You are on page 1of 13

Cement and Concrete Research 95 (2017) 257–269

Contents lists available at ScienceDirect

Cement and Concrete Research

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

Effect of crack openings on carbonation-induced corrosion


Rita Maria Ghantous a,c,⁎, Stéphane Poyet a, Valérie L'Hostis a, Nhu-Cuong Tran b, Raoul François c
a
Den-Service d'Etude du Comportement des Radionucléides (SECR), CEA, Université Paris-Saclay, F-91191, Gif-sur-Yvette, France
b
EDF, R&D, MMC, F-77818 Moret-sur-Loing Cedex, France
c
LMDC, Université de Toulouse, INSA, UPS, Toulouse, France

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

Article history: Reinforced concrete is widely used in the construction of buildings, historical monuments, infrastructures and
Received 9 November 2016 nuclear power plants. For a variety of reasons, many concrete structures are subject to unavoidable cracks that
Received in revised form 14 February 2017 accelerate the diffusion of atmospheric carbon dioxide to the steel/concrete interface. Carbonation at the inter-
Accepted 23 February 2017
face induces steel corrosion that could cause the development of new cracks in the structure, a determining factor
for its durability. The aim of this article is to study the effect of existing cracks on the development of carbonation-
Keywords:
Reinforced concrete
induced corrosion. The results indicate that, after the initiation phase, the corrosion kinetics decreases with time
Cracks and the free corrosion potential increases independently of the crack opening. In addition, the corroded zone
Carbonation-induced corrosion matches the carbonated one. The interpretation of these results allows the authors to conclude that, during the
Environmental conditions corrosion process, corrosion products seal the crack and act as a barrier to oxygen and water diffusion. Conse-
quently, the influence of crack opening on corrosion development is masked and the corrosion development is
limited.
© 2017 Elsevier Ltd. All rights reserved.

1. Introduction alkalization of carbonated mortar. In the same study, results of experi-


ments exposing cracked and carbonated specimens to different relative
Corrosion of the reinforcing steel is the main pathology affecting re- humidities, validate the repassivation assumption. Tremper [19] inves-
inforced concrete structures and is a determining factor for their dura- tigated cracked specimens after 10 years of outdoor exposure and iden-
bility. These structures are subject to unavoidable cracks which create tified the presence of corrosion near the cracks and their immediate
a pathway for atmospheric carbon dioxide, oxygen, water and chlorides surroundings. He also added that the corrosion detected did not have
to the steel/concrete interface and subsequently reduce the time to cor- deleterious consequences because of its minor degree. Additionally,
rosion initiation. Both laboratory studies [1–6] and in situ observations Vesikari [20] investigated the corrosion development in the cracked
[7–11] have noted earlier corrosion initiation in cracked concrete struc- concrete of 50 year old bridges and found that very few cracks had
tures. Load-induced cracking is accompanied by interfacial slipping and given rise to serious corrosion damage. This observation supports the
separation between concrete and steel [12–15]. It has also been shown idea that a reduction of corrosion kinetics occurs after the initiation
that both chloride- and carbonation-induced corrosion start and devel- phase, due to the corrosion products developed deep in the crack.
op for a few millimeters around the rebar at its interception with the On the other hand, Dang et al. [21], studying carbonation-induced
preexisting crack [16,17]. While there is a general consensus in the sci- corrosion propagation, used ring-shaped mortar specimens with
entific community on the deleterious effects of cracks and steel/mortar preexisting cracks induced mechanically by internal pressure (expan-
interface quality on the initiation of reinforcement corrosion, the effects sive core method). After accelerated carbonation (50% CO2 ‐ 65% RH),
of such phenomena on the propagation of reinforcement corrosion in- the specimens were subjected to humidification/drying cycles to accel-
duced by carbonation are still open to debate. erate corrosion development. It was observed that the CO2 spread along
Tuutti [18] proposed a corrosion model for cracked and carbonated the total length of the steel/mortar interface and induced corrosion ini-
concrete, which supposes that steel repassivation is possible due to tiation on the entire circumference of the rebar. Thereafter, new cracks
the sealing properties of the corrosion products and to a potential re- were detected in the concrete cover characterized by a low tensile
strength [22]. Nevertheless, because corrosion was initiated on the en-
tire length of the rebar, it was not possible to conclude whether carbon-
⁎ Corresponding author at: Den-Service d'Etude du Comportement des Radionucléides ation-induced corrosion in cracked structures would lead to corrosion
(SECR), CEA, Université Paris-Saclay, F-91191, Gif-sur-Yvette, France.
E-mail addresses: rita-maria.ghantous@yncrea.fr (R.M. Ghantous),
cracks and thus to corrosion propagation.
stephane.poyet@cea.fr (S. Poyet), valerie.lhostis@cea.fr (V. L'Hostis), In the field of nuclear energy, reinforced concrete is used for struc-
nhu-cuong.tran@edf.fr (N.-C. Tran), Raoul.francois@insa-toulouse.fr (R. François). tures that have an impact on safety (containment buildings, cooling

http://dx.doi.org/10.1016/j.cemconres.2017.02.014
0008-8846/© 2017 Elsevier Ltd. All rights reserved.
258 R.M. Ghantous et al. / Cement and Concrete Research 95 (2017) 257–269

Fig. 1. Schematic representation of the reinforced mortar specimen.

