You are on page 1of 12

ACI MATERIALS JOURNAL TECHNICAL PAPER

Title No. 113-M44

Preconditioning Method for Accelerated Testing of


Concrete under Sulfate Attack
by Hocine Siad, Mohamed Lachemi, Mustafa Sahmaran,
¸ and Khandaker M. Anwar Hossain

This paper aims to recommend a new supplementary precondi- play important roles in any transport process in concrete.
tioning procedure that can accelerate the degradation process of Additionally, varying types of concrete require that sulfate
concrete specimens when exposed to sulfate attack. Several spec- resistance testing be performed on concrete specimens, not
imens obtained from 30 and 60 MPa (4.4 and 8.7 ksi) strength the equivalent mortar or paste specimens. Although concrete
concretes were tested. Expansion, mass change, and ultrasonic
deterioration due to sulfate ions has been widely studied, the
pulse velocity measurements were conducted for specimens
mechanism of degradation by sulfate attack and the long-
subjected to two primary methods of presaturation using sulfate
solution and desiccators, then immersed in sulfate solution. A term duration of laboratory tests remain real obstacles to
complementary microstructural analysis was also carried out to emphasizing the difference in concrete composition. The
define the degradation mechanism. When compared to the control ASTM C1012 test, performed on mortar specimens with
method, which had the same immersion conditions without primary dimensions of 25.4 x 25.4 x 285 mm (1 x 1 x 11.22 in.),
presaturation, test results showed an important acceleration of the takes 12 to 18 months of immersion in 5% sodium sulfate
degradation process with the use of supplementary preconditioning solution, which means that testing on concrete samples may
in desiccators. In addition to the remarkable acceleration in the require a very long period of time (in regards to minimum
degradation process, the microstructure study confirmed that the dimensions required for concrete samples).
degradation mechanism of concrete specimens subjected to supple- There are examples (not exhaustive) of proposed methods
mentary preconditioning by desiccators reflected the field observa-
that attempt to deal with the problem of testing time in sulfate
tions in sulfate attack environments.
attack by accelerating the degradation process of concrete.
Keywords: concrete degradation; expansion; microstructure; sulfate attack; Based on ASTM procedures, Mehta12 proposed automati-
test method. cally controlling the PH of the solution rather than renewing it
weekly. However, the complexity of the test setup has inhib-
INTRODUCTION ited its use as a standard method.13 Several subsequent accel-
Sulfate attack can be classified as an aggressive natural erated test methods rely on: 1) a high sulfate concentration14
threat to concrete structures due to the large presence of sulfate (instead of the 5% suggested in ASTM C1012); 2) dry/wet
in soil, groundwater, seawater, and industrial effluents.1-3 It is cycles15 or partial immersion16 as a replacement for full
classified as a “very severe attack” in the ACI 318-99 building immersion; or 3) storage at high temperatures17 instead of
code,4 and research shows that concrete is susceptible to 25oC (77°F), as proposed by ASTM International. Due to the
ingress by sulfate ions from the environment. The interaction change in the degradation mechanism that does not reflect
between concrete and sulfate ions usually produces gypsum the field observations, these test methods have all undergone
and ettringite. The rapid precipitation of ettringite in concrete numerous criticisms and have been deemed inadequate for
can cause volume change, cracking, disintegration, and assessing concrete resistance to sulfate. According to Bell-
strength loss, which can lead to structural failure. mann et al.9 and De Belie,18 in high sulfate concentrations,
Degradation of concrete due to external sulfate attack gypsum is the main product, whereas under realistic field
progresses slowly in the field, occurring over decades.5 The conditions, the primary product is ettringite or a combina-
main challenges in assessing concrete performance are simu- tion of gypsum and ettringite. For wet/dry cycles or partial
lating natural environmental conditions in the laboratory, immersion methods, Clifton et al.19 showed that concrete
and limiting testing time to the minimum. The absence of degradation has a physical process due to the concentra-
standard accelerated procedures for testing the vulnerability tion of sulfate salts on outer concrete surfaces. Assaad et
of concrete samples to sulfate attack has been observed in al.16 observed that the partial immersion test can be better
the literature. Most traditional accelerated methods, based used to determine resistance against corrosion due to sulfate
on the ASTM C10126 standard, evaluate external sulfate attack. For the method that employs storage at high tempera-
attack on cementitious materials using mortar or paste spec- tures, Damidot and Glasser20 showed that, at 25°C (77°F),
imens. Regardless of the criticisms of this method made ettringite and gypsum are stable, and monosulfoaluminate is
by researchers,7-10 tests involving cement paste or mortar
samples are not necessarily good indicators on concrete ACI Materials Journal, V. 113, No. 4, July-August 2016.
MS No. M-2015-298.R1, doi: 10.14359/51688705, received September 1, 2015, and
behavior. According to Crumbie11 and Chabrelie,5 both reviewed under Institute publication policies. Copyright © 2016, American Concrete
Institute. All rights reserved, including the making of copies unless permission is
the interface between aggregates and cement paste (the obtained from the copyright proprietors. Pertinent discussion including author’s
so-called ITZ), and the nature and content of the aggregate, closure, if any, will be published ten months from this journal’s date if the discussion
is received within four months of the paper’s print publication.

