Вы находитесь на странице: 1из 12

Cement and Concrete Research 116 (2019) 134–145

Contents lists available at ScienceDirect

Cement and Concrete Research


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

Durability of cement pastes exposed to external sulfate attack and leaching: T


Physical and chemical aspects

Rim Ragouga, Othman Omikrine Metalssib, , Fabien Barberonc, Jean-Michel Torrentid,
Nicolas Roussele, Loïc Divetf, Jean-Baptiste d'Espinose de Lacaillerieg
a
CESI,93 Boulevard de la Seine, 92000 Nanterre, France
b
Civil Engineering at the Research Unit: Experimental and Numerical Analysis of Structures, IFSTTAR-Paris-Est University, IFSTTAR, 14-20 Bd Newton, 77447, Champs/
Marne, MLV Cedex 2, France
c
EQIOM, 49 Avenue Georges Pompidou, 92300 Levallois-Perret, France
d
Materials & Structures Department, MASTat IFSTTAR-PARIS/Paris-Est University, IFSTTAR, 14-20 Boulevard Newton, Cité Descartes Champs sur Marne, F-77447
Marne la Vallée, Cedex 2, France
e
Civil Engineering at the Research Unit: NAVIER, IFSTTAR, IFSTTAR, 14-20 Boulevard Newton, Cité Descartes Champs sur Marne, F-77447 Marne la Vallée, Cedex 2,
France
f
Physical-Chemical Behaviour and Durability of Materials Laboratory, IFSTTAR, 14-20 Boulevard Newton, Cité Descartes Champs sur Marne, F-77447 Marne la Vallée,
Cedex 2, France
g
ESPCI-CNRS, École Supérieure de Physique et de Chimie Industrielles de la Ville de Paris (ESPCI Paris), PSL Research University, Soft Matter Sciences and Engineering
Laboratory, CNRS UMR 7615, 10 rue Vauquelin, F-75005 Paris, France

A R T I C LE I N FO A B S T R A C T

Keywords: Sulfate ions in seawater or underground can attack the cement paste leading to expansion and strength loss. This
External sulfate attack expansion is usually related to ettringite and gypsum precipitation. This study aims to characterize the possible
Structure durability altered zone and to identify the mechanisms of degradation under such conditions. For this study, cement pastes
Sulfate ingress manufactured using CEM I with two w/c ratios (0.45 and 0.60) were exposed to sodium sulfate solutions (semi-
Microstructure
immersion at a controlled pH = 8.0 ± 0.1) after one year of curing in water. The analysis of the physical aspect
Leaching
of this phenomenon shows that two modes of transfer occur: transfer of sulfate ions to the cement matrix and
leaching of calcium ions to the external solution. The analysis of the chemical composition of the affected
material highlights the progressive consumption of portlandite at the surface, the decalcification of C-S-H and
the formation of AFt from both Afm and aluminium incorporated in C-S-H.

1. Introduction processes associated with ESA and their consequences have been ex-
tensively studied in different laboratories; however they are often
The durability of concrete structures can be significantly improved presented as being complex and dependent on several variables in-
if the effects of the surrounding environment are taken into account trinsic to the material or to its surrounding environment [1,13–15].
upstream of the material formulation and of the structure di- Moreover, most previous studies have investigated the mechanisms of
mensioning. Despite numerous studies and expertise, some degradation ESA and its macroscopic consequences by studying the effect of this
mechanisms remain unclear and controversial. This is the case for ex- phenomenon on loss of mass, loss of mechanical resistance and possible
ample of external sulfate attack (ESA) [1–7] which can be defined as a expansions [13,16,17]. Recent works are increasingly interested in the
chemical reaction between the sulfate ions, having penetrated into the transfer aspect in ESA and the coupled chemistry-transport aspect
concrete by a transfer mechanism, and the mineral components of the [10,18]. Despite differences in the literature, all studies conclude on the
hardened cement paste, mainly tricalcium aluminate [8–10]. It can severity of this type of attack on the durability of the concrete structures
cause a macroscopic swelling of the material with an appearance of since it can induce very high damage after a short period (10 to
cracks inducing a loss of the mechanical resistance [11,12]. Indeed, the 15 years).


Corresponding author.
E-mail addresses: rragoug@cesi.fr (R. Ragoug), othman.omikrine-metalssi@ifsttar.fr (O.O. Metalssi), fabien.barberon@eqiom.com (F. Barberon),
jean-michel.torrenti@ifsttar.fr (J.-M. Torrenti), nicolas.roussel@ifsttar.fr (N. Roussel), loic.divet@ifsttar.fr (L. Divet),
jean-baptiste.despinose@espci.fr (J.-B. d'Espinose de Lacaillerie).

https://doi.org/10.1016/j.cemconres.2018.11.006
Received 15 May 2018; Received in revised form 7 November 2018; Accepted 8 November 2018
Available online 24 November 2018
0008-8846/ © 2018 Published by Elsevier Ltd.
R. Ragoug et al. Cement and Concrete Research 116 (2019) 134–145