towers), it is therefore important to complete the literature and under- 2. Experimental program
stand the effect of cracking and steel/concrete interface damage, in
terms of both corrosion initiation and corrosion propagation in cracked 2.1. Materials and specimens
specimens. In this project, specimens were exposed to raining/drying
cycles to simulate the outside environment and accelerate corrosion The specimens tested were 70 × 70 × 280 mm prisms. A 6 mm cor-
naturally. The ultimate goal was to find answers to the following rugated rebar was positioned in the middle of each specimen as shown
questions: in Fig. 1. All the specimens were prepared with a mortar mix containing
three parts sand, two parts cement and one part water. Ordinary
• Does crack opening influence the kinetics of carbonation-induced cor- Portland cement (CEM I 52.5) and siliceous sand (according to CEN EN
rosion? 196-1) were used. The yield strength of the steel used was 500 MPa.
• Will carbonation-induced corrosion deep in the crack induce the de- Mortar was poured in prismatic molds in two layers, each of them
velopment of corrosion cracks and threaten the durability of the struc- being vibrated to eliminate air bubbles. After 24 h, the specimens
ture or it will seal the cracks and slow down corrosion propagation? were unmolded and then cured for 28 days in water with calcium hy-
droxide. The mechanical properties of the mortar mix measured follow-
ing current standards and recommendations are shown in Table 1. The
In the following sections, the experimental protocols developed to specimens tested in this study were made from 2 consecutive batches
provide answers to the above questions are detailed and then the re- of 60 l each.
sults obtained are discussed.
2.2. Cracking

Table 1 Immediately after curing, cracks of different widths were generated


Mechanical characteristics of the mortar mix. at mid-span of the prismatic specimens using a three-point bending
Compressive strength fc (MPa) 55 test. The cracking protocol consisted in applying the load at the mid
Tensile strength fctk (MPa) 3.5 span of the simply supported specimen using a hydraulic pump as
Young’s Modulus E (GPa) 33.3 shown in Fig. 2(a). The crack was usually obtained in the cross section
Poisson’s ratio ν 0.21
subjected to the maximal tensile stress (mid-span). In order to quantify

Fig. 2. Three-point bending test on 70 × 70 × 280 mm specimens.


R.M. Ghantous et al. / Cement and Concrete Research 95 (2017) 257–269 259

Fig. 3. Range of residual crack openings obtained by the three point bending test.

the crack opening, a Linear Variable Differential Transformer (LVDT) openings, of 100 μm, 300 μm and 500 μm, were performed on several
was placed where the crack was expected to appear as shown in Fig. prismatic specimens.
2(b). It measured the horizontal displacement between two points
located 20 mm from the specimen center on either side. Throughout 2.3. Corrosion
the test, the applied load and the displacement measured by the LVDT
were logged. Note that the displacement registered by the LVDT To trigger corrosion, the specimens were kept in an accelerated car-
before the appearance of the crack was due to flexural deformation of bonation chamber at 3% CO2, 55% relative humidity and 25 °C for one
the beam and cannot be considered as a crack opening. Therefore, to month. It had been verified that these accelerated carbonation condi-
calculate the crack opening, these initial displacements were set equal tions were representative of the natural conditions in a previous study
to zero. [23,24], in which it was also shown that the carbonation duration was
At the beginning of the bending test, the load was increased slowly long enough to ensure carbonation of the alteration zone [23,24]. Car-
until a crack was detected. Then, by performing loading\unloading cy- bonation induces depassivation of the steel intercepting a crack and
cles and by increasing the maximum applied load from one cycle to an- thus active corrosion could start once the presence of the required re-
other, the residual crack opening was increased in a controlled way and agents (water and oxygen) was ensured. It is important to note that, be-
a range of crack opening values was obtained (Fig. 3). fore carbonation, the cracked specimens were pre-conditioned at (20 ±
It was important, here, to limit the length of the zone in which the 1 °C; 60 ± 5% RH) for one month. This step was important to reduce the
bond between the steel and the mortar was altered (to prevent corro- water content of the cementitious materials and thus ensure CO2 diffu-
sion initiation all along the steel/mortar interface as in [22]). A previous sion in them during the carbonation test.
study proved that the three point bending test did not lead to total steel/ After carbonation, active corrosion was accelerated by performing
mortar interface damage [23]. In the present study, three residual crack multiple simulations of rain/drying cycles. To this end, a custom rain
chamber was designed and manufactured. It was equipped with three
rows, in order to test the required number of specimens. The ‘rainfall’
was ensured using a micro drip system (Fig. 5(b)). On each row,
water was recovered in a collector (yellow in Fig. 4) and was evacuated
from the three rows through a common pipe (grey pipe linked to the
yellow collector in Fig. 4). Tap water was used to represent rainfall. To
avoid clogging of the sprinklers due to hard water, the water was fil-
tered to remove impurities (filters in green in Fig. 4). The pH of the
water was 7.5 and its chemical composition is given in Table 2.
Raining/drying cycles were programmed by an automated control
unit. In each row of each rain chamber and in the laboratory, a ther-
mo-hygrometer continuously measured temperature and relative hu-
midity. Dedicated software was used to log all the data each minute. A
photo of the small chamber is shown in Fig. 5(a). This experiment was
performed in an air-conditioned laboratory in which the temperature
was regulated at 20 °C ± 1 °C and the RH at 60 ± 5%.
Each raining phase lasted 30 min and was followed by a 72 h drying
phase. During the raining period, the RH value increased to 95% ± 5%.
Once the drying phase started, the RH took around 12 h to drop to
Fig. 4. Custom rain chambers. 50% ± 10% (Fig. 6).
260 R.M. Ghantous et al. / Cement and Concrete Research 95 (2017) 257–269

Table 2
Elements present in water.