ACI Materials Journal/July-August 2016 493


a metastable phase. Increased temperature drastically influ- energy-dispersive X-ray spectroscopy (EDS) and micro­
ences the CaO-A12O3-CaSO4-H2O system; at temperatures analyses on polished surfaces.
above 45°C (113°F), monosulfoaluminate becomes increas-
ingly stable at the expense of ettringite. Other examples of RESEARCH SIGNIFICANCE
proposed methods of accelerating sulfate attack can be also Most laboratory methods aimed at studying the sulfate
cited,21-23 however, the minimum recommended time may attack mechanism are based on accelerated mortar tests.
exceed 2 years for conservation and monitoring. However, using these tests to evaluate concrete perfor-
A new testing procedure needs to be developed to accel- mance is impractical because deterioration will not occur in
erate testing of the resistance of concrete samples against concrete within a reasonable period of time, as in the mortar
sulfate attack. This may be achieved by accelerating the sample. The aim of the new test method is to develop an
degradation of concrete samples. The objective of this study accelerated and reproducible procedure for monitoring the
is to suggest a faster and more realistic procedure to accel- resistance of different types of concrete to external sulfate
erate the rate of degradation and, therefore, test duration, to attack. The test results showed that an important acceleration
limit testing time to a few months while depicting long-term of the degradation process of concrete exposed to external
influence. To develop a realistic test procedure, the test setup sulfate attack occurred with supplementary pre-conditioning
must be simple and the required equipment must be econom- in desiccators under vacuum.
ical and accessible to most material laboratories. The experi-
mental study conducted in this work was based on the accel- EXPERIMENTAL PROGRAM
eration of the diffusion of sulfate ions into concrete. Two Materials, mixture characteristics, and specimen
new methods of pre-conditioning were proposed and tested preparation
based on preliminary saturation by sulfate solution. Using General-use (GU) portland cement complying with ASTM
desiccators and under vacuum, sulfate ions were forced C150 Type I cement24 and Type High Early (HE) cement
to migrate through the concrete specimens. The reliability conforming with Canadian Standards in CSA A3000-0825
of the proposed methods was evaluated by monitoring the were used to produce normal and high-strength concrete
expansion, mass change, and ultrasonic pulse velocity of (30  and 60 MPa [4.4 and 8.7 ksi] compressive strength),
normal and high-strength concretes (30 and 60 MPa [4.4 and respectively. Siliceous natural sand with a maximum grain
8.7 ksi] compressive strength). Mineralogical, chemical, size of 5 mm (0.2 in.) and silico-calcareous crushed gravel
and microstructural studies were also carried out on several with 15 mm (0.59 in.) maximum size were used as fine and
layers of degraded samples using X-ray diffraction (XRD) coarse aggregates. Specific gravity and water absorption
and scanning electronic microscopy (SEM) coupled with properties were 2.65% and 0.8%, and 2.7 and 0.5%, respec-
tively. Silica fume (SF) was used as a mineral admixture
Table 1—Chemical composition and physical in the production of the 60 MPa (8.7 ksi) concrete. Table 1
properties of cementitious materials describes the physical and chemical characteristics of
Chemical component, % GU HE Silica fume cements and silica fume, and Table 2 provides the mixture
SiO2 19.2 21.2 95.6 proportions of the two concretes used in this study.
For each mixture (Mix 1 and Mix 2), Ø100 x 200 mm
Fe2O3 3.13 3.4 0.07
(3.94 x 7.87 in) (diameter x length) cylindrical specimens
Al2O3 4.9 5.6 0.14 were used to determine compressive strength. Prismatic
CaO 62.3 64.6 0.38 75 x 75 x 285 mm (2.95 x 2.95 x 11.22 in.) specimens were
MgO 2.7 2.1 0.29 prepared to measure expansion, mass change, and ultrasonic
pulse velocity before and after exposure to sulfate attack,
SO3 3.73 1.20 0.26
and the average values from three identical specimens were
K 2O 0.54 2.51 0.53 used. All specimens were demolded after 24 hours and cured
Na2O 0.24 0.5 0.15 at 21°C (69.8°F) and 95% relative humidity. After 28 days,
Physical properties the initial compressive strength, mass, length, and ultrasonic
pulse velocity of the specimens were determined before
Specific gravity 3.15 3.15 2.35
exposure to sulfate attack.
Blaine fineness, m /kg2
411 490 620

Note: 1 m /kg = 0.54 yd /lb.


2 2

Table 2—Mixture proportions of investigated concretes


Composition, kg/m3 Cement Silica fume Sand Gravel Water HRWRA* w/b† 28-day compressive strength, MPa
Mixture ID
Mix 1 285 — 930 930 172 — 0.60 31.1
Mix 2 415 25 890 810 171 5 0.39 60.4
*
High-range water-reducing admixture.

Water-binder (cement + silica fume) ratio.
Notes: 1 kg/m3 = 1.685 lb/yd3; 1 MPa = 0.145 ksi.