It seems that two main processes are responsible for ESA: a physical Table 2
process of transfer of the outer solution to the porous media of the The main cement clinker phases according to Bogue's analysis.
cement matrix [19–23] and a chemical interaction between sulfate ions Cement Anhydrous phases (%)
having penetrated into the pore solution and the hydrates of the cement
paste [24]. The first process obviously depends on the transfer para- C3S C2S C3A C4AF
meters and the second one on the initial chemical composition of the
CEM I 65.2 8.8 7.9 8.9
material. As a consequence of the transfer of sulfate ions through the
interconnected capillary porosity, supersaturation conditions can be
achieved and the mineralogy of the cement paste drastically affected. during 24 h by using a slow rotation device shown in Fig. 1-a. This
As known in the hydration mechanisms of the cement paste, some hy- device limits bleeding and segregation during setting and produce
drates are sensitive to the presence of sulfate ions in solution [1,8]. homogenous samples especially for higher w/c ratios. After 24 h, all
Essentially, the balance of the calcium mono-sulfoaluminate (Afm) can specimens were removed from their moulds and cut into 5 cm thick
be very disturbed and can cause the re-precipitation of ettringite (Aft). cylinders. Then, they were cured in 200 L of non-renewed tap water for
However, the molar volume of AFt is about twice as large as that of one year (matured case). After this long curing period (a total hydration
AFm (707 cm3·mol−1 for 309 cm3·mol−1) [25,26]. Therefore, the pre- of cement is then assumed to be reached), samples were placed in the
cipitation of secondary ettringite in a hardened cement paste can im- sulfate solution tanks. Beforehand, an epoxy resin was applied on the
pose a significant internal pressure on the microstructure [26]. peripheral surface of the samples in order to insure unidirectional dif-
On the other hand, leaching phenomenon in neutral solution (e.g. fusion of sulfate ions (Fig. 1-b). Only one side of the specimens was in
seawater) is accelerated at the surfaces [27,28] and can affect the in- contact with the test solution (semi-immersion), which was prepared
gress of aggressive ions into the material. This phenomenon leads to with 15 g/L of sodium sulfate in deionized water (10 g/L of sulfate). The
dissolution of hydrated products particularly portlandite out of the surface area of samples to the solution volume ratio was 25. The pH of
material causing consequently a change in the microstructure allowing the sulfate solution was adjusted to 8.0 ± 0.1, as explained later using
an increase of the porosity and inevitably the transport properties in the the pH regulator. Finally, some samples were kept in 200 L of tap water
porous media [29–31]. It leads also to the formation of a layer of de- as reference cases.
posits at some micros of the sample surface. Therefore, it is necessary to
take this phenomenon into account for the structures exposed to the
ESA. 2.2. Accelerated test for ESA
This work aims to describe the results of the exposition of a mature
cement pastes with two w/c ratios (0.45 and 0.60), assumed to be to- As described above, the constant pH value of 8.0 ± 0.1 (pH lower
tally hydrated, to a sodium sulfate solution of constant pH 8.0 ± 0.1 than the one of the pore solution) of the cementitious materials appears
automatically controlled and under semi-immersion conditions. The to be a desirable feature of testing for ESA resistance, since it permits
objective of this work is, by mimicking in the laboratory exposure to simulation of field exposure conditions (e.g. seawater) and it allows
seawater, to characterize the possible altered zone and to identify the accelerating the sulfate degradation process [19]. Thus, the experi-
physical and chemical mechanisms of degradation under such condi- mental conditions provided in our proposed accelerated test method are
tions (case of bridge piers partially immersed in seawater). Here, the more representative of field conditions because the sulfate concentra-
results concerning ion transport, namely the transfer of sulfate ions into tion and the pH of the solution remain constant and similar to field
the cement matrix and calcium ions leaching in the surrounding solu- conditions. A schematic and a photo of the test apparatus are given in
tion are presented. Then, the samples altered or not, are characterized Fig. 2.
along the sulfate profile by a variety of investigation techniques such as The device consists of one big tank containing approximately 60 L of
XRD, ICP, NMR (27Al and 29Si), TGA and DTA. sulfate solution with a concentration of 15 g/L and acting as a reservoir.
A pump with a maximum flow rate of 750 L/h was immersed in this
tank in order to circulate the sulfate solution towards a series of smaller
2. Experimental study specimens' tanks containing 10 L each (Fig. 2-a and -b). At the output of
these tanks, the solution crosses a cylindrical tube (process 1) where the
2.1. Materials and exposure conditions pH of the solution is constantly monitored and regulated to 8.0 ± 0.1
by means of a continuous titration with sulfuric acid H2SO4 (0.05%)
Cement paste samples were obtained by hydration of CEM I (CEM I before it flows back into the reservoir tank (process 2). The added
52.5 N CE CP2), which satisfied the requirements of the European sulfuric acid is well mixed in the sulfate solution as this one is con-
Standard EN 197-1 [32], where the composition of the clinker is given tinuously moving between the sulfate reservoir and the specimens tanks
in Table 1. This composition was determined using TGA and ICP (process 3).
techniques. The clinker phase contents are calculated according to In the small tanks, semi-immersed samples (1 cm thickness was
Bogue's approach and shown in Table 2. This calculation highlights the immersed in the sulfate solution) of similar chemistry are arranged on a
high quantity of aluminates in C3A and C4AF favouring the formation of mesh. The volume ratio of the specimens to the sulfate solution in each
secondary ettringite after reaction with a high quantity of sulfate ions. tank is adjusted to disturb as little as possible the variation of the cir-
For the laboratory specimen's fabrication, cylindrical samples of culating sulfate concentration. This parameter remains approximately
10 cm diameter and 14 cm initial length were cast from cement pastes constant because of the weekly renewing of the solution. Moreover, the
prepared with two water-to-cement ratios (w/c equal to 0.45 and 0.6). use of sulfuric acid instead of another acid in the titration ensures that
Cement paste mixtures were cast into moulds maintained in rotation the sulfate concentration of the solution remains almost constant over