Element NH4-N NH4 NO2 HCO3 Cl NO3 SO4 CO3

Concentration (mg/L) b0.039 b0.05 b0.01 200 12 17 45 b1

Fig. 5. Real device for raining/drying cycles.

2.4. Corrosion analyses rebar was inspected visually. Corroded lengths were measured on its
upper and lower parts with respect to the casting direction.
After various exposure durations, three specimens were used for After steel extraction, phenolphthalein was pulverized onto the
physical-chemical analysis and one other specimen was used for fresh surface to measure the carbonated length along the steel/mortar
surface analysis (optical microscopy and Raman micro-spectroscopy). interface and compare it to the corroded length (Fig. 7).
Specimens were analyzed after 0, 15, 30, 60, 119 raining/drying cycles
corresponding to 0, 1.5, 3, 6, 12 months respectively. The free corrosion b). Gravimetric measurements
potential of the rebars was monitored continuously during the experi-
ment on other specimens exposed to the same corrosion conditions. The corrosion products attached to the extracted rebar were removed
Each of these analysis methods was applied to specimens having three by chemical attack (Sb2O3 + SnCl2 + HCl) following the standard ISO
different residual crack openings (100, 300 and 500 μm). 8407 [25]. The difference between the initial mass of the steel (before cor-
The experimental program was summarized in Table 3 keeping in rosion) and its final mass (after chemical treatment) gave the steel mass
view the parameters of the specimens studied and the corrosion analy- lost during corrosion (ΔmS). In this study, gravimetric measurements
ses methods. were repeated on three replicates.

2.4.1. Physico-chemical analysis c). Estimation of corrosion kinetics

a). Extraction of rebar and visual inspection The corrosion kinetics was calculated according to Eq. (1) [26].

The specimen was first split into two parts. Before the reinforcement mm ΔmS 3:65  10−2
¼  ð1Þ
was extracted, the distribution of corrosion products along the steel year ρS  SSL Δt

Fig. 6. Relative humidity variation during raining/drying cycles in the reference protocol.
R.M. Ghantous et al. / Cement and Concrete Research 95 (2017) 257–269 261

Table 3
Summary of the experimental program.

Specimens Crack Number of Time period for Method of


characteristics openings specimens corrosion analysis
(μm) analysis

70 × 70 × 280 100 22 0, 1.5, 3, 6, 12 - Gravimetric


mm (CEM I 300 22 measure-
mortar) 500 22 ments
- Optical mi-
croscopy Fig. 9. Electrochemical measurement.
- Raman
spectrosco-
py
- Free corro-
sion 2.4.2. Surface analysis
potential Optical microscopy and Raman micro-spectroscopy were used to
identify and to measure the thickness of the corrosion products formed
at various locations in the specimens. From each specimen, a sample of
120 × 70 × 15 mm was extracted as shown in Fig. 8(a) and mounted in
epoxy resin at room temperature (20 °C). The sample was then polished
until the mid-section of the steel/reinforcement was reached (according
to the polishing direction shown in Fig. 8(a). Then, the surface of this
section was polished by grinding with SiC paper, grades 80 to 4000
and with 3-μm diamond paste under ethanol (Fig. 8(b)).
The morphology of the corrosion products layer and its thickness
along the entire length of the rebar were obtained by optical microsco-
py. Raman micro-spectroscopy analyses were performed on two differ-
ent zones: deep in the crack (called zone 2 in Fig. 8(b)) and at the end of
Fig. 7. Steel corrosion length with respect to the length of the carbonated steel/mortar
interface.
the corrosion layer (called zone 1 in Fig. 8(b)). The phases were identi-
fied by comparison with spectra in the literature [27–29].

where:ΔmS is the iron mass lost (mg);ρS is the volumetric mass density
2.4.3. Electrochemical analysis
of steel (g/cm3).SSL is the surface area of the corroded steel reinforce-
The evolution of the free corrosion potential was monitored by
ment (dm2).Δt is the duration of the active corrosion period (days).
means of an embedded manganese dioxide reference electrode (ERE
X-ray Computed Tomography (X-ray CT) was performed on the cor-
20) that was placed in the mortar cover of some specimens before cast-
rugated 6 mm steel rebar in order to determine its external area. The
ing. It was positioned at 50 mm from the crack (Fig. 9). In this configu-
rebar was examined using the 3D μCT scanner at BAM (Federal Institute
ration, the reference electrode measured a coupled corrosion potential
for Materials Research and Testing in Berlin). This scanner is equipped
(between active and passive parts of the steel bar), which was continu-
with a 225 kV micro focus X-ray tube and a flat panel detector with
ously logged using a data acquisition switch unit.
2048 × 2048 pixels. The spatial resolution was approximately 3 μm spe-
cial voxel (volumetric picture element) when scanning the 6 mm corru-
gated rebar. The surface area obtained by μCT was 19.54 ± 3. Results
0.08 mm2/mm.
Consequently, the area of the corroded steel reinforcement was ob- 3.1. Distribution of the corrosion products
tained according to Eq. (2). In this equation, Lcorr is the average corro-
sion length determined during the visual inspection of the corroded a). Length of the corrosion product layer
rebar (mm).
The evolution of the average corrosion length measured on three rep-
SSL ¼ 19:54  Lcorr  10ð−4Þ ð2Þ licates with respect to the number of raining/drying cycles is given in

Fig. 8. Sample preparation.


