Next Article in Journal
In Situ and Ex Situ Raman Studies of Cysteine’s Behavior on a Titanium Surface in Buffer Solution
Previous Article in Journal
High Temperature Oxidation and Oxyacetylene Ablation Properties of ZrB2-ZrC-SiC Ultra-High Temperature Composite Ceramic Coatings Deposited on C/C Composites by Laser Cladding
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

A Review on the Transport-Chemo-Mechanical Behavior in Concrete under External Sulfate Attack

1
School of Civil and Transportation Engineering, Ningbo University of Technology, Ningbo 315211, China
2
College of Art and Design, Nanjing Forestry University, Nanjing 210037, China
3
School of Science, Nanjing University of Science and Technology, Nanjing 210094, China
4
School of Civil Engineering, Yangzhou Ploytechnic College, Yangzhou 225009, China
*
Author to whom correspondence should be addressed.
Coatings 2023, 13(1), 174; https://doi.org/10.3390/coatings13010174
Submission received: 21 December 2022 / Revised: 10 January 2023 / Accepted: 11 January 2023 / Published: 12 January 2023

Abstract

:
Cementitious concrete structures serving in sulfate environments suffer from serious durability challenges caused by chemical sulfate attacks (CSA), which lead to the volume expansion, cracking, and spalling of concrete and the early failure of structures. CSA on concrete involves the behaviors of ion transport, chemical reactions, the crystallization of reaction products, microstructural damage to the cement matrix, and the macroscopic deterioration of concrete, namely the transport-chemo-mechanical behaviors. This paper first introduces the reaction products, such as gypsum, ettringite, brucite, and thaumasite, between sulfate and concrete under different environmental conditions and their formation mechanism. Then, aiming at the ettringite type CSA, the theories of volume increase and crystallization pressure are elaborated to explain it-induced concrete degradation. Additionally, the crystallization pressure theory is used to describe the cracking behavior in the microstructure slurry caused by the ettringite crystal filling pore. Finally, a series of transport-chemo-mechanical models for ettringite type CSA are displaced module by module. It includes the sulfate diffusion-reaction model, the free expansion of concrete, and equivalent expansive force in concrete related to the reaction behavior: the model for chemo-mechanical behavior in concrete caused by CSA. These models can be used to analyze the distribution of sulfate ions and the reaction product content, expansive stress and strain in the concrete, and the cracking and spalling degree of the concrete, which is beneficial to evaluate the durability of concrete structures serving permanently in a sulfate environment.

1. Introduction

External sulfate attack (ESA) is an important environmental factor for durability degradation and service-life reduction in concrete structures [1,2]. According to the survey from the “China National Geographic Website” [3], the sulfate service environment in China is widely distributed. In offshore China, the sulfate content is about 1400~2700 mg/L, which causes underground soil or groundwater in coastal cities, such as Liaoning, Shandong, Jiangsu, and Zhejiang provinces, to contain a large number of sulfate ions. As a result, the service life of concrete structures such as seaports and wharves in these areas is only 25 years [4], and some of them have been severely damaged for between 7 and 20 years [5]. In Xinjiang, Qinghai, Ningxia, Gansu, Inner Mongolia, and other western regions, the sulfate content in some saline soils has reached more than 4200 mg/L, especially in the Korla area of Xinjiang, where the sulfate concentration in the groundwater is as high as 21,299 mg/L. In the Chaerhan area of Qinhai, the service life of concrete structures, such as roads, bridges, and underground pipelines, without anti-corrosion measures is only 3–5 years or even 1 year [6].
In reality, ESA is a global problem in civil engineering. As early as 1915, Wig and Williams (American Bureau of Standards) published an academic paper to report that [7] several concrete structures exposed to saline-alkali soil or water environments in the western barren environment of the USA were to be destructed after 7 years of operation. Up to now, the research on ESA has a history of more than 120 years [8], but Adam Neville [9] pointed out in 2004 that this research was still in a confused state. He summarized two definitions of sulfate attack on concrete: one is that, as long as the interaction between the sulfate and concrete is involved, this behavior can be defined as a sulfate attack. Another is that when the chemical reaction between the sulfate ions and cement hydration products occurs and causes the expansion damage of cement paste or the durability degradation of concrete components, this behavior can be defined as a sulfate attack [10]. Obviously, the first view has a shortcoming if the sulfate in concrete does not cause any damage to the cement paste, the definition of “sulfate attack” is contrary to the essence of an “attack”. Bensted [11] considered that the generation of sulfate products and the it-induced damage of concrete are two different concepts, which should be treated differently, and therefore, Neville [9] proposed that the behaviors, without causing the degradation of concrete durability, cannot be called a sulfate attack. The second view is more reasonable in that the definition proposes two key conditions, namely, ESA inducement (sulfate reaction products) and ESA consequence (expansion and degradation of concrete). As Harrison and Cooke [12] pointed out, although there is a large amount of sulfate ion entering into concrete or reaction products generated in the concrete, it does not mean the occurrence of ESA. Only strength decline or other damage characteristics in the concrete can be further observed, which can provide effective evidence for ESA.
However, it can be known from the above definition that the ESA refers in particular to the chemical sulfate attack (CSA), and correspondingly, another kind of ESA is the physical sulfate attack (PSA). Under specific environmental conditions similar to the drying-wetting alternation and periodic temperature change, the repeated dissolution-crystallization action of sulfate, which causes the fatigue failure of concrete [8]. This sulfate crystallization behavior does not involve the chemical reaction of substances in concrete.
In this paper, the behavior (sulfate transport-chemo-mechanical behavior) of CSA on concrete is the main discussion object. The degradation of concrete structures caused by CSA is the result of long-term corrosion, and the model prediction is a necessary analysis method for the CSA-induced durability problem. To clearly describe the current research status of the CSA model, the reaction products of CSA, their formation mechanism, and them-induced failure forms of concrete are first described simply. Most of these studies were completed between the 1980s and 1990s, and a few were developed after the 2000s. Then, the numerical model for CSA behaviors is presented module by module. It involves the diffusion of sulfate ions, chemical reactions, the it-induced volume expansion of concrete, crystallization pressure in the pore solution, and mechanical responses (expansive stress, strain, damage degree) in the concrete, and most of these works were carried out after 2000. The framework of this paper is shown in Figure 1.

2. Chemical Reaction Products of Sulfate Attack

The important characterization of CSA on concrete is the microstructure damage of cement paste caused by the growth of sulfate reaction products. Modern microscopic measurements, such as scanning electron microscope (SEM), X-ray diffraction (XRD), and mercury intrusion porosimetry (MIP), are utilized to observe the microstructure evolution of cement paste. The results show that the sulfate products form in micropores, including gypsum ( C S ¯ H 2 ) [13,14], ettringite ( C 6 A S ¯ 3 H 32 ) [15,16,17], and thaumasite ( C 3 S C ¯ S ¯ H 15 ) [18,19,20], accompanied by the propagation of microcracks.
In reality, the types of sulfate products are affected by corrosion conditions, such as the solution concentration, pH value, ambient temperature, and cation type. Biczok [21] found that if the concentration of the sodium sulfate solution is lower than 1000 mg/L, the main product is ettringite, while it is mainly gypsum if the concentration is higher than 8000 mg/L. Bellmann [22] found that the greater the pH value of the solution, the higher the sulfate concentration required for gypsum formation. The minimum sulfate concentration required for gypsum formation is about 1400 mg/L, and the corresponding pH of the solution is 12.45. When the pH value rises to 12.9, sulfate ions hardly react with calcium hydroxide to form gypsum. Some studies [19,23,24] pointed out that thaumasite will be formed preferentially in a low-temperature ( 15   ° C ) humid environment with a pH value higher than 10.5 and the appropriate amount of CO 2 or CO 3 2 . However, it is observed in some on-site engineering investigations with a temperature 20   ° C [25,26] that, for cation type, the products of magnesium sulfate attack are different from that of other common basic ions such as Na+/K+ [27,28,29], as shown in Figure 2 [30]. Besides gypsum and ettringite, the insoluble brucite, Mg OH 2 , is also formed to cause a decrease in the pH value of pore solutions and the decalcification of C-S-H gel, which becomes the cohesionless silica gel, SiO 2 x H 2 O , or magnesium silicate, 3 MgO 2 SiO 2 2 H 2 O . In short, sulfate products are different in different corrosion environments.

3. Formation Mechanism of Reaction Products

The formation mechanism of sulfate reaction products, including gypsum, ettringite, thaumasite, and brucite is different, even controversial. In different references, the description of their reaction mechanism is inconsistent, similar to ion–ion reactions for gypsum, through-solution [31,32] or the topochemical [33,34] mechanism for ettringite, the direct [35] or indirect [36] reaction for thaumasite, and ion–ion reactions for brucite, presented in Table 1. It can be seen from the chemical reaction equations in this table that the essence of through-solution and direct reaction is the ion–ion reactions in pore solution, and that of topochemical and indirect reaction is the solid–solid reaction in the cement matrix.

4. Failure Forms Caused by Chemical Reaction Attack

According to the difference in sulfate products, the CSA can be classically divided into gypsum type, ettringite type, thaumasite type CSA in the case of Na+/K+, and brucite type CSA in the case of Mg2+. The different types of CSA present different failure forms of concrete, including softening, cohesiveness, volume expansion, and its-induced cracking/spalling [37,38]. Table 2 shows the relationship between cation types of sulfate solution, CSA classification, and failure forms of concrete. It should be pointed out that the third failure form is the most common among them. At the initial stage of CSA, there are two major disputes about the degradation of concrete exposed to the sulfate environment, namely “degradation cause” and “degradation mechanism”. The former refers to which reaction product (gypsum or ettringite) is the main cause of concrete expansion/cracking, while the latter indicates how CSA causes concrete degradation.

