
Institute (TCI), South Africa, National Research Foundation (NRF), Building Materials 159 (20) (2018) 652–671.
180 M. Otieno et al. / Construction and Building Materials 198 (2019) 172–181

[15] www.weathersa.co.za Accessed: 15th May, 2018. [46] A.A. Torres-Acosta, A. Sagues, Concrete Cracking by Localized Steel Corrosion -
[16] M. Castellote, L. Fernandez, C. Andrade, C. Alonso, Chemical changes and phase Geometric Effect, ACI Materials Journal 101 (6) (2004) 501–507.
analysis of OPC pastes carbonated at different CO2 concentrations, Materials [47] A. Jamali, U. Angst, B. Adey, B. Elsener, Modeling of corrosion-induced concrete
and Structures 42 (4) (2009) 515–525. cover cracking: A critical analysis, Construction and Building Materials 42
[17] Y. Ballim, B.J. Lampacher, Long-term carbonation of concrete structures in the (2013) 225–237.
Johannesburg environment, Journal of the South African Institution of Civil [48] Z. Bažant, Physical model for steel corrosion in concrete sea structures -
Engineering (SAICE) 38 (1) (1996) 5–9. Theory and application, Journal of The Structural Division 105 (ST6) (1979)
[18] J. Wuhan, J. Shen, Y. Ma, S. Lai, Comparison of concrete carbonation process 1137–1166.
under natural condition and high CO2 concentration environments, Journal of [49] Y. Liu, Modelling the time-to-corrosion cracking of the cover concrete in
Wuhan University of Technology - Materials Science Education 25 (3) (2010) chloride-contaminated reinforced concrete structures. PhD Thesis,
515–522. Department of civil engineering, Virginia Polytechnic Institute and State
[19] AM. Neville, Properties of Concrete (5th Edition). Pearson Education Limited, University, 1996.
Edinburg Gate, Harlow, Essex CM20 2JE, England, 2011, ISBN: 978-0-273- [50] T. El Maaddawy, K. Soudki, A model for prediction of time from corrosion
75580-7. initiation to corrosion cracking, Cement & Concrete Composites 29 (3) (2007)
[20] B. Šavija, M. Luković, Carbonation of cement paste: Understanding, challenges, 168–175.
and opportunities, Construction and Building Materials 116 (2016) 285–301. [51] F. Chen, H. Baji, C.-Q. Li, A comparative study on factors affecting time to cover
[21] G.W. Groves, A. Brogh, I.G. Richardson, C.M. Dobson, Progressive changes in the cracking as a service life indicator, Construction and Building Materials 163
structure of hardened C3S cement pastes due to carbonation, Journal of (28) (2018) 681–694.
American Ceramic Society 74 (11) (1991) 2891–2896. [52] S.S. Rasheeduzzafar, A.S. Al-Saadoun, A. Al-Gahtani, Corrosion cracking in
[22] G.W. Groves, D. Rodway, I. Richardson, The carbonation of hardened cement relation to bar diameter, concrete cover and quality, Journal of Materials in
pastes, Advanced Cement Research 3 (11) (1990) 117–125. Civil Engineering 4 (1992) 327–342.
[23] A.E. Morandeau, C.E. White, In situ X-ray pair distribution function analysis of [53] W. Ashraf, Carbonation of cement-based materials: Challenges and
accelerated carbonation of a synthetic calcium–silicate–hydrate gel, Journal of opportunities, Construction and Building Materials 120 (1) (2016) 558–570.
Materials Chemistry A 3 (16) (2015) 8597–8605. [54] M. Auroy, S. Poyet, P.L. Bescop, J.-M. Torrenti, T. Charpentier, M. Moskura, X.
[24] V. Papadakis, M. Fardis, C. Vayenas, Effect of composition, environmental Bourbon, Comparison between natural and accelerated carbonation (3% CO2):
factors and cement-lime mortar coating on concrete carbonation, Materials Impact on mineralogy, microstructure, water retention and cracking, Cement
and Structures 25 (5) (1992) 293–304. and Concrete Research 109 (2018) 64–80.
[25] N. Huet, V. L’Hostis, F. Miserque, H. Idrissi, Electrochemical behavior of mild [55] SANS-3001-CO3-2,, Civil engineering test methods: Part CO3-2: Concrete
steel in concrete: influence of pH and carbonate content of concrete pore durability index testing - Oxygen permeability test, South African Bureau of
solution, Electrochimica Acta 51 (1) (2005) 172–180. Standards - Standards Division, Pretoria, South Africa, 2015.
