You are on page 1of 18

Construction and Building Materials 293 (2021) 123550

Contents lists available at ScienceDirect

Construction and Building Materials


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

Compressive strength assessment of sulfate-attacked concrete by using


sulfate ions distributions
Hanbin Cheng, Tiejun Liu, Dujian Zou ⇑, Ao Zhou
School of Civil and Environmental Engineering, Harbin Institute of Technology, Shenzhen 518055, PR China

h i g h l i g h t s

 The performance of concrete under sulfate attack and dry-wet cycles is investigated.
 A newly defined index is proposed to describe the progress of sulfate attack.
 A novel method is proposed to assess the performance of sulfate-attacked concrete.

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

Article history: Sulfate attack is a major cause of concrete durability deteriorations. Mass loss, strength reductions, and
Received 28 January 2021 expansive strain of concrete specimens are generally used in laboratory testing to identify the resistance
Received in revised form 9 April 2021 of concrete materials to sulfate attack. However, these indicators cannot be directly used to quantita-
Accepted 30 April 2021
tively predict the bearing capacity of actual concrete structures under sulfate attack. There exists a sig-
Available online 10 May 2021
nificant size effect between laboratory and engineering size concrete components. In this study, the
durability performance of concrete specimens, exposed to sulfate attack and dry–wet cycles, was inves-
Keywords:
tigated. Mass loss, dynamic elastic modulus, compressive strength, and sulfate ions distributions of dete-
Concrete material
Sulfate attack
riorated concrete were measured over time. Test results indicate that the newly defined integral area of
Non-uniform deteriorations sulfate ions distributions is a suitable index to describe the non-uniform deteriorations behavior of sul-
Homogenization’s theory fate–attacked concrete; and a novel method based on the homogenizations theory is proposed to predict
Sulfate ions distributions the deteriorations level of components of attacked concrete structures, which provides a potential use in
Strength assessment assessing the loading capacity of actual concrete structures based on accelerated test results in a
laboratory.
Ó 2021 Elsevier Ltd. All rights reserved.

1. Introductions loss that occur [13–24]. Therefore, mass loss and expansive strain
are generally recommended as measuring indices by current codes
Sulfate attack is a major contributing factor to the deteriora- to assess the deteriorations level of sulfate-attacked concrete in
tions of cement-based materials [1–5]. Sulfates are naturally laboratory testing [25,26]. Based on the systematic study on the
occurring minerals found in saline-alkali soil, groundwater, and change of mesoscopic pore structures induced by chemical sulfate
coastal areas [6]. Considering the actual exposure environment, attack and physical sulfate attack, pore structures can also express
sulfate ions are not always the only erosion chemical products, the degradations process of sulfate attack inside concrete [27,28].
degradations process induced by sulfate attack usually can be Furthermore, according to the analysis of phase compositions of
accelerated by other ions, including chlorides [7,8] or magnesium corrosions products, the damage level of sulfate attacked concrete
[9]. The deteriorations of concrete structures exposed to sulfates could also be revealed by using XRD and SEM-EDS [13–17]. In addi-
is caused by the formations of ettringite and gypsum resulting tions, depth of sulfate attack could also be used as the indicator to
from the chemical reactions of sulfate ions and cement hydrate express the deteriorations of concrete under the attack of sulfate
products [10–12]. Two typical manifestations of sulfate attack on [13–18]. Although the above discussed indicators and methods
concrete structures are the disruptive expansive cracks and mass are straightforward and reliable in identifying the resistance of dif-
ferent concrete materials exposed to sulfate attack, the compres-
⇑ Corresponding author. sive strength of sulfate attacked concrete is one of the important
E-mail addresses: chenghanbin1994@gmail.com (H. Cheng), liutiejun@hit.edu.cn points of concern [29].
(T. Liu), zoudujian@163.com (D. Zou), zhouao@hit.edu.cn (A. Zhou).

https://doi.org/10.1016/j.conbuildmat.2021.123550
0950-0618/Ó 2021 Elsevier Ltd. All rights reserved.
H. Cheng, T. Liu, D. Zou et al. Construction and Building Materials 293 (2021) 123550

The mechanical property or constitutive relations of deterio- 1470 kg/m3, fineness modulus is 2.7, silt content is 0.8%. Coarse
rated concrete is critically important to assess the safety of con- aggregate comes from Huizhou, with needle-like content of 2.0%,
crete structures in aggressive environments. Many experimental crushing index of 8.0%, silt content of 0.5%, and maximum particle
investigations have been reported in the study of the compressive size of 25 mm. STD-1 retarding superplasticizer was used in this
strength and flexure strength of sulfate-attacked concrete [4,30– experiment.
39]. Meanwhile, recent studies focus on using or dynamic elastic Commercial concrete, supplied by Shenzhen Tiandi Concrete
modulus [29] or static elastic modulus [40] to conduct assessment Corporations, was used for testing in this study. Three water–bin-
on sulfate attacked. Considering many actual structures bearing der ratios (W/Bs) of concrete were considered, and the detailed mix
shear stress, other experimental investigation uses shear strength proportions are presented in Table 2. Sulfate solutions with mass
as the main indicator to express damage level of sulfate-attacked fractions of 3%, 5%, and 10% were used to accelerate the deteriora-
concrete [41]. However, these tests primarily focus on the acceler- tions process of concrete. A total of 288 concrete specimens were
ated deteriorated tests using small-scale concrete specimens. cast (Table 3) and 45 of these concrete specimens were prepared
However, as to sulfate attacked concrete, the deteriorations is not for the control test curing in water. All concrete specimens were
uniform across the sections area of concrete. The exterior part of cured for 28 days in a standard environment with a temperature
concrete may be severely deteriorated while the inner part of con- of 20 °C and relative humidity of 95%.
crete could remain healthy [42–47]. Therefore, the isotropy
assumptions, that generally employed to study the constitutive 2.2. Deteriorations of concrete specimens
relations of normal concrete, cannot be applied to the analysis of
concrete under sulfate attack in actual conditions [48,49]. In addi- Two-dimensional sulfate attack was considered in the test. The
tions, the size effect between concrete specimens of various sizes top and bottom surfaces of all concrete specimens were sealed
under the same environmental conditions is not fully investigated with paraffin as illustrated in Fig. 1. Concrete specimens were first
[29]. For the identical deterioration depth of concrete, the decrease fully immersed in the sulfate solutions at room temperature for
in bearing capacity of concrete members varies between the tested seven days and dried at room temperature for eight days, which
size and engineering size of concrete members. Based on the above represents a dry–wet cycle. When the concrete specimens were
discussions, the accelerated test results cannot be used to predict fully immersed in the sulfate solutions, the pools were covered
the damage level of full-scale concrete members in actual with plastic film to prevent volatilizations of sulfate solutions,
conditions. and the solutions was replaced every month.
Considering the existing research, it is a prerequisite to propose
a new indicator and method to predict and assess strength of sul- 2.3. Test devices and methods
fate attacked concrete. In this study, based on the homogeniza-
tion’s theory, the attacked concrete was considered to be divided 2.3.1. Mass loss
into two sections, namely the damaged area and healthy area. All concrete specimens for each mixture were transferred to the
The strength of healthy area of concrete was assumed to be equal exposure conditions after measuring their initial mass by using a
to that of normal concrete. Therefore, the key to strength assess- balance with an accuracy of 0.1 g. Considering the influence of
ment of sulfate attacked concrete lies on the strength predictions water content inside concrete, all concrete specimens were dried
of attacked area of concrete. Based on experimental results, the at lab room temperature until mass reached to stable state before
sulfate ions distribution was used to identify the attacked area being measured initial mass. Except for the control concrete spec-
and the integral areas of sulfate ions distributions at various times imens, all the concrete specimens were immersed in a sulfate solu-
were calculated to express the deteriorations process of concrete tion for seven days and dried at room temperature for eight days
under sulfate attack. Then, a preliminary relationship between until a constant mass was reached, which constitutes one cycle
the integral area of sulfate ions distributions and the compressive of sulfate attack. After each cycle, the mass for each concrete spec-
strength of attacked area of concrete was established by using imen was measured. The specific measuring method followed the
robust regressions analysis. Finally, a novel methodology based standards outlined in GB/T50082-2012 [26], and the mass loss
on homogenizations and sulfate ions distributions was proposed ratio was calculated according to Eq. (1):
to assess the compressive strength of attacked concrete members.
Wt  W0
DW t ¼ 100  ð1Þ
W0
2. Experimental investigations
where DWt represents the mass change fractions, W0 indicates the
2.1. Materials and concrete specimens initial mass of the concrete specimen, and Wt represents the mass
of the concrete specimen at the deteriorations time of t.
The 42.5R ordinary Portland cement of Guangdong Huarun
Cement (Fengkai) Co., Ltd. was used in this test. It has a specific 2.3.2. Mechanical tests
surface area of 362 m2/kg, and its stability is qualified. The detailed The dynamic elastic modulus was determined by using a
chemical compositions are listed in the Table. 1. dynamic elastic modulus measurement machine (Model DT-
The fly ash used in the test is the product of Shajiao C Power 20W) was the setup. The measuring method followed the standard
Plant of Guangdong Yuedian Group Co. Ltd. The fineness of fly test method of mechanical properties for ordinary concrete [49].
ash is 21.30. The water demand ratio is 99% and the ignition loss The relative dynamic elastic modulus (RDEM) was used to evaluate
is 1.64%. The fine aggregate used in this test is river sand produced the compactness of the concrete specimens under sulfate attack,
in Dongguan. Apparent density is 2570 kg/m3, bulk density is and it was calculated by:

Table.1
Cement compositions.

Composition Na2O MgO Al2O3 SiO2 SO3 K2O CaO Fe2O3


Content (%) 0.46 0.86 4.99 26.01 0.55 0.48 65.78 2.49

2
H. Cheng, T. Liu, D. Zou et al. Construction and Building Materials 293 (2021) 123550

Table.2
Concrete Mix Proportions.

Strength grade Slump (mm) W/C Coarse aggregate (kg/m3) Sand (kg/m3) Cement (kg/m3) Fly ash (kg/m3) Water reducing agent (%) Water (kg/m3)
C20 140–180 0.57 1030 870 190 80 2.75 167
C30 140–180 0.47 1040 815 260 80 2.80 166
C45 140–200 0.37 1050 740 345 80 2.90 159

Table.3
Deteriorations Groups of Specimens.

Exposure conditions W/C Concentrations (%) Prism 1 Prism 2 Compressive strength at 28 days (MPa)
Na2SO4 0.37 3% 23 4 38.803
5% 23 4
10% 23 4
0.47 3% 23 4 23.021
5% 23 4
10% 23 4
0.57 3% 23 4 14.039
5% 23 4
10% 23 4
H2O 0.37 – 15 0 38.803
0.47 – 15 0 23.021
0.57 – 15 0 14.039

Note: Prism 1: 150 mm  150 mm  300 mm; Prism 2: 100 mm  100 mm  400 mm

Fig. 1. Sulfate attack test.

Erd ðtÞ ¼ Ed ðtÞ=Ed ð0Þ ð2Þ


where Erd(t) represents the RDEM of concrete specimens after t
days of sulfate attack; Ed(t) denotes the dynamic elastic modulus
of concrete specimens after t days of sulfate attack; and Ed(0) indi-
cates the initial dynamic elastic modulus of concrete specimens.
A servo-hydraulic testing machine with a capacity of 5000 kN
(YAS-5000) was used to conduct the compressive testing. The dis-
placement loading method was employed at a rate of 0.01 mm/s.
The detailed testing method followed the GB 50081-2019
guidelines.

2.3.3. Determinations of sulfate ions distributions


Fig. 2. Concrete core.
The distribution of sulfate ions at different depths in concrete
were determined using core samples of concrete specimens as
depicted in Fig. 2 at various periods of 60, 120, 180, 240, and 3 min. Besides, 25 pieces of diamond cutting blade were prepared
300 days. The sampling depth was set to 20 mm, and the corre- for blade replacement. The concrete slices were first dried in a dry-
sponding thickness for each slice was 2 mm. In this test, a precision ing oven at the temperature of 40 °C for 24 h and then crushed in a
cutting machine was used to cut the concrete slices. During the ceramic mortar. The larger particles, including sand or stones, were
process, water is needed to cool down the machine, in order to removed, and the remaining particles were ground repeatedly until
avoid sulfates dissolved and leached away, limited amount of all the particles could pass through a sieve with a diameter of
water was carefully sprayed on the blade to cool down every 2– 0.045 mm.

3
H. Cheng, T. Liu, D. Zou et al. Construction and Building Materials 293 (2021) 123550

The sulfate ions content in each layer of the four core samples of ucts, namely gypsum and ettringite; and (3) the depositions of sul-
concrete powder was determined by chemical analysis based on fate salt crystals resulting from water evaporations during the dry–
ASTM C114-15 [50]. The steps for this analysis are presented in wet cycle. In the second stage, the cement hydration is completed,
detail in Fig. 3. First, the weighed concrete powder was dissolved and the mass increase is attributed to the continuations of sulfate
in hydrochloric acid; then, barium chloride was added to the solu- attack and depositions of sulfate salt crystals. When there are too
tions to obtain barium sulfate precipitations. The suctions filtering excessive deposited expansive products concentrating in concrete
machine was used to obtain the barium sulfate precipitations from initial pores and cracks, a tension develops which is greater than
the mixture solutions. The mean value of the four measurements the tensile strength of the concrete material. Under this condition
was taken as the sulfate ions content of each layer. However, the in the experiment, new cracks were generated, and the scaling and
value, exceeding 15% of the mean value, was invalid. The percent- peeling of cement produced a decrease in the mass of the concrete
age of sulfate ions content at different depths of attacked concrete specimens. However, for the concrete with a larger W/B (0.57), the
can be calculated according to Eq. (3): mass loss occurred in a two-stage change, namely a rapidly
increasing period followed by a falling period. This difference
0:412  ðm3  m2 Þ stems from the rapid diffusions of sulfate ions caused by the larger
w¼ ð3Þ
m1 W/B. During the first stage for the concrete with a larger W/B
where w represents the percentage of SO2- (0.57), the increase in mass was also attributed to further cement
4 content (%); 0.412 indi-
cates the conversions coefficient factor of the relative molecular hydrations. Also, a large amount of ettringite and sulfate salt crys-
mass between barium sulfate and sulfate tetraoxide; m3 represents tal was already deposited in the initial pores or cracks of the con-
the total mass of barium sulfate and filter membrane; m2 denotes crete, shown in the Fig. 4. With further development of expansive
the mass of filter membrane; and m1 represents the mass of con- products and more deposited salt crystal, there was a rapid forma-
crete powder. tion of cracks leading to spalling and peeling off, and the mass loss
reached the falling stage. [51] It can be concluded that a larger W/B
can cause a greater mass change at both the increasing and falling
3. Results and discussions stages. Because a larger W/B provide more pathways for sulfate
ions diffusions.
3.1. Mass loss Another influential factor during the sulfate attack process is
the concentrations of the sulfate solutions. Under general condi-
Fig. 5 illustrates the development of mass loss with sulfate tions, a larger concentrations of sulfate solutions could promote
attack time. According to Fig. 5a and b, the mass loss for concrete the diffusions of sulfate ions into the concrete specimens. How-
with W/B of 0.37 and 0.47 consists of three main stages, including ever, during the dry–wet cycle, more sulfate crystals could be
a rapidly rising stage, a slow rising stage, and a falling stage. During formed when the solutions evaporate. The combined influence of
the first stage, there are three main reasons for the rapid increase more generated chemical products and deposited salt crystals
of the mass of concrete specimens: (1) the further hydrations of accelerates the process of sulfate attack in terms of chemical and
cement; (2) the new generations of sulfate attack reactions prod-

Fig. 3. Sulfate ions titrations experiment.

4
H. Cheng, T. Liu, D. Zou et al. Construction and Building Materials 293 (2021) 123550

Fig. 4. Sulfate attack chemical products.

Fig. 5. Mass loss ratio of concrete under dry-wetting sulfate attack.

physical perspectives [51–53]. As illustrated in Fig. 5, it is interest- 3.2. Dynamic elastic modulus
ing to observe that the mass increase at the early stage is not
always proportional to the sulfate solutions concentrations. A Fig. 6 presents the variations of the dynamic elastic modulus for
higher solution concentration does not always lead to greater each dry–wet cycle. For concrete with W/B of 0.37, 0.47, and 0.57,
weight gain at the early increasing stage. However, the descent the dynamic elastic modulus undergoes three stages: a rapidly
rate of the mass loss ratio at the falling stage is positively related increasing stage, a steady stage, and a falling stage. Prior to 75 days
to the solutions concentrations. It can thus be concluded that a sul- of exposure, cement hydrations contributed to the compactness of
fate solution with a higher concentration can cause severe damage the concrete. During this time, the sulfate attack was mainly con-
to concrete specimens in the falling stage. centrated on the surface area of the concrete. The depositions of

5
H. Cheng, T. Liu, D. Zou et al. Construction and Building Materials 293 (2021) 123550

ettringite and gypsum caused the concrete to condense, which trend. Except for the control concrete group with a W/B of 0.37, the
increased the stiffness of the concrete. Additionally, the deposited compressive strength decreases from 41.90 MPa to 37.84 MPa
sulfate salt crystal during the dry–wet cycle also improved the when the time increases from 240 days to 300 days. It can also
compactness of the sulfate-attacked concrete. Therefore, during be observed that the compressive strength of the control concrete
the first stage, the dynamic elastic modulus increased quickly. increases continuously with time, while the compressive strength
Combined with continuously generated expansive products and of attacked concrete first increases and then decreases with time.
the crystallizations of sodium sulfate during the drying cycle, The inconsistent test results for the control concrete group may
micro cracks developed, and the dynamic elastic modulus began be caused by the variations of raw materials or the fabrications
to decrease rapidly [51–54]. process of the specimens.
As presented in Fig. 6a, the influence of solutions concentrations Based on the Fig. 7, the compressive strength of concrete spec-
on the dynamic elastic modulus is similar to the influence of solu- imens in water is almost continuously increasing, which means
tions concentrations on mass loss. At the dropping stage, the des- that the cement hydration is generally throughout the whole
cent rate of dynamic elastic modulus with higher solutions attack test span. Therefore, as to concrete specimens under sulfate
concentrations is significantly larger than that of lower solutions attack, the strength change is jointly influenced by cement hydra-
concentrations. It is also observed that the dynamic modulus of tions and sulfate attack. To avoid the influence of cement hydra-
concrete with a higher W/B began to decrease at an earlier age, tions on compressive strength degradations and more accurately
which indicates that the expansive damage caused by chemical analyze the influence of sulfate attack on compressive strength of
reactions and crystallizations first occurred in concrete with a concrete specimens, a deteriorations factor, Kf(t), is introduced to
higher W/B. express the deteriorations of compressive strength. Kf(t) can be
defined as follows:

3.3. Compressive strength f s ðtÞ


Kf ðtÞ ¼ ð4Þ
f o ðtÞ
Fig. 7 presents the evolutions of the compressive strength for all
groups of concrete specimens after each testing period. The com- where Kf(t) represents the deteriorations factor of compressive
pressive strength of the three groups of concrete presents a similar strength after t days of sulfate attack; fs(t) indicates the compres-

Fig. 6. Dynamic elastic modulus of concrete under dry-wetting sulfate attack.

6
H. Cheng, T. Liu, D. Zou et al. Construction and Building Materials 293 (2021) 123550

Fig. 7. Compressive strength of concrete under dry-wetting sulfate attack.

sive strength of concrete specimens under sulfate attack after t of deteriorations indicated by fs(t) and Kf(t) are not consistent. For
days; and f0 denotes the compressive strength of concrete speci- instance, as shown in Fig. 8(a), the Kf(t) of the concrete specimen
mens curing in water for t days. with a W/B = 0.37 in the 10% sulfate solutions is less than 1 after
The deteriorations factor is the ratio of strength of attacked con- been attacked for 120 days, which indicates that sulfate attack
crete to strength at the same soaked time. As discussed before, the no longer strengthens the compressive strength of concrete but
strength of sulfate attacked concrete is jointly influenced by begins to cause damage. However, as presented in Fig. 7(a), the
cement hydrations and sulfate attack, so, strength of sulfate compressive strength of attacked concrete begins to drop after
attacked concrete could be expressed as the follow: 180 days of exposure. This indicates that although the sulfate
attack began to damage the concrete after 120 days, the strength
f s ðtÞ ¼ f 0 ðtÞ þ f sulfate ðtÞ ð5Þ
increase caused by cement hydrations was greater than the
where, fsulfate(t) refers to the strength change contributed by sulfate strength decrease caused by sulfate attack until 180 days. The main
attack. And Kf(t) could be expressed as the following equations: reaseon for this phenomenon is that, from 120 days to 180 days,
the strength rediction due to sulfate attack was always accumulat-
f 0 ðtÞ þ f sulfate ðtÞ ing until the point when strength reduction is equal to the strength
Kf ðtÞ ¼ ð6Þ
f 0 ðtÞ increase resulted from cement hydration. After this point, 180 days,
sulfate attack was intensified which leads to the strength reduction
Therefore, when Kf(t) is larger than 1.0, it refers that fsulfate(t) is
which is larger than strength gain from cement hydration. Under
larger than 0 Mpa, which indicates sulfate attack play a positive
this condition, the turning point of compressive strength appeared,
influence on strength of concrete due to deposited chemical prod-
the compressive strength started to decrease. As for the influence
ucts. When Kf(t) is equal to 1.0, it refers that fsulfate(t) is equal to
of solutions concentrations on Kf(t), it can be clearly observed in
0 MPa, which indicates a critical state that sulfate attack no longer
Fig. 8 that the variations of Kf(t) with solutions concentrations all
strengthens the compressive strength of concrete and is about to
increase at the early stage and decrease later. The larger the solu-
lead strength reductions. When Kf(t) is smaller than 1.0, it refers
tions concentrations, the greater the gain of Kf(t) at the increasing
that fsulfate(t) is minus, which indicates that strength reduction
stage, and the greater the loss of K(t) at the dropping stage. It is
was resulted from sulfate attack.
noted that as to concrete specimens with W/B = 0.37, the increase
Fig. 8 presents the variations of Kf(t) for all attacked concrete
of strength and K(t) of specimens in 10% sulfate solutions is not lar-
specimens. Generally, the Kf(t) of attacked concrete presents a sim-
ger than that of specimens in 3% sulfate solutions, as shown in the
ilar change as that of compressive strength, but the turning points
7
H. Cheng, T. Liu, D. Zou et al. Construction and Building Materials 293 (2021) 123550

Fig. 8. Kf(t) of concrete under dry-wetting sulfate attack.

Fig. 7(a) and Fig. 8(a). The main reason could be attributed to that
the compressive strength of concrete in 10% sulfate solutions did
reach to peak on one day within the range from 60 days to
120 days. And the strength peak of concrete in 10% sulfate solu-
tions could be truly larger than that of concrete in 3% sulfate solu-
tions. However, the peak strength could not be measured due to
compressive strength is measured every 60 days.
Fig. 9 presents the variations of Kf(t) with different W/Bs
exposed to 10% Na2SO4 solutions after each testing period. Prior
to 60 days, the Kf(t) increases with the W/B. In general, concrete
with a higher W/B has more paths and pores connected to the
external environment. As can be observed in Fig. 9, at the same
deteriorations time, more sulfate ions diffused into the concrete
specimens and reacted with the cement to produce ettringite.
The ettringite performed a filling effect on the concrete, and an
increase in densifications introduced a larger Kf(t). It is worth not-
ing that the Kf(t) of concrete with a W/B = 0.37 first started to less
than 1, which does not indicate that concrete with a lower W/B is
Fig. 9. Kf(t) of concrete with different W/C under dry-wetting sulfate attack.
more susceptible to severe damage. The concrete with a low W/B
generally has fewer initial pores and cracks. Thus, the capacity
for the accommodations of deposited products is limited, which
concrete with a W/B = 0.47, which indicates that a larger W/B
leads to a smaller increase in compressive strength. The test results
can accelerate the sulfate attack. After 300 days, the sulfate attack
indicate that when erosions product accumulates, the new cracks
damage increased with the W/B. The W/B is the primary factor
generate quickly, which causes a decrease in Kf(t). The Kf(t) of
influencing the deteriorations progress of sulfate attack such that
concrete with a W/B of 0.57 dropped below 1 sooner than that of
a larger W/B could accelerate the progress of sulfate attack. On

8
H. Cheng, T. Liu, D. Zou et al. Construction and Building Materials 293 (2021) 123550

the other hand, a smaller W/B could decrease the permeability of in the first layer is due to the formations of peak concentrations
concrete, which improves the resistance of concrete to sulfate of sulfate ions. The peak concentrations of sulfate ions may be
attack. attributed to the precipitations of solid sulfates at a certain depth
such as ettringite and gypsum formed by the chemical reactions
between the hydrates and penetrating sulfate ions. By increasing
3.4. Distributions of sulfate ions in concrete the immersions time, more sulfate ions penetrate the concrete,
and the penetrations depth increases with the diffusions process.
After removing the initial sulfate ions content, the distributions Meanwhile, the peak concentrations of sulfate ions increases and
of sulfate ions in all groups of concrete with different attack ages is shifts to greater depths [55,56]. As to concrete specimens with a
presented in Figs. 10–12. The initial sulfate ions content for con- lower water-binder ratio, the sulfate attack progress is hindered
crete specimens with W/Bs of 0.37, 0.47, and 0.57 are 1.0%, by the denser and more compact concrete matrix, the ions peak
0.91%, and 0.73%, respectively. Fig. 10 indicates that the overall concentrations layer hardly penetrates the concrete inner part,
trend of sulfate ions distributions in concrete with a W/B of 0.37 only located at the first layer.
in different concentrations of sulfate solutions is similar. The closer Table 4 lists the attack depths of all groups of concrete speci-
to the concrete surface, the greater the sulfate ions content is. The mens. The sulfate attack depth is defined as the point where sulfate
sulfate ions content in the first layer is significantly higher than ions concentrations is 0%. It is generally shown that the attack
that of the second layer, as illustrated by a sudden dropping phe- depth of all attacked concrete gradually increases with deteriora-
nomenon. With the attack time, the sulfate ions content in all lay- tions progress. However, further analysis indicates that there are
ers increased, indicating the progress of deterioration. Figs. 11 and several limitations of using sulfate attack depth as the main indica-
12 show the time-varying distributions of sulfate ions in concrete tor of sulfate–attacked concrete.
with W/B of 0.47 and 0.57, respectively. The sulfate ions were con- First, the attack depth based on experimental results can only
centrated in the first layer and decreased as the depth increased at roughly describe the progress of sulfate attack. For instance, the
the early ages from 60 days to 180 days, as observed in Fig. 10. As attack depth of concrete with a W/B of 0.47 in the 5% Na2SO4 solu-
the sulfate attack progressed, the sulfate ions content in each layer tions does not increase from 120 days to 180 days. Whereas, at the
of concrete increased. However, after 240 days, the sulfate ions same time, the sulfate attack deteriorations continuously progress.
content in the first layer dropped and became concentrated in Moreover, sulfate attack depth cannot identify the influence of
the second layer rather than the first layer. This phenomenon is different factors on sulfate attack deteriorations. For instance, after
quite different from the distribution’s behaviors shown in the 240 days of exposure, the same attack depth was detected in the
Fig. 10. The main reason for the changing behavior of sulfate ions

Fig. 10. Sulfate ions distributions of concrete with w/b = 0.37.

9
H. Cheng, T. Liu, D. Zou et al. Construction and Building Materials 293 (2021) 123550

Fig. 11. Sulfate ions distributions of concrete with w/b = 0.47.

concrete specimens with a W/B of 0.57 in both 5% and 10% Na2SO4. ions distributions curve, mass ratio, and depth axes as illustrated
However, the deteriorations rate of these two groups of concrete in Fig. 13. Table 5 presents the integral area of all attacked concrete
specimens is different. The compressive strengths of concrete spec- with soak time. The integral area increases with the attack time. At
imens in 5% and 10% sulfate solutions are 21.18 MPa and 19.4 MPa, the same deteriorations period, the integral area increases monoto-
respectively, and the mass loss ratios for these two groups of con- nously with the W/B and sulfate solutions concentrations. There-
crete specimens are 0.75% and 0.62%. Furthermore, the dynamic fore, the integral area of sulfate ions distributions can more
elastic modulus is 36.61 GPa and 35.83 GPa. Therefore, although specifically describe the progress of sulfate attack.
equal sulfate attack depths are detected in these two conditions The time-varying model of the integral area of sulfate ions has
of concrete, the deterioration of specimens in 10% sulfate solutions been established by using multivariate fitting functions in Origi-
is more serious. nLab Pro 9.1. The test data was fitted to formulate the following
Finally, another limitation of sulfate depth is that aggregates equations:
could interfere the diffusions of sulfate ions. Merely based on sul- 4:345
fate attack depth, the damage level of sulfate-attacked concrete SðtÞ ¼ ð0:882  1:177  w=bÞ  1:371  ðw=bÞ
could not be revealed. For instance, after 300 days of exposure, ð1:544 þ 10:912  scÞ  ð93:913 þ 1:497  tÞ ðR  S ¼ 0:91Þ
for concrete specimens with a W/B of 0.57 in both 5% and 10% sul- ð7Þ
fate solutions, the depth for specimens in 5% sulfate solutions is
19 mm, whereas the depth for concrete in 10% sulfate solutions where S(t) represents the integral area of sulfate ions distributions
is 17 mm. However, the deteriorations of concrete in 10% sulfate after t days of sulfate attack; w/b represents the W/B of the concrete
solutions is more severe. The compressive strength for these two specimen; and sc indicates the concentrations of external sulfate
groups of concrete is 17.8 MPa and 17.1 MPa, respectively. The solutions.
mass loss ratios are 0.04% and 0.23%, and the dynamic elastic
moduli are 32.88 GPa and 31.85 GPa, respectively. The same situ- 4. Deterioration analysis of the damaged area of concrete
ations can also be found in the concrete specimens with a W/B of
0.37 in 3% and 5% Na2SO4 after 240 days of sulfate attack. 4.1. Definitions and basic assumptions of the damaged area
Based on the above discussions, by only using the sulfate attack
depth, the deteriorations process cannot be fully expressed, and The deteriorations level is not uniformly distributed in the
the influence of various factors on the sulfate–attack process can- sulfate-attacked concrete. Generally, the deterioration begins at
not be identified. To conduct further quantitative analysis, the inte- the surface and diminishes as depth increases. For simplifications,
gral area of sulfate ions distributions is introduced. The integral the cross–section of attacked concrete is divided into a deterio-
area of sulfate ions is defined as the area enclosed by the sulfate rated area and a healthy area as presented in Fig. 14. The load bear-
10
H. Cheng, T. Liu, D. Zou et al. Construction and Building Materials 293 (2021) 123550

Fig. 12. Sulfate ions distributions of concrete with w/b = 0.57.

Table 4
The depth of sulfate attack (mm).

W/C Concentrations (%) 60 d 120 d 180 d 240 d 300 d


0.37 3% 3 7 9 13 15
5% 3 7 9 11 19
10% 5 11 13 19 19
0.47 3% 7 9 11 11 17
5% 11 13 13 17 17
10% 13 13 15 17 17
0.57 3% 13 13 13 17 19
5% 15 15 15 17 19
10% 15 17 17 17 17

ing capacity of the attacked–concrete specimen (fT  A) is consid-


ered as a summation of the bearing capacity of the deteriorated
concrete (fc  Ac) and the healthy concrete (fss  As).
The relationship between the three strengths is described by Eq.
(6) according to a Taylor–type homogenization [57,58]:

f t  A ¼ f c  Ac þ f ss  As ð8Þ
Fig. 13. Integral area of sulfate ions distributions.
Therefore, the average compressive strength of the deterio-
rated–area concrete can be calculated according to Eq. (7):
tively. In this study, the value of fc is assumed to be equal to the
f ss ¼ ðf T  A  f c  Ac Þ=As ð9Þ compressive strength measured from the compressive testing on
the concrete specimens soaked in water.
where fT, fc, and fss represent the compressive strengths of the
In this study, the sulfate ions mainly concentrate in the outer
whole attacked–concrete specimen, healthy area, and deteriorated
regions, and the sulfate attack depth is always within 20 mm.
area, respectively, and A, Ac, and As represents the total area,
For simplifications, the range from concrete surface to a depth of
healthy area, and deteriorated area of the attacked concrete, respec-
11
H. Cheng, T. Liu, D. Zou et al. Construction and Building Materials 293 (2021) 123550

Table 5
The integral areas of sulfate ions distributions (mm%).

W/C Concentrations (%) 60 d 120 d 180 d 240 d 300 d


0.37 3% 0.424 1.115 1.5632 3.809 6.941
5% 1.477 3.405 3.6867 6.403 12.948
10% 2.197 5.316 8.4291 11.100 17.387
0.47 3% 3.006 8.000 10.918 11.993 17.960
5% 4.864 8.881 11.420 16.566 21.529
10% 9.344 12.628 16.728 18.498 26.074
0.57 3% 7.264 12.321 14.348 25.538 28.693
5% 14.266 16.171 17.486 22.066 28.908
10% 14.401 20.004 24.106 29.472 31.315

Fig. 14. Schematic diagram of sulfate-attacked concrete.

20 mm is considered as the deteriorations area. Thus, the integral crete matrix thus destroying the bonding strength among aggre-
area of sulfate ions distributions within 5 mm to 20 mm is calcu- gates and cement matrix [13,14]. Meanwhile, sulfate ions react
lated to identify the deteriorations level of concrete specimens. with the cement hydrations product to form soft and soluble gyp-
sum. subsequently, this gypsum reacts with the alumina to form
4.2. Degradations of mechanical properties ettringite. Initially, ettringite crystals mainly deposit in the initial
air voids of concrete. With the further reactions, too excessive
The degradations of the compressive strength of deteriorated– amount of ettringite lead to expansive stress and cracks [15–17].
area concrete is presented in Fig. 15. The compressive strength of The newly generated cracks make the concrete more porous, sus-
all groups of concrete specimens that underwent sulfate attack ceptible to further sulfate attack, which leads to a high-rate reduc-
presents a similar change. It initially increases with attack time tion in strength [20]. The deterioration is not uniform across in
and decreases as further deteriorations occurs. The concrete spec- concrete, namely, the damage decreases along with depth. There-
imens with a higher W/B and in a higher concentrations of sulfate fore, concrete under sulfate attack could be divided into two main
solutions show a greater increase of compressive strength initially parts, including attacked part and healthy part shown in the
followed by a greater decrease during the falling stage. Fig. 16. Because sulfate attack only concentrates in attacked part
of concrete, so the strength of healthy part of concrete can be
assumed to be equal to the strength of normal concrete. Thus,
5. A novel strength assessment method based on sulfate ions the key to strength assessment of deteriorated concrete is to esti-
distributions mate the compressive strength of the damaged area of concrete.
Based on the analysis presented in the previous sections, the
5.1. Basic assumptions and indicator integral area of sulfate ions distributions can be used as a better
indicator to express the process of sulfate attack and reveal the
Sulfate ions from outside environment diffuse into concrete influence of the W/B and concentrations of sulfate solutions on
based on the Fick’s law, combining with the C-S-H gel of the con-
12
H. Cheng, T. Liu, D. Zou et al. Construction and Building Materials 293 (2021) 123550

Fig. 15. Compressive strength of deteriorated area of concrete.

Fig. 18. Schematic diagram of assessment method of sulfate attacked concrete.

13
H. Cheng, T. Liu, D. Zou et al. Construction and Building Materials 293 (2021) 123550

Fig. 16. The process of sulfate attack.

pared to currently assessing indicators, the distributions of sulfate


ions can be directly measured by sampling small portions of con-
crete. Therefore, the integral area of sulfate ions distributions is
considered as the primary indicator to assess the compressive
strength of attacked concrete.
In this experiment, the deteriorated area is considered as the
range from the surface of the concrete to the depth of 20 mm,
where sulfate ions are mainly concentrated. The formulations
between the integral area of sulfate ions distributions and com-
pressive strength is established to for strength assessment. Even
though the sulfate ions content of each layer is determined by
using the average sulfate content measured in the powder of four
concrete samples, the influence of randomly distributed aggregate
on sulfate ions distributions cannot be overlooked. Thus, in this
study, in order to mitigate this negative influence, random sample
consensus (RANSAC) was used to establish the relationship
between the integral area of sulfate ions content and compressive
strength of attacked part of concrete shown in the Fig. 17. The
Fig. 17. The relationship between integral area and strength. RANSAC algorithm is a parameter estimations approach used to
fit a model to experimental data with many outliers. The algorithm
generates solutions using a minimum number of random data
the attack process. Furthermore, compared to mass loss ratio and points that are required to estimate the model parameters. The
dynamic elastic modulus, this indicator is in a continuous accumu- model identifies how many points are classified as inliers (i.e.
lations process without an initial increasing stage or a falling stage, points that fit the model) according to a predetermined tolerance.
which can more accurately identify the deteriorations level of con- The fractions of inliers over the total number of points in the set is
crete. Most importantly, in the actual applications conditions, com- used to select the model with larger consensus through an iterative

14
H. Cheng, T. Liu, D. Zou et al. Construction and Building Materials 293 (2021) 123550

Fig. 19. Strength comparison between M1, M2, and M3.

process where the data is resampled a fixed number of times sulfate ions distributions. The residual_threshold is maximum
[59,60]. In this study, RANSAC regressor algorithm from scikit- residual for a data sample to be classified as an inlier. By default,
learn was carried out to establish the relationship between com- the threshold is chosen as the MAD (median absolute deviations)
pressive strength of attacked area of concrete and integral area of of the target dataset (compressive strength of attacked area of con-

Fig. 20. verifications test.

15
H. Cheng, T. Liu, D. Zou et al. Construction and Building Materials 293 (2021) 123550

Table.6
Comparison of various assessment methods.

Experiment M1 M2 M3
fs (MPa) 17.97 19.95 19.72 19.27
Error (%) 0 11.02 9.74 7.2

crete). In this study, the MAD is determined by automated multi- surface concrete was severely damaged, the sulfate ions became
round calculation, which is 7.73 MPa. The outlier refers to the inte- concentrated in the third layer. From 5 mm to 19 mm, the sulfate
gral area of sulfate ions distributions which is heavily influenced ions content decreased with depth. The integral area of sulfate ions
by randomly distributed aggregates. distributions is 30.57 (mm%).
The equations for assessing the deteriorations of compressive Further comparison is presented in Table 6. Two additional pre-
strength is: dictions methods are used to predict the compressive strength of
concrete specimens. In this verifications process, the sulfate attack
f ss ¼ 1:0796  S þ 43:8714 ð10Þ time is 360 days, the W/B is 0.47, the concentration of the sulfate
where fss represents the compressive strength of attacked-area con- solutions is 10%, and the compressive state is 0 MPa. M3 represents
crete; S represents the integral area of sulfate ions distributions. the method proposed in the present study. As shown in Table 5, the
errors for each method are 11.02%, 9.74%, and 7.20%, respectively.
Notably, M3 can fit the measurements well and has a smaller rel-
5.2. Assessment methodology
ative error. It can thus be concluded that the proposed method
based on the sulfate ions distributions and homogenizations the-
The deteriorated concrete is divided into two areas, including
ory can be used as a more accurate method to predict the compres-
the deteriorated area and healthy area. The strength of the healthy
sive strength of concrete under sulfate attack in a long term.
area of concrete is assumed to equal to the strength of concrete
Most importantly, compared to other strength assessment
specimens in water. The strength of the damaged area of concrete
methods, the only requirement for this proposed method is to
can be calculated using Eq. (10) by using measured integral area of
measure sulfate ions distribution, which makes this method easier
sulfate ions distributions. Based on the theory of homogenizations,
and more practicable than other assessment methods. Because, in
the strength of attacked concrete can be calculated using the com-
actual conditions, it is difficult to determine accurate attack time,
binations of healthy and damaged parts. The detailed procedures
elastic modulus and peak strain. Therefore, the newly proposed
are presented in Fig. 18.
assessment method based on sulfate ions distributions and
homogenizations provides a potential application in assessing the
5.3. Verifications of assessment methodology loading capacity of actual concrete structures exposed to aggres-
sive environments.
To carry out verification analysis on the proposed method, con-
crete specimens with water-binder ratio 0.37, subjected to 3%, 5%
and 10% sulfate solution, are used to prove the accuracy and feasi- 6. Conclusions
bility of the new method (denoted as M3). Two additional predic-
tions methods are also used to assess the compressive strength of This study investigated the deteriorations progress of concrete
concrete specimens. M1 represents the method proposed by Jiang exposed to sulfate attack and dry–wet cycle. The non–uniform
[61]. The compressive strength of sulfate-attacked concrete could damage characteristics of sulfate–attacked concrete were dis-
be predicted by the sulfate attack time, the concentrations of sul- cussed by dividing the concrete specimens into deteriorated area,
fate solutions, and the W/B. M2 refers to the method established and healthy area. A new indicator, the integral area of sulfate ions
by Cao [62]. In his study, the time-dependent stress strain relation- distributions, was proposed to express the deteriorations progress
ship of sulfate-attacked concrete in a compressive state and dry– of attacked concrete, and a novel assessment method based on sul-
wet cycles were proposed. The compressive strength of sulfate- fate ions distributions was established. The conclusions are sum-
attacked concrete could be calculated by determining the attack marized as follows:
time, W/B, sulfate solutions concentrations, elastic modulus, peak
strain, and compressive state. Based on the strength comparison (1) The mass loss ratio and dynamic elastic modulus of sulfate–
between M1, M2, and M3 shown in the Fig. 19, M3 achieves the attacked concrete present three main stages: a rapidly
smallest error in all 15 cases considered. Therefore, it can conclude increasing stage, a steadily increasing stage, and a dropping
that the proposed method, M3 is not only reliable but also more stage. Additionally, the compressive strength presents two
accurate for short- and long-term strength assessment of sulfate main stages: an initially increasing stage and a continuously
attacked concrete. dropping stage. Notably, a higher solutions concentration
A further verification was carried out by conduction a long-term does not always lead to a greater increase in mass or
(360 days) strength prediction. Both mechanical tests and chemical dynamic modulus at the early increasing stage.
titrations experiments were conducted on the concrete specimens (2) Compared to the sulfate attack depth, the integral area of
with W/B of 0.47 after 360 days of sulfate attack exposed to 10% sulfate ions distributions is a better indicator to describe
Na2SO4 solutions. The strength of the healthy area of concrete is the progress of sulfate attack. The index of the integral area
assumed to equal to the strength of concrete in water after increases monotonously with deteriorations time.
300 days. According to the experimental results of concrete in (3) The established formulation between compressive strength
water, the hydrations process was almost complete after 300 days; and integral area of sulphate ions distributions can be used
the compressive strength is 27.5 MPa. The compressive strength of to determine the deterioration of sulfate–attacked concrete.
sulfate–attacked concrete after 360 days of exposure is 17.97 MPa, The proposed assessment method can be used to accurately
as presented in Fig. 20(a). The sulfate ions distributions of sulfate– and practically assess the loading capacity of actual concrete
attacked specimens is presented in Fig. 20(b). Given that the structures exposed to aggressive environments.
16
H. Cheng, T. Liu, D. Zou et al. Construction and Building Materials 293 (2021) 123550

[24] B. Tian, M.D. Cohen, Does gypsum formations during sulfate attack on concrete
lead to expansions?, Cem. Concr. Res. 30 (2000) 117–123.
Declaration of Competing Interest [25] American Society for Testing and Materials, Standard Test Method for Length
Change of Hydraulic-Cement Mortars Exposed to a Sulfate Solutions, ASTM
The authors declare that they have no known competing finan- C1012. West Conshohocken, United States, 2015.
[26] Chinese Nationsal Standard, Standard for Test Method of Long-term
cial interests or personal relationships that could have appeared Performance and Durability of Ordinary concrete, GB 50082-2012. Beijing,
to influence the work reported in this paper. China, 2012.
[27] Z. Zhang, J. Zhou, J. Yang, Y. Zou, Z. Wang, Understanding of the deterioration
characteristic of concrete exposed to external sulfate attack: insight into
Acknowledgments mesoscopic pore structures, Constr. Build. Mater. 260 (2020) 119932.
[28] Z.Y. Zhang, J.T. Zhou, J. Yang, Y. Zou, Z.S. Wang, Cracking characteristics and
This study was financially supported by the National key pore development in concrete due to physical attack, Mater. Struct. 53 (2020)
1–13.
research and development program of China (Grant No. [29] D.J. Zou, H.B. Cheng, T.J. Liu, S.S. Qin, T.H. Yi, Monitoring of concrete structure
2019YFC1511005), National Science Fund for Distinguished Young damage caused by sulfate attack with the use of embedded piezoelectric
Scholars (Grant No. 52025081), National Natural Science Founda- transducers, Smart Mater. Struct. 28 (2019) 1275–1296.
[30] L. Guo, Y. Wu, P. Duan, Z. Zhang, Improving sulfate attack resistance of
tions of China (Grant No. 51978211), Key Research and Develop- concrete by using calcined Mg-Al-CO3 LDHs: Adsorptions behavior and
ment Program of Guangdong Province (Grant No. mechanism, Constr. Build. Mater. 232 (2020) 117256.
2019B111107001), and Shenzhen Science and Technology Program [31] D. Mostofinejad, S.M. Hosseini, F. Nosouhian, T. Ozbakkaloglu, B.N. Tehrania,
Durability of concrete containing recycled concrete coarse and fine aggregates
(Grant No. RCYX20200714114525013). and milled waste glass in magnesium sulfate environment, J. Build. Eng. 29
(2020) 101182.
References [32] A.J. Boyd, S. Mindess, The use of tensions testing to investigate the effect of W/
C ratio and cement type on the resistance of concrete to sulfate attack, Cem.
Concr. Res. 34 (2004) 373–377.
[1] J. Geng, D. Easterbrook, L.-Y. Li, L.-W. Mo, The stability of bound chlorides in
[33] D.M. Burgos, Á. Guzmán, S. Delvasto, Assessment of the performance of SCC
cement paste with sulfate attack, Cem. Concr. Res. 68 (2015) 211–222.
incorporating volcanic materials in a sodium sulfate environment, Constr.
[2] N. Cefis, C. Comi, Chemo-mechanical modelling of the external sulfate attack in
Build. Mater. 195 (2019) 52–65.
concrete, Cem. Concr. Res. 93 (2017) 57–70.
[34] Y. Tan, H. Yu, H. Ma, Study on the micro-crack evolutions of concrete subjected
[3] W. Wongprachum, M. Sappakittipakorn, P. Sukontasukkul, P. Chindaprasirt, N.
to stress corrosions and magnesium sulfate, Constr. Build. Mater. 141 (2017)
Banthia, Resistance to sulfate attack and underwater abrasions of fiber
453–460.
reinforced cement mortar, Constr. Build. Mater. 189 (2018) 686–694.
[35] Y. L, R.J. Wang, S.Y. Li, Y.Bao, Y. Q, Resistance of recycled aggregate concrete
[4] C.L. Zhang, W.K. Chen, S. Mu, Numerical investigation of external sulfate attack
containing low- and high-volume fly ash against the combined actions of
and its effect on chloride binding and diffusion in concrete, Constr. Build.
freeze–thaw cycles and sulfate attack, Constr. Build. Mater. 166 (2018) 23-
Mater. 285 (2021), https://doi.org/10.1016/j.conbuildmat.2021.122806.
34.
[5] J. Haufe, A. Vollpracht, Tensile strength of concrete exposed to sulfate attack,
[36] Zhuo Tang, Wengui Li, Guojun Ke, John L. Zhou, Vivian W.Y. Tam, Sulfate attack
Cem. Concr. Res. 116 (2019) 81–88.
resistance of sustainable concrete incorporating various industrial solid
[6] A. Neville, The confused world of sulfate attack on concrete, Cem. Concr. Res.
wastes, J. Clean. Prod. 218 (2019) 810–822.
34 (8) (2004) 1275–1296.
[37] Q.H. Xiao, Z.Y. Cao, X. Guan, Q. Li, X.L. Liu, Damage to recycled concrete with
[7] G. Zhao, J. Li, W. Shao, Effect of mixed chlorides on the degradation and sulfate
different aggregate substitutions rates from the coupled actions of freeze-thaw
diffusion of cast-in-situ concrete due to sulfate attack, Constr. Build. Mater.
cycles and sulfate attack, Constr. Build. Mater. 221 (2019) 74–83.
181 (2018) 49–58.
[38] G. Zhao, J. Li, F. Han, M. Shi, H. Fan, Sulfate-induced degradations of cast-in-situ
[8] C. Liang, Z. Cai, H. Wu, J. Xiao, Y. Zhang, Z. Ma, Chloride transport and induced
concrete influenced by magnesium, Constr. Build. Mater. 199 (2019) 194–206.
steel corrosion in recycled aggregate concrete: a review, Constr. Build. Mater.
[39] Q.H. Xiao, Q. Li, Z.Y. Cao, W.Y. Tian, The deteriorations law of recycled concrete
282 (2021) 122547, https://doi.org/10.1016/j.conbuildmat.2021.122547.
under the combined effects of freeze-thaw and sulfate attack, Constr. Build.
[9] G. Zhao, J. Li, F. Han, M. Shi, H. Fan, Sulfate-induced degradation of cast-in-situ
Mater. 200 (2019) 344–355.
concrete influenced by magnesium, Constr. Build. Mater. 199 (2019) 194–206.
[40] Tiejun Liu, Dujian Zou, Jun Teng, Guilan Yan, The influence of sulfate attack on
[10] T. Schmidt, B. Lothenbach, M. Romer, J. Neuenschwander, K. Scrivener, Physical
the dynamic properties of concrete column, Constr. Build. Mater. 38 (2012)
and microstructural aspects of sulfate attack on ordinary and limestone
201–207.
blended Portland cements, Cem. Concr. Res. 39 (2009) 1111–1121.
[41] Z.Y. Zhang, J.T. Zhou, Y. Zou, J. Yang, J. Bi, Change on shear strength of concrete
[11] R. Ragoug, O.O. Metalssi, F. Barberon, J.-M. Torrenti, N. Roussel, L. Divet, J.-B.
fully immersed in sulfate solutions, Constr. Build. Mater. 235 (2020) 117463.
d’Espinose de Lacaillerie, Durability of cement pastes exposed to external
[42] A.E. Idiart, C.M. López, I. Carol, Chemo-mechanical analysis of concrete
sulfate attack and leaching: Physical and chemical aspects, Cem. Concr. Res.
cracking and degradations due to external sulfate attack: a mesoscale model,
116 (2019) 134–145.
Cem. Concr. Compos. 33 (2011) 411–423.
[12] L. Wei, J. Xiao-Guang, Z. Zhong-Ya, Triaxial test on concrete material
[43] X. Zuo, W. Sun, C. Yu, Numerical investigations on expansive volume strain in
containing accelerators under physical sulphate attack, Constr. Build. Mater.
concrete subjected to sulfate attack, Constr. Build. Mater. 36 (2012) 404–410.
206 (2019) 641–654.
[44] F. Xie, J.P. Li, L. Li, G.W. Zhao, M.B. Yao, Numerical solutions and damage
[13] G. Zhao, J. Li, M. Shi, H. Fan, J. Cui, F. Xie, Degradations mechanisms of cast-in-
evaluations for cast-in-situ piles exposed to external sulfate attack, Constr.
situ concrete subjected to internal-external combined sulfate attack, Constr.
Build. Mater. 214 (2019) 269–279.
Build. Mater. 248 (2020) 118683.
[45] S. Sarkar, S. Mahadevan, J.C.L. Meeussen, Numerical simulations of
[14] J.P. Li, M.B. Yao, W. Shao, Diffusions-reactions model of stochastically mixed
cementitious materials degradations under external sulfate attack, Cem.
sulfate in cast-in-situ piles, Constr. Build. Mater. 115 (2016) 662–668.
Concr. Compos. 138 (2010) 241–252.
[15] C. Sun, J. Chen, J. Zhu, M. Zhang, J. Ye, A new diffusions model of sulfates in
[46] Keshu Wan, Lin Li, Wei Sun, Li L, Sun W, Solid-liquid equilibrium curve of
concrete, Constr. Build. Mater. 39 (2013) 39–45.
calcium in 6 mol/L ammonium nitrate solutions, Cem. Concr. Res. 53 (2013)
[16] S. Sarkar, S. Mahadevan, J.C.L. Meeussen, H.V.D. Sloot, D.S. Kosson, Numerical
44–50.
simulations of cementitious materials degradations under external sulfate
[47] W.L. Qiu, F. Teng, S.S. Pan, Damage constitutive model of concrete under
attack, Cem. Concr. Comp. 32 (2012) 241–252.
repeated load after seawater freeze-thaw cycles, Constr. Build. Mater. 236
[17] S.C. Devi, R.A. Khan, Influence of graphene oxide on sulfate attack and
(2020) 117560.
carbonations of concrete containing recycled concrete aggregate, Constr. Build.
[48] B. Wang, F. Wang, Q. Wang, Damage constitutive models of concrete under the
Mater. 250 (2020) 118883.
coupling actions of freeze–thaw cycles and load based on Lemaitre
[18] F. Liu, Z. You, A. Diab, Z. Liu, C. Zhang, S. Guo, External sulfate attack on
assumptions, Constr. Build. Mater. 173 (2018) 332–341.
concrete under combined effects of flexural fatigue loading and drying-
[49] Chinese Nationsal Standard, Standard for Test Method of Mechanical
wetting cycles, Constr. Build. Mater. 249 (2020) 118224.
Properties on Ordinary Concrete, GB 50081-2002. Beijing, China, 2002.
[19] M.H. Alyami, R.S. Alrashidi, H. Mosavi, M.A. Almarshoud, K.A. Riding, Potential
[50] American Society for Testing and Materials, Standard Test Methods for
accelerated test methods for physical sulfate attack on concrete, Constr. Build.
Chemical Analysis of Hydraulic Cement, ASTM C114 – 15. West
Mater. 229 (2019) 116920.
Conshohocken United States, 2015.
[20] D. Sun, K. Wu, H. Shi, S. Miramini, L. Zhang, Deformations behavior of concrete
[51] X. Jiang, S. Mu, Z.Q. Yang, J.H. Tang, T. Li, Effect of temperature on durability of
materials under the sulfate attack, Constr. Build. Mater. 210 (2019) 232–241.
cement-based material to physical sulfate attack, Constr. Build. Mater. 266
[21] M. Santhanam, M.D. Cohen, J. Olek, Mechanism of sulfate attack: a fresh look:
(2021) 120936.
part 1: summary of experimental results, Cem. Concr. Res. 32 (2002) 915–921.
[52] P. Liu, Y. Chen, W. Wang, Z. Yu, Effect of physical and chemical sulfate attack on
[22] M. Santhanam, M.D. Cohen, J. Olek, Mechanism of sulfate attack: a fresh look:
performance degradation of concrete under different conditions, Chem. Phys.
part 2. Proposed mechanisms, Cem. Concr. Res. 33 (2003) 341–346.
Lett. 745 (2020) 137254.
[23] Z. Zhang, X. Jin, W. Luo, Long-term behaviors of concrete under low-
[53] G.L. Kalousek, L.C. Porter, E.J. Benton, Concrete for long-time service in
concentrations sulfate attack subjected to natural variations of
sulphate environment, Cem. Concr. Res. 2 (1972) 79–89.
environmental climate conditionss, Cem. Concr. Res. 116 (2019) 217–230.

17
H. Cheng, T. Liu, D. Zou et al. Construction and Building Materials 293 (2021) 123550

[54] R. Douglas Hooton, Current developments and future needs in standards for [59] M.A. Fischler, R.C. Bolles, Random sample consensus: a paradigm for model
cementitious materials, Cem. Concr. Res. 78 (2015) 165–177. fitting with applications to image analysis and automated cartography,
[55] D.J. Zou, S.S. Qin, T.J. Liu, Experimental and numerical study of the effects of Commun. ACM 24 (1981) 381–395.
solution concentration and temperature on concrete under external sulfate [60] Jiangtao Peng, Silong Peng, Yong Hu, Partial least squares and random sample
attack, Cem. Concr. Res. 139 (2021) 106284. consensus in outlier detection, Anal. Chim. Acta 719 (2012) 24–29.
[56] S.S. Qin, D.J. Zou, T.J. Liu, A chemo-transport-damage model for concrete under [61] L. Jiang, D.T. Niu, Study of constitutive relations of concrete under sulfate
external sulfate attack, Cem. Concr. Res. 132 (2020) 106048. attack and dry-wetting cycles, J. China Univ. Min. Technol. 46 (2017) 66–73.
[57] G. Sachs, Plasticity problems in metals, J. Chem. Soc. Faraday. Trans. 24 (1928) [62] J. Cao, Long-term performance of concrete under axial compressions coupled
84–92. effects of sulfate attack and dry-wet cycle, Beijing Jiaotong Univ. (2013).
[58] B. Klusemann, B. Svendsen, Homogenization modeling of thin-layer-type
microstructures, Int. J. Solids. Struct. 49 (2012) 1828–1838.

18

You might also like