4.1. Degradation Cause of Concrete

In the early period of research on CSA, many scholars, such as Mehta, Tian and Cohen, Mather, Hansen, Collepard [39], Brown and Taylor et al. [31], believed that the volume expansion of concrete exposed to a sulfate environment was caused by the ettringite growth. Subsequently, some scholars, including Metha, Cohen [14], and Santhanam et al. [13], carried out the corrosion experiment of cement paste/mortar specimens immersed in the sulfate solution to investigate the effect of gypsum formation on the failure of cement-based materials. In these studies [13,14], C3S cement with low C3A content was used to reduce the source of the aluminum phase required for ettringite formation, thereby minimizing the amount of ettringite formation. The results showed that the C3S cement specimen soaked in sulfate solution for a long time occurred macroscopic damage phenomena, such as volume expansion and surface cracking, mass, and strength loss. Additionally, the expansion deformation of the C3S cement specimen is obviously greater than that of the Portland cement specimen. Therefore, the formation of ettringite and gypsum is the main cause of the volume expansion or cracking of concrete subjected to CSA.

4.2. Degradation Mechanism of Concrete

Although the solid volume increase to 1.2 times that of the original form of calcium hydroxide transforming into gypsum, the research focus at that time was still on the expansion and deterioration of concrete caused by ettringite. It was believed that the gypsum formation mainly played a role in softening concrete and filling micro-cracks. So, the degradation mechanisms of concrete under CSA, including the theories of volume increase [40], topochemical reaction [41,42], water swelling [43], and crystallization pressure [44,45], are mostly proposed for ettringite type CSA.
The theory of volume increase considers that when the calcium aluminate (CA) in concrete is converted into ettringite, the volume of the solid phase increases 2.5 times, causing the significant expansion of the concrete. This theory seems to be the simplest degradation mechanism, so it is adopted by most scholars to model the behavior of CSA [46]. However, through thermodynamic calculation, Lothernbach [47] and Kunther [48] found that the total volume of products (ettringite) did not exceed the total volume of reactants (for the case of 2.5 times, water was consumed as the reactant). Additionally, there is no experimental evidence to prove the direct relationship between ettringite production and the volume expansion of concrete [37,49]. The theory of topochemical reaction refers to that; ettringite is directly converted from the surface of CA crystal through solid–solid reactions. However, the crystal structure of both hydrated and unhydrated C3A is irrelevant to that of ettringite, which means that it is difficult to achieve the above transformation at room temperature [37,50]. In reality, SEM observation shows that ettringite is not found at the location of C3A hydrated products and on the surface of unhydrated C3A particles. Water swelling refers to the fact that, in the condition of the saturated lime solution, the gel-like and small ettringite colloid can be formed in concrete. This kind of colloid with a large specific surface area can absorb a large number of water molecules, resulting in the volume expansion of concrete [43]. However, Scrivener [50] believes that ettringite generated in concrete subjected to a sulfate attack is a typical crystal, and it is impossible to absorb a lot of water.
At present, the theory of crystallization pressure developed in recent years can reasonably explain the CSA-induced expansion and damage of concrete. As early as 1949, based on thermodynamic theory, Correns deduced the calculation formula of salt crystallization pressure [51], expressed as Equation (1). Subsequently, Scherer [42,43], Steiger [52,53], Espinosa [54], and Komiorczyk [55] et al. developed the theory of crystallization pressure and pointed out that the “supersaturated concentration of solution” and “relatively closed small space” are two key factors for the generation of crystallization pressure. Flatt et al. [56] found that when ettringite was formed in a pore with a size smaller than about 100 nm, the generated crystallization pressure due to the supersaturated solution caused the cracking of cement paste. Müllauer et al. [57] found that, under the immersion of the 30 g/L sodium sulfate solution, the maximum crystallization pressure in cement mortar reached 7–8 MPa, while the pore diameter in the mortar was about 23–27 nm.
P c = R T v mol ln β
where P c is the crystallization pressure; R is the ideal gas constant; T is the absolute temperature; v mol is the molar volume of ettringite crystal; β is the solution supersaturation.
Furthermore, Scherer [44,45], Steiger [52,53], and Müllauer [57] also pointed out that the crystallization pressure in a single pore, exceeding the materials’ tensile strength, cannot cause the destruction of porous materials such as cement mortar and concrete. However, the equivalent expansion stress in the local area of porous materials caused by crystallization pressure in multiple pores will result in the macroscopic expansion and cracking of a porous material. In the process of CSA, there are several connected pores with different space sizes in the local area of concrete, while their solution supersaturations are the same. According to the crystallization pressure theory, the crystals in these pores with the same supersaturation pore solution have the same curvatures/sizes in the equilibrium state, and a smaller crystal (with a large curvature) needs higher supersaturation to maintain its size [57]. If there are two crystals with different sizes in the same pore solution, the small crystal dissolves first, and the solute molecules transfer to the large crystal, resulting in an increase in its size [58]. Thus, with the increase in pore solution saturation, ettringite first crystallizes in big pores. After the big pore is filled to a certain extent, the crystals continue to fill into small pores; that is, small pores are filled after big ones [57], as shown in Figure 3. Meanwhile, the crystallization pressure increases to cause macroscopic damage to concrete.
(In equilibrium state: the higher the solution supersaturation required to maintain smaller crystals, the crystallization pressure driven by supersaturated solution is greater.)

5. Model for Chemical Sulfate Attack

5.1. Diffusion-Reaction Model of Sulfate

Due to the complexity of CSA and the limitations of its mechanism at the early stage of this research, it is difficult to directly establish a diffusion-chemo-mechanical model for numerically simulating the whole failure process of CSA on concrete. So, the scholars first obtained the distribution of sulfate ion concentration in concrete through EDTA titration and EDS quantitative analysis. Then, based on the experimental data, the diffusion-reaction model of sulfate ions in concrete was built up to analyze the behavior of CSA by using Fick’s law, chemical reaction kinetics, and thermodynamics. The basic equation is expressed as Equation (2), which is summarized from the works by Yin et al. [59], Guan et al. [60], Zhao et al. [61], Ran et al. [62], Zou et al. [63], Yu et al. [64], and Wang et al. [65].
C t = J + k r C , J = D C
in which C is the sulfate ion concentration in concrete, J is the sulfate ion flow in concrete, D is the effective diffusion coefficient of the sulfate ion in concrete, k r C respects the reaction consumption of sulfate ion, and is the Hamilton operator. It should be pointed out that the diffusion coefficient D is a spatiotemporal variable, so Equation (2) is a non-steady equation and can be solved by a numerical solution [59], such as the finite difference method, finite volume method, or finite element method, etc. In the works of Guan et al. [60] and Zhao et al. [61], D is regarded to be a constant, so Equation (2) can be expressed as
C t = D Δ C + k r C
where Δ is the Laplace operator, and Equation (3) can be solved analytically by using Laplace transform.
Equation (2) or (3) is only applied to the diffusion behavior of the sulfate ion in saturation concrete, driven by the concentration gradient, and the term of J = D C represents the diffusion behavior. Furthermore, the sulfate ion transport is affected by other factors [63,64,65], such as liquid advection, ion–ion interactions, the electrical field, temperature effect, etc. So, the term of J = D C in Equation (2) is usually corrected as in Table 3.
Gospodinov [66,67] built up a two-dimensional non-steady diffusion model of sulfate ions by considering the effect of reaction consumption between sulfate ions and cement hydration products on the transport process. Later, he [68] further considered the influence of the sulfate products’ precipitation and liquid pressure in pores on the ion effective diffusivity and established a three-dimensional non-steady diffusion model. Marchand [69] considered the coupling effect of ion transport and solution advection in saturated concrete and the chemical equilibrium of solid phase decomposition and established a STADIUM mathematical model. This model can be used to analyze the effects of the water-cement ratio, cement type, solution concentration, and environmental humidity on sulfate ion transport behavior.
Besides the ion transport form, the effective diffusivity D is also a major parameter of the transport process of ions in concrete. It is mainly related to the microstructure characteristics of concrete (such as porosity φ and tortuosity τ) and the ion diffusivity in water, while they are also affected by various factors, including the temperature, the sulfate products filling in the pore [70] (as shown in Figure 4), the damage degree, and the ESA-induced microcrack effect, etc., as listed in Table 3. Zuo et al. [71,72] considered the influence of external load on the tortuosity and porosity of concrete, established two- and three-dimensional non-steady diffusion-reaction models of sulfate ions, and used the ADI scheme of the finite difference method for its numerical solution. Sun et al. [73] carried out experiments to investigate the influence of the CSA-induced damage degree on the transport process of sulfate ions in concrete and integrated it into a new ion diffusion model. The above model can numerically calculate the time-varying concentration of sulfate ion in concrete but cannot analyze the CSA-induced mechanical response and damage degree of the concrete.
In the above transport model, the concept of the effective diffusion coefficient is introduced from the perspective of macro homogeneity materials to facilitate modeling. In reality, concrete is a heterogeneous random multiphase material with different ion diffusion properties in each phase, such as cement paste, aggregate, and the interface transition zone (ITZ). In recent years, there has been considerable research into the influence of ITZ on ion diffusivity in concrete. Yang et al. [74] found that the thickness of ITZ and the water cement ratio has an obvious effect on the ion diffusivity. When the thicknesses of ITZ is 20, 40, and 50 μm, the diffusion coefficients of chloride ion in it are 2.83, 1.75, and 1.55 times them in the cement matrix. In the cases of water-cement ratio 0.35, 0.45, and 0.55, the diffusion coefficients in ITZ were 27.1, 29.8, and 34 times them in the cement matrix [75]. Additionally, Huang et al. [76] found that the water transfer performance in ITZ is also related to the type and material of aggregate, and its ratio between ITZ and cement matrix is about 90–175, and it decreases with the increase in the sand cement ratio. By considering the difference in the transport property of ITZ and the cement matrix, the progress of water transport, ion diffusion, or chemical sulfate attack in concrete can be better predicted.