[26] M. Arandigoyen, B. Bicer-Simsir, J.I. Alvarez, D.A. Lange, Variation of [56] Y. Ballim, Towards an early age index for the durability of concrete, in:
microstructure with carbonation in lime and blended pastes, Applied Surface Proceedings of international conference, Concrete 2000 - Economic and
Science 252 (20) (2006) 7562–7571. durable construction through excellence. 7-9 September, 1993, Dundee,
[27] B. Johannesson, P. Utgenannt, Microstructural changes caused by carbonation Scotland, Volume 2, E & FN Spon, London. pp. 1003–1012.
of cement mortar, Cement and Concrete Research 31 (6) (2001) 925–931. [57] M.G. Alexander, Y. Ballim, K. Stanish, A framework for use of durability indexes
[28] V. Ngala, C. Page, Effects of carbonation on pore structure and diffusional in performance-based design and specifications for reinforced concrete
properties of hydrated cement pastes, Cement and Concrete Research 27 (7) structures, Materials and Structures 41 (5) (2008) 921–936.
(1997) 995–1007. [58] C.Q. Li, W. Lawanwisut, J.J. Zheng, W. Kijawatworawet, Crack Width Due to
[29] P.H. Borges, J.O. Costa, N.B. Milestone, C.J. Lynsdale, R.E. Streatfield, Corroded Bar in Reinforced Concrete Structures, International Journal of
Carbonation of CH and C-S-H in composite cement pastes containing high Materials & Structural Reliability 3 (2) (2005) 87–94.
amounts of BFS, Cement and Concrete Research 40 (2) (2010) 284–292. [59] C.Q. Li, S.T. Yang, Prediction of concrete crack width under combined
[30] Y.F. Houst, F.H. Wittmann, Influence of porosity and water content on the reinforcement corrosion and applied load, ASCE Journal of Engineering
diffusivity of CO2 and O2 through hydrated cement paste, Cement and Mechanics 137 (11) (2011) 722–731.
Concrete Research 24 (6) (1994) 1165–1176. [60] ASTM-G1-03, Standard practice for preparing, cleaning, and evaluating
[31] J.J. Chen, J.J. Thomas, H.M. Jennings, Decalcification shrinkage of cement paste, corrosion test specimens. ASTM International, ASTM International, 100 Barr
Cement and Concrete Research 36 (5) (2006) 801–809. Harbour Drive, West Conshohocken, PA 19428-2959, USA, 2011.
[32] B.G. Salvoldi, H. Beushausen, M.G. Alexander, Oxygen permeability of concrete [61] N. Ananmalay, Characterising the strength and durability performance of
and its relation to carbonation, Construction and Building Materials 85 (2015) South African Silica fume concretes. MSc dissertation, University of the
30–37. Witwatersrand, Johannesburg, South Africa, 1996.
[33] Y.A. Alhassan, The effects of materials and micro-climate variations on [62] E.J. Garboczi, Microstructure and transport properties of concrete, in: Kropp J,
prediction of carbonation rate in reinforced concrete in the inland Hilsdorf H., editors. Perform. Criteria Concr. Durab., London: E & FN SPON;
environment, School of Civil and Environmental Engineering, University of 1995, pp. 198-212.
the Witwatersrand, South Africa, Johannesburg, 2014. [63] P.K. Mehta, P.J.M. Monteiro, Concrete – microstructure, properties and
[34] Y. Liu, R.E. Weyers, Modelling the time-to-corrosion cracking in chloride materials, 3rd Edition., McGraw Hill, New York, 2006.
contaminated reinforced concrete structures, ACI Materials Journal 95 (6) [64] P.K. Mehta, Pozzolanic and cementitious by-products as mineral admixtures
(1998) 675–681. for concrete - a critical review, in: Proc First ACI/CANMET International
[35] S. Ahmad, Reinforcement corrosion in concrete structures, its monitoring and conference on the use fly ash, silica fume, slag, and other mineral by-products
service life prediction - a review, Cement and Concrete Composites 25 (4–5) in concrete, Montebello, Canada, Vol 1, ACI SP-79, pp. 1-46, 1987.
(2003) 459–471. [65] M.G. Alexander, Y. Ballim, J.R. Mackechnie, Use of durability indexes to achieve
[36] J.P. Broomfield, Corrosion of steel in concrete - understanding, investigation durable cover concrete in reinforced concrete structures, Materials Science of
and repair, 2nd Edition., Taylor & Francis, Oxford, United Kingdom, 2007. Concrete, VI, 2001, pp. 483–511.