Table 1
Chemical composition of CEM I cement by using ICP-AES and TGA.
Constituents (%)

CaO SiO2 Al2O3 Fe2 O3 CaO (free) MgO SO3 S K2O Na2O Ignition loss

62.81 20.22 4.85 2.92 1.58 0.84 2.88 0 0.77 0.34 2.59

135
R. Ragoug et al. Cement and Concrete Research 116 (2019) 134–145

Fig. 1. Specimens used for the study (a) and a slow rotation device used during the setting (b).

time. ettringite, from those expanding in free volumes and well-crystallized


that causes no damage at the microstructure scale.
Furthermore, the X-ray diffraction (XRD) test was performed on
2.3. Experimental techniques for investigation
some samples exposed to the ESA. The X-ray patterns were obtained on
a D8 advance diffractometer from Bruker operating at 35 kV and
Several test methods were used in this study to characterize the
40 mA, using CoK (1.79 A) radiation at an angular step of 0.01 °C per
sound materials and to investigate the kinetics of sulfate ingress in the
second between 3.0° and 80° [2θ] angles. This technique is a reliable
porous media of the cement pastes and its consequences on their mi-
method for the identification of cement paste hydrates and especially
crostructures. Indeed, Inductively Coupled Plasma, ICP-AES was used to
crystallographic phases as portlandite and ettringite. At each mea-
quantify sulfate content and/or sulfate ingress into the samples. The
surement time, powders, previously drawn from exposed samples, are
output from this experimental method is the total sulfur concentration,
manually ground and sieved to obtain a uniform particle size less than
from which sulfate concentration is computed. In this study, specimens
80 μm. These powders are then analysed to draw the mineralogical
were grinded starting from the exposed surface. The obtained amount
profile and to identify the evolution of different mineralogical phases
of powder corresponding to a 1 mm thick sample displayed an average
depending on the type of exposure. Another method, also added in this
particle size of 315 μm. All sulfates present in this powder were ionized
investigation to confirm the results given by XRD is Nuclear Magnetic
by an acid attack. The resulting solution was then tested using ICP-AES.
Resonance (NMR) of 27Al and of 29Si. This technique aims at de-
Otherwise, thermogravimetric analysis (TGA) coupled to differ-
termining the quantity of aluminate and the change in C-S-H structure
ential thermal analysis (DTA) tests were performed by using a
in the cementitious materials and then, to deduce the quantity of dif-
METTLER TOLEDO TGA/DSC1 with a balance accuracy of 0.1 μg. These
ferent phases (AFm and Aft) in the cement after the spectra decom-
methods were used as techniques for quantitative characterization of
position.
cement pastes during the ESA. In this work, these techniques could
Finally, some visual observations of swelling, cracks or material loss
explain the mechanism of the AFt formation from other hydrates of
were carried out on samples after different exposure durations to the
cement.
sulfate solution.
Moreover, Environmental Scanning Electron Microscopy (Env. SEM,
Quanta 400 from FEI company) (SEM) was also used as a technique for
qualitative characterization of cement pastes to obtain information on
the spatial distribution of different cement paste phases. Even though
this technique provided only qualitative information, the SEM images
can distinguish the phases crystallizing in confined pores as compressed

Fig. 2. Schematic of pH-control for ESA accelerated test (a), Photo of the device (b).

136
R. Ragoug et al. Cement and Concrete Research 116 (2019) 134–145

Fig. 3. Sulfate profiles measured using ICP-AES, for different exposure duration to the sulfate solution.