5.2. Volume Expansion or Crystallization Pressure

5.2.1. Free Volume Expansion of Concrete Caused by Sulfate Products

The volume expansion of concrete is always aimed at ettringite type CSA. As described above, the viewpoint of the ettringite formation mechanism has two kinds, namely the solid–solid reaction and ion–ion reaction, while most of the computational analysis is based on the former. In early research, Atkinson and Hearned [77] put forward an empirical model to describe the relationship between the expansion deformation of the concrete and the amount of ettringite formation by summarizing the experimental data. Clifton and Pommershein [46] studied the volume change caused by various chemical reactions between sulfate ion and cement hydration products and also proposed a prediction model for the volume expansion of cement-based materials subjected to CSA, in which the expansion degree has a linear relationship with the amount of ettringite generated.
In order to facilitate the calculation of ettringite production, the chemical reactions of ettringite formation, namely Equations (a)–(c) in Table 1, are simplified into a unified formula, expressed as Equation (4) [78].
CA + q C   S ¯ H 2 C 6 A S ¯ 3 H 32
in which CA respects the hydrated calcium aluminate ( C 4 A H 13 and C 4 A S ¯ H 12 ) in cement hydration products and unhydrated tricalcium aluminate (C3A). q is the equivalent reaction coefficient of ettringite produced by gypsum consumption, which is equal to 8/3. In reality, the gypsum formation from the sulfate ion and calcium ion, provided by the dissolution of solid calcium (calcium hydroxide and C-S-H gel), also causes the volume expansion of concrete.
The coefficient of the volume change for each chemical reaction to generate ettringite can be calculated by the molar volume of materials before and after the reaction, as in Equation (5). The coefficient value of the volume change in concrete is given in Table 4 [78].
Δ V i V i = ν mol - Ett ν mol - CA i + γ i ν mol - Gyp 1
In which i = 1 ,   2 ,   3 correspond, in turn, to the reaction Equations (a)–(c) in Table 1. Δ V i and V i are the volume change caused by the chemical reaction I and the total volume of substances before reaction, respectively. ν mol - Ett , ν mol - CA i , and ν mol - Gyp represent the molar volumes of ettringite, calcium aluminate, and gypsum; γ i i = 1 ,   2 ,   3 are the reaction coefficients of reaction Equations (a)–(c), respectively.
Table 4. Coefficients of volume change in concrete caused by chemical reactions.
Table 4. Coefficients of volume change in concrete caused by chemical reactions.
Chemical Reactions Δ V i / V i
Equation   ( a ) :   C 4 AH 13 + 3 C S ¯ H 2 + 14 H 2 O C 6 A S ¯ 3 H 32 + CH 0.48
Equation   ( b ) :   C 4 A S ¯ H 12 + 2 C S ¯ H 2 + 16 H 2 O C 6 A S ¯ 3 H 32 0.51
Equation   ( c ) :   C 3 A + 3 C S ¯ H 2 + 26 H 2 O C 6 A S ¯ 3 H 32 1.26
Due to the existence of capillary pores in the concrete, sulfate products first fill the pore space. However, a large number of SEM observations on the microstructure of failure concrete subjected to sulfate attacks found that the sulfate products such as gypsum and ettringite do not fill the entire cracked pores. In other words, if the pores are filled to a certain extent, the continuous generation of sulfate products causes the volume expansion of the concrete. In many references, the parameters of the critical filling volume fraction of pore, f, are introduced to describe the above filling effect, and its empirical value may be 0.05–0.45 by Tixier and Mobasher [78], 0.5 by Basista and Wegnewski [79], 0.05 by Idiart [80], 0.1 by Ikumi [81], and 0.3 by Yu [64]. So, the free volume expansion of concrete caused by sulfate products [64,81], without considering the restraint effect of the hole wall, can be calculated as:
ε V = max 1 q i = 1 3 Δ V i V i 1 ν mol - CA i γ i ν mol - CA i C CA f φ 0 , 0
where CCA is the concentration of CA in the concrete; φ 0 is the initial porosity in the concrete.

5.2.2. Equivalent Expansive Force Generated by Crystallization Pressure

The crystallization pressure driven by pore solution supersaturation is the alternative view for concrete degradation under CSA. It should be pointed out that the volume expansion for concrete failure is a phenomenological viewpoint, while the crystallization pressure, based on the thermodynamic theory, seems to be more scientific and systematic [82]. However, the mechanical analysis of concrete under CSA uses the crystallization pressure theory. A possible reason is the lack of such a bridge between local damage around the pores and the global failure of concrete [83].
As pointed out by Scherer [45], the concrete matrix around the pore will crack once the crystallization pressure on the pore wall is beyond its tensile strength, but only the synergistic damage of the multiple pores’ local area will cause the failure of the whole element or component. Therefore, it is necessary to think about the equivalent expansive force generated by crystallization pressure in the concrete element or component. In the authors’ incomplete statistics, Bary [84] (published in 2008, after Sherer 1999 [44], 2004 [45]) first analyzed the mechanical behavior of concrete caused by CSA from the perspective of crystallization pressure. In this research, the expansion-induced cracking of the concrete is attributed to the crystallization pressure generated by the ettringite growth in the pores, and the equivalent expansive force, P eff , in concrete is calculated by crystallization pressure [84] as:
P eff = α AFm p c
where α AFm is the interaction coefficient between the ettringite crystallization growth and concrete matrix constraint.
Feng et al. [85] analyzed the microstructure mechanical response in the cement paste by crystallization pressure and established the quantitative relation among the crystallization pressure, localized stress-free strain, and equivalent expansive stress, namely P eff   ~   ε free   ~   p c . Basista and Weglewski [79] presented the calculation formula of the equivalent expansive force to describe the propagation of cracks in concrete, but the P eff is obtained from the expansion strain of sulfate product growth, namely P eff   ~   ε V [calculated as Equation (6)].

5.3. Chemo-Mechanical Model of Sulfate Attack

On the basis of more clear degradation mechanisms of CSA and the sulfate diffusion-reaction model, scholars have established a series of chemo-mechanical models to describe the whole process of the expansion-induced degradation of concrete under CSA by using continuum damage mechanics, micropore mechanics, and thermodynamic theory, etc. Table 5 lists some achievements in the research on chemo-mechanical models in recent years.
In the above research, the model proposed by Tixier and Mobasher [78,86] is the most classic, in which the growth of ettringite in the cement paste is regarded as an inclusion problem. Using the theory of volume increase, they obtained the free expansion strain caused by the ettringite growth and analyzed the macroscopic stress and strain in the concrete in combination with the simplified uniaxial tensile stress–strain relationship. In addition, by taking the residual calcium aluminate (CA) concentration at the initial cracking moment as the critical threshold concentration, the corresponding boundary movement criterion was proposed.
On the basis of Tixier and Mobasher [78,86], Sarkar et al. [88] established a mathematical model to simulate the deterioration process of concrete under CSA. In the model, the diffusion behavior of ions from outside to inside and the calcium leaching effect from inside to outside were synchronously considered as reflecting this coupled influence on the chemical reaction balance. The basic mechanical equation or constitutive equation of the model is written as Equation (9).
Bary et al. [84] attributed the concrete damage to the crystallization pressure generated by the growth of ettringite and gypsum and the volume expansion of cement paste around the micropores. By simultaneously introducing a free volume expansion strain and crystallization pressure, they established a more reasonable transport-chemo-mechanical model, and the basic mechanical equation is expressed as Equation (10).
Cefis and Comi [92] studied the mechanical response of partially or fully saturated concrete structures under sulfate attack and proposed a weak coupling analysis method. First, the water content in the cement paste was obtained through a simplified diffusion model; secondly, using the diffusion-reaction model, the concentrations of the sulfate ions and sulfate products were calculated. Finally, considering the water pressure on the pore wall, a multiphase elastic-damage model coupled with chemo-mechanical damage was established to analyze the nonlinear mechanical response of concrete under CSA, presented in Equation (11).
Other chemo-mechanical models are much the same, and the basic framework is also improved on the basis of the Tixier and Mobasher model [78,86]. The specific mechanical equations are shown in Table 6.

5.4. Damage Characterization of Concrete

In the chemo-mechanical models, the damage variables are introduced in the basic mechanical equations (in Table 6) to describe the actual mechanical response in the concrete, such as the effect of microcracking on expansion strain and microstructure damage on stress relaxation. The damage variables in the examples of previous literature can be summarized into three types. The first one is associated with chemical reactions for the filling of sulfate products and the decalcification of cement-hydrated phases, called chemical damage dc. The second and third one is both for the stress-induced microcracks, while the sources of stress generated in the concrete are different. In this review, the stress-induced damage related to the interaction between the expansion of sulfate products and the constraint of the cement matrix is called mechanical damage dm. The stress-induced damage caused by the external loading on the concrete surface is called load damage dl.
In many studies [92,100,101,103,104], the expansion of the total damage coupled with dc, dm, and dl is proposed according to the theory of continuum damage mechanics as:
d coup = 1 1 d c 1 d m 1 d l

5.4.1. Chemical Damage

For the dc, Saetta et al. [100,101] earlier carried out research on the mechanical response of concrete structures under coupled load action and environmental corrosion at 1998 and 1999. They assumed that the chemical damage of concrete caused by an environmental attack (not only for sulfate attack, but also for chloride corrosion, calcium dissolution, etc.) mainly depends on the pollutant concentration of the corrosion ion C and the residual strength of degraded concrete f dam , defined as Equation (19). It is an empirical equation, and the f C C ref reflects the degree of chemical reaction, calculated by the ratio of ion concentration C to the reference concentration Cref ( f C C ref = C / C ref ). The η R - re reflects the relative residual strength of the concrete caused by chemical reactions, calculated by the ratio of f dam to the initial strength f ini ( η R - re = f dam / f ini ).
Cefis and Comi [92] considered that the decalcification of cement hydrated phases caused by sulfate chemical reactions causes the formation of diffuse microcracks, resulting in the degradation of concrete mechanical properties. In the model, an isotropic chemical damage variable dc is introduced, which is only related to the extent of the chemical reaction ξ, expressed as Equation (20). Other similar expressions of dc are shown in Table 7.