494 ACI Materials Journal/July-August 2016


Fig. 1—Vacuum desiccator used for Method A and Method B preconditioning.
Proposed accelerated preconditioning methods imately 90 mm [3.54 in.]) in the first stage (Fig. 1(b)). The
Degradation of concrete by external sulfate ions involves vacuum pump was kept running for an additional 24 hours.
both physical and chemical aspects. The physical aspect is The sulfate solution was then drained slowly into the desic-
related to the ingress of sulfate ions into concrete pores, cator container for the second time to cover two-thirds of the
especially by diffusion. The influence of the chemical aspect length (around 180 mm [7.09 in.]), and the vacuum was kept
depends on the profile of penetration of sulfate ions.26 running for another 24 hours (Fig. 1(c)). The sulfate solution
Based on the acceleration of the diffusion process of sulfate was drained into the desiccator for the third time to cover the
ions into concrete, two methods of preconditioning were specimens’ length completely (Fig. 1(d)). The vacuum pump
proposed and tested, and sulfate ions were forced to migrate was run for an additional 24 hours for maximum penetration
through the concrete specimens via preliminary saturation of sulfate solution into the concrete prisms.
by vacuum desiccators. After applying the new methods In the control method, no special preconditioning was
of supplementary preconditioning, all concrete specimens applied to the concrete specimens. After 28 days of curing,
were immersed in sulfate solution for the duration of the the samples were immersed directly in the 5% sodium
test. To facilitate the presentation and discussion of results, sulfate solution.
preconditioning procedures were divided into Method A, After supplementary preconditioning using the control
Method B, and the control method. method, Method A, and Method B, testing was followed by
In Methods A and B, concrete prisms were dried in an immersing concrete prisms in tanks containing 5% sodium
oven after 28 days of curing. To avoid pore structure damage sulfate solution at a temperature of 23 to 27°C (73.4 to
that can be caused by the drying procedure,27 a 50°C (122°F) 80.6°F). The volume of solution used was calculated as
drying temperature was used up to a constant weight, usually five times the concrete prisms’ volume for all concretes.
between 1 and 3 weeks.28 After the oven drying time, the According to ASTM C1012-04,6 the solution must be
concrete prisms were placed vertically in the desiccator renewed each week, which requires large amounts of sodium
(described by ASTM C120229 as one requiring two hose sulfate. Mehta12 recommended controlling the pH within a
connections). The prisms were kept under vacuum for range of 6.0 to 8.0 by adding a suitable amount of sulfuric
24  hours to remove the maximum amount of air, and a acid solution (0.1 N H2SO4); this study combined the two
5% sodium sulfate solution (by weight) was prepared. The methods. The pH of the sulfate solution was corrected daily
difference between Methods A and B can be observed in the (between 6 and 7) during immersion time, and the aggres-
procedure of presaturation by sulfate solution. sive solution was completely renewed every 6 weeks.
In Method A, after 24 hours of dried vacuuming with The efficiency of supplementary preconditioning via
the desiccator with vacuum pump and pressure main- Methods A and B was tested by assessing the expansion,
tained between –635 mm Hg and –760 mm Hg (0.084 and mass, and ultrasonic pulse velocity changes of concrete
0.1 MPa), the sulfate solution was drained slowly into the samples. The results were compared with the control
desiccator container to cover the entire surface and length method (which had no preliminary saturation) using the
of specimens. Draining stopped after levels reached 10 to same immersion conditions. Thirty and 60 MPa [4.4 and
20 mm (039 to 0.79 in.) above the tops of the samples. The 8.7 ksi] medium and high sulfate-resistant) concretes were
vacuum pump was kept running for another 72 hours, with tested. To investigate the most effective method for accel-
full immersion of specimens. erating degradation by sulfate attack, the control method,
In Method B (Fig. 1), after 24 hours of dried vacu- Method A, and Method B results were compared. Miner-
uming with the desiccator (Fig. 1(a)) with the running alogical (XRD) investigations were performed to clearly
vacuum pump and pressure running between –635 and understand the mechanism of attack using the new method
–760 mm  Hg (0.084  and 0.1 MPa), the sulfate solution of preconditioning, and chemical and microstructural
was drained slowly into the desiccator container to cover studies were carried out using SEM coupled to EDS and
approximately one-third of the specimen’s length (approx- microanalyses of polished surfaces.

ACI Materials Journal/July-August 2016 495


Fig. 2—Expansion of concrete specimens in 5% Na2SO4 solution after being subjected to different preconditioning methods.
RESULTS AND DISCUSSION the first cause of expansion and damage of concrete under
Expansion measurements sulfate attack.30 According to Perruchot et al.,31 ettringite
Figure 2 shows the 1-year expansion of concrete speci- crystals formed by the reaction between sulfate ions and
mens exposed to sulfate attack after being subjected to aluminates (sulfoaluminate or carboaluminates) may cause
the control and new methods of supplementary precondi- crystallization pressure related to super-saturation and pore
tioning. The length changes of the 30 MPa (4.4 ksi) class size, which generate significant expansive pressure. The
concrete samples presented in Fig. 2(a) show a difference in high expansion in specimens subjected to preconditioning
expansion trend between test methods. Unlike the 120-day via Method B can be explained by the accelerated diffu-
dormant period in the case of the control method, expan- sion of sulfate ions throughout the core of concrete spec-
sion increased systematically for specimens preconditioned imens up to super-saturation. Early super-saturation and
via Methods A and B, which may prove the efficiency of ettringite formation caused early expansion that increased
presaturation using desiccators. The specimens subjected to dramatically with time. The continuous increase of expan-
supplementary preconditioning using Method B showed the sion for 60 MPa (8.7 ksi) class specimens subjected to
highest degree of expansion from the first week of immer- Method A and Method B can be observed from the results
sion in sodium sulfate solution. The difference in expansion in Fig. 2(b). Negligible (0.004%) expansion was recorded
appears to have been more pronounced with time, especially for specimens subjected to the control method. At the end
when comparing Method B and the control method. Speci- of 1 year, expansions of 0.18% and 0.08% were recorded
mens subjected to Method A consistently exhibited higher for Methods  B and A, respectively. In previous studies of
expansions than those subjected to the control method; high-strength concretes (high-sulfate-resistant concrete),
however, they showed a lower rate of expansion evolution first expansion started after at least 2 years of exposure to
compared to those exposed to Method B. At 1 year, expan- sulfate attack.26 Using desiccators to presaturate the speci-
sions of 0.11%, 0.31%, and 0.82% were measured using the mens using Method A or B resulted in expansion, starting
control, A, and B preconditioning procedures, respectively. from the first weeks of immersion in sulfate solution. This
Method B provided expansion 7.5 times higher than the result explains the importance of supplementary precondi-
control method, and 2.6 times higher than Method A. As tioning by presaturation on the acceleration of expansion in
cited in the literature, ettringite formation is believed to be sulfate attack environments.

496 ACI Materials Journal/July-August 2016


Fig. 3—Mass change of concrete specimens in 5% Na2SO4 solution after being subjected to different preconditioning.
Mass measurements Khelifa et al.,32 the mass gain may be due to water filling
Mass change results for specimens subjected to supple- cracks and to the mass of water used to precipitate the
mentary preconditioning using control, A, and B methods hydrated phase as ettringite. Mass decrease occurred due
and immersed in 5% sulfate solution are shown in Fig. 3. to degradation by sulfate attack, starting by spalling and
The effects of preconditioning methods can be observed cracking at corners and edges.
from the trends in mass change (Fig. 3(a)). Starting with the Figure 3(b) shows that for the 60 MPa (8.7 ksi) strength
first week of immersion, specimens subjected to Method B class, at 360 days of exposure to sulfate attack, mass change
gained greater mass than their counterparts subjected to showed an increasing trend for specimens subjected to
Method A or the control method, and Method A speci- preconditioning using the control method or Method A.
mens gained greater mass than those exposed to the control Specimens preconditioned using Method B presented
method. Unlike the systematic gain of mass evident with greater mass gain starting from the first week of immer-
the control method, specimens exposed to Methods A and sion. As in the 30 MPa (4.4 ksi) class, this trend in mass
B showed three stages of change: a rapid increase of mass, gain can be attributed to the higher expansion in Method B
a short stabilization period, and finally a decrease in mass. results, and also to accelerated ettringite precipitation and
Concrete specimens subjected to preconditioning using gypsum formation from sulfate reaction, which need more
Method B had the shortest mass-gain time and earlier and solution absorption. The superiority of Method B (as accel-
greater mass decrease. The greater mass gain with Method B erated test) compared to Method A and the control method,
specimens can be related to the higher expansion from the can be explained by the fact that, in Method B, the trend of
first week of exposure to sulfate attack. In addition, the use decrease in mass change (which can be considered as initi-
of presaturation, especially with Method B, may accelerate ation of deterioration due to sulfate attack) was observed
ettringite formation, and, consequently, water absorption. within 360 days.
Indeed, mass gain in sulfate attack is usually attributed to
the hydration of anhydrous cement, and to the formation of Ultrasonic pulse velocity measurements
gypsum and secondary ettringite due to reactivity between Figure 4 shows the variation of ultrasonic pulse velocity
hydrates and sodium sulfate.26 In addition, according to (UPV) with time for different concrete samples subjected to

ACI Materials Journal/July-August 2016 497


Fig. 4—Ultrasonic pulse velocity change of concrete in 5% Na2SO4 solution after being subjected to different precondi-
tioning methods.
preconditioning using the control method and Methods A eration of damage process stemming from presaturation
and B, and immersed in 5% Na2SO4 solution. Figure 4(a) according to Method A or B. However, it is evident that
shows that in the first 120 days of exposure to sulfate attack, Method B speeds up the damage process of concrete speci-
using new Method A or B preconditioning procedures led to mens more significantly.
higher velocity gain than using the control method, with the Fig. 4(b) shows that the UPV results of 60 MPa (8.7 ksi)
pre-conditioning method showing a significant influence class concrete specimens were not severely affected by the
on velocity variation over time. Starting from 135 days preconditioning method used. However, after 220 days of
of immersion, the pulse velocity of samples subjected to immersion in sulfate solution, the UPV values of specimens
Method B dropped with time. Those subjected to Method A subjected to Method B decreased slowly with immersion
showed a decrement in pulse velocity after roughly time. These results provide information about changes in
240 days of immersion, and the control method showed a microstructure and the beginning of internal deteriora-
systematic increment of velocity with time. At 360 days of tion due to accelerated sulfate ingress with presaturation
exposure to sulfate attack, pulse velocities of 8%, –2.1%, according to Method B. Method A preconditioning showed
and –10.6% were recorded for specimens subjected to a small decrease of velocity after 320 days of immersion,
Control, A, and B methods, respectively. In terms of reduc- while UPV values of specimens subjected to the control
tion in pulse velocity (which exhibits the concrete deterio- method showed no decrements up to 360 days of exposure.
ration), the results demonstrate that preconditioning using This illustrates the inability of the control method to initiate
Method B is much more dominant than Method A and the sulfate deterioration within 360 days, and confirms the
control method. Because all samples of 30 MPa (4.4 ksi) acceptability of Method B as an accelerated test. The dense
class concrete were taken from the same mixture and had microstructure of the 60 MPa (8.7 ksi) concrete prevented
the same initial properties, and because the pulse velocity initiation and spread of deterioration. However, a compar-
test offers information about the internal microcracking and ison of UPV curves confirms that the supplementary precon-
the soundness of the materials microstructure,33 the reason ditioning methods can significantly affect sulfate vulnera-
for pulse velocity differences is likely due to the accel- bility testing.

498 ACI Materials Journal/July-August 2016


Fig. 5—X-ray diffraction spectra at different layers after 1-year immersion in sulfate solution (Note: C is calcite; D is dolomite;
E is ettringite; G is gypsum; P is portlandite; Q is quartz; 1 mm = 0.0394 in.)
Microstructural observations is in agreement with the realistic degradation in sulfate envi-
To better understand the effects of different precondi- ronments shown in literature. Indeed, according to Bellmann
tioning methods on degradation levels of concrete, 30 MPa et al.,9 in field conditions where sulfate attack is present,
(4.4 ksi) class specimens subjected to preconditioning using ettringite or both gypsum and ettringite can be observed as
the Control, A, and B methods were examined by XRD after results of concrete damage. A relatively higher quantity and
1 year of immersion in sodium sulfate solution. Measurements strong peaks of ettringite can be noted in the surface layer of
were performed starting from the surface layer (0 to 2 mm samples subjected to preconditioning by Methods A and B
[0 to 0.079 in.]), moving toward the core (8 to 10, 18 to 20, than those exposed to the control method, especially at around
28 to 30 mm [0.31 to 0.39, 0.71 to 0.79, 1.1 to 1.2 in.]). Figure 5 22.04 deg 2θ and 45.15 deg 2θ. Furthermore, an ettringite
presents the XRD spectra at different layers for concretes peak of around 30.55 deg 2θ was present only in the surfaces
subjected to each type of preconditioning. Patterns were of the Method B samples. Important peaks of gypsum can be
obtained using an X-ray diffractometer with Cu Ka radiation seen on all surfaces; however, at around 11.6 deg 2θ, they
and 2θ scanning, ranging between 5 and 80 degrees of 2θ. were more significant in the Method B results. There was no
The superposition of the XRD patterns of different layers detectable trace of portlandite on all surfaces, which indi-
shows dominant peaks of ettringite and gypsum on all cates total leaching of this phase in the surfaces that were in
surfaces exposed to the sodium sulfate solution. This finding permanent content with the sulfate solution.

ACI Materials Journal/July-August 2016 499


Fig. 6—Sulfate profile of concrete samples exposed to sulfate attack after preconditioning. (Note: 1 mm = 0.0394 in.)
Greater differences between the data of each layer were sulfate ions up to this layer, which can lead to the formation
observed by moving into the specimen cores. In spite of of gypsum or ettringite.
the presence of ettringite, gypsum, and portlandite, in all According to Siad et al.,26,35 looking at the sulfate profile
8 to 10 mm (0.31 to 0.39 in.) layers, XRD results showed on polished surfaces is a reliable method of determining the
higher intensity of ettringite peaks in specimens subjected level of degradation by sulfate attack. To better understand
to Method B, especially in the principal calcite peaks, the role of the preconditioning method on the acceleration
located at around 22.04 deg 2θ, 35.5 deg 2θ, 36.84 deg 2θ, of degradation, sulfur variation was investigated from the
and 45.15 deg 2θ. Starting from a 18 mm (0.71 in.) depth, surface toward the core of specimens. The concretes inves-
strong ettringite peaks were noted only in concrete samples tigated by SEM were examined using EDS analysis on
subjected to Method B, though with fewer peaks in the inner polished surfaces. Line profiles avoided maximum aggre-
layer (28 to 30 mm [1.1 to 1.2 in.]). For specimens subjected gates. Figure 6 presents the sulfate profile of the samples
to preconditioning by Method A, ettringite patterns were subjected to different methods of preconditioning and
also detected; however, the amount and intensity were lower immersed for 1 year in sodium sulfate solution.
than for those subjected to Method B. Very weak peaks of As seen from the figure, the sulfur was not uniformly
ettringite were seen in the 18 to 20 mm (0.71 to 0.79 in.) and dispersed in the concrete pastes. Profiles presented in
28 to 30 mm (1.1 to 1.2 in.) layers of samples preconditioned Fig. 6 follow three stages: a maximum density within a few
using the control method. millimeters of the exterior layer; a gradual increase; and
Better-defined portlandite peaks were seen, starting from a different extent of the stable amount. A small difference
the 8 to 10 mm (0.31 to 0.39 in.) layer of the control speci- in sulfur content was seen in the exterior surface exposed
mens, than in the Method A samples. There was a detectable directly to sulfate attack. Within 1 mm (0.04 in.) of depth,
trace of portlandite, starting only from 18 mm (0.71 in.) of maximum amounts of sulfur of approximately 13.1, 13.5,
the samples preconditioned by Method B, indicating signif- and 14.6% were recorded for the control method, Method A,
icant leaching of calcium hydroxide before this phase. It is and Method B specimens, respectively. Starting from 1 mm
well known that CH reacts with sulfate to create more expan- (0.4 in.) toward the core, an important difference was notice-
sive secondary ettringite and gypsum,26 which is also in able. When compared to the control results, the amount of
agreement with the higher ettringite peaks in the Method B sulfur seemed to stabilize at approximately 2% density.
results. In comparing the gypsum content of each layer, Using this value to characterize the unaffected zone, the
unlike the important peaks showing in the exterior layer of depth of degradation was found to be equal to 7.8, 13.8,
all the specimens, only a few gypsum patterns were notice- and 25 mm (0.31, 0.54, and 0.98 in.), for the control, A, and
able in the 8 to 10 mm (0.31 to 0.39 in.) layer, with a slightly B specimens, respectively. Moreover, the density of sulfur
better defined peak at around 11.65 deg 2θ in Method B spec- in concrete specimens subjected to Method B showed a
imens. Gypsum peaks were totally absent in the 18 to 20 mm minimum amount of roughly 8% (up to 23 mm [0.91 in.] in
(0.71 to 0.79 in.) layer of concretes preconditioned by the depth), which is four times the density found in the control
control method or Method A; however, a trace of gypsum specimens. These results are consistent with XRD patterns,
was seen in those preconditioned by Method B. Regardless which show strong ettringite peaks starting from the surface
of preconditioning method, the presence of gypsum within and reaching the 18 to 20 mm (0.71 to 0.79 in.) layer of
a few millimeters of the surface of specimens immersed in Method B specimens. The presence of high amounts of
sulfate solution has been noted in literature. According to ettringite can explain the high expansion and damage shown
Santhanam et al.,34 in concrete damaged by sulfate attack, above in the Method B results.
gypsum is primarily detected close to the surface, especially To detect and quantify the form of sulfur (ettringite,
in cracks and voids. The trace of gypsum, which appeared gypsum, or monosulfoaluminate) shown in the sulfate profiles
even in inner (18 to 20 mm [0.71 to 0.79 in.]) layers of spec- section, a complementary analysis of the microstructure was
imens from Method B, can be attributed to the penetration of performed according to the Scrivener correlation.36 Figure 7
presents the changes in the phase assemblage from the surface

500 ACI Materials Journal/July-August 2016


toward the core (0 to 8 mm [0 to 0.31 in.] and 18 to 26 mm Microanalysis of the external zone (0 to 8 mm [0 to
[0.71 to 1.02 in.] zones) of the element ratios S/Ca as a func- 0.31 in.]) showed the presence of ettringite partly intermixed
tion of Al/Ca. The plots of S/Ca versus Al/Ca suggest that with C-S-H in all specimens. High amounts of ettringite
the ettringite phase has S/Ca = 0.5 and Al/Ca = 0.33, and were noticed in specimens subjected to Method B, with less
the S/Ca and Al/Ca ratios of 1 and 0 correspond respec- in Method A specimens and an even smaller amount in the
tively to gypsum and monosulfoaluminate, with S/Ca = control specimens. Gypsum was also identified in all the
0.25 and Al/Ca = 0.5. external zones exposed to sulfate attack, with higher content
in Method B specimens. Greater amounts of monosulfoalu-
minate intermixed with the C-S-H phase were detected in
the control specimens. Furthermore, portlandite was clearly
absent among the external zone of Method B specimens, but
was present in the Method A and control samples.
The phase assemblage across the cores of concrete
samples revealed the presence of high amounts of ettringite
in Method B specimens, partly intermixed with C-S-H,
even between 18 and 26 mm (0.71 and 1.02 in.) of depth. A
slight quantity of ettringite was detected in Method A speci-
mens, and even less was observed in the control specimens.
Portlandite was detected in Method B specimens, however,
with lower amounts compared to the Method A and control
samples. An insignificant difference was noted for monosul-
foaluminate intermixed with the C-S-H phase of different
specimen cores.
SEM-EDS analyses were also conducted to investigate the
form of ettringite and gypsum produced by sulfate reaction.
Figures 8 through 10 show SEM micrographs and EDS
analysis of concrete samples subjected to different methods
of preconditioning. The gypsum micrographs (Fig. 8(a), 9(a),
and 10(a)) are taken from areas close to the surface layer,
while ettringite micrographs (Fig. 8(b), 9(b), and 10(b)) are
taken from 10 to 15 mm (0.39 to 0.59 in.) inside the core.
Fig. 7—Phase assemblages of concrete specimens exposed SEM micrographs confirmed the presence of significant
to sulfate attack after preconditioning (Atom ratio plot of amounts of compressed (Fig. 10(b)(1)), massive (Fig. 10(b)
S/Ca versus Al/Ca for surface zone and core.) (Note: G is (2)), and needle (Fig. 10(b)(3)) forms of ettringite dispersed
gypsum; Ett is ettringite; Ms is monosulfoaluminate; CH is along cement paste/aggregate interfaces and in the cavities
portlandite; 1 mm = 0.0394 in.)

Fig. 8—SEM and EDS data for concrete specimens subjected to preconditioning using control method.

ACI Materials Journal/July-August 2016 501


Fig. 9—SEM and EDS data for concrete specimens subjected to preconditioning using Method A.

Fig. 10—SEM and EDS data for concrete specimens subjected to preconditioning using Method B.
left by disintegrated aggregates of samples subjected to confirmed to be expansive. These results can explain the
preconditioning by Method B. This type of ettringite has acceleration method of degradation by using part-by-part
been noted in literature as an expansive form.37,38 According saturation in primary preconditioning of concrete speci-
to Divet and Pavoine,38 the abundance of the massive form mens. Indeed, the use of desiccators to saturate specimens
of ettringite at the cement paste/aggregate interface gives under vacuum forced the sulfate ions to migrate through
the impression that its formation will have generated suffi- their cores.
ciently large internal stresses to cause deformation and Ettringite detected in the cores of samples subjected to
cracking within the material. In addition to the large quan- Method A (Fig. 9) also showed massive (Fig. 9(b)(1)) and
tity of ettringite throughout the cores of samples subjected needle (Fig. 9(b)(3)) forms, mainly present around cavi-
to preconditioning by Method B, they also had higher ties left by the disintegration of aggregates during prepa-
sulfate profiles and phase assemblage sections; the quality ration. However, although saturation by complete immer-
of ettringite generated from the reaction with sulfate attack sion (Method A) accelerated sulfate penetration inside the