3. Results and discussion in the cement matrix. In this coupled chemistry-transport process, dif-
fusion acts as a limiting phenomenon.
3.1. Physical aspects of degradation On the other hand, the transfer profiles are almost similar in terms
of penetration depth for cement pastes with the two w/c ratios. Indeed,
3.1.1. Sulfate profiles as reported earlier, the penetration front is about 4.5 mm (affected
Fig. 3 illustrates the effect of the w/c ratio on the sulfate ion zone) and only the amount of accumulated sulfate ions in the subsur-
transport profiles obtained using ICP-AES. Two clearly distinct areas are face zone [0–2 mm] differs between the two cement pastes. The pene-
highlighted. The first is characterized by a penetration front of sulfate tration front kinetics of the sulfate ions is rapidly slowed down after
ions corresponding to an affected zone of depth of about 4.5 mm. The 1 month of exposure to the surrounding solution for both cement pastes
second one is the zone beyond this front (sound zone) where the sulfate due to self-deceleration. Indeed, the chemical interaction with the hy-
concentration corresponds to the amount of sulfate present in the re- drates can induce the formation of sulfated products in the available
ference cement paste (of the order of 3% in g/g of anhydrous cement). porosity of the cement paste. This could slow the diffusion of ions into
The initiation of ESA occurred therefore by a transfer process of sulfate the interstitial solution as long as chemical fixation is possible [33,34].
ions into the cement matrix. In addition, the average sulfate content, This interaction occurs as long as the concentrations of calcium, alu-
from the surface (of the order of 14% in g/g of anhydrous cement) is minium and sulfate in the porosity allow it. Once these elements are
well above that of what one would expect from a simple equilibrium exhausted, the sulfates can penetrate to the following depths.
between the internal pore solution and the external sodium sulfate Fig. 4 compares the measured amounts of sulfates in the case of w/
solution (of the order of 0.4% in g/g of anhydrous cement). Thus, the c = 0.60 to the theoretical one assuming that all C3A and C4AF react
measured sulfate content does not reflect the sulfate in the pore solu- with sulfate to form ettringite (the reaction with dissolved calcium to
tion, but rather represents the solid phase content containing sulfate in form gypsum is neglected in this simulation) and considering the sulfate
the matrix (CS H2; C6AS 3H32; C3AS H12), as highlighted in previous stu- ions in the pore solution. In Fig. 4-a, only C3A is assumed to be totally
dies [31]. In other words, the content of physically or chemically bound consumed to form Aft. In this case, a maximum of 6% of sulfate in g/g
sulfates is higher than the free sulfates in the interstitial solution [33]. of cement paste could be chemically bound to form AFt. Since about 9%
Consequently, the energy involved in the fixation process is extremely was actually measured, this means that the amount of physically ad-
high. On the other hand, the limiting parameter of adsorption is the C/S sorbed sulfate is about 3% in g/g of cement paste or that the sulfate can
ratio more than the free sulfate content. This C/S ratio is directly af- be fixed in gypsum. Another explanation is that aluminium can come
fected by the leaching phenomenon that could have a significant effect from the C4AF phase and allow a chemical fixation of additional sulfate
on the ESA mechanisms. ions, up to about 9.5% g/g of cement paste (Fig. 4-b). This amount
Furthermore, the sulfate profiles are similar near the surface after corresponds to the maximum value measured by ICP on the surface.
one and two months of exposure. Thus, the important part of the However, this approximation depends on the local equilibrium and the
transfer of sulfate ions to the cement matrix occurs during the first thermodynamics of the system. In areas very close to the surface, the pH
month of exposure. Therefore, this observation suggests that the sulfate is low favouring the dissolution of the AFt and the likely formation of
fixation kinetics is faster than the one of the ion transfer in the porous the gypsum if the concentrations of calcium and sulfate ions allow it
network. Even if a diffusion process due to a concentration gradient is [9,35,36]. As already mentioned in this study, the pH is 8.0 ± 0.1 and
initially responsible for the penetration of sulfate ions into the ce- the ettringite could then decompose to form gypsum. Once the alumi-
mentitious matrix, the physicochemical interaction with the hydrates nate source is exhausted, ettringite formation stops, so sulfates may
seems to have higher kinetics. Consequently, ESA is characterized by react with calcium ions to form gypsum.
the coexistence of diffusion and physicochemical fixation of sulfate ions

137
R. Ragoug et al. Cement and Concrete Research 116 (2019) 134–145

Fig. 4. Different sulfate contents of cement paste with w/c = 0.60 after 1 month of exposure to ESA: (a) reaction of only C3A (b) reaction of both C3A and C4AF.

3.1.2. Calcium leaching


In order to quantify the leached calcium in the solution before its
renewal (each week), ICP measurements of calcium concentration in
the baths were performed (Fig. 5-a). The calcium leaching is almost
similar for the two cases of cement pastes. The shape of the curves
reflects a diffusive phenomenon as the concentration of leached calcium
is a function of the square root of time ([Ca2 +] = a· t + b ) as found in
[31]. However, there are two diffusive regimes with different kinetics
for the case of cement pastes with w/c = 0.60: slower and then higher
kinetics. This increase of leaching kinetics can be explained by the high
quantity of calcium ions coming from the CH, the decalcification of C-S-
H or possibly dissolved gypsum. Therefore, the increase of calcium
leaching is favoured by the low value of pH of the exposure solution.
Indeed, this value is lower than that of the interstitial solution equili-
brium which causes dissolution of CH leading to the increase of Ca2+
and OH− ions in the surrounding solution.
On the other hand, Fig. 5-b presents the amount of leached calcium
Fig. 6. CaO content for different depths after 2 months of sulfate exposure.
in solution as a function of the total measured sulfate content for the
case of w/c = 0.60. The curve shows that the value measured at the
surface of about 9% in g/g of cement paste (14% in g/g of anhydrous after 2 months of exposure to the sodium sulfate solution, is measured
cement) corresponds to a saturation value of sulfate ions. Indeed, at this for the first 6 mm of depth. The results presented in Fig. 6 show a de-
value, only the calcium leaching increases without additional fixation crease of the CaO content for the first 4 mm. This result is in line with
process of sulfate. the sulfate front shown in Fig. 3. However, the CaO content for the first
Furthermore, CaO content of the cement paste with 0.60 w/c ratio, depth (0–1 mm), although lower than the core content of the sample,
stays higher than the content for the other “affected” depths. This

Fig. 5. Variation of leached calcium vs. (a) time and (b) sulfate concentration at the surface.

138
R. Ragoug et al. Cement and Concrete Research 116 (2019) 134–145

Fig. 7. Degradation of cement paste with w/c = 0.60 by ESA: (a) after 2 months, (b–c) after 3 months of exposure to sulfate solution.