5.4.2. Mechanical Damage

For the characterization of dm, the work performed by Tixier and Mobasher [78] is the earliest and most meaningful, and was developed by many later studies, such as Sarkar et al. [88,89], Yu et al. [64,98,105], Qin et al. [106], Wang et al. [107], Yin et al. [59,96] and Li et al. [108] and is expressed as Equation (22). In this work, a simplified uniaxial tensile stress–strain law was adopted to quantitatively describe the evolution of the damage degree in the process of CSA. The dm is associated with the free expansive strain ε V in Equation (6) caused by the formation of sulfate products, namely d m ε V . The uniaxial tensile stress–strain curve is divided into three stages, including the linear ascending stage (LAS), pre-peak non-linear ascending stage (PNAS), and post-peak non-linear descending stage (PNDS). In the CSA process, the initial ε V is in LAS, meaning that the concrete is undamaged, namely d m = 0 . When the increasing ε V is into PNAS, the microcracks begin to form and spread in the concrete, and the d m increases linearly. Later, the ε V increases into PNDS, the microcracks converge into macrocracks, and the d m has a non-linear rapid increase.
On this basis, Sarkar et al. [88,89] took d m d crit ( d crit is a critical value of concrete failure) as the criterion judge of whether the concrete happens to show surface failure or boundary movement. This work is superior to that of Tixier and Mobasher [78], in which the threshold concentration of reacted calcium aluminate CA was selected to differentiate the cracked and uncracked regions in the concrete. Unfortunately, there is no systematic mechanical analysis in the above research to reveal the essence of mechanical damage, that is, the mechanical interaction between the expansion of sulfate products and the constraint of the concrete matrix.
Therefore, the quantitative method of load damage based on the elastoplastic damage mechanics is used to characterize the stress-induced mechanical damage of concrete under CSA, such as Bary et al. [84,87], Yin et al. [96,102], in which the basic expression of damage is similar to that in Mazars’ damage model [109]. The specific equations of mechanical damage are also presented in Table 7. As shown in the table, the difference in the calculation method of I type dm (dm-I) and II type dm (dm-II) is whether the calculation can be decoupled. dm-I can be directly calculated by the free volume expansion ε V from the growth of sulfate products, namely ε V ξ d m - I . dm-II needs to be numerically solved by coupling with stress and strain, and they all need to meet the unified equations, including the equilibrium differential equation, constitutive equation, and geometric equation, as shown in Figure 5.

5.4.3. LOAD DAMAGE

The research on the characterization method of the load-induced damage degree is earlier than the above two, and even part of the research on mechanical damage (namely dm-II) is based on load damage dl. The achievements of dl have been quite rich, and they have proposed along with the plasticity or damage constitutive model for the mechanical response in concrete under an external load. The typical constitutive models include Mazars’ damage model [109], Lubliner’s damage model [110], Faria’s rate-independent plasticity damage constitutive model [111], Grassl’s damage-plastic model [112,113], and Wu’s plastic damage model based on the energy release rate [114]. In these models, the load damage is determined by the damage loading function, the evolution law for the damage variable, and the loading-unloading conditions, as Equations (16)–(18).
F d - ς ε p , κ d - ς = ε ˜ ς ε p κ d - ς
d l - ς = g d κ d - ς
F d - ς 0 ,   κ ˙ d - ς 0 , κ ˙ d - ς F d - ς = 0
where the subscript ς = t ,   c represents the tension and compression condition. F d - ς is the damage loading function. ε ˜ ς is the effective stress. κ d - ς is the damage-driving variable. g d is the damage function to describe the evolution of the load damage variable.
Generally, two damage parameter variables are adopted to, respectively, describe the damage degree of concrete in two different forms under uniaxial tension and compression. The reason is the different nonlinear (damage) characteristics of concrete under single tension and single compression loads [114]. Therefore, there are two damage driving variables corresponding to the damage variables, namely κ d - t for d l - t and κ d - c for d d - c . The calculation idea of load damage is that the damage-driving variable κ d - ς is first determined through the damage loading function of Equation (16), in which the plastic strain ε p can be obtained by the plasticity model. Then, Equation (17) is used to judge the evolution of the damage-driving variable in the whole loading-unloading process. Finally, the load damage degree can be calculated according to the damage function of Equation (17). The expressions of g d are roughly similar in different constitutive models, as presented in Table 7.
Table 7. Expressions of damage degree in concrete caused by CSA.
Table 7. Expressions of damage degree in concrete caused by CSA.
AuthorBasic Mechanical EquationNumber
Chemical damage dc
Saetta et al. [110,111] d c = 1 η R - re 1 1 1 + 2 f C C ref 2 Related   to   ion   concentration   C (19)
Cefis and Comi [92] d c = 1 exp A 1 ξ 1 + exp A 1 ξ + A 2 Related   to   reaction   extent A 3 (20)
Sun et al. [73]
Zhang et al. [115]
d c = 1 φ 0 1 exp A 4 t t 0 Related   to   diffuse   time 1 1 1 + A 5 C C es 0 A 6 Related   to   ion   concentration   C (21)
t / t 0 is related to the diffuse time or hydrated time. C / C 0 is the ratio of ion concentration in the concrete to that in the external solution.
Mechanical damage dm
ITixier and Mobasher [78] d m = r 7 1 ε th ε r 8 Related   to   volume   expansion , ε = 1 3 ε V (22)
ε th is a threshold strain for the initiation of microcracks or damage.
Wang et al. [116] d m = 1 1 1 + ς 1 ε / ε m ς / 1 ς Related   to   volume   expansion , ς = E 0 f m / ε m (23)
E 0 is the initial elastic modulus. ε m is the ultimate tensile strain corresponding to the ultimate tensile stress f m at the peak of the stress–strain curve.
IIBary et al. [84,87] d m = 1 1 A 9 ε ˜ / ε th A 9 exp A 10 ε ˜ ε th (24)
ε ˜ is the equivalent strain.
Yin et al. [96,102] d m = σ ¯ + d m - t + σ ¯ d m - c σ ¯ d m - c   or   - t = 1 1 A 11 - c   or   - t 1 + f ε ˜ / ε 0 - c   or   - t A 11 - c   or   - t exp A 12 - c   or   - t f ε ˜ / ε 0 - c   or   - t (25)
σ ¯ +   and   σ ¯ are the positive and negative spectral decomposition parts of the effective stress tensors σ ¯ . d m - c   or   - t is the compressive or tensile stress-induced mechanical damage.
Load damage dl
Grassl et al. [112] d l = 1 exp κ d ε f (26)
ε f is the parameter that controls the slope of the softening curve.
Wu et al. [114] d l + = 1 r 0 + r + exp A 13 1 r + r 0 + ,   d l = 1 r 0 r 1 A 14 A 14 exp A 15 1 r r 0 (27)
d l +   and   d l are the tensile damage and shear damage corresponding to positive and negative stress components.
Mazars et al. [109]
Zheng et al. [117]
d m = 1 1 s t d l - t 1 s c d l - c d l - c   or   - t = 1 1 A 16 - c   or   - t 1 + κ d - c   or   - t / ε 0 - c   or   - t A 16 - c   or   - t exp A 17 - c   or   - t κ d - c   or   - t / ε 0 - c   or   - t (28)
s t and s c represent the stiffness recovery effects of tension and compression.
A 1 ~ 16 are empirical parameters.

6. Challenges

The above research is aimed at the CSA (transport-chemo-mechanical behavior) on saturated Portland concrete. However, in many service environments, there exists the wetting and drying alternation caused by temperature change, groundwater rising, and falling or tidal fluctuation, resulting in the PSA (transport-physical-mechanical behavior) on unsaturated concrete [118,119]. So, one future challenge is to model the PSA behavior, and it differs from the CSA model in the repeated crystallization-dissolution behavior of sulfate crystal.
The other challenge is to evaluate the CSA-induced durability of cementitious green building materials (CGBMs), such as 3D-printed concrete and recycled aggregate concrete. Due to the printing method, mineral type and content, recycled aggregate quality (loose structure or high sulfate content), 3D-printed concrete [120,121], and recycled aggregate concrete [122,123], low mechanical properties and high porosity may show poor sulfate-resistant properties. At present, relevant research focuses on the effect of these material properties on structural characteristics, while the models to evaluate the durability of CGBMs subjected to CSA are few reported. However, the latter is the key ensuring that the life of CGBMs structures service in sulfate environments.

7. Conclusions

By consulting a lot of research in the literature about chemical sulfate attacks (CSA) on concrete, this review introduces the chemical reaction products of sulfate attack, the formation mechanism of reaction products, them-induced failure forms of concrete, and the model for the whole process of CSA on concrete, step by step. In part of the model, the diffusion-reaction model, chemo-mechanical model, and CSA-induced damage characterization are elaborately detailed. The main conclusions derived from the review are presented as follows:
For the Na+/K+ basic ions type CSA, ettringite, gypsum, and thaumasite are the major reaction products. Additionally, the formation mechanism of ettringite and thaumasite is controversial, namely the dispute between the ion–ion reaction and solid–solid reaction.
According to the kind of reaction products, CSA can be classified as gypsum type, ettringite type, and thaumasite type in the case of Na+/K+ basic ions and brucite type in the case of Mg2+ basic ions. The different types of CSA present different failure forms of the concrete, including softening, cohesiveness, volume expansion, and its-induced cracking/spalling.
For the ettringite type CSA, the theories of volume increase and crystallization pressure are proposed to reveal the degradation mechanism of the concrete. Although the latter, based on thermodynamics, is more scientific, the former, despite lacking a scientific basis, is more widely used in the CSA model because it builds the relationship between the volume expansion of concrete and the content of reaction products.
In order to model the behavior of CSA, the diffusion-reaction models of sulfate are first established to simulate the sulfate transport in the concrete, in which the transport mechanisms, including the ion diffusion, liquid advection, ion–ion interaction, electrical field, and temperature, etc., are considered. However, the diffusion-reaction models cannot analyze the mechanical response in concrete caused by CSA.
On the basis of the diffusion-reaction model and CSA-induced degradation mechanism, the chemo-mechanical models, in which the variables of chemical damage and mechanical damage are introduced, are developed to describe the whole damage process of concrete under CSA. These models can provide a basis for the service life prediction of concrete structures in a sulfate environment. However, the limitation of chemo-mechanical models reported in this review is that these models only apply to the ettringite type CSA on saturated concrete. The model for other types of CSA on unsaturated concrete needs more scholars to invest in this research.