[37] D. Frangopol, K. Soudki, Reliability of reinforced concrete girders under [66] M.B. Otieno, H.D. Beushausen, M.G. Alexander, Effect of chemical composition
corrosion attack, Journal of Structural Engineering 123 (3) (1997) 286–297. of slag on chloride penetration resistance of concrete, Cement and Concrete
[38] K. Vu, M. Stewart, Predicting the likelihood and extent of reinforced concrete Composites 46 (2014), https://doi.org/10.1016/j.cemconcomp.2013.11.003,
corrosion-induced cracking, Journal of Structural Engineering 131 (11) (2005) pp. 56-64.
1681–1689. [67] Y. Ballim, Curing and the durability of OPC, fly ash and blast-furnace slag
[39] C. Andrade, C. Alonso, F.J. Molina, Cover cracking as a function of bar corrosion: concretes, Materials and Structures 26 (1993) 238–244.
Part I-Experimental test, Materials and Structures 26 (8) (1993) 453–464. [68] B.G. Salvoldi, Modelling the carbonation of concrete using early age oxygen
[40] A. Beeby, Cracking, cover and corrosion of reinforcement, Concrete permeability index tests. MSc Thesis, Department of civil engineering,
International 5 (2) (1983) 35–40. University of Cape Town, 2010.
[41] S. Care, Q.T. Nguyen, K. Beddiar, Y. Berthaud, Times to cracking in reinforced [69] H.S. Wong, Y.X. Zhao, A.R. Karimi, N.R. Buenfeld, W.L. Jin, On the penetration of
mortar beams subjected to accelerated corrosion tests, Materials and corrosion products from reinforcing steel into concrete due to chloride-
Structures 43 (1–2) (2010) 107–124. induced corrosion, Corrosion Science 52 (7) (2010) 2469–2480.
[42] J.A. Mullard, M.G. Stewart, Corrosion-Induced cover cracking: new test data [70] K. Vu, M.G. Stewart, J. Mullard, Corrosion-induced cracking: experimental data
and predictive models, ACI Structural Journal 108 (1) (2011) 71–79. and predictive models, ACI Structural Journal 102 (5) (2005) 719–726.
[43] C. Alonso, C. Andrade, J. Rodriguez, J.M. Diez, Factors controlling cracking of [71] A. Michel, B.J. Pease, A. Peterová, M.R. Geiker, H. Stang, A.E.A. Thybo,
concrete affected by reinforcement corrosion, Materials and Structures 31 (7) Penetration of corrosion products and corrosion-induced cracking in
(1998) 435–441. reinforced cementitious materials: Experimental investigations and
[44] J. Rodriguez, L. Ortega, J. Casal, J. Diez, Corrosion of reinforcement and service numerical simulations, Cement and Concrete Composites 47 (2014) 75–86.
life of concrete structures’, in: C. Sjostrom (Ed.), Durability of Building [72] L.J. Parrott, A study of carbonation-induced corrosion, Magazine of Concrete
Materials and Components 7, London, UK, E & FN SPON, 1996, pp. 117–126. Research 46 (166) (1994) 23–28.
[45] S. Morinaga, Prediction of service lives of reinforced concrete buildings based [73] M. Thiéry, P. Faure, J. Bouteloup, Effect of carbonation on the microstructure
on rate of corrosion of reinforcing steel, in: Proceedings of the 5th and moisture properties of cement - based materials, in: International
International conference on durability of building materials and conference on Durability of Building Materials and Components, 12-15 April,
components, 1990. 7-9 November, Brighton, UK. Porto-Portugal, pp. 1-8, 2011.
M. Otieno et al. / Construction and Building Materials 198 (2019) 172–181 181

[74] J.H.M. Visser, Accelerated carbonation testing of mortar with supplementary [76] M. Stefanoni, U. Angst, B. Elsener, Corrosion rate of carbon steel in carbonated
cementing materials - limitation of the acceleration due to drying, Heron 57 concrete – a critical review, Cement and Concrete Research 103 (2018) 35–48.
(2012). [77] M. Otieno, Sensitivity of chloride-induced corrosion rate of steel in concrete to
[75] A. Morandeau, M. Thiery, P. Dangla, Impact of accelerated carbonation on OPC cover depth, crack width and concrete quality, Materials and Structures 50 (9)
cement paste blended with fly ash, Cement and Concrete Research 67 (2015) (2017) 1–10.
226–236.

You might also like