502 ACI Materials Journal/July-August 2016


specimens, preconditioning using part-by-part saturation 3. Preconditioning methods had a significant influence
(Method B) caused maximum diffusion. This can lead to on UPV variation, with early and higher variations noted
super-saturation by sulfate ions and accelerate the formation in Method B specimens. For instance, the 30 MPa (4.4 ksi)
and precipitation of ettringite needles, which generate local- concrete specimens showed UPV decrements after 135 and
ized high pressure and consequently lead to early expansion. 240 days for B and A preconditioning methods, respec-
For the control specimens, it was difficult to identify the tively. The control method showed a systematic increment
ettringite phase in the sample cores (Fig. 8(b)). However, of velocity with time. It is evident that Method B sped up the
some agglomerations (massive [Fig. 8(b)(2)] and needle damage process and is therefore more effective as an accel-
[Fig. 8(b)(3)] forms of ettringite) were noticed randomly erated test.
scattered in the paste around aggregates. Although the expan- 4. Microstructural analysis showed that sulfate profiles
sive ettringite was present inside the control specimens, its and phase assemblages were consistent with expansion
abundance prevented the generation of high pressure and, results. Significant amounts of ettringite throughout the
therefore, expansion occurred later than in Method A speci- surfaces and cores of Method B specimens were observed.
mens and especially in Method B specimens. The depth of sulfur was equal to 7.8, 13.8, and 25 mm (0.31,
Gypsum from the control (Fig. 8(a), A (Fig. 8(b)), and B 0.54, and 0.98 in.), for the control, B, and A method spec-
(Fig. 8(c)) specimens was also detected at the inner surface, imens, respectively. Density of sulfur in specimen cores
close to regions where maximum ettringite was observed. subjected to Method B preconditioning was four times that
However, it was more pronounced and denser for Method B of the control specimens. SEM micrographs and EDS anal-
and A specimens. Despite the debate over the expansivity yses confirmed the expansive quality of ettringite found in
of gypsum caused by sulfate attack,34,39,40 the clearly high the cores of Method B specimens. The mechanism of degra-
quantity of this product related to the high expansion and dation of Method B specimens confirmed those observed in
damage in specimens preconditioned by Methods B and A field conditions.
offers information about the possible role of gypsum forma-
tion in the acceleration of damage caused by sulfate attack. AUTHOR BIOS
Hocine Siad is a Postdoctoral Researcher in the Department of Civil Engi-
neering at Ryerson University, Toronto, ON, Canada. His research interests
CONCLUSIONS include the development and assessment of new materials and methods for
This paper presents the results of a study aimed to recom- concrete durability.
mend a new method of accelerating the degradation of
Mohamed Lachemi is a Professor in the Department of Civil Engineering
concrete specimens exposed to external sulfate attack. The at Ryerson University. His research interests include the use of high-
study is important, considering the absence of a standard performance materials and self-consolidating concrete, including the
accelerated test to inspect concrete sulfate resistance. The development and use of reactive powder in concretes.
goal was to reduce testing time while simulating degrada- Mustafa Şahmaran is an Associate Professor in the Department of Civil
tion mechanisms that reflect field conditions. The reliability Engineering at Gazi University, Ankara, Turkey. His research interests
and performance of methods based on immersion in sulfate include advanced materials technology and composite materials develop-
ment for sustainable infrastructure.
solution coupled with primary preconditioning were inves-
tigated. The study compared two new methods involving Khandaker M. Anwar Hossain is a Professor in the Department of Civil
presaturation (A and B) with a control method using no Engineering at Ryerson University. His research interests include the devel-
opment of construction materials with volcanic materials/industrial wastes
special preconditioning. In Method A, preconditioning and composite materials for structural applications.
involved a one-time complete immersion inside the desic-
cator under vacuum, while in Method B, presaturation was REFERENCES
performed with part-by-part presaturation under vacuum. 1. Sahmaran, M.; Erdem, T. K.; and Yaman, I. O., “Sulfate Resistance
The efficiency of the newly proposed methods was studied of Plain and Blended Cements Exposed to Wetting-Drying and Heating-
Cooling Environments,” Construction and Building Materials, V. 21, No. 8,
by assessing the expansion evolution, mass change, and UPV 2007, pp. 1771-1778. doi: 10.1016/j.conbuildmat.2006.05.012
of the concrete specimens during the immersion period. The 2. Sahmaran, M.; Kasap, O.; Duru, K.; and Yaman, I. O., “Effects of
following conclusions were drawn from the study: Mix Composition and Water-Cement Ratio on the Sulfate Resistance of
Blended Cements,” Cement and Concrete Composites, V. 29, No. 3, 2007,
1. Higher expansion was observed in specimens subjected pp. 159-167. doi: 10.1016/j.cemconcomp.2006.11.007
to preconditioning by Method B than by Method A and the 3. Girardi, F.; Vaona, W.; and Di Maggio, R., “Resistance of Different
control method. For the 30 MPa (4.4 ksi) strength concrete, Types of Concretes to Cyclic Sulfuric Acid and Sodium Sulfate Attack,”
Cement and Concrete Composites, V. 32, No. 8, 2010, pp. 595-602. doi:
Method B provided 7.5 and 2.5 times more expansion than 10.1016/j.cemconcomp.2010.07.002
the control method and Method A, respectively, at the end 4. ACI Committee 318, “Building Code Requirements for Structural
of a 1-year exposure period. For the high-strength concrete Concrete (ACI 318-99) and Commentary,” American Concrete Institute,
Farmington Hills, MI, 1999, 353 pp.
(60 MPa [8.7 ksi]), negligible (0.004%), 0.08%, and 0.18% 5. Chabrelie, A., “Mechanisms of Degradation of Concrete by External
expansion was recorded for specimens made with the Sulfate Ions under Laboratory and Field Conditions,” PhD thesis, Ecole
control, A, and B methods, respectively. Polytechnique Fédérale de Lausanne, Lausanne, Switzerland, 2010, 186 pp.
6. ASTM C1012-95a, “Standard Test Method for Length Change of
2. Although no mass loss was measured during immersion, Hydraulic-Cement Mortars Exposed to a Sulfate Solution,” ASTM Interna-
differences in mass change influenced by preconditioning tional, West Conshohocken, PA, 1995, 5 pp.
method were noted. Method B samples showed higher mass 7. Hooton, R. D., and Brown, P. W., “Development of Test Methods
to Address the Various Mechanisms of Sulphate Attack,” Proceedings of
gain, and early and higher mass decrements than the control Concrete in Aggressive Aqueous Environments, Performance, Testing and
and A methods. Modeling, Toulouse, France, 2009, pp. 280-297.

ACI Materials Journal/July-August 2016 503


8. Aleksic, M., “Development and Standardization of the NIST Rapid 25. CAN/CSA A3000-08, “Cementitious Materials Compendium,”
Sulphate Resistance Test,” master’s thesis, Department of Civil Engi- Canadian Standards Association, Rexdale, ON, Canada, 2008, 282 pp.
neering, University of Toronto, Toronto, ON, Canada, 2010, 178 pp. 26. Siad, H.; Kamali-Bernard, S.; Mesbah, H. A.; Escadeillas, G.;
9. Bellmann, F.; Moser, B.; and Stark, J., “Influence of Sulfate Solution Mouli, M.; and Khelafi, H., “Characterization of the Degradation of
Concentration on the Formation of Gypsum in Sulfate Resistance Test Spec- Self-Compacting Concretes in Sodium Sulfate Environment: Influence
imen,” Cement and Concrete Research, V. 36, No. 2, 2006, pp. 358-363. of Different Mineral Admixtures,” Construction and Building Materials,
doi: 10.1016/j.cemconres.2005.04.006 V. 47, Oct. 2013, pp. 1188-1200. doi: 10.1016/j.conbuildmat.2013.05.086
10. Suleiman, A. R., “Physical Sulphate Attack on Concrete,” master’s 27. Neville, A. M., Properties of Concrete, third edition, Longman
thesis, Department of Civil and Environmental Engineering, The University Scientific and Technical, Essex, UK, 1981, pp. 338-344.
of Western Ontario, London, ON, Canada, 2014, 127 pp. 28. Sanjuán, M. A., and Muñoz-Martialay, R., “Oven-Drying as a Precon-
11. Crumbie, A. K., “Characterization of the Microstructure of Concrete,” ditioning Method for Air Permeability Test on Concrete,” Materials Letters,
PhD thesis, University of London, Imperial College, London, UK, 1994, V. 27, No. 4-5, 1996, pp. 263-268. doi: 10.1016/0167-577X(95)00283-9
pp. 253-258. 29. ASTM C1202-12, “Standard Test Method for Electrical Indication
12. Mehta, P. K., “Evaluation of Sulfate Resisting Cements by a New Test of Concrete’s Ability to Resist Chloride Ion Penetration,” ASTM Interna-
Method,” ACI Journal Proceedings, V. 72, No. 10, Oct. 1975, pp. 573-575. tional, West Conshohocken, PA, 2012, 7 pp.
13. Drimalas, T., “Laboratory and Field Evaluations of External Sulfate 30. ACI Committee 201, “Guide to Durable Concrete (ACI 201.2R-92),”
Attack,” PhD thesis, University of Texas at Austin, Austin, TX, 2007, 187 pp. American Concrete Institute, Farmington Hills, MI, 1992, 41 pp.
14. Cohen, M. D., and Mather, B., “Sulfate Attack on Concrete— 31. Perruchot, C.; Chehimi, M. M.; Vaulay, M. J.; and Benzarti, K.,
Research Needs,” ACI Materials Journal, V. 88, No. 1, Jan.-Feb. 1991, “Characterization of the Surface Thermodynamic Properties of Cement
pp. 62-69. Components by Inverse Gas Chromatography at Infinite Dilution,” Cement
15. Almeida, I. R., “Resistance of High-Strength Concrete to Sulfate and Concrete Research, V. 36, No. 2, 2006, pp. 305-319. doi: 10.1016/j.
Attack: Soaking and Drying Test,” Durability of Concrete, SP-100, cemconres.2005.02.005
J. M. Scanlon, ed., American Concrete Institute, Farmington Hills, MI, 32. Khelifa, R.; Brunetaud, X.; Chabil, H.; and Al-Mukhtar, M.,
1991, pp. 1073-1092. “Conséquences Mécaniques de L’attaque Sulfatique Externe sur les BAP,”
16. Assaad, V. F.; Jofriet, J. C.; Negi, S. C.; and Hayward, G. L., “Labo- SciTechnol Journal, V. 28, 2008, pp. 23-28.
ratory Tests of Sulphide and Sulphate Corrosion of Reinforced Concrete,” 33. Skaropoulou, A.; Kakali, G.; and Tsivilis, S., “Thaumasite Form
Proceedings of the Vth International Symposium on Concrete for a Sustain- of Sulfate Attack in Limestone Cement Concrete: The Effect of Cement
able Agriculture, El Escorial, Spain, 2005, pp. 179-190. Composition, Sand Type and Exposure Temperature,” Construction
17. Damidot, D., and Glasser, F. P., “Thermodynamic Investi- and Building Materials, V. 36, Nov. 2012, pp. 527-533. doi: 10.1016/j.
gation of the CaO-Al2O3-CaSO4-H2O System at 50°C and 85°C,” conbuildmat.2012.06.048
Cement and Concrete Research, V. 22, No. 6, 1992, pp. 1179-1191. doi: 34. Santhanam, M.; Cohen, M. D.; and Olek, J., “Effects of Gypsum
10.1016/0008-8846(92)90047-Y Formation on the Performance of Cement Mortars during External Sulfate
18. De Belie, N., “Evaluation of Methods for Testing Concrete Degra- Attack,” Cement and Concrete Research, V. 33, No. 3, 2003, pp. 325-332.
dation in Aggressive Solutions,” Workshop on Performance of Cement- doi: 10.1016/S0008-8846(02)00955-9
Based Materials in Aggressive Aqueous Environments—Characterization, 35. Siad, H.; Lachemi, M.; Bernard, S. K.; Sahmaran, M.; and Hossain,
Modeling, Test Methods and Engineering Aspects, N. De Belie, ed., Ghent, A., “Assessment of the Long-Term Performance of SCC Incorporating
Belgium, 2007, pp. 63-74. Different Mineral Admixtures in a Magnesium Sulphate Environment,”
19. Clifton, J. R.; Frohnsdorff, G.; and Ferraris, C., “Standards for Eval- Construction and Building Materials, V. 80, Apr., 2015, pp. 141-154. doi:
uating the Susceptibility of Cement-Based Materials to External Sulfate 10.1016/j.conbuildmat.2015.01.067
Attack,” Materials Science of Concrete, Sulfate Attack Mechanisms, 36. Scrivener, K. L., “Backscattered Electron Imaging of Cementi-
Special Volume, 1999, pp. 337-355. tious Microstructures Understanding and Quantification,” Cement and
20. Damidot, D., and Glasser, F. P., “Thermodynamic Investiga- Concrete Composites, V. 26, No. 8, 2004, pp. 935-945. doi: 10.1016/j.
tion of the CaO-Al2O3-CaSO4-K2O-H2O System at 25°C,” Cement cemconcomp.2004.02.029
and Concrete Research, V. 23, No. 5, 1993, pp. 1195-1204. doi: 37. Cohen, M. D., “Modeling of Expansive Cements,” Cement
10.1016/0008-8846(93)90180-H and Concrete Research, V. 13, No. 4, 1983, pp. 519-528. doi:
21. Higgins, D. D., “Increased Sulfate Resistance of GGBS Concrete 10.1016/0008-8846(83)90011-X
in the Presence of Carbonate,” Cement and Concrete Composites, V. 25, 38. Divet, L., and Pavoine, A., “Delayed Ettringite Formation in Massive
No. 8, 2003, pp. 913-919. doi: 10.1016/S0958-9465(03)00148-3 Concrete Structures: An Account of Some Studies of Degraded Bridges,”
22. Osborne, G. J., “Durability of Portland Blast-Furnace Slag Cement International RILEM TC-186 ISA, Switzerland, 2002, pp. 98-126.
Concrete,” Cement and Concrete Composites, V. 21, No. 1, 1999, pp. 11-21. 39. Tian, B., and Cohen, M. D., “Does Gypsum Formation during Sulfate
doi: 10.1016/S0958-9465(98)00032-8 Attack on Concrete Lead to Expansion,” Cement and Concrete Research,
23. Neville, A., “The Confused World of Sulfate Attack on Concrete,” V. 30, No. 1, 2000, pp. 117-123. doi: 10.1016/S0008-8846(99)00211-2
Cement and Concrete Research, V. 34, No. 8, 2004, pp. 1275-1296. doi: 40. Mather, B., “Discussion of the Process of Sulfate Attack on
10.1016/j.cemconres.2004.04.004 Cement Mortars,” Advanced Cement Based Materials, V. 5, No. 3, 1996,
24. ASTM C150/C150M-12, “Standard Specification for Portland pp. 109-110.
Cement,” ASTM International, West Conshohocken, PA, 2012, 9 pp.

504 ACI Materials Journal/July-August 2016

You might also like