Author Contributions

Writing—Original draft, G.-J.Y. and Y.-J.T.; Foundation support, G.-J.Y., X.-D.W. and Y.-J.T.; Theme determination, X.-D.W.; Investigation, L.M.; Table and Figure Editing, D.C.; Original manuscript modification, X.-B.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Zhejiang Provincial Natural Science Foundation of China (LQ21E080008), National Natural Science Foundation of China (51569035), Natural Science Foundation of the Jiangsu Higher Education Institutions of China (20KJB430030), Ningbo Scientific and Technological Innovation Major Project in 2025 (2020Z056) and Scientific Research Start-up Foundation of Ningbo University of Technology in 2020.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Chen, Z.; Wu, L.; Bindiganavile, V.; Yi, C.F. Coupled models to describe the combined diffusion-reaction behaviour of chloride and sulphate ions in cement-based systems. Constr. Build. Mater. 2020, 243, 118232. [Google Scholar] [CrossRef]
  2. Martins, M.C.; Langaro, E.A.; Macioski, G.; Medeiros, M.H.F. External ammonium sulfate attack in concrete: Analysis of the current methodology. Constr. Build. Mater. 2021, 277, 122252. [Google Scholar] [CrossRef]
  3. Salt Lake: Accelerating Chinese civilization and creating magnificent scenery. Chin. Natl. Geogr. 2011, 605, 13.
  4. Sun, W. Durability and service life of structure concrete under load and environment coupling effects. J. Southeast Univ. (Nat. Sci. Ed.) 2006, 36 (Suppl. SII), 7–14. [Google Scholar]
  5. Liu, Z.Y. Study on Methods of Accelerated Testing of Marine Concrete Durability Based on Simulating Environment and Service Life Prediction; Southeast University: Nanjing, China, 2006. [Google Scholar]
  6. Zhou, G.; Li, S.R.; Wang, Z.J.; Wang, C.F. Investigation and analysis on corrosion situation of concrete in saline soil region. J. Archit. Civ. Eng. 2011, 4, 121–126. [Google Scholar]
  7. Mehta, P.K. Sulfate attack on concrete: Separating myths from reality. Concr. Int. 2000, 22, 57–61. [Google Scholar]
  8. Hime, W.G.; Mather, B. “Sulfate attack,” or is it. Cem. Concr. Res. 1999, 29, 789–791. [Google Scholar] [CrossRef]
  9. Neville, A. The confused world of sulfate attack on concrete. Cem. Concr. Res. 2004, 34, 1275–1296. [Google Scholar] [CrossRef]
  10. Neville, A. Consideration of durability of concrete structure: Past, present, and future. Mater. Struct. 2001, 34, 114–118. [Google Scholar] [CrossRef]
  11. Bensted, J. Thaumasite-direct, woodfordite and other possible formation routes. Cem. Concr. Compos. 2003, 25, 873–877. [Google Scholar] [CrossRef]
  12. Harrison, W.H.; Cooke, R.W. Sulfate resistance of buried concrete. Proc. Inst. Civ. Eng. 1981, 70, 871–874. [Google Scholar]
  13. Santhanam, M.; Cohen, M.D.; Olek, J. Effects of gypsum formation on the performance of cement mortars during external sulfate attack. Cem. Concr. Res. 2003, 33, 325–332. [Google Scholar] [CrossRef]
  14. Tian, B.; Cohen, M.D. Does gypsum formation during sulfate attack on concrete lead to expansion? Cem. Concr. Res. 2000, 30, 117–123. [Google Scholar] [CrossRef]
  15. Lee, H.; Cody, R.D.; Cody, A.M.; Spry, P.G. The formation and role of ettringite in Iowa highway concrete deterioration. Cem. Concr. Res. 2005, 35, 332–343. [Google Scholar] [CrossRef]
  16. Wang, X.; Pan, Z.; Shen, X.; Liu, W.Q. Stability and decomposition mechanism of ettringite in presence of ammonium sulfate solution. Constr. Build. Mater. 2016, 124, 786–793. [Google Scholar] [CrossRef]
  17. Pavoine, A.; Brunetaud, X.; Divet, L. The impact of cement parameters on delayed ettringite formation. Cem. Concr. Compos. 2012, 34, 521–528. [Google Scholar] [CrossRef]
  18. Abualgasem, J.M.; Cripps, J.C.; Lynsdale, C.J. Effects of wetting and drying cycles on thaumasite formation in cement mortars. J. Mater. Civ. Eng. 2015, 27, 1–6. [Google Scholar] [CrossRef]
  19. Rahman, M.M.; Bassuoni, M.T. Thaumasite sulfate attack on concrete: Mechanisms, influential factors and mitigation. Constr. Build. Mater. 2014, 73, 652–662. [Google Scholar] [CrossRef]
  20. Bellmann, F.; Stark, J. Prevention of thaumasite formation in concrete exposed to sulphate attack. Cem. Concr. Res. 2007, 37, 1215–1222. [Google Scholar] [CrossRef]
  21. Biczok, I. Concrete Corrosion Concrete Protection; Chemical Publishing Company: New York, NY, USA, 1967. [Google Scholar]
  22. Bellmann, F.; Möser, B.; Stark, J. Influence of sulfate solution concentration on the formation of gypsum in sulfate resistance test specimen. Cem. Concr. Res. 2006, 36, 358–363. [Google Scholar] [CrossRef]
  23. Ma, B.G.; Luo, Z.T.; Li, X.G.; Gao, X.J.; Zhang, M.X. Microscopic structure and growth mechanism of the corrosion products including thaumasite. J. Chin. Ceram. Soc. 2006, 34, 1503–1507. [Google Scholar]
  24. Santhanam, M.; Cohen, M.D.; Olek, J. Sulfate attack research—Whither now? Cem. Concr. Res. 2001, 31, 845–851. [Google Scholar] [CrossRef]
  25. Bassuoni, M.T.; Nehdi, M.L. Durability of self-consolidating concrete to different exposure regimes of sodium sulfate attack. Mater. Struct. 2009, 42, 1039–1057. [Google Scholar] [CrossRef]
  26. Collepardi, M. Thaumasite formation and deterioration in historic buildings. Cem. Concr. Compos. 1999, 21, 147–154. [Google Scholar] [CrossRef]
  27. Bonen, D.; Cohen, M.D. Magnesium sulfate attack on Portland cement paste—II. Chemical and mineralogical analyses. Cem. Concr. Res. 1992, 22, 707–718. [Google Scholar] [CrossRef]
  28. Durgun, M.Y.; Sevinc, A.H. Determination of the effectiveness of various mineral additives against sodium and magnesium sulfate attack in concrete by Taguchi method. J. Build. Eng. 2022, 57, 104849. [Google Scholar] [CrossRef]
  29. Huang, Q.; Zhu, X.H.; Xiong, G.Q.; Zhang, M.T.; Deng, J.X.; Zhao, M.; Zhao, L. Will the magnesium sulfate attack of cement mortars always be inhibited by incorporating nanosilica. Constr. Build. Mater. 2021, 305, 124695. [Google Scholar] [CrossRef]
  30. Marchand, J.; Ivan Older Skalny, J.P. Sulfate Attack on Concrete (Morden Concrete Technology); Spon Press: London, UK; New York, NY, USA, 2002. [Google Scholar]
  31. Brown, P.W.; Taylor, H.F.W. The Role of Ettringite in External Sulfate Attack: Material Science of Concrete—Sulfate Attack Mechanism; American Ceramic Society: Westerville, OH, USA, 1999. [Google Scholar]
  32. Peng, J.H.; Lou, Z.H. Study Mech. Ettringite Formation. J. Chin. Ceram. Soc. 2000, 28, 511–515. [Google Scholar]
  33. Evju, C.; Hansen, S. The kinetics of ettringite formation and dilatation in a blended cement with β-hemihydrate and anhydrite as calcium sulfate. Cem. Concr. Res. 2005, 35, 2310–2321. [Google Scholar] [CrossRef]
  34. Silva, D.A.; Monteiro, P.J.M. Early Formation of Ettringite in Tricalcium Aluminate-Calcium Hydroxide-Gypsum Dispersions. J. Am. Ceram. Soc. 2007, 90, 614–617. [Google Scholar] [CrossRef]
  35. Skalny, J.; Marchand, J.; Odler, I. Sulfate Attack on Concrete; Spon Press: London, UK, 2002. [Google Scholar]
  36. Schmidt, T. Sulfate Attack and the Role of Internal Carbonate on the Formation of Thaumasite; EPFL: Lausanne, Switzerland, 2007. [Google Scholar]
  37. Yu, C.; Sun, W.; Scrivener, K. Mechanism of expansion of mortars immersed in sodium sulfate solutions. Cem. Concr. Res. 2013, 43, 105–111. [Google Scholar] [CrossRef]
  38. Souza, D.J.D.; Medeiros, M.H.F.D.; Filho, J.H. Evaluation of external sulfate attack (Na2SO4 and MgSO4): Portland cement mortars containing fillers. Rev. IBRACON Estrut. E Mater. 2020, 13, 644–655. [Google Scholar] [CrossRef]
  39. Collepardi, M. A state-of-the-art review on delayed ettringite attack on concrete. Cem. Concr. Compos. 2003, 25, 401–407. [Google Scholar] [CrossRef]
  40. Polivka, M. Factors influencing expansion of expansive cement concretes. Int. Concr. Abstr. Portal 1973, 38, 239–250. [Google Scholar]
  41. Cohen, M.D. Modeling of expansive cements. Cem. Concr. Res. 1983, 13, 519–528. [Google Scholar] [CrossRef]
  42. Cohen, M.D. Theories of expansion in sulfoaluminate—Type expansive cements: Schools of thought. Cem. Concr. Res. 1983, 13, 809–818. [Google Scholar] [CrossRef]
  43. Mehta, P.K. Mechanism of expansion associated with ettringite formation. Cem. Concr. Res. 1973, 3, 1–6. [Google Scholar] [CrossRef]