finding may be explained by the possible interaction between the cal- related to the process of interaction of the microstructure with sulfate
cium and sulfate ions at the surface to form gypsum and AFt, which ions. Here, the pressure exerted by AFt and/or gypsum would be dif-
leads to an overconcentration of calcium ions necessary for these re- ferent, depending on where the crystal exerting this pressure is formed
actions. This may also explain the increase in leaching at this time, (in which range of porosity).
since an additional concentration of calcium ions is needed to interact To explain the mechanisms of the degradations of cementitious
with the surplus sulfate ions at the surface. materials by ESA illustrated in Figs. 7 and 8, different theories exist: (i)
the molar volumes of any precipitates (gypsum and AFt) forming in a
hardened paste are about twice as large as the molar volumes of the
3.1.3. Macroscopic degradation initial phases (portlandite and AFm). This simplistic theory implies that
Fig. 7 illustrates the degradation observed in the case of cement a material with low porosity should degrade more quickly (less space
paste with w/c = 0.60. Longitudinal and transverse cracks appeared available to contain the extra volume of hydrates). A comparison of the
after just two months of exposure to the sodium sulfate solution (Fig. 7- observed degradations on the two cases of cement paste in our study
a). Then after three months of testing, the cement paste has lost all eliminates this hypothesis. (ii) The second theory is based on the
cohesion and the damage is total (Fig. 7-b and -c). Degradation oc- principle of crystallization pressure [37]. According to this theory, two
curred at the periphery of the cement paste cylinder while the core conditions are necessary to induce mechanisms of expansion within the
remained unaffected. The appearance of transverse cracks is caused by hardened material. The first is that the crystal develops from a super-
significant tensile stresses that are rather at the origin of this de- saturated solution resulting in sufficient energy to induce expansion.
gradation. The second condition obeying the mechanics of porous media is that the
Otherwise, some visual monitoring were carried out on cement crystal develops in a confined space (small enough pore size), and that
paste with w/c = 0.45 (Fig. 8). Degradation is also observed but is it exerts an expansive strength in a pore size range fine enough to in-
delayed in time. Indeed, the beginning of the degradation is observed duce higher stresses than ultimate tensile strength.
with the appearance of some cracks and swelling at the edge of the Therefore, the precipitated ettringite or gypsum content resulting
sample after 6 months of testing. Then, the cracking is followed by a from the sulfate ion supply is not the only cause of the degradation
chipping causing a significant and rapid loss of material. After 8 months observed on cement pastes. This seems to be related to the range of pore
of exposure, the cement paste is totally damaged (Fig. 8-c and -d). Thus, and the energy released by the development of the precipitated mineral
the degradation is faster in the case of cement paste with a higher w/c to explain the observed cracks.
ratio. What is more striking in these observations is the different me-
chanical behaviour of the two cement pastes. This difference suggests
that the mechanical response of the material occurs differently de-
pending on the porosity of the material. The behaviour is possibly

139
R. Ragoug et al. Cement and Concrete Research 116 (2019) 134–145

Fig. 8. Degradation of cement paste with w/c = 0.45 by ESA: (a) after 2 months, (b) after 6 months, (c–d) after 8 months of exposure to sulfate solution.

3.2. Chemical aspects of degradation of water loss for CSH and AFt (maximum respective endothermic peaks
of DTA, at 144.7 °C and 140.7 °C), an important difference is measured
3.2.1. Characterization of sound materials between the two curing times. Indeed, about 8% increase in this loss of
As previously described, some investigation techniques were used to water is measured after 1 year of curing. This increase may be caused by
determine the effect of the long curing in water on the reference ma- both a decalcification of the C-S-H and an increase in the AFt content at
terial. TGA was used to quantify the possible surface leaching while the the surface. Otherwise, the mass losses relative to the AFm phases
29
Si NMR was performed to determine the degree of hydration and (maximum endothermic peaks of DTA, respectively 199.1 °C and
structure of the C-S-H chains. Moreover, XRD and 27Al NMR tests are 198.7 °C) seem to have a slight difference of 0.17%, which might sug-
also carried out for a more complete monitoring of the chemical com- gest that the large increase in water loss corresponding to (CSH + AFt)
position of the material. is rather driven by decalcification of C-S-H. Finally, no change in the
Fig. 9 presents the results given by XRD analysis for cement paste mass loss related to the calcite (endothermic peaks of 807.8 °C and
with 0.60 w/c ratio after one month and one year of curing in water. 814.5 °C respectively). Indeed, this mass loss is about 2.61% and 2.64%
Investigations using this technique were performed on three depths: between 1 month and 1 year of curing in water.
0–1 mm, 2–3 mm and 4–5 mm. The main crystallized phase is the por- Furthermore, low pH and sulfate ingress could affect C-S-H by
tlandite whose peaks become more intense with depth. Traces of et- changing the C/S ratios and chain lengths [38]. It is therefore necessary
tringite are also detected with an additional presence of peaks after to quantify the C-S-H in order to observe the evolution of the cement
1 year of curing in water. Moreover, some traces of calcite are high- matrix before and after contamination by the aggressive solution. The
lighted for all the analysed depths, resulting from low carbonation of results of TGA/DTA exposed above, provide qualitative information on
the samples despite the precautions taken to avoid it. Finally, a poorly this evolution of the C-S-H. However, it is necessary to know before-
defined peak appeared for all the powders tested, corresponding to the hand the quantities of ettringite that can be formed by the calcium of
presence of C-S-H. the C-S-H for a quantitative synthesis. It is for this reason that the 29Si
On the other hand, TGA-DTA curves obtained for the sound cement NMR is particularly suitable in this case of study for characterizing and
paste after one month and one year of curing in water are plotted in quantifying the C-S-H phase [39]. Indeed, this technique makes it
Fig. 10. These techniques give water losses for C-S-H + AFt, AFm, CH possible to determine the chain length change in the C-S-H, their
and CC for cementitious materials. The results show a slight difference quantity and the degree of hydration of the anhydrous phases. In ad-
in the thermal losses (about 0.76%) between the two curing times. dition, the proportion of aluminium incorporated into the C-S-H can be
Thus, dissolution of portlandite occurs between 1 month and 1 year of also detected by 27Al NMR [39]. Then, by combining these techniques,
curing. However, this dissolution is negligible and can be explained by it will be possible to calculate the mass percentage of C-S-H in cement
the long duration of the curing in water. On the other hand, in the case paste.