262 R.M. Ghantous et al. / Cement and Concrete Research 95 (2017) 257–269

Fig. 10. Corrosion length evolution with respect to raining/drying cycles and residual crack width (visual inspection).

Fig. 10. It should be noted that, in each specimen, there was no difference According to Glasser and Matschei [30], the carbonated area detected by
in the corrosion lengths measured on the upper and the lower part of the the phenolphthalein test is limited to the totally carbonated zone where
rebar with respect to the casting direction. Consequently, the corroded the pH is lower than 8 and where portlandite and C-S-H are depleted.
length in each rebar was considered to be equal to the average value of This pH is lower than the value needed to depassivate the steel, which
these two corrosion lengths measured on its upper and lower part. is between 9.4 and 10 [31]. Parrott et al. [32] divided the carbonated
The corrosion length was observed to increase with the residual depth in two zones: the first totally carbonated and the second a
crack opening and tended to stabilize after 30–60 raining/drying cycles. mixed zone where calcium carbonate and portlandite existed. A color
This observation is consistent with the study by Tremper [19] in which change in the phenolphthalein solution was visible in the first zone
the size of the corroded area increased with the crack width. and covered only half of the mixed zone. Moreover, by detecting car-
bonation products in zones where the phenolphthalein result was neg-
b). Development of corrosion products ative, Demoulin et al. [31] confirmed the limited ability of the
phenolphthalein test to detect carbonation, especially at its beginning.
Fig. 11 shows the development of the corrosion to carbonation Therefore, in this study the carbonation length measured using the
length ratios with respect to raining/drying cycles applied to specimens phenolphthalein test may have underestimated the real carbonated
having 100, 300 and 500 μm residual crack openings. Each point corre- length and this could explain the presence of some corrosion to carbon-
sponds to the average corrosion length to carbonation length ratio mea- ation ratios that were slightly higher than 1.
sured on the upper and lower steel/mortar interfaces of different
specimens subjected to the same exposure conditions for the same c). Thickness of the layer of corrosion products
duration.
It can be seen that the corrosion products filled the carbonated steel/ Fig. 12 provides information on the corrosion layer thickness on ei-
mortar interface zone in the first thirty raining/drying cycles and did not ther side of the crack in specimens showing different residual crack
spread farther thereafter. Moreover, some ratios were greater than 1, openings (100, 300 and 500 μm) and subjected to different numbers
reaching a value of 1.6. This may have been related to the accuracy of of raining/drying cycles (30, 60 cycles). The letter “S” on each graph rep-
the phenolphthalein pH indicator test and not to corrosion propagation. resents the area between the curve and the x-axis. It corresponds to the

Fig. 11. Evolution of the corrosion over carbonation length ratio with respect to raining/drying cycles and residual crack widths (visual inspection).
R.M. Ghantous et al. / Cement and Concrete Research 95 (2017) 257–269 263

Fig. 12. Corrosion products distribution obtained on specimens showing different residual crack openings (optical microscopy).

overall corroded surface measured on the upper part of the steel/mortar triplicates subjected to the same conditions. An increase in the iron
interface and could be used as an indicator of the amount of corrosion mass loss with the number of cycles and with crack opening is obvious.
products having developed at this part of the interface. Whatever the crack opening, the iron mass loss tends to stabilize after
It can be noted that the corrosion layer thickness measured on the 30–60 raining/drying cycles. This stabilization in the loss of iron mass
upper steel/mortar interface increases with the residual crack opening. is consistent with the observation by Tuutti [18] (who exposed cracked
It is obvious that the length of the corrosion product layer also increases specimens to corrosion at different relative humidities). Corrosion at-
with the residual crack opening, which is in agreement with Fig. 10. More- tack was fast at the beginning of the corrosion process, then was seen
over, the amount of corrosion products increases with the residual crack to stabilize after this time and remained the same until the end of the
opening. Concerning the corrosion evolution, no significant difference in test.
the corrosion length is visible between 30 and 60 cycles for the three re-
sidual crack openings. This is in agreement with Fig. 10 and Fig. 11. b). Corrosion rate

It is clear from Fig. 14 that the corrosion kinetics decreased with the
3.2. Gravimetric measurements number of cycles for the three residual crack openings. This was due to
the formation of porous oxide layer at the steel surface. The decrease in
a). Iron mass loss the corrosion kinetics could not be related to an autogenous healing of
cracks in concrete because the occurrence of the latter phenomena re-
Fig. 13 presents the results of the gravimetric measurements. Each quests a fully saturated crack [33,34] which is not the case in this
point in this figure is an average value of iron mass loss measured on study. In addition, the optical microscopy analyses performed along

Fig. 13. Iron mass losses in specimens having different residual crack openings with respect to raining/drying cycles.
264 R.M. Ghantous et al. / Cement and Concrete Research 95 (2017) 257–269

Fig. 14. Corrosion rate evolution in specimens having different residual crack openings with respect to raining/drying cycles.

the total duration of the experiments confirm the absence of a calcite rebar. However, substances that increase the pH value can diffuse
crystal growth in the crack with the raining/drying cycles. For instance, from the mortar to the steel, leading to a repassivation. This assumption
Fig. 15 shows an optical microscopy photo of a crack in a specimen hav- was also proposed by Schießl [10]. The phenomenon was verified here
ing a 100 μm surface residual crack opening (smallest crack width tested by monitoring the evolution of the carbonated steel/mortar interface
in this study) after 6 months of exposure to raining/drying cycles. It can length at different times during the corrosion test. It was found that
be seen that no calcite crystal seems to plug the crack path while the the carbonated length remained unchanged for the entire exposure du-
corrosion kinetics after this exposure duration is reduced. Therefore, ration and did not decrease (Fig. 16). This does not exactly fit the theory
the reduction in the corrosion kinetics is certainly due to corrosion of Tuutti and Schießl and the explanation could be linked to the neutral
products that seal the crack and not to an autogenous healing of crack water that fills the cracked area every three days, during the raining
in concrete. phase. The absorption of water by the mortar may inhibit a possible in-
It can also be noted that, near 60 cycles, the corrosion rate seems crease in its pH in the cracked zone surrounding the rebar. In addition,
to be independent of the residual crack opening and is around 10 μm/ this could also be related to the macrocell process that leads to the cre-
year after 119 cycles. In addition, the difference between the ation of hydroxyl ions outside the carbonated (and corroded) area. Nev-
corrosion kinetics measured at 6 and 12 months is negligible. ertheless, despite all these assumptions, an increase in pH cannot be
Consequently, an estimation of the corrosion kinetics at 12 months totally discounted because it could occur without bringing the pH
relative to that at 6 months should logically give a value comparable above 9, which would make it undetectable by the phenolphthalein
to the passive corrosion kinetics (0.1–1 μm/year [35,36]). However, pH indicator test. The theory of Tuutti and Schießl would probably
it was not possible to determine the exact value in this study because apply for long exposure durations (more than 12 months).
the specimens analyzed at 6 months were not the same as those
analyzed at 12 months. 3.3. Free corrosion potential measurements
The observed decrease in the corrosion kinetics corresponds to the
decrease proposed in Tuutti's model [18]. According to Tuutti, this is The evolution of the measured corrosion potentials is given in
linked to the steel repassivation and he assumes that the corrosion Fig. 17. Two distinct periods can be observed in the corrosion process.
products seal cracks and inhibit the access of carbon dioxide to the
• The first period, during the first month of exposure, corresponds to
the initiation of corrosion, in which a drop in the corrosion potential
from −70 mV/SCE to −350 mV/SCE indicates corrosion initiation at
the crack location.
At the beginning of the corrosion test, the steel deep in the crack
intercepted carbonated mortar. Therefore, this part of the steel was
able to be oxidized and the corrosion potential thus fell.
In this project, the rebar was surrounded by a heterogeneous mortar
(carbonated and non-carbonated). This heterogeneity induced a differ-
ence in the potential and led to the formation of a cell between these
two zones. The passivated steel intercepting a non-carbonated mortar
and located far from the crack could act as a cathode while the steel
intercepting a carbonated mortar and located deep in the crack could
act as an anode. Therefore, during corrosion initiation, in addition to
the microcell mechanism, a macrocell corrosion mechanism directly de-
pendent on the chemical characteristics of the surrounding electrolyte
(pH of the mortar) also came into play. According to [37], it is likely
that macrocell corrosion would predominate.
• The second part of the process concerns corrosion propagation deep in
Fig. 15. Optical microscopy photo of a crack in a specimen showing 100 μm residual crack the crack. In this process, the corrosion potential starts to increase from
opening and exposed to 60 raining/drying cycles. - 350 mV/SCE and almost reaches the corrosion potential of the
R.M. Ghantous et al. / Cement and Concrete Research 95 (2017) 257–269 265

Fig. 16. Evolution of the carbonated steel/mortar interface length in specimens having different residual crack openings and corroded in the reference test.

uncracked specimen (around −70 mV/SCE), which is an indication of process at the crack location due to the oxide layer developed deep in
the reduction of the macro-cell process of corrosion. This increase in the crack. The difference in the potential between the zone of the
the corrosion potential could be related to the limitation of the anodic steel intercepting a carbonated mortar and the one intercepting a

Fig. 17. Variation of free corrosion potential (versus SCE) in specimens showing different residual crack openings and exposed to raining/drying cycles for 310 days.
266 R.M. Ghantous et al. / Cement and Concrete Research 95 (2017) 257–269

Fig. 18. Representative Raman spectra of corrosion products obtained on specimens having three different residual crack openings.

non-carbonated mortar would then be strongly reduced. At this mo- crystallized goethite is also detected and, as pointed out in the study
ment, the macrocell corrosion process would be limited and the by Monnier et al. [41], the electrochemical reactivity of this phase is
microcell corrosion process would dominate. linked to its crystallinity. Therefore, the goethite detected in the corrod-
ed specimens can also be a reactive phase. Lepidocrocite is also detected
in the corroded specimens and, according to Misawa et al. [42], this
3.4. Type of corrosion products phase appears in more aerated conditions. Despite the fact that this
phase is not usually shown in the corrosion of steel in concrete [31,
Representative Raman spectra obtained in three different zones of 43], it has been detected in long-term corrosion layers developed on re-
the upper part of the steel/mortar interface are given in Fig. 18. A sche- bars embedded in concrete but only where the steel/concrete interface
matic summary of the types of corrosion products identified is given in contains air bubbles [44]. The presence of these voids may increase the
Fig. 19. Ferrihydrite (5Fe2O3.9H2O), a poorly crystallized phase whose aeration, which may justify the appearance of lepidocrocite in [44]. Its
chemical formula is still under discussion [38,39], is always present. In presence in this study could be correlated with the presence of cracks
addition to ferrihydrite, maghemite is also detected. These phases are in the concrete. It may also have been an artifact due to the preparation
known to be chemically and electrochemically reactive1 [40]. Poorly of the sample. The cause of the lepidocrocite detection could not be con-
firmed but, independently of its origin during the wetting stages of
1
In this case, reactivity means the ability of identified phases constituting the rust layer raining/drying cycles, it could be reduced and be involved in the follow-
to be reduced ing cycle as proven in [45]. Antony et al. [46,47] confirm that
R.M. Ghantous et al. / Cement and Concrete Research 95 (2017) 257–269 267

Fig. 19. Summary of the corrosion products identified in different zones of specimens having different residual crack openings.

lepidocrocite can be one of the reactive phases but they add that ferrihy- It can be seen that the types of corrosion products are independent
drite is much more reactive. of the residual crack openings.
In some Raman spectra, a double peak is visible at 340, 395 cm−1. It
was impossible to determine the corresponding phase on the basis of 4. Discussion and conclusions
our literature review. It usually appears with ferrihydrite and it may
be a precursor phase developing during the formation or the transfor- In this research study, the effect of crack opening on the initiation
mation of ferrihydrite. and propagation of carbonation-induced corrosion was tested and

Fig. 20. Corrosion model supposing corrosion propagation induced by a total carbonation of the concrete cover (corresponding to the corrosion model proposed by François & Arliguie
[50]).
268 R.M. Ghantous et al. / Cement and Concrete Research 95 (2017) 257–269

Fig. 21. Corrosion model supposing corrosion propagation induced by a corrosion crack.

analyzed. The results indicate a decrease in the corrosion kinetics with specimens after four and ten years of exposure and reported the pres-
the number of raining/drying cycles. The effect of the crack opening ence of corrosion in the uncracked zones due to the penetration of the
on the corrosion kinetics was masked after several such cycles and the carbonation front to the steel.
corrosion products did not spread beyond the limits of the carbonated • The second possibility assumes a creation of corrosion cracks before
zone. No corrosion cracks were visible on the outside surface of the the total carbonation of the concrete cover. In this condition, the cor-
specimens. Furthermore, an increase in the free corrosion potential rosion cracks may be the underlying cause of the corrosion propaga-
was observed after several raining/drying cycles. All these results tion process. However, it should be noted that the new corrosion
allow the authors to infer that the corrosion products seal the cracks cracks intercept rebar embedded in an alkaline medium. Therefore,
and limit the access of oxygen and water to the rebar, thus slowing a certain duration of carbonation is needed before the corrosion can
down the propagation of corrosion This result is in agreement with propagate along the rebar. In fact, corrosion cracks are generally par-
that obtained by a numerical simulation in the study by Millard et al. allel to the rebar and this promotes homogenous carbonation deep
[48]. It has also been proven that even a crack of 500 μm residual in the crack. Therefore, a microcell corrosion mechanism may exist
width (in un-loaded conditions) does not lead to a dangerous corrosion during the propagation phase. A schematic sketch of steel corrosion
state. This result was unexpected since the European standard EN 1992- development in this second possibility is given in Fig. 21. It should
1-1 recommends a maximum crack opening of 300 μm (in loaded con- be noted that this second possibility has a small probability of occur-
ditions) for the XC4 class. rence, especially in structures that are not subjected to dynamic
Additionally, the free corrosion potential measurements suggest loading.
that two corrosion mechanisms may occur consecutively. The first
mechanism is globally macrocell corrosion that occurs during the initi- Finally, concerning the types of corrosion products, it can be de-
ation phase. Afterwards, the second mechanism, which acts during the duced that there is no influence of the crack opening on the type of cor-
propagation phase, corresponds mainly to microcell corrosion. rosion products. The main observation is that the corrosion products
The corrosion rate measured in this study decreased with time and, developed are known to be reactive phases that are electrochemically
for the duration of this study, the corrosion was in its induction phase and chemically unstable and have tendency to be reduced
(reduction in the corrosion rate induced by the development of a porous (lepidocrocite, ferrihydrite, poorly crystallized goethite, etc.). Therefore,
corrosion layer on the steel surface). The duration of this phase cannot these phases may act as a cathode during corrosion due to the microcell
be predicted and, in the long term, two possibilities exist: mechanism, in which the iron is the anode.
• The first is that corrosion will not have any detrimental effect on the
durability of structures (Fig. 20(a)).
Acknowledgment
In this case, the parameter most harmful for the sustainability of a
structure is the concrete cover carbonation. The model proposed by
The experiments reported in this paper were conducted in the
Tuutti [18] to describe the steel corrosion sequence in uncracked con-
LECBA Laboratory at CEA Saclay in collaboration with the LMDC Labora-
crete could be applicable in the uncracked zones. It is presented in Fig.
tory of INSA Toulouse. The authors gratefully acknowledge the support
20(b) and consists of two phases (initiation and propagation). During
of the industrial partner EDF. The authors also gratefully acknowledge
the initiation phase, carbon dioxide penetrates the concrete cover,
the Federal Institute for Materials Research and Testing in Berlin for
eventually leading, at a certain depth of carbonation, to a situation
the X-ray Computed Tomography measurement. Finally, there is also
where corrosion initiation is thermodynamically favored. Thereafter,
an acknowledgment for Etienne Amblard, from the LECBA laboratory,
the corrosion propagation phase begins and is marked by an acceler-
for his help during the realization of the experiments.
ation in the corrosion rate and consequently in the deterioration
rate. The duration of the initiation and the propagation phases are de-
References
pendent on the concrete cover properties (thickness, w/c ratio) [6,10,
11,19,49]. [1] C. Arya, F.K. Ofori-Darko, Influence of crack frequency on reinforcement corrosion in
By combining, the corrosion processes deep in the crack and far from concrete, Cem. Concr. Res. 26 (1996) 345–353.
[2] N.S. Berke, M.P. Dallaire, M.C. Hicks, R.J. Hoopes, Corrosion of steel in cracked con-
the crack, it can be deduced that the corrosion model proposed in the
crete, Corros. Sci. 49 (1993) 934–943.
study by François & Arliguie [50] may describe the global behavior of a [3] R. Francois, G. Arliguie, Effect of micro-cracking and cracking on the development of
structure (Fig. 20(c)). A similar corrosion model for cracked and car- corrosion in reinforced concrete members, Mag. Concr. Res. 51 (1999) 143–150.
bonated concrete is also proposed by Tuutti [18]. [4] O. Gautefall, O. Vennesland, Effects of cracks on the corrosion of embedded steel in
silica-concrete compared to ordinary concrete, Nord. Concr. Res. 2 (1983) 17–28.
The results shown in Schießl's study [10] support this first hypothesis [5] T.U. Mohammed, N. Otsuki, M. Hisada, T. Shibata, Effect of crack width and bar types
concerning corrosion propagation. The author investigated cracked on corrosion of steel in concrete, J. Mater. Civ. Eng. 13 (2001) 194–201.
R.M. Ghantous et al. / Cement and Concrete Research 95 (2017) 257–269 269

[6] P. Schießl, M. Raupach, Laboratory studies and calculations on the influence of crack [29] M. Hanesh, Raman spectroscopy of iron oxides and (oxy)hydroxides at low laser
width on chloride-induced corrosion of steel in concrete, ACI Mater. J. 94 (1997) power and possible applications in environmental magnetic studies, Geophys. J.
56–61. Int. 177 (2009) 941–948.
[7] P. Fidjestol, N. Nilson, Field test of reinforcement corrosion in concrete, ACI Spec. [30] F. Glasser, T. Matschei, Interactions between Portland cement and carbon dioxide,
Publ. 65 (1980) 205–217. Proceedings of the 12th International Congress on the Chemistry of Cements, 2007.
[8] K. Katawaki, Corrosion of steel in the concrete exposed to seawater spray zone, [31] A. Demoulin, C. Trigance, D. Neff, E. Foy, P. Dillmann, V. L'hostis, The evolution of the
Symp. Proc. Crack. Concr. Struct. (1977) 133–136. corrosion of iron in hydraulic binders analysed from 46- and 260-year-old buildings,
[9] G. Rehm, H. Moll, Versuche zum studium des einflußes der rissbreite auf die Corros. Sci. 52 (2010) 3168–3179.
rostbilding an der bewehrung von stahlbeton-bauteilen, Deutscher Ausschuss für [32] L. Parrott, D. Killoh, Carbonation in a 36 year old, in-situ concrete, Cem. Concr. Res.
Stahlbeton 1964, p. 169. 19 (1989) 649–656.
[10] P. Schießl, Zur frage der zulassigen rissbreite und der enforderlichen betondeckung [33] C. Edvardsen, Water permeability and autogenous healing of cracks in concrete, ACI
im stahlbetonbau unter besonderer berucksichtigung der karbonatisierung des Mater. J. 96 (1999) 448–454.
betons, Deutscher Ausschuss für Stahlbeton 1976, p. 255. [34] N. Hearn, Self-sealing, autogenous healing and continued hydration: what is the dif-
[11] E.F. O'Neil, Study of reinforced concrete beams exposed to marine environment, ACI ference? Mater. Struct. 31 (1998) 563–567.
Spec. Publ. 65 (1980) 113–132. [35] A. Rosenberg, C. Hansson, C. Andrade, Mechanisms of corrosion of steel in concrete,
[12] R. François, I. Khan, N.A. Vu, H. Mercado, A. Castel, Study of the impact of localised Mater. Sci. Concr. 1 (1989) 285–314.
cracks on the corrosion mechanism, Eur. J. Environ. Civ. Eng. 16 (2012) 392–401. [36] C.M. Hansson, Comments on electrochemical measurements of the rate of corrosion
[13] A. Michel, A.O.S. Solgaard, B.J. Pease, M.R. Geiker, H. Stang, J.F. Olesen, Experimental of steel in concrete, Cem. Concr. Res. 14 (1984) 574–584.
investigation of the relation between damage at the concrete-steel interface and ini- [37] P. Bamforth, Probalistic performance based durability design of concrete structures,
tiation of reinforcement corrosion in plain and fibre reinforced concrete, Corros. Sci. Management of Concrete Structures for Long-Term Serviceability, Proceedings of
77 (2013) 308–321. the Int. Seminar. Sheffield, UK, 1997.
[14] B.J. Pease, Influence of Concrete Cracking on Ingress and Reinforcement [38] U. Cornell, Rochelle M. Schwertmann, The Iron Oxides: Structure, Properties, Reac-
Corrosion(Thesis) Technical University of Denmark, 2010. tions, Occurrences and Uses, John Wiley & Sons, Weinheim, 2003.
[15] B. Pease, M. Geiker, H. Stang, J. Weiss, The design of an instrumented rebar for as- [39] U. Schwertmann, R.M. Cornell, Iron Oxides in the Laboratory: Preparation and Char-
sessment of corrosion in cracked reinforced concrete, Mater. Struct. 44 (2011) acterization, John Wiley & Sons, Weinheim, 2008.
1259–1271. [40] R.M. Cornell, U. Schwertmann, The Iron Oxides: Structure, Properties, Reactions, Oc-
[16] R. Durton, A. Mommens, Corrosion des armatures dans le béton armé, 1964. currences and Uses, Wiley-VCH, 2003.
[17] R. François, J.C. Maso, Effect of damage in reinforced concrete on carbonation or [41] J. Monnier, et al., A corrosion study of the ferrous medieval reinforcement of the
chloride penetration, Cem. Concr. Res. 18 (1988) 961–970. Amiens cathedral. Phase characterisation and localisation by various microprobes
[18] K. Tuutti, Corrosion of Steel in Concrete, Cement and Concrete Research Institute, techniques, Corros. Sci. 52 (2010) 695–710.
1982. [42] T. Misawa, K. Hashimoto, S. Shimodaira, The mechanism of formation of iron oxide
[19] B. Tremper, The corrosion of reinforcing steel in cracked concrete, J. Proc. 43 (1947). and oxyhydroxides in aqueous solutions at room temperature, Corros. Sci. 14
[20] E. Vesikari, Corrosion of reinforcing steels at cracks in concrete, NASA STI/Recon (1974) 131–149.
Tech. Rep. N 83 (1981). [43] B. Huet, Comportement à la corrosion des armatures dans un béton carbonaté. Influ-
[21] V.H. Dang, R. François, V. L'Hostis, D. Meinel, Propagation of corrosion in pre-cracked ence de la chimie de la solution interstitielle et d ‘une barrière de transport(PhD the-
carbonated reinforced mortar, Mater. Struct. 48 (2015) 2575–2595. sis) Insitutat National des Sciences Appliquées de Lyon, 2005.
[22] V.H. Dang, Initiation and Propagation Phases of re-Bars Corrosion in Pre-Cracked Re- [44] V. L'Hostis, L. Vincent, V. Praca, D. Neff, L. Bellot-Gurlet, P. Dillmann, Characterization
inforced Concrete Exposed to Carbonation or Chloride Environment(PhD thesis) of long-term corrosion of rebars embedded in concretes from French Historical
Institut National des Sciences Appliquées de Toulouse, 2013. Buildings aged from 50 to 80 years, Eurocorr 2007, pp. 1–10.
[23] R.M. Ghantous, A. Millard, S. Poyet, R. François, V. L'Hostis, N.C. Tran, Experimental [45] M. Stratmann, K. Hoffmann, In situ Mössbauer spectroscopic study of reactions
and numerical characterization of load-induced damage in reinforced concrete within rust layers, Corros. Sci. 29 (1989) 1329–1352.
members9th International Conference on Fracture Mechanics of Concrete and Con- [46] H. Antony, L. Legrand, L. Maréchal, S. Perrin, P. Dillmann, A. Chaussé, Study of
crete Structures 2016, pp. 1–7. lepidocrocite γ-FeOOH electrochemical reduction in neutral and slightly alkaline so-
[24] R. M. Ghantous, S. Poyet, V. L'Hostis, N. C. Tran, and R. François, “Effect of accelerated lutions at 25 °C, Electrochim. Acta 51 (2005) 745–753.
carbonation conditions on the characterization of load-induced damage in rein- [47] H. Antony, S. Perrin, P. Dillmann, L. Legrand, A. Chaussé, Electrochemical study of in-
forced concrete members,” Submitt. to Mater. Struct., 2017. door atmospheric corrosion layers formed on ancient iron artefacts, Electrochim.
[25] NF EN ISO 8407, Corrosion des métaux et alliages - Élimination des produits de cor- Acta 52 (2007) 7754–7759.
rosion sur les éprouvettes d'essai de corrosion, 2014. [48] A. Millard, V. L'Hostis, Modelling the effects of steel corrosion in concrete, induced
[26] D. Landolt, Corrosion and Surface Chemistry of Metals, CRC Press, 2007. by carbon dioxide penetration, Eur. J. Environ. Civ. Eng. 16 (2012) 375–391.
[27] D. Neff, P. Dillmann, L. Bellot-Gurlet, G. Beranger, Corrosion of iron archaeological ar- [49] R. François, G. Arliguie, Influence of service cracking on reinforcement steel corro-
tefacts in soil: characterisation of the corrosion system, Corros. Sci. 47 (2005) sion, J. Mater. Civ. Eng. 10 (1998) 14–20.
515–535. [50] R. François, G. Arliguie, Durability of loaded reinforced concrete in chloride environ-
[28] D. Neff, S. Reguer, L. Bellot-Gurlet, P. Dillmann, R. Berthelon, Structural characteriza- ment, ACI Spec. Publ. 145 (1994) 573–596.
tion of corrosion products on archeological iron: an integrated analytical approach
to establish corrosion forms, J. Raman Spectrosc. 35 (2004) 739–745.

You might also like