  44. Scherer, G.W. Crystallization in pores. Cem. Concr. Res. 1999, 29, 1347–1358. [Google Scholar] [CrossRef]
  45. Scherer, G.W. Stress from crystallization of salt. Cem. Concr. Res. 2004, 34, 1613–1624. [Google Scholar] [CrossRef]
  46. Clifton, J.R.; Ponnersheim, J.M. Sulfate Attack of Cementitious Materials: Volumetric Relations and Expansions; Building and Fire Research, National Institute of Standards and Technology: Gaithersburg, MD, USA, 1994.
  47. Lothenbach, B.; Bary, B.; Bescop, P.L.; Schmidt, T.; Leterrier, N. Sulfate Ingress in Portland Cement. Cem. Concr. Res. 2010, 40, 1211–1225. [Google Scholar] [CrossRef]
  48. Kunther, W.; Lothenbach, B.; Scrivener, K.L. On the relevance of volume increase for the length changes of mortar bars in sulfate solution s. Cem. Concr. Res. 2013, 46, 23–29. [Google Scholar] [CrossRef]
  49. Taylor, H.F.W.; Famy, C.; Scrivener, K.L. Delayed ettringite formation. Cem. Concr. Res. 2001, 31, 683–693. [Google Scholar] [CrossRef]
  50. Bizzpzero, J.; Gpsselin, C.; Scrivener, K.L. Expansion mechanisms in calcium aluminate and sulfoaluminate systems with calcium sulfate. Cem. Concr. Res. 2014, 56, 190–202. [Google Scholar] [CrossRef]
  51. Metha, P.K. Evaluation of sulfate-resisting cements by a new test method. J. ACI 1975, 72, 573–575. [Google Scholar]
  52. Steiger, M. Crystal growth in porous materials—II: Influence of crystal size on the crystallization pressure. J. Cryst. Growth 2005, 282, 470–481. [Google Scholar] [CrossRef]
  53. Steiger, M. Crystal growth in porous materials—I: The crystallization pressure of large crystals. J. Cryst. Growth 2005, 282, 455–469. [Google Scholar] [CrossRef]
  54. Espinosa, R.M.; Franke, L.; Deckelmann, G. Model for the mechanical stress due to the salt crystallization in porous materials. Constr. Build. Mater. 2008, 22, 1350–1367. [Google Scholar] [CrossRef]
  55. Koniorczyk, M.; Gawin, D. Modelling of salt crystallization in building materials with microstructure—Poromechanical approach. Constr. Build. Mater. 2012, 36, 860–873. [Google Scholar] [CrossRef]
  56. Flatt, R.J. Salt damage in porous materials: How high supersaturations are generated. J. Cryst. Growth 2002, 242, 435–454. [Google Scholar] [CrossRef]
  57. Müllauer, W.; Beddoe, R.E.; Heinz, D. Sulfate attack expansion mechanisms. Cem. Concr. Res. 2013, 52, 208–215. [Google Scholar] [CrossRef]
  58. Ju, X.D.; Feng, W.J.; Zhang, Y.J. Crystallization stress in brittle porous media. Chin. J. Geotech. Eng. 2016, 38, 1253–1264. [Google Scholar]
  59. Yin, G.J.; Shan, Z.Q.; Miao, L.; Tang, Y.J.; Zuo, X.B.; Wen, X.D. Finite element analysis on the diffusion-reaction-damage behavior in concrete subjected to sodium sulfate attack. Eng. Fail. Anal. 2022, 137, 106278. [Google Scholar] [CrossRef]
  60. Guan, B.W.; Liu, J.N.; Wu, J.W.; Chen, X.H.; Hu, Y.; Zhang, L.Q. Transport behavior of sulfate ions in concrete with attack damage. Bull. Chin. Ceram. Soc. 2020, 39, 3169–3174. [Google Scholar]
  61. Zhao, S.B.; Yang, X.M. Experimental study on regularity of sulfate-ion diffusion and distribution in concrete attacked by sulfate. China Harb. Eng. 2009, 161, 26–29. [Google Scholar]
  62. Ran, B.; Li, K.F.; Teddy, F.C.; Omikrine-Metalssi, O.; Dangka, P. Spalling rate of concretes subject to combined leaching and external sulfate attack. Cem. Concr. Res. 2022, 162, 106951. [Google Scholar] [CrossRef]
  63. Zou, D.J.; Qin, S.S.; Liu, T.J.; Jivkov, A. Experimental and numerical study of the effect of solution concentration and temperature on concrete under external sulfate attack. Cem. Concr. Res. 2021, 139, 106284. [Google Scholar] [CrossRef]
  64. Yu, Y.; Zhang, Y.X. Numerical modelling of mechanical deterioration of cement mortar under sulfate attack. Constr. Build. Mater. 2018, 158, 490–502. [Google Scholar] [CrossRef]
  65. Shazail, M.A.; Baluch, M.H.; Al-Gadhib, A.H. Predicting residual strength in unsaturated concrete exposed to sulfate attack. J. Mater. Civil Eng. 2006, 18, 343–354. [Google Scholar] [CrossRef]
  66. Gospodinov, P.; Kazandjiev, R.; Mironova, M. The effect of sulfate ion diffusion on the structure of cement stone. Cem. Concr. Compos. 1996, 18, 401–407. [Google Scholar] [CrossRef]
  67. Gospodinov, P.N.; Kazandjiev, R.F.; Partalin, T.A.; Mironovac, M.K. Diffusion of sulfate ions into cement stone regarding simultaneous chemical reactions and resulting effects. Cem. Concr. Res. 1999, 29, 1591–1596. [Google Scholar] [CrossRef]
  68. Gospodinov, P.N. Numerical simulation of 3D sulfate ion diffusion and liquid push out of the material capillaries in cement composites. Cem. Concr. Res. 2005, 35, 520–526. [Google Scholar] [CrossRef]
  69. Marchand, J.; Samson, E.; Maltais, Y.; Beaudoin, J.J. Theoretical analysis of the effect of weak sodium sulfate solutions on the durability of concrete. Cem. Concr. Compos. 2002, 24, 317–329. [Google Scholar] [CrossRef]
  70. Liu, X.; Liu, Q.W.; Wang, S.G.; Xu, F.; Han, J.D. Recent development on theoretical models of sulfate attack on concrete. Mater. Rev. 2014, 28, 89–95. [Google Scholar]
  71. Zuo, X.B.; Sun, W.; Li, H.; Zhao, Y.K. Modeling of diffusion-reaction behavior of sulfate ion in concrete under sulfate environments. Comput. Concr. 2012, 10, 47–51. [Google Scholar] [CrossRef]
  72. Sun, W.; Zuo, X.B. Numerical simulation of sulfate diffusivity in concrete under combination of mechanical loading and sulfate environments. J. Sustain. Cem.-Based Mater. 2012, 1, 46–55. [Google Scholar] [CrossRef]
  73. Sun, C.; Chen, J.K.; Zhu, J.; Zhang, M.H.; Ye, J. A new diffusion model of sulfate ions in concrete. Constr. Build. Mater. 2013, 39, 39–45. [Google Scholar] [CrossRef]
  74. Yang, C.C.; Su, J.K. Approximate migration coefficient of interfacial transition zone and the effect of aggregate content on the migration coefficient of mortar. Cem. Concr. Res. 2002, 32, 1559–1565. [Google Scholar] [CrossRef]
  75. Yang, C.C.; Cho, S.W. Approximate migration coefficient of percolated interfacial transition zone by using the accelerated chloride migration test. Cem. Concr. Res. 2005, 35, 344–350. [Google Scholar] [CrossRef]
  76. Huang, Q.H.; Zhou, C.Z.; Gu, X.L.; Zhang, W.P. Experimental study on moisture transport property of interfacial transition zone in concrete. J. Build. Struct. 2019, 40, 174–180. [Google Scholar]
  77. Atkinson, A.; Hearne, J.A. Mechanistic model for the durability of concrete barriers exposed to sulphate-bearing groundwaters. Mrs. Proc. 1989, 176, 149–156. [Google Scholar] [CrossRef]
  78. Tixier, R.; Mobasher, B. Modeling of Damage in cement-based materials subjected to external sulfate attack. I: Formulation. J. Mater. Civ. Eng. 2003, 15, 305–313. [Google Scholar] [CrossRef]
  79. Basista, M.; Weglewski, W. Chemically Assisted Damage of Concrete: A Model of Expansion Under External Sulfate Attack. Int. J. Damage Mech. 2009, 18, 155–175. [Google Scholar] [CrossRef]
  80. Idiart, A.E.; López, C.M.; Carol, I. Chemo-mechanical analysis of concrete cracking and degradation due to external sulfate attack: A meso-scale model. Cem. Concr. Compos. 2011, 33, 411–423. [Google Scholar] [CrossRef]
  81. Ikumi, T.; Cavalaro, S.H.P.; Segura, I.; Aguado, A. Alternative methodology to consider damage and expansions in external sulfate attack modeling. Cem. Concr. Res. 2014, 63, 105–116. [Google Scholar] [CrossRef]
  82. Yu, C.; Sun, W.; Scrivener, K.L. Degradation mechanism of slag blended mortars immersed in sodium sulfate solution. Cem. Concr. Res. 2015, 72, 37–47. [Google Scholar] [CrossRef]
  83. Yin, G.J.; Zuo, X.B.; Sun, X.; Tang, Y.J. Macro-microscopically numerical analysis on expansion response of hardened cement paste under external sulfate attack. Constr. Build. Mater. 2019, 207, 600–615. [Google Scholar] [CrossRef]
  84. Bary, B. Simplified coupled chemo-mechanical modeling of cement pastes behavior subjected to combined leaching and external sulfate attack. Int. J. Numer. Anal. Methods Geomech. 2008, 32, 1791–1816. [Google Scholar] [CrossRef]
  85. Feng, P.; Bullard, J.W.; Garboczi, E.J. A multiscale microstructure model of cement paste sulfate attack by crystallization pressure. Modelling Simul. Mater. Sci. Eng. 2017, 25, 65013. [Google Scholar]