140
R. Ragoug et al. Cement and Concrete Research 116 (2019) 134–145

Fig. 9. XRD analysis for cement paste with 0.60 w/c ratio: (a) one month of curing, (b) one year of curing.

The hydration rate of the anhydrous phases concerns only the increase suggests a change in the structure and chemical balance of C-S-
progress of the formation of CSH (represented by Q1 and Q2 in 29Si H. This is consistent with the observations of the 29Si NMR results.
NMR tests) resulting from the anhydrous phases (represented by Q0). Indeed, the decalcification of C-S-H causes an increase in the level of
The 29Si NMR results, shown in Fig. 11-a and concern only the first calcium in solution. Therefore, the C/S ratio and the C/Al ratio are
depth of the sample (0–1 mm) taken from the surface. The evolution of modified as well as the C-S-H structure.
the CSH structure (Q1 + Q2) after 24 h of setting, after 3 months and
after 1 year of curing is presented by the various 29Si NMR spectra. The 3.2.2. Chemical fixation of sulfate ions
hydration of the cement paste is almost complete after 1 year of curing XRD analysis was coupled with 27Al NMR to follow the evolution of
in water (at 3 months a peak relative to anhydrous Q0 is no longer the aluminate phases after 2 months of exposure to the sodium sulfate
present). The structure of the C-S-H shows an increase in the peak of the solution (Fig. 12). The results show a change in the mineralogical
Q2 chains (long chains) due to a condensation of these hydrates after composition of the cement matrix. In comparison with the sound ma-
curing in water. This result confirms a decalcification of C-S-H by a long terial, an increase in the intensities of the AFt peaks is highlighted.
curing in water, which is in line with the results of TGA/DTA observed Ettringite is thus the main crystallized phase. On the other hand, the
previously. This increase in long chains is reflected in the chemical peaks corresponding to the portlandite have their intensity greatly re-
equilibrium by a decrease in the C/S ratio in the stoichiometry of C-S-H duced. This suggests its dissolution in the interstitial solution leading to
[38]. the formation of AFt. In addition, a peak corresponding to gypsum
The results of DRX and TGA/DTA exposed above provide informa- appeared confirming the coexistence of gypsum and ettringite in the
tion on the presence of aluminate phases. However, these techniques do first zone of the material from the exposed surface. These results are
not make it possible to quantify the distribution of these phases. That is confirmed by 27Al NMR where the spectrum shows the disappearance of
why the 27Al NMR was used. the AFm peak and the appearance of the AFt one. This confirms that
The evolution of the proportions of the aluminate phases is de- ettringite is formed after dissolution of AFm given the availability of
scribed in Fig. 11-b. In general, the hydration of the CEM I gives a calcium and sulfate with adequate concentrations in the solution
majority of AFm phases. The AFt rate decreases with curing time in (Fig. 13).
favour of AFm formation and therefore the AFt/AFm ratio decreases Fig. 14 shows the comparison of the microstructure between a
over time. Silicon-substituted aluminium in C-S-H (Al (IV)) and cal- sound and affected cement paste by ESA given by SEM technique. The
cium-substituted aluminium in C-S-H are rebalanced over the course of application of this investigation method confirms the coexistence of
the water curing with a slight increase after 1 year of curing. This gypsum and ettringite in the affected samples as detected by other

Fig. 10. TGA-DTA analysis for cement paste with 0.60 w/c ratio: (a) one month of curing, (b) one year of curing.

141
R. Ragoug et al. Cement and Concrete Research 116 (2019) 134–145

29 27
Fig. 11. NMR spectra for cement paste with 0.60 w/c ratio: (a) Si NMR, (b) Al NMR.

Fig. 12. Determination of phases newly formed by NMR (a) and XRD (b) after exposure of cement paste samples to the sulfate solution.

27
Fig. 13. Al NMR spectrum of the AFm and Aft content during water curing and sulfate exposure.

methods. Moreover, the observations confirm the precipitation of the low pH of the surrounding solution, is an important process in the ESA
pre-stressed AFt in the pores leading to the expansion of the material study. Indeed, this leaching contributes to changes in the local ther-
and the creation of some micro-cracks in the cement matrix (see Fig. 14- modynamic equilibrium of the cement matrix. These solid phase
b). changes are also observed according to the XRD analysis for the sound
cement paste. Fig. 15-a presents the DTA curves for reference and af-
fected samples at the surface zone. Portlandite disappears after
3.2.3. Dissolution of CH in presence of sulfate ions 2 months of sodium sulfate exposure (Fig. 15-b). On the other hand, the
The leaching of calcium ions from the cement paste to the sur- endothermic peak corresponding to the chemically bound water of C-S-
rounding solution, due to the calcium concentration gradient and the

142
R. Ragoug et al. Cement and Concrete Research 116 (2019) 134–145

Fig. 14. SEM images: (a) sound cement matrix; (b) ESA-affected cement matrix [(1): Pre-stressed Aft, (2): Micro-cracks].

Fig. 15. Quantification of hydrates during ESA: (a) DTA characterization, (b) CH content.

Fig. 16. Changes of C-S-H structure investigated by NMR: (a) spectrum, (b) average lengths of C-S-H chains.

H and AFt increases, highlighting the decalcification of C-S-H and also reference and the exposed samples to the sodium sulfate solution. A
the increase of the AFt content. In addition, the DTA curve has a decrease of the peak of short chains Q1 and an increase of that of the
shoulder around 200 °C confirming the presence of AFm in the case of long chains Q2 are highlighted. This change therefore highlights a
sample cured in water. This shoulder disappears after 2 months of modification of the C-S-H structures. The average length of the C-S-H
contact with the sodium sulfate solution, which confirms the XRD and chains increases to about the value of 10 leading to polycondensation in
NMR analysis (Figs. 12 and 13). This finding also asserts that AFt's the C-S-H structure (Fig. 16-b). This fact is also characteristic of acid
formation mechanism is carried out in part from AFm. As illustrated in attacks on cement pastes, as shown in [34,40]. Indeed, the dec-
Fig. 15-b, a progressive consumption of calcium ions from the dis- alcification causes a decrease of the C/S ratio and therefore an increase
solution of the CH over the depth of the sulfate ingress is highlighted. in the length of the C-S-H chains. This effect is not observed in the case
of exposure to sodium chloride solution [41].
Furthermore, the solid phase's distribution is given in Fig. 17. As
3.2.4. Change of C-S-H structure during ESA
clearly highlighted, the AFt dominates the aluminates phases (total
The 29Si NMR spectra are shown in Fig. 16-a for both cases: the

143
R. Ragoug et al. Cement and Concrete Research 116 (2019) 134–145

Fig. 17. Solid phase's distribution before and after sulfate exposure.

disappearance of AFm). Indeed, the mass content of ettringite is [8] M. Santhanam, et al., Mechanism of sulfate attack: a fresh look part 2. Proposed
therefore three times higher after 2 months of samples exposure to the mechanisms, Cem. Concr. Res. 33 (2003) 341–346.
[9] E.F. Irassar, et al., Microstructural study of sulfate attack on ordinary and limestone
sodium sulfate solution. On the other hand, the CH content becomes portland cements at ambient temperature, Cem. Concr. Res. 33 (2003) 31–41.
zero with a 12% increase in the “other” phases that would correspond [10] C. Yu, K. Scrivener, Mechanism of expansion of mortars immersed in sodium sul-
to the transformation of portlandite to form gypsum. phate solution, Cem. Concr. Res. 43 (2013) 105–111.
[11] C. Ayora, et al., Weathering of iron sulfides and concrete alteration: thermodynamic
model and observation in dams from central Pyrenees, Spain, Cem. Concr. Res. 28
4. Conclusions (1998) 591–603.
[12] J.S. Chinchon, et al., Influence of weathering of iron sulfides contained in ag-
gregates on concrete durability, Cem. Concr. Res. 25 (1995) 1264–1272.
- The analysis of the physical aspect of ESA shows that two modes of [13] P.K. Mehta, Mechanism of expansion associated with ettringite formation, Cem.
transfer occur. Sulfate ion transfer to the cement matrix and Concr. Res. 3 (1973) 1–6.
leaching of calcium ions to the outer solution which deposits on the [14] J. Marchand, J. Scalny, Materials of Concrete: Sulfate Attack Mechanisms,
American Ceramic Society (1999).
surface. Moreover, this phenomenon is characterized by the coex-
[15] Ch. Xiong, L. Jiang, Y. Xu, H. Chu, Y. Zhang, Deterioration of pastes exposed to
istence of diffusion and physicochemical fixation of sulfate ions in leaching, external sulfate attack and dual actions, Constr. Build. Mater. 116 (2016)
the cement matrix. In this coupled chemistry-transport process, 52–62.
diffusion acts as a pilot phenomenon. Indeed, a concentration gra- [16] P.K. Mehta, S. Wang, Expansion of ettringite by water adsorption, Cem. Concr. Res.
12 (1) (1982) 121–122.
dient transfer occurs initially. Then, the sulfate ion transfer is ra- [17] W. Kunther, B. Lothenbach, K.L. Scrivener, On the relevance of volume increase for
pidly stabilized and only surface accumulation of sulfate ions is the length changes of mortar bars in sulfate solutions, Cem. Concr. Res. 46 (2013)
observed, suggesting that fixation occurs causing self-deceleration of 23–29.
[18] B. Lothenbach, B. Bary, P. Le Bescop, T. Schmidt, N. Leterrier, Sulfate ingress in
the ingress as long as the sulfates can react with the cement hy- Portland cement, Cem. Concr. Res. 40 (2010) 1211–1225.
drates. [19] D. Bonen, M.D. Cohen, Magnesium sulfate attack on Portland cement paste-I.
- On the other hand, the analysis of the chemical composition made it Microstrcutyure analysis, Cem. Concr. Res. 22 (1) (1992) 707–718.
[20] D. Bonen, M.D. Cohen, Magnesium sulfate attack on Portland cement paste-II.
possible to highlight the consumption of the surface portlandite, Microstrcutyure analysis, Cem. Concr. Res. 22 (1) (1992) 169–180.
until a total consumption after 2 months of exposure to the sulfate [21] D. Planel, J. Sercombe, P. Le Bescop, F. Adenot, J.-M. Torrenti, Long-term perfor-
solution. In addition, the decalcification of C-S-H with highly sig- mance of cement paste during combined calcium leaching-sulfate attack: kinetics
and size effect, Cem. Concr. Res. 36 (2006) 137–143.
nificant chain extension is confirmed by the 29Si NMR results. As for [22] R.S. Gollop, H.F.W. Taylor, Microstructural and microanalitical studies of sulfate
ettringite, it is formed from both AFm and aluminium incorporated attack III. Sulfate-resisting Portland cement: reaction with sodium and magnesium
in C-S-H. sulfate solutions, Cem. Concr. Res. 25 (7) (1995) 1581–1590.
[23] P.K. Mehta, Concrete: Structure, Properties and Materials, Prentice-Hall, 1986, pp.
- Finally, in the case of low pH exposure, the simultaneous formation
105–169.
of gypsum and ettringite can be highlighted. The relative propor- [24] S. Sarkar, S. Mahadevan, J.C.L. Van der Sloot, H.D.S. Kosson, Numerical simulation
tions of those hydrates are a function of both the free sulfate con- of cementitious materials degradation under external sulfate attack, Cem Concr
centration in the pore solution and aluminates and calcium contents. Com 32 (3) (2010) 241–252.
[25] T. Matschei, B. Lothenbach, F.P. Glasser, The Distribution of Sulfate in Hydrated
Portland Cement Paste, 12th International Congress on the Chemistry of Cement,
Acknowledgements Montréal, Canada, (2007).
[26] W.G. Hime, B. Mather, Sulfate attack, or is it? Cem. Concr. Res. 29 (1999) 789–791.
[27] C. Carde, R. François, Modelling the loss of strength and porosity increase due to the
The authors would like to thank the members of Bouygues Society leaching of cement pastes, Cem. Concr. Compos. 21 (1999) 181–188.
for its financial and technical support. [28] M. Mainguy, C. Tognazzi, J.-M. Torrenti, F. Adenot, Modelling of leaching in pure
cement paste and mortar, Cem. Concr. Res. 30 (2000) 83–90.
[29] F. Adenot, M. Buil, Modelling of the corrosion of the cement paste by deionized
References water, Cem. Concr. Res. 22 (1992) 259–272.
[30] S. Kamali, B. Gérard, M. Moranville, Modelling the leaching kinetics of cement
[1] A. Neville, The confused world of sulfate attack on concrete, Cem. Concr. Res. 34 based materials — influence of materials and environment, Cem. Concr. Compos.
(2004) 1275–1296. 25 (2003) 451–458.
[2] P. Feng, E. Garboczi, C. Miao, J.W. Bullard, Microstructural origins of cement paste [31] E. Rozière, A. Loukili, R. El Hachem, F. Grondin, Durability of concrete exposed to
degradation by external sulfate attack, Constr. Build. Mater. 96 (2015) 391–403. leaching and external sulphate attacks, Cem. Concr. Res. 39 (12) (2009)
[3] P.J.M. Monteiro, Scaling and saturation laws for the expansion of concrete exposed 1188–1198.
to sulfate attack, Proc. Natl. Acad. Sci. U. S. A. 103 (31) (2006) 11467–11472. [32] EN 197, Cement – Part 1: Composition, Specifications and Conformity Criteria for
[4] M. Santhanam, et al., Effects of gypsum formation on the performance of cement Common Cements, European Standards, Brussels, (2011).
mortars during external sulfate attack, Cem. Concr. Res. 33 (2003) 325–332. [33] C. Yu, W. Sun, K. Scrivener, Degradation mechanism of slag blended mortars im-
[5] N.M. Al-Akhras, Durability of metakaolin concrete to sulfate attack, Cem. Concr. mersed in sodium sulfate solution, Cem. Concr. Res. 72 (2015) 37–47.
Res. 36 (2006) 1727–1734. [34] T. Schmidt, B. Lothenbach, M. Romer, J. Neuenschwander, K. Scrivener, Physical
[6] S.L. Sarkar, et al., Numerical simulation of cementitious materials degradation and microstructural aspects of sulfate attack on ordinary and limestone blended
under external sulfate attack, Cem. Concr. Compos. 32 (2010) 241–252. Portland cements, Cem. Concr. Res. 39 (2009) 1111–1121.
[7] P.J.M. Monteiro, K.E. Kurtis, Time to failure for concrete exposed to severe sulfate [35] P.K. Mehta, D. Pirtz, M. Polivka, Properties of alite cements, Cem. Concr. Res. 9
attack, Cem. Concr. Res. 33 (2003) 987–993. (1979) 439–450.
[36] R.P. Khatri, V. Sirivivatnanon, J.L. Yang, Role of permeability in sulfate attack,

144
R. Ragoug et al. Cement and Concrete Research 116 (2019) 134–145

Cem. Concr. Res. 27 (8) (1997) 1179–1189. Am. Ceram. 76 (1993) 2285–2288.
[37] X. Ping, J. Beaudoin, Mechanism of sulfate expansion. I-Thermodynamic principle [40] R. El-Hachem, E. Rozière, F. Grondin, A. Loukili, Multi-criteria analysis of the
of crystallization pressure, Cem. Concr. Res. 22 (1992) 631–640. mechanism of degradation of Portland cement based mortars exposed to external
[38] P. Faucon, Durabilité des bétons: physico-chimie de l'altération par l'eau (PhD sulfate attack, Cem. Concr. Res. 42 (2012) 1327–1335.
Thesis), University of Cergy Pontoise, 1997. [41] M. Saillio, Interactions physique et chimiques ions-matrice dans les bétons sains et
[39] I. Richardson, A. Brough, R. Brydson, G.W. Roves, C.M. Dobson, Location of alu- carbonatés, Influence sur le transport ionique (PhD Thesis), University of Paris-Est-
minium in substitued CSH gels as determined by 29Si et 27Al NMR and EELS, J. Marne-La-Vallée, 2012.

145

Вам также может понравиться