  86. Tixier, R.; Mobasher, B. Modeling of Damage in Cement-Based Materials Subjected to External Sulfate Attack. II: Comparison with experiments. J. Mater. Civ. Eng. 2003, 15, 314–322. [Google Scholar] [CrossRef]
  87. Bary, B.; Leterrier, N.; Deville, E.; Bescop, P. Coupled chemo-transport-mechanical modelling and numerical simulation of external sulfate attack in mortar. Cem. Concr. Compos. 2014, 49, 70–83. [Google Scholar] [CrossRef]
  88. Sarkar, S.; Mahadevan, S.; Meeussen, J.C.L.; Sloot, H.V.D.; Kosson, D.S. Numerical simulation of cementitious materials degradation under external sulfate attack. Cem. Concr. Compos. 2010, 32, 241–252. [Google Scholar] [CrossRef] [Green Version]
  89. Sarkar, S.; Mahadevan, S.; Meeussen, J.C.L.; Sloot, S.V.D.; Kosson, D.S. Sensitivity Analysis of Damage in Cement Materials under Sulfate Attack and Calcium Leaching. J. Mater. Civ. Eng. 2012, 24, 430–440. [Google Scholar] [CrossRef]
  90. Ikumi, T.; Cavalaro, S.H.P.; Segura, I.; Fuente, A.D.L.; Aguado, A. Simplified methodology to evaluate the external sulfate attack in concrete structures. Mater. Des. 2016, 89, 1147–1160. [Google Scholar] [CrossRef]
  91. Cefis, N.; Comi, C. Damage modelling in concrete subject to sulfate attack. Frat. Integrità Strutt. 2014, 8, 222–229. [Google Scholar] [CrossRef] [Green Version]
  92. Cefis, N.; Comi, C. Chemo-mechanical modelling of the external sulfate attack in concrete. Cem. Concr. Res. 2017, 93, 57–70. [Google Scholar] [CrossRef]
  93. Zuo, X.B.; Sun, W. Full process analysis of damage and failure of concrete subjected to external sulfate attack. J. Chin. Ceram. Soc. 2009, 37, 1063–1067. [Google Scholar]
  94. Zuo, X.B.; Sun, W.; Yu, C. Numerical investigation on expansive volume strain in concrete subjected to sulfate attac k. Constr. Build. Mater. 2012, 36, 404–410. [Google Scholar] [CrossRef]
  95. Nie, Q.; Zhou, C.; Li, H.; Sun, X.; Gong, H.; Huang, B. Numerical simulation of fly ash concrete under sulfate attack. Constr. Build. Mater. 2015, 84, 261–268. [Google Scholar] [CrossRef]
  96. Yin, G.J.; Zuo, X.B.; Tang, Y.; Tang, Y.J. Numerical simulation on time-dependent mechanical behavior of concrete under coupled axial loading and sulfate attack. Ocean Eng. 2017, 142, 115–124. [Google Scholar] [CrossRef]
  97. Yin, G.J.; Zuo, X.B.; Sun, X.; Tang, Y.J. Numerical investigation on ESA-induced expansion response of cement paste by using crystallization pressure. Model. Simul. Mater. Sci. Eng. 2019, 27, 25006. [Google Scholar] [CrossRef]
  98. Yu, Y.G.; Gao, W.; Feng, Y.; Gastel, A.; Chen, X.J.; Liu, A.R. On the competitive antagonism effect in combined chloride-sulfate attack A numerical exploration. Cem. Concr. Res. 2021, 144, 106406. [Google Scholar] [CrossRef]
  99. Yi, C.F.; Zheng, C.; Vivek, B. A non-homogeneous model to predict the service life of concrete subjected to external sulfate attack. Constr. Build. Mater. 2019, 212, 254–265. [Google Scholar] [CrossRef]
  100. Saetta, A.; Scotta, R.; Vitaliani, R. Mechanical Behavior of Concrete under Physical-Chemical Attacks. J. Eng. Mech. 1998, 124, 1100–1109. [Google Scholar] [CrossRef]
  101. Saetta, A.; Scotta, R.; Vitaliani, R. Coupled Environmental-Mechanical Damage Model of RC Structures. J. Eng. Mech. 1999, 125, 930–940. [Google Scholar] [CrossRef]
  102. Yin, G.J.; Zuo, X.B.; Li, X.N.; Zou, Y.X. An integrated macro-microscopic model for concrete deterioration under external sulfate attack. Eng. Fract. Mech. 2020, 240, 107345. [Google Scholar] [CrossRef]
  103. Li, R.T.; Li, X.K. A coupled chemo-elastoplastic-damage constitutive model for plain concrete subjected to high temperature. Int. J. Damage Mech. 2010, 19, 971–1000. [Google Scholar]
  104. Yin, G.J.; Zuo, X.B.; Tang, Y.; Olawale, A.; Ding, D.N. Modeling of time-varying stress in concrete under axial loading and sulfate attack. Comput. Concr. 2017, 19, 143–152. [Google Scholar] [CrossRef]
  105. Yu, Y.G.; Gao, W.; Castel, A.; Liu, A.; Chen, X.J.; Liu, M. Assessing external sulfate attack on thin-shell artificial reef structures under uncertainty. Ocean Eng. 2020, 207, 107397. [Google Scholar] [CrossRef]
  106. Qin, S.; Zhou, D.; Liu, T.; Jiv, A. A chemo-transport-damage model for concrete under external attack. Cem. Concr. Res. 2020, 132, 106048. [Google Scholar] [CrossRef]
  107. Wang, P.; Mo, R.; Li, S.; Xu, J.; Jin, Z.Q.; Zhao, T.J.; Wang, D.Z. A chemo-damage-transport model for chloride ions diffusion in cement-based materials: Combined effects of sulfate attack and temperature. Constr. Build. Mater. 2021, 288, 123121. [Google Scholar] [CrossRef]
  108. Li, J.P.; Xie, F.; Zhao, G.W.; Li, L. Experimental and numerical investigation of cast-in-situ concrete under external sulfate attack and drying-wetting cycles. Constr. Build. Mater. 2020, 249, 118789. [Google Scholar] [CrossRef]
  109. Mazars, J.; Pyaudier-Cabot, G. Continuum damage theory-application to concrete. J. Eng. Mech. 1989, 115, 345–365. [Google Scholar] [CrossRef]
  110. Lubliner, J.; Oliver, J.; Oller, S.; Onate, E. A plastic-damage model for concrete. Int. J. Solids Struct. 1989, 25, 299–326. [Google Scholar] [CrossRef]
  111. Faria, R.; Oliver, J.; Cervera, M. A strain-based plastic viscous-damage model for massive concrete structures. Int. J. Solids Struct. 1998, 35, 1533–1558. [Google Scholar] [CrossRef]
  112. Grassl, P.; Jirásek, M. Damage-plastic model for concrete failure. Int. J. Solids Struct. 2006, 43, 7166–7196. [Google Scholar] [CrossRef] [Green Version]
  113. Grassl, P.; Xenos, D.; Nyström, U.; Rempling, R.; Gylltoft, K. CDPM2, A damage-plasticity approach to modelling the failure of concrete. Int. J. Solids Struct. 2013, 50, 3805–3816. [Google Scholar] [CrossRef] [Green Version]
  114. Wu, J.Y.; Li, J.; Faria, R. An energy release rate-based plastic-damage model for concrete. Int. J. Solids Struct. 2006, 43, 583–612. [Google Scholar] [CrossRef]
  115. Zhang, C.L.; Chen, W.K.; Mu, S.; Savija, B.; Liu, Q.F. Numerical investigation of external sulfate attack and its effect on chloride binding and diffusion in concrete. Constr. Build. Mater. 2021, 285, 122806. [Google Scholar] [CrossRef]
  116. Wang, H.L.; Chen, Z.; Li, H.; Sun, X.Y. Numerical simulation of external sulphate attack in concrete considering coupled chemo-diffusion-mechanical effect. Constr. Build. Mater. 2021, 292, 123325. [Google Scholar] [CrossRef]
  117. Zheng, F.; Wu, Z.; Gu, C.; Bao, T.F.; Hu, J. A plastic damage model for concrete structure cracks with two damage variables. Sci. China Technol. Sci. 2012, 55, 2971–2980. [Google Scholar] [CrossRef]
  118. Silva, D.; Fajardo-San-Miguel, G.; Escadeillas, G.; Cruz-Moreno, D. Surface treatment with silicon-based nanoparticles in Portland cement specimens subjected to physical sulfate attack. Case Stud. Constr. Mater. 2023, 18, e01795. [Google Scholar] [CrossRef]
  119. Esselami, R.; Wilson, W.; Tagni-Hamou, A. An accelerated physical sulfate attack test using an induction period and heat drying: First applications to concrete with different binders including ground glass pozzolan and limestone filler. Constr. Build. Mater. 2022, 345, 128046. [Google Scholar] [CrossRef]
  120. Singh, A.; Liu, Q.; Xiao, J.Z.; Lyu, Q.F. Mechanical and macrostructural properties of 3D printed concrete dosed with steel fibers under different loading direction. Constr. Build. Mater. 2022, 323, 126616. [Google Scholar] [CrossRef]
  121. Khosravani, M.R.; Haghighi, A. Large-scale automated additive construction: Overview, robotic solutions, sustainability, and future prospect. Sustainability 2022, 14, 9872. [Google Scholar] [CrossRef]
  122. Waseem, S.A.; Singh, B. An experimental study on shear capacity of interfaces in recycled aggregate concrete under multiaxial compression. Struct. Concr. 2018, 19, 230–245. [Google Scholar] [CrossRef] [Green Version]
  123. Zhou, J.H.; Kang, T.B.; Wang, F.C. Pore structure and strength of waste fiber recycled concrete. J. Eng. Fibers Fabr. 2019, 14, 1–10. [Google Scholar] [CrossRef]
Figure 1. Framework of the review.
Figure 1. Framework of the review.
Coatings 13 00174 g001
Figure 2. Chemical reactions in concrete under different cation sulfates [30].
Figure 2. Chemical reactions in concrete under different cation sulfates [30].
Coatings 13 00174 g002
Figure 3. Filling behavior of ettringite crystal in pore system of local concrete.
Figure 3. Filling behavior of ettringite crystal in pore system of local concrete.
Coatings 13 00174 g003
Figure 4. Filling of sulfate products in pore [70].
Figure 4. Filling of sulfate products in pore [70].
Coatings 13 00174 g004
Figure 5. Expressions of damage degree in concrete caused by CSA. Note: f cons , f geom , and f dama are the functions of stress–strain relationship, strain–displacement relationship, and damage degree. F is an external load; u is the displacement.
Figure 5. Expressions of damage degree in concrete caused by CSA. Note: f cons , f geom , and f dama are the functions of stress–strain relationship, strain–displacement relationship, and damage degree. F is an external load; u is the displacement.
Coatings 13 00174 g005
Table 1. Reaction equations and formation mechanisms of products in concrete.
Table 1. Reaction equations and formation mechanisms of products in concrete.
Cation TypeSulfate ProductFormation MechanismChemical Reaction
Na+/K+Gypsum C S ¯ H 2 Ion–ion reaction Ca 2 + + SO 4 2 + 2 H 2 O C S ¯ H 2
Ettringite C 6 A S ¯ 3 H 32 Topochemical mechanism
(Solid–solid reaction)
Equation ( a ) :   C 4 AH 13 + 3 C S ¯ H 2 + 14 H 2 O C 6 A S ¯ 3 H 32 + CH Equation ( b ) :   C 4 A S ¯ H 12 + 2 C S ¯ H 2 + 16 H 2 O C 6 A S ¯ 3 H 32 Equation ( c ) :   C 3 A + 3 C S ¯ H 2 + 26 H 2 O C 6 A S ¯ 3 H 32
Through-solution mechanism
(Ion–ion reaction)
Al OH 4 + 2 OH Al OH 6 3 Al OH 6 3 + 3 Ca 2 + + 12 H 2 O Ca 3 Al OH 6 12 H 2 O 3 + 2 Ca 3 Al OH 6 12 H 2 O 3 + + 3 SO 4 2 + 2 H 2 O C 6 A S ¯ 3 H 32
Thaumasite C 3 S C ¯ S ¯ H 15 Direct reaction
(Ion–ion reaction)
SO 4 2 + 3 Ca 2 + + CO 3 2 + SiO 3 2 + 15 H 2 O C 3 S C ¯ S ¯ H 15
Indirect reaction
(Solid–solid reaction)
C 3 S 2 H 3 + C 6 A S ¯ 3 H 32 + 2 C C ¯ + 4 H 2 C 3 S C ¯ S ¯ H 15 + C S ¯ H 2 + AH 3 + 4 CH
Mg2+Brucite Mg OH 2 Ion–ion reaction Mg 2 + + SO 4 2 + Ca 2 + + 2 OH + 2 H 2 O MH + C S ¯ H 2 a Mg 2 + + a SO 4 2 + a CaO S i O 2 x H 2 O + 2 H 2 O a MH + a C S ¯ H 2 + S H x 2 a Mg 2 + + 2 a SO 4 2 + 2 a CaO Si O 2 x H 2 O + 6 a 1 2 x H 2 O 2 a 3 MH + 2 a C S ¯ H 2 + M 3 S 2 H 2
H 2 O H ; Ca OH 2 CaO H 2 O CH ; 2 Al OH 3 Al 2 O 3 3 H 2 O AH 3 ; CaSO 4 2 H 2 O CaO SO 3 2 H 2 O C S ¯ H 2 ; CaCO 3 CaO C O 2 C C ¯ ; Mg OH 2 MgO H 2 O M H ; SiO 2 x H 2 O S H x ; 3 MgO 2 SiO 2 2 H 2 O M 3 S 2 H 2 .
Table 2. Relationship among cation type of sulfate solution, CSA classification, and failure forms.
Table 2. Relationship among cation type of sulfate solution, CSA classification, and failure forms.
Cation TypeCSA Type/Sulfate Product
ThaumasiteGypsumEttringiteBrucite
Na+/K+   ( CO 2   or   C O 3 2 )√ (main)√ (main)
Mg2+ √ (main)
Failure formCohesiveness
Softening
Volume expansion and its-induced cracking/spalling
Table 3. Influence factors of sulfate ion transport in concrete.
Table 3. Influence factors of sulfate ion transport in concrete.
Basic Form C t = J Ion   tan sform + k r C Chemical   consumption
Ion flowDiffusion behavior J = D C
Mutual restriction effect between charged ions J = D C Ψ
Influence of ionic chemical activity J = D C ln γ
Ion migration effect with solution convection J = C v
Temperature effect J = D C T T
   Coupled by the above factors J = D C + D C Ψ + D C ln γ + C v
Effective diffusivityTemperature effect D = D c min + D c 0 T   Tempeture D c min Product   filling f φ Porosity or   D cr σ Crack   effect   ( stress ) or   D w φ σ Load τ σ , φ 1 Load ,   tortuosity or   D c 0 φ w hydration + d C , t ESA - induced   damage
Sulfate products filling in pore
ESA-induced microcrack effect
Damage degree
Cement hydration
J is the ion flow in concrete; D is the effective diffusivity of sulfate ion in concrete; C is the sulfate ion concentration in concrete; the term k r C is the consumption of sulfate ion due to chemical reaction in concrete; the term Ψ reflects the effect of Electric field; the term ln γ reflects the influence of ionic chemical activity; v is the solution flow rate; d is the damage degree; φ is the porosity; τ is the tortuosity; T is the temperature; σ is the stress.
Table 5. Chemo-mechanical models for concrete subjected to CSA.
Table 5. Chemo-mechanical models for concrete subjected to CSA.
AuthorPublished TimeDegradation CauseDegradation Mechanism
Tixier and Mobasher [78,86]2003EttringiteVolume increase theory
Bary [84] 2008Ettringite, GypsumCrystallization pressure theory
Bary et al. [87]2014EttringiteVolume increase theory, crystallization pressure theory
Basista and Weglewski [79]2009EttringiteVolume increase theory
Sarkar et al. [88]2010EttringiteVolume increase theory
Sarkar et al. [89]2012Ettringite, GypsumVolume increase theory
Idiart et al. [80]2011EttringiteVolume increase theory
Ikumi et al. [81]2014EttringiteVolume increase theory
Ikumi et al. [90]2016EttringiteVolume increase theory
Cefis and Comi [91]2014EttringiteVolume increase theory
Cefis and Comi [92]2017EttringiteVolume increase theory
Zuo et al. [93]2009EttringiteVolume increase theory
Zuo et al. [94]2012EttringiteVolume increase theory
Nie et al. [95]2015EttringiteVolume increase theory
Yin et al. [96] 2017Ettringite, GypsumVolume increase theory
Yin et al. [97]2019 EttringiteCrystallization pressure theory
Yu et al. [64]2018EttringiteVolume increase theory
Yu et al. [98]2021EttringiteVolume increase theory
Yi et al. [99]2019EttringiteVolume increase theory
Table 6. Basic constitutive equations in the chemo-mechanical models.
Table 6. Basic constitutive equations in the chemo-mechanical models.
AuthorsBasic Mechanical EquationNumber
Saetta et al. [100,101] σ = 1 d c Chemical   damage 1 d l Load   damage 0 : ε (8)
dc is the CSA-induced chemical damage. d l is the loading damage caused by external loading. This model can be unsed to analyze the stress in concrete under the external load and environment corrosion.
Sarkar et al. [88] σ = 1 d c Chemical   damage E 0 ε Related   to   free   expansion   ε V , d c = f ε (9)
In sraker’s model, the ε is calclulated by the free expnasion and ε V is caused by the growth of sulfate products.
Bary et al. [87] σ = 1 d c Chemical   damage 0 : ε ε es Related   to   free   expansion   ε V α AFm p c 1 Crystallization   pressure (10)
The effects of crystallization pressure and free expansion are considered together in the model.
Cefis and Comi [92] σ = 1 d c Chemical   damage 1 d m Mechanical   damage 2 G e + K tr ε 1 b p w 1 Hydrostatic   pressure (11)
In this model, the CSA-induced mechanical and chemical damages are considered. Additionally, the hydrostatic pressure pw is introduced in the basic constitutive equation. So, this model is applicable to the case of sulfate attack on unsaturated concrete.
Ikumi et al. [81] σ = 1 d c Chemical   damage E 0 ε ε non - mech Related   to   free   expansion   ε V (12)
Ikumi’s model is similar to that of Sarkar, but he considered the influence of pore size on ε V .
Yin et al. [96] σ = 1 d l Load   damage 1 d c Chemical   damage 0 : ε ε p (13)
Yin et al. [102] σ = 1 d c Chemical   damage 0 : ε ε p ε EG Related   to   free   expansion   ε V (14)
In Yin’s models, the chemical damage dc is analyzed by using the elastoplastic damage mechanics, not determined by the empirical formula related to the content of sulfate ion or the reaction product.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Yin, G.-J.; Wen, X.-D.; Miao, L.; Cui, D.; Zuo, X.-B.; Tang, Y.-J. A Review on the Transport-Chemo-Mechanical Behavior in Concrete under External Sulfate Attack. Coatings 2023, 13, 174. https://doi.org/10.3390/coatings13010174

AMA Style

Yin G-J, Wen X-D, Miao L, Cui D, Zuo X-B, Tang Y-J. A Review on the Transport-Chemo-Mechanical Behavior in Concrete under External Sulfate Attack. Coatings. 2023; 13(1):174. https://doi.org/10.3390/coatings13010174

Chicago/Turabian Style

Yin, Guang-Ji, Xiao-Dong Wen, Ling Miao, Dong Cui, Xiao-Bao Zuo, and Yu-Juan Tang. 2023. "A Review on the Transport-Chemo-Mechanical Behavior in Concrete under External Sulfate Attack" Coatings 13, no. 1: 174. https://doi.org/10.3390/coatings13010174

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop