Next Article in Journal
Comparison between Mechanical Properties and Joint Performance of AA 2024-T351 Aluminum Alloy Welded by Friction Stir Welding, Metal Inert Gas and Tungsten Inert Gas Processes
Previous Article in Journal
Ferrovalley and Quantum Anomalous Hall Effect in Janus TiTeCl Monolayer
Previous Article in Special Issue
Feasibility Study of Pervious Concrete with Ceramsite as Aggregate Considering Mechanical Properties, Permeability, and Durability
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Modeling of Concrete Deterioration under External Sulfate Attack and Drying–Wetting Cycles: A Review

1
School of Construction Engineering, Shenzhen Polytechnic University, Shenzhen 518055, China
2
School of Civil and Environmental Engineering, Harbin Institute of Technology, Shenzhen, Shenzhen 518055, China
*
Author to whom correspondence should be addressed.
Materials 2024, 17(13), 3334; https://doi.org/10.3390/ma17133334
Submission received: 31 May 2024 / Revised: 1 July 2024 / Accepted: 3 July 2024 / Published: 5 July 2024
(This article belongs to the Special Issue New Advances in Cement and Concrete Research2nd Edition)

Abstract

:
This paper comprehensively summarizes moisture transport, ion transport, and mechanical damage models applied to concrete under sulfate attack and drying–wetting cycles. It highlights the essential aspects and principles of each model, emphasizing their significance in understanding the movement of moisture and ions, as well as the resulting mechanical damage within the concrete during these degradation processes. The paper critically analyzes the assumptions made in each model, shedding light on their limitations and implications for prediction accuracy. Two primary challenges faced by current models under sulfate attack and drying–wetting cycles are identified: the limited consideration of the coupled effects of chemical and physical attacks from sulfate, and the unclear mechanism of the sulfate attacks. Future research directions are proposed, focusing on exploring the transport mechanism of sulfate ions under various driving forces and further clarifying the crystallization process and expansion damage mechanism in concrete pores. Addressing these research directions will advance our understanding of sulfate attack under drying–wetting cycles, leading to improved models and mitigation strategies for enhancing the durability and performance of concrete structures.

1. Introduction

Concrete structures in saline–alkali soils, salt lakes, mining sites, and coastal areas face significant challenges due to the high concentrations of sulfate ions in these regions. The presence of sulfate ions can trigger reactions with the hydration products of concrete, forming expansive products such as ettringite and gypsum. These products exert pressure on the walls of capillary pores within concrete, resulting in cracking, spalling, and softening.
Furthermore, dynamic weather conditions, groundwater fluctuations, and tidal movements introduce drying–wetting cycles to the concrete structures. Extensive studies have demonstrated that exposure to drying–wetting cycles significantly accelerates the degradation rate of concrete compared to full immersion conditions [1,2,3]. During drying–wetting cycles, moisture continuously moves in and out of the concrete, facilitating the penetration of sulfate ions into its inner layers. This increased penetration rate intensifies the potential for sulfate attack. Once sulfate ions infiltrate the concrete, they initiate chemical reactions with the cementitious materials, forming expansive products. Simultaneously, the sulfate ions induce severe physical crystallization damage due to moisture variations. This combined chemical and physical attack accelerates the expansion and cracking of concrete, further exacerbating the degradation of concrete structures [4,5,6].
Extensive research [7,8,9,10,11,12,13,14,15] has been conducted on the degradation of concrete fully submerged in sulfate solutions, providing valuable insights into the influence of sulfate attack on concrete’s mechanical properties. Studies [16,17] have shown that regardless of whether concrete is fully or partially immersed, the volume of pore structures initially decreases during the early stages of sulfate attack. However, in the middle and later stages, the volume increases. This is attributed to the formation of products like chemical reaction products or sodium sulfate crystals, which initially reduce the available pore space. However, as sulfate attack progresses, the destruction of aggregate–paste structures or the precipitation of sodium sulfate crystals leads to the development of cracks, causing an increase in pore volume during the middle and later stages of sulfate attack. The compressive strength, elastic modulus, and tensile strength of concrete under sulfate attack also typically depict a dual-phase trend, characterized by an initial slow increase followed by a significant decrease [18,19,20,21,22]. These changes in mechanical properties can be attributed to the alterations in pore structures caused by sulfate attack. The reduction in pore volume indicates a densification of the concrete, leading to an increase in its mechanical properties. Conversely, when the pore volume increases, the concrete becomes more porous and less dense, resulting in a decrease in its mechanical properties. The expansion trend induced by sulfate attack in concrete is characterized by a slow initial growth, followed by a rapid increase, before finally reaching a stable state [23,24,25,26,27]. This progression occurs as the expansion products gradually fill the pores during the early stage, resulting in a gradual expansion. Once a certain portion of the pores is completely filled, the ongoing increase in expansion products significantly accelerates the expansion rate. When the expansion strain exceeds the tensile strength of the material, cracks are initiated and propagated, providing additional space for further expansion and ultimately leading to a stable state. Furthermore, several studies [28,29,30,31] have proposed that the depth of the sulfate attack can serve as an indicator for assessing concrete degradation. This depth represents the extent of expansion and damage in the affected area and exhibits a similar trend to the expansion behavior.
In order to simulate the damage and deterioration caused by sulfate attack, numerical models have been developed, incorporating processes such as ion transport, chemical reactions, and expansive deformation [7,32,33,34]. However, the process of sulfate attack under drying–wetting cycles presents a significantly more intricate scenario. During the drying–wetting cycles, ion movement in concrete is influenced by diffusion, driven by concentration gradients, and convection, driven by saturation gradients. This combination introduces a more complex ion transport process compared to fully saturated conditions. Both diffusion and convection play significant roles in determining ions’ movement and distribution within the concrete [35]. In drying–wetting cycles, a sulfate attack on concrete involves chemical reactions and physical crystallization. Chemical reactions between sulfate ions and the hydration products of concrete result in the formation of expansive products such as ettringite [5,35,36,37,38,39,40]. Simultaneously, the physical attack occurs as sulfate ions induce crystalline growth due to changes in saturation within the concrete, resulting in expansion and cracking [25,26,41,42,43,44,45,46,47]. However, the precise interplay between these coupled mechanisms during sulfate attack under drying–wetting cycles remains unclear. There is currently a lack of consensus among researchers regarding the specific crystallization processes and types of sulfate within the concrete pores.
Further research is imperative to enhance our understanding of the intricate deterioration process of concrete subjected to sulfate attack and drying–wetting cycles. The present study provides a comprehensive analysis of the numerical modeling of concrete deterioration under sulfate attack and drying–wetting cycles. It offers a systematic overview of the assumptions and limitations associated with the models of moisture transport, ions transport, chemical reaction, physical crystallization, and expansion damage. This analysis serves as a valuable reference for future researchers interested in conducting studies related to deterioration models for concrete under sulfate attack and drying–wetting cycles. By understanding the assumptions and limitations of these models, researchers can build upon existing knowledge and develop more accurate and comprehensive models to better predict the deterioration behavior of concrete under sulfate attack and drying–wetting cycles.

2. Moisture Transport Model

Moisture transport within concrete can be effectively described using Darcy’s law and mass conservation principle. These fundamental principles establish a relationship between the flow of moisture through porous media and the permeability coefficients, as follows:
d s d x = D m s d s d x
where s denotes the saturation of pores, which indicates the water content within the concrete. Dm is the moisture transport coefficient. To establish a comprehensive moisture transport model for concrete, defining the moisture transport coefficient is crucial to establish a relationship between the moisture content and the pressure gradient. This relationship serves as a key component in understanding how pressure changes influence moisture movement within the concrete. Several researchers have proposed different models for moisture transport in concrete based on various assumptions. These models can be broadly categorized into two main types: theoretical models and empirical models.

2.1. Theoretical Model

Theoretical models are based on fundamental principles and mathematical equations derived from the governing equations of fluid flow and transport phenomena. These models aim to describe the physical processes and mechanisms involved in moisture transport within the concrete matrix.
Under drying–wetting cycles, the moisture state within concrete undergoes continuous changes. Pel et al. [48] observed that concrete’s water vapor transport coefficient varies with internal saturation levels. Their findings indicate that the coefficient increases with higher internal saturation levels and decreases with increasing saturation at lower levels. This behavior can be attributed to the change in the primary form of water transport within the concrete. In highly saturated conditions, liquid water becomes the dominant mode of water transport. As the amount of liquid water increases, water transfer efficiency also increases, resulting in a higher water vapor transport coefficient. However, as the saturation level falls below a critical threshold, gaseous water vapor replaces liquid water as the primary mode of transport. In this case, water vapor density increases as saturation decreases. This increase in water vapor density leads to faster water vapor transport through the concrete. To capture the water transport phenomenon under drying–wetting cycles, Zhang et al. [49] proposed a model that considers two components of water within the concrete: gaseous water vapor and liquid water. They separately investigated the contributions of these two states to the water vapor transport coefficient, as represented by Equation (2).
D m s = D l s + D v s = k l + k v p s
where Dl and Dv represent the transport coefficients of liquid water and water vapor, respectively. kl and kv indicate the permeability coefficients of liquid water and water vapor, respectively. s denotes the saturation of pores within the concrete, indicating the water content. p is the pore pressure, which is associated with relative humidity (RH) and represents the energy state of water within the concrete [50].
The balance in the relationship between the moisture content (s) within concrete and the energy (p) or relative humidity (RH) is a crucial aspect in the study of moisture migration within concrete. The moisture transport process within concrete exhibits a hysteresis effect, as depicted in Figure 1. During the drying process, as the RH decreases, the moisture within the larger pores tends to evaporate first. However, as the RH drops further, the moisture in the smaller pores may become trapped and unable to freely evaporate due to the blockage of the evaporative pathway by the surrounding smaller pores. This hindering of moisture movement results in the drying curve in the relationship between moisture content (s) and RH being higher than the wetting curve. In other words, for a given RH, the concrete will have a higher moisture content during the drying stage compared to the wetting stage. To account for the hysteresis effect, Zhang et al. [49] developed separate equations for moisture content during the drying and wetting processes, as shown in Equations (3) and (4):
Drying process:
s d = 1 exp ( B r c )
Wetting process:
s w = 1 exp ( B r c ) × 1 ln 1 exp ( B r c )
where B represents the parameter of the Rayleigh–Ritz distribution (R-R distribution), which can be obtained from mercury intrusion porosimetry tests. rc denotes the modified critical pore radius, taking into account the adsorbed water film on the surface of the pores, which is adsorbed by the pore solution, as shown in Figure 2. The relationship between the modified critical pore radius (rc) and the pore pressure (p) can be expressed as follows:
r c = 2 C γ cos θ p
where C represents the regression parameter, based on the modified BET (Brunauer–Emmett–Teller) theory [51], taking 2.15. γ is the liquid surface tension, with a value of 0.073 N/m. θ denotes the contact angle between water and cement, assuming complete wetting of concrete, so θ is 0. By substituting Equation (5) into Equations (3) and (4), the relationship between moisture content (s) and pore pressure (p) is established as follows:
Drying process:
s d = 1 exp 2 B C γ cos θ p
Wetting process:
s w = 1 exp 2 B C γ cos θ p 1 ln 1 exp 2 B C γ cos θ p
For the permeability coefficients of liquid water (kl) in Equation (2), assuming that the capillary pores within the concrete are all interconnected cylindrical pores, and the convective flow of the pore solution occurs only within saturated pores while neglecting the transport of water vapor in unsaturated pores, then the coefficient kl can be simplified as [52]:
k l = ρ l φ 2 50 B 2 η 1 1 2 B C γ cos ( θ ) p exp ( 2 B C γ cos ( θ ) / p ) 2
where ρl represents the density of liquid water (kg/m3). φ is the porosity of concrete. η is the viscosity coefficient of the pore solution (Pa·s).
Different researchers have proposed various models for the permeability coefficients of water vapor (kv) in Equation (2). Zhang [49] developed a calculation model for the saturation degree (s) and the relative humidity (RH) within the concrete, expressed as follows:
k v 1 = ρ v φ D v ( π / 2 ) 2 1 s 1 + l m / 2 r m t m M w ρ l R T exp 2 p M w ρ l R T
where ρv represents the density of water vapor(kg/m3). Dv is the free diffusion coefficient of water vapor (m2/s). lm denotes the mean free path of water vapor molecules (m). rm represents the average radius of unsaturated pores (m). tm is the thickness of the adsorbed liquid water layer in a pore of size rm (m). Mw indicates the molar mass of water (18 g/mol). R is the universal gas constant (8.3143 J/mol/K). T represents the temperature (K). It is noted that this equation encompasses many parameters and is complex to solve. To simplify it, Wu et al. [53] adopted the equation of kv2 with relative humidity as a variable, as follows:
k v 2 = D v 1 + 1 + RH 4 1 + RH 0 4 1 RH p
where RH0 represents the critical relative humidity (75%). However, in the moisture transport models, the variable is generally the moisture content (s). Sun et al. [35], considering the influence of porosity and temperature, provided a relationship between moisture content (s) and permeability coefficients of water vapor (kv3) as follows:
k v 3 = 0.217 p a t m p g T T 0 1.88 φ a ( 1 s ) b ρ v M w ρ l 2 R T φ
where Patm is the reference atmospheric pressure (Pa). Pg represents the vapor pressure (Pa). T0 denotes the reference temperature (K). a and b are the fitting parameters.
After establishing the relationship between moisture content (s) and pore pressure (p) and determining the parameters for permeability coefficients of liquid water (kl) and water vapor (kv), the moisture transport coefficient models for both drying and wetting conditions can be obtained by substituting these parameters into Equation (1). This model is developed based on an understanding of the internal moisture migration mechanism within concrete, considering the separate transport of gaseous water vapor and liquid water to accurately represent the drying and wetting processes. By utilizing this model, researchers can better understand the complex moisture transport phenomena occurring within concrete. This includes the pore structure, saturation levels, and the hysteresis effect observed between drying and wetting. The model provides insights into how different parameters influence moisture transport and helps predict moisture content under various conditions. While the theoretical model for moisture transport in concrete offers a comprehensive understanding of the underlying mechanisms, it can present challenges when applied in practical engineering applications. The model equations can be complex, involving numerous parameters. This complexity can make implementation challenging, both in terms of computational feasibility and obtaining accurate parameter values.

2.2. Empirical Model

Empirical models are constructed by analyzing experimental observations and data. These models use statistical correlations and regression analyses to establish relationships between input parameters and moisture transport behavior. Empirical models are simpler to implement and require fewer input parameters compared to theoretical models. Table 1 presents empirical models for moisture transport coefficients during the drying and wetting processes obtained by different researchers.
Li et al. [50] utilized theoretical and empirical models for moisture transport coefficients to predict water loss and absorption in concrete during the drying and wetting processes. These models were compared with the experimental results to assess their accuracy in capturing moisture transport behavior. The findings indicated that both the theoretical and empirical models effectively predict concrete moisture transport. However, the empirical model demonstrated greater accuracy in predicting moisture distribution during wetting. Similarly, Sun et al. [35] conducted experiments to investigate moisture transport in concrete and compared the empirical model to the experimental data. Their results showed a close alignment between the empirical model and the experimental data, particularly during the drying process. This close agreement between the empirical model and the experimental results further supports the effectiveness of the empirical approach in capturing moisture transport behavior in concrete.
In summary, both empirical and theoretical models for moisture transport coefficients can effectively capture the moisture transport behavior in concrete during drying and wetting processes. However, empirical models have distinct advantages in practical applications. Empirical models demonstrate a good fitting performance, accurately representing the moisture transport behavior observed in the experimental data. They are user-friendly and easy to use, requiring minimal theoretical knowledge for implementation. Additionally, empirical models offer fast calculations, enabling efficient assessments and prompt decision-making. Due to their simplicity, accuracy, and efficiency, numerous researchers favor empirical models for practical applications. As for the modeling approach, empirical models often employ an exponential relationship to describe moisture transport during wetting. This relationship captures the non-linear nature of moisture absorption and adsorption phenomena during wetting. During drying processes, the moisture transport coefficient is often assumed to be constant in empirical models. This assumption simplifies the modeling process and provides a reasonable approximation for drying conditions where the moisture content decreases.

3. Ion Transport Model

Under saturated conditions, the transport model for sulfate ions in concrete is typically based on Fick’s law, which describes diffusion as the predominant mechanism for ion transport. However, under conditions of drying–wetting cycles, the driving force for ion transport changes, and the convective effect becomes influential [37]. The convective effect refers to the movement of ions through the concrete due to the flow of water, which is particularly important in the surface layer of concrete subjected to drying–wetting cycles. This highlights the need to consider diffusion and convection mechanisms in the transport model to predict ion ingress and assess ion-induced damage in concrete accurately.
Liu et al. [59] established a transport model for sulfate ions under drying–wetting cycles, assuming that the amount of ion substance entering the concrete during each wetting process is proportional to the moisture content. After k drying–wetting cycles, the amount of sulfate ion substance in the pore solution, denoted as nk, equals the following:
n k = n k 1 + c p v p Δ s , t h e w e t t i n g p r o c e s s n k 1 , t h e d r y i n g p r o c e s s
where the cp is the concentration of sulfate ions in the pore solution (mol/m3), vp is the volume of pores in the concrete (m3), and Δs is the difference in the saturation degree of the pore solution between the initial and subsequent moments. In the described model, the total amount of sulfate ion is assumed to remain constant during the drying process. As moisture evaporates from the concrete, the pore solution volume decreases, increasing the concentration of sulfate ions. This change in pore solution concentration reflects the evolving characteristics of sulfate ion concentration throughout the drying process. However, this assumption oversimplifies the situation by neglecting ion diffusion and the consumption of sulfate ions through chemical reactions with cementitious materials. As a result, the model implies that sulfate ion levels in the concrete only increase with the number of drying–wetting cycles, disregarding the potential depletion of sulfate ion concentration caused by chemical reactions.
Zheng et al. [60] developed a model based on Fick’s law to investigate the transport of sulfate ions in concrete. They incorporated alternating ion concentration boundary conditions to account for the influence of drying–wetting cycles on sulfate ion transport. The formulation of their model is as follows:
c S O t = x φ D S O c S O x k c S O c C A
where cSO and cCA represent the concentration of sulfate ions and calcium aluminates in the matrix (mol/m3), respectively. DSO is the diffusion coefficient of sulfate ions (m2/s). k denotes the chemical reaction rate constant (m3/mol/s). t indicates the diffusion time. By considering these varying conditions, we aimed to enhance the understanding of how sulfate ions are transported in concrete under different moisture conditions. However, it should be noted that this model does not account for factors such as moisture content and convection.
Zhang et al. [61] introduced a connection between moisture transport and sulfate diffusion in their model by incorporating the saturation degree. They accounted for the hysteresis effect of pore solution evaporation and changes in porosity due to chemical reactions. The ion transport equation is expressed as follows:
c S O t = div ( D S O φ s c S O ) + c d t
( φ s ) t = div D m d ( φ , s ) ( φ s )
( φ s ) t = div D m w ( φ , s ) ( φ s )
where cd is the concentration of sulfate ions consumed by chemical reactions with cement materials. Dmd and Dmw indicate the moisture transport coefficient Dm during the drying and wetting processes, respectively. On this basis, Li et al. [62] considered convection as follows:
c S O t = 1 r r ( D S O s c S O + D m ( s ) ( s ) ) c d t
where r is the radial distance. This comprehensive model considers various factors influencing ion transport, including moisture transport, chemical reactions, and convective effects.
Shan et al. [63] considered diffusion, convection, and chemical reactions in their model for sulfate ion transport under drying–wetting cycles. They incorporated factors such as porosity, tortuosity, sulfate attack-induced damage, and humidity in the diffusivity of sulfate ions. The model is expressed as follows:
φ s c S O t + D S O c S O + φ s c S O s t + φ s c d t = 0
D S O = f d d c f 1 s φ τ D s 0
where Ds0 is the diffusion coefficient of sulfate ions in water. τ is the tortuosity. fd(dc) and f1(s) represent the influence of sulfate attack-induced damage and pore moisture content on diffusion coefficient, respectively. Additionally, Yin et al. [57] extended the model by including the influence of temperature on the diffusion coefficient, expressed as follows:
f T T = exp 2170 1 T r e f 1 T
where Tref is the reference temperature.
These sulfate transport models highlight the importance of considering multiple factors, such as diffusion, convection, chemical reactions, moisture content, porosity, and temperature, to accurately predict sulfate ion transport and assess its impact on concrete structures. By incorporating these factors, researchers strive to develop comprehensive models that capture the complexities of sulfate ion transport phenomena and improve our understanding of sulfate attack mechanisms in concrete.

4. Expansion Damage Mechanism

4.1. Microscopic Expansion Mechanisms

The mechanism of concrete damage caused by sulfate attack continues to be a topic of substantial debate among researchers [5,36]. Despite numerous studies and research efforts, there is still no consensus on the exact mechanisms underlying sulfate attack on concrete. The most widely applied theories in the field are the volume increase theory and the salt crystallization theory. The volume increase theory suggests that during the process of a sulfate attack, the generation of expansive products leads to an additional increase in volume, resulting in expansion. This theory quantifies the expansion by calculating the difference in volume between the solid reaction products and the solid reactants:
ν = V s V r V r
where ν is the expansion coefficient. ∑Vs is the total volume of solid reaction products. ∑Vr is the total volume of solid reactants. The volume increase theory provides a quantitative method for assessing the volumetric changes during a sulfate attack on concrete. It establishes a direct relationship between the macroscopic expansion of concrete and the amount of expansion products formed [27]. This approach simplifies the calculation by focusing solely on the volume occupied by the solid reactants and products. However, it is crucial to acknowledge that this method lacks adequate experimental evidence to substantiate its accuracy [25,36,64]. In addition, thermodynamic calculations have revealed that when considering water consumption, the total volume of reactants can be smaller than the total volume of products [27]. Some researchers have also pointed out that the additional volume formed by ettringite and gypsum is typically lower than the available free pore space [65]. These findings highlight the importance of considering factors such as water consumption and available pore space when evaluating the volume changes associated with sulfate attack.
The salt crystallization theory suggests that the expansion observed in concrete is attributed to the crystallization of sulfates from supersaturated solutions. This crystallization process exerts pressure on the pore walls, expressed as follows:
P = R T V c r y ln Q K
where P is pressure (Pa). Vcry represents the molar volume of the crystal (m3/mol). K indicates the equilibrium constant. Q is the ionic activity product. In the given equation, the generation of salt crystallization pressure requires the fulfillment of two necessary conditions. Firstly, the ionic activity product must be larger than the equilibrium constant, where Q > K; otherwise, the calculated salt crystallization pressure would be zero or negative. Secondly, the growing crystals must be restricted within the confined space of the concrete pores. As they continue to crystallize, the crystals exert pressure against the pore walls, causing them to expand. This is why not all formed crystals exhibit expansion, as the crystals need to be constrained to generate crystallization pressure. Some researchers have assessed the influence of crystal size and shape on salt crystallization pressure under constrained conditions. Scherer et al. [41,42,43,66] demonstrated an inverse relationship between the pressure exerted and the size of the pore. It is suggested that smaller pores may experience higher pressure from salt crystallization. Some researchers [42,43,64,67] thought that the morphology of crystals determines the energy balance between the surface atoms of the crystal and the concentration of ions in the surrounding solution. Smaller crystals with higher curvature can generate greater crystallization pressure in equilibrium conditions. Accordingly, the locations of ettringite growth within the microstructure have a greater influence on the development of expansive forces during sulfate attack, rather than the total volume of ettringite formed [5]. The crystallization pressure theory has been extensively discussed from a theoretical perspective in various publications, and it is widely recognized as the most plausible explanation. However, researchers have faced challenges in obtaining conclusive experimental data to directly support this theory in the context of sulfate attack [25].
Bary et al. [46] introduced a chemo–transport–mechanical model, proposing that the observed expansion during sulfate attack is a consequence of the combined effects of volume increase and salt crystallization. They found that the calculated free expansion rate aligns closely with experimental measurements. Furthermore, their research suggests that the contribution of crystallization pressure can be neglected when compared to the strain induced by the increased volume resulting from ettringite formation. This indicates that volume increase plays a dominant role in macroscopic strain development. Ikumi et al. [26] also pointed out that the volume increase and salt crystallization theories may be compatible, as they represent two different sulfate attack stages. According to their perspective, when sulfate ions enter and form ettringite, eventually reaching the solubility limit, the system tends to restore equilibrium via precipitating ettringite. However, if the release of energy through crystal precipitation is hindered, it is released as pressure on the pore walls, resulting in the formation of microcracks. These microcracks alleviate the pressure conditions within the pores, leading to ettringite precipitation near the cracks. Consequently, a proportional increase in the macroscopic free strain corresponds to the amount of precipitated ettringite. This suggests that the initial generation of macroscopic strain is primarily attributed to the action of crystallization pressure, while macroscopic free expansion is explained by a volume increase.

4.2. Macroscopic Expansion Calculation

Numerous scholars have employed a variety of theories to conduct numerical simulations of the mechanical degradation and damage process of concrete induced by sulfate attack and drying–wetting cycles. Chen [68] presented a methodology that utilizes crystallization pressure and micropore mechanics to convert the internal crystallization pressure into strains within concrete. The approach can be summarized as follows:
ε = P l + S c R T V m ln β × v p × V a V c / k e
where ε represents the strain in concrete. Pl denotes the pressure generated by the salt solution (MPa). ke is the elastic modulus after a certain period of degradation (MPa). Sc indicates the volume fraction of crystals. Vm stands for the molar volume of the crystals. β is the supersaturation of the pore solution (cm3/mol). Va/Vc represents the volume fraction of aggregates. vp is the coefficient that accounts for the influence of pores. In this model, the relationship between physical crystallization stress and ion concentration in concrete under drying–wetting cycles is established using sodium nitrate, while the damage caused by chemical attack is determined through full immersion with sodium sulfate. The physical and chemical stress obtained from these experiments are combined to calculate the total erosion strain caused by sulfate attack under drying–wetting cycles. The assumption underlying this approach is that, under identical concentration and wet–dry cycle conditions, the ion concentration distribution and physical crystallization stress generated by sodium nitrate are equivalent to those of sodium sulfate. However, it is important to note that the crystallization pressure of sodium nitrate is significantly lower than that of sodium sulfate.
Yang [69] assumed that the precipitation of Na2SO4 · 10H2O during the wetting process is the primary cause of physical crystallization damage. Based on the theory of porous media mechanics, the strains generated in concrete during the wetting process were characterized as follows:
ε = b p l + b c R T V m ln β K
where K represents the bulk modulus of concrete (MPa). b and bc denote the Biot coefficient of concrete and crystals, respectively. However, this model overlooks the influence of sulfate ions reacting with cementitious materials. The formation of expansive products can lead to additional internal stresses and damage within the concrete, significantly impacting the mechanical degradation and damage process.
Ren et al. [70] developed a numerical model that considers both the chemical and physical effects of sulfate attack under drying–wetting cycles. By analyzing the volume fractions of CaSO4 · 2H2O and Na2SO4 · 10H2O within the pores following various numbers of drying–wetting cycles, the macroscopic tensile stress PC is calculated as follows:
P C = R T v M ln K S P , T ln K S P , M + 10 ln R H s a t , T 100 + R T v G ln Q r e a c   K r e a c  
where RHsat,T represents the relative humidity. KSP,i is the thermodynamic solubility product for the corresponding crystal i. Qreac denotes the ionic activity of the reactants. Kreac indicates the chemical reaction equilibrium constant. In this model, only the formation of gypsum during the chemical reaction is considered, and a linear increase in the concentration of sulfate ions within the concrete is assumed, starting from an initial concentration. However, this assumption contradicts the actual ion transport behavior observed in the process [71].

5. Future Work

The aforementioned works provide a comprehensive analysis of numerical models, including the moisture transport model, ion transport model, and mechanical damage model. These models are specifically designed to assess the degradation of concrete under the conditions of sulfate attack and drying–wetting cycles. Overall, the current models for sulfate attack under drying–wetting cycles face two primary challenges. Firstly, many existing mechanical damage models tend to linearly combine the effects of sulfate chemical attacks and physical attacks, overlooking a comprehensive consideration of their coupled effects. Secondly, the mechanism of sulfate attack under drying–wetting cycles remains unclear.
To address these limitations, exploring the transport mechanism of sulfate ions within concrete under various driving forces, including concentration gradient, saturation gradients, calcium leaching, temperature effects, and loading effects, is crucial for gaining valuable insights into the deterioration process. Investigating these factors, either individually or in combination, will enhance our understanding of how sulfate ions migrate and interact within the concrete matrix. This can help to identify the pathways and mechanisms through which sulfates penetrate the concrete and initiate the degradation process.
Moreover, it is advisable to conduct experimental investigations to separately clarify the effects of sulfate chemical attack, physical attack, and the coupled chemical–physical attack on the expansion and mechanical properties degradation of concrete. Microscopic analysis techniques such as Scanning Electron Microscopy (SEM), X-ray Diffraction (XRD), Mercury Intrusion Porosimetry (MIP), and Thermogravimetric Analysis (TGA) can be employed to observe and analyze the microstructural changes, the pore structures, and the formation of expansive products during the attacking process. Establishing an experimental relationship between expansion and microstructural evolution can help to reveal the degradation mechanism of different sulfate attack modes at macro- and micro-levels. This relationship can help to clarify the crystallization process and expansion damage mechanism of sulfates in concrete pores, and help to understand the contributions of both sulfate chemical and physical attacks.
Furthermore, once the relationship between sulfate attack and the mechanical degradation of concrete is established, it becomes possible to develop a mechanical degradation model specifically tailored to sulfate-attacked concrete. This model can simulate the deterioration process of concrete by incorporating approaches such as smeared crack approaches, cohesive zone approaches, and the moving mesh technique [72,73]. These methods allow for the simulation of crack propagation and the associated mechanical damage caused by sulfate attack.
Finally, conducting a study on the durability failure of sulfate-attacked concrete would be valuable. This study aims to predict the lifespan of concrete structures under sulfate attack. By integrating the mechanical degradation model with other factors, such as environmental conditions, loading conditions, and structural design, researchers can assess the long-term durability and performance of concrete structures affected by sulfate attack. This prediction of the durability lifespan can guide decision-making processes for maintenance, repair, and replacement strategies, ultimately ensuring the longevity and safety of concrete structures in real-world construction scenarios.

6. Conclusions

This paper offers a comprehensive summary of the moisture transport, ion transport, and mechanical damage models applied to concrete under sulfate attack and drying–wetting cycles. It provides an overview of the essential aspects and principles of each model, emphasizing their significance in comprehending the movement of moisture and ions, along with the consequent mechanical damage within the concrete caused by these degradation processes. This paper also critically analyzes the assumptions made in each model, shedding light on their limitations and the potential implications for the accuracy of the predictions. In addition, the paper suggests future research directions to further enhance our understanding of sulfate attacks on concrete. One important area of focus is the exploration of the transport mechanism of sulfate ions under various driving forces. Another is the need to clarify the crystallization process and expansion damage mechanism within concrete pores and to establish an experimental relationship between expansion and microstructural evolution to reveal the degradation mechanism at both the macro- and micro-levels. By addressing these research directions, future studies can contribute to advancing our knowledge of sulfate attack under drying–wetting cycles. This will lead to the development of more accurate and comprehensive models that capture the complex interactions and mechanisms involved in sulfate attack-induced damage. Ultimately, this research will aid in the development of improved mitigation strategies to enhance the durability and performance of concrete structures.

Funding

This research was funded by the National Natural Science Foundations of China (Grant No. 52108234), the Shenzhen Polytechnic University start-up project (Grant No. 6022312014), Shenzhen Polytechnic University Research Fund (Grant No. 6024310037K).

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Elahi, M.M.A.; Shearer, C.R.; Reza, A.N.R.; Saha, A.K.; Khan, M.N.N.; Hossain, M.M.; Sarker, P.K. Improving the sulfate attack resistance of concrete by using supplementary cementitious materials (SCMs): A review. Constr. Build. Mater. 2021, 281, 122628. [Google Scholar] [CrossRef]
  2. Ting, M.Z.Y.; Wong, K.S.; Rahman, M.E.; Meheron, S.J. Deterioration of marine concrete exposed to wetting-drying action. J. Clean Prod. 2021, 278, 123383. [Google Scholar] [CrossRef]
  3. He, R.; Zheng, S.; Gan, V.J.; Wang, Z.; Fang, J.; Shao, Y. Damage mechanism and interfacial transition zone characteristics of concrete under sulfate erosion and Dry-Wet cycles. Constr. Build. Mater. 2020, 255, 119340. [Google Scholar] [CrossRef]
  4. Santhanam, M.; Cohen, M.D.; Olek, J. Mechanism of sulfate attack: A fresh look: Part 2. Proposed mechanisms. Cem. Concr. Res. 2003, 33, 341–346. [Google Scholar] [CrossRef]
  5. Ikumi, T.; Segura, I. Numerical assessment of external sulfate attack in concrete structures. A review. Cem. Concr. Res. 2019, 121, 91–105. [Google Scholar] [CrossRef]
  6. Chen, H.; Huang, H.; Qian, C. Study on the deterioration process of cement-based materials under sulfate attack and drying-wetting cycles. Struct. Concr. 2018, 19, 1225–1234. [Google Scholar] [CrossRef]
  7. Cefis, N.; Comi, C. Chemo-mechanical modelling of the external sulfate attack in concrete. Cem. Concr. Res. 2017, 93, 57–70. [Google Scholar] [CrossRef]
  8. Samson, E.; Marchand, J. Modeling the transport of ions in unsaturated cement-based materials. Comput. Struct. 2007, 85, 1740–1756. [Google Scholar] [CrossRef]
  9. 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]
  10. 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]
  11. Cefis, N.; Comi, C. Degradation of Concrete Structures due to External Sulfate Attack. Key Eng. Mater. 2016, 711, 310–318. [Google Scholar] [CrossRef]
  12. Schmidt-Döhl, F.; Rostásy, F.S. A model for the calculation of combined chemical reactions and transport processes and its application to the corrosion of mineral-building materials Part II. Experimental verification. Cem. Concr. Res. 1999, 29, 1047–1053. [Google Scholar] [CrossRef]
  13. Sarkar, S.; Mahadevan, S.; Meeussen, J.C.L.; Van der Sloot, H.; Kosson, D.S. Numerical simulation of cementitious materials degradation under external sulfate attack. Cem. Concr. Compos. 2010, 32, 241–252. [Google Scholar] [CrossRef]
  14. Sarkar, S.; Mahadevan, S.; Meeussen, J.C.L.; van Der Sloot, H.; 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]
  15. 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]
  16. Zhang, Z.; Zhou, J.; Yang, J.; Zou, Y.; Wang, Z. Understanding of the deterioration characteristic of concrete exposed to external sulfate attack: Insight into mesoscopic pore structures. Constr. Build. Mater. 2020, 260, 119932. [Google Scholar] [CrossRef]
  17. Zhang, Z.; Zhou, J.; Yang, J.; Zou, Y.; Wang, Z. Cracking characteristics and pore development in concrete due to physical attack. Mater. Struct. 2020, 53, 104. [Google Scholar] [CrossRef]
  18. Haufe, J.; Vollpracht, A. Tensile strength of concrete exposed to sulfate attack. Cem. Concr. Res. 2019, 116, 81–88. [Google Scholar] [CrossRef]
  19. Nehdi, M.L.; Suleiman, A.R.; Soliman, A.M. Investigation of concrete exposed to dual sulfate attack. Cem. Concr. Res. 2014, 64, 42–53. [Google Scholar] [CrossRef]
  20. Zhang, Z.; Jin, X.; Luo, W. Long-term behaviors of concrete under low-concentration sulfate attack subjected to natural variation of environmental climate conditions. Cem. Concr. Res. 2019, 116, 217–230. [Google Scholar] [CrossRef]
  21. Yuan, J.; Liu, Y.; Tan, Z.; Zhang, B. Investigating the failure process of concrete under the coupled actions between sulfate attack and drying–wetting cycles by using X-ray CT. Constr. Build. Mater. 2016, 108, 129–138. [Google Scholar] [CrossRef]
  22. Chen, Y.; Gao, J.; Tang, L.; Li, X. Resistance of concrete against combined attack of chloride and sulfate under drying–wetting cycles. Constr. Build. Mater. 2016, 106, 650–658. [Google Scholar] [CrossRef]
  23. Gu, Y.; Martin, R.P.; Metalssi, O.O.; Fen-Chong, T.; Dangla, P. Pore size analyses of cement paste exposed to external sulfate attack and delayed ettringite formation. Cem. Concr. Res. 2019, 123, 105766. [Google Scholar] [CrossRef]
  24. Hossack, A.M.; Thomas, M.D. The effect of temperature on the rate of sulfate attack of Portland cement blended mortars in Na2SO4 solution. Cem. Concr. Res. 2015, 73, 136–142. [Google Scholar] [CrossRef]
  25. 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]
  26. Ikumi, T.; Cavalaro, S.H.; 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]
  27. Kunther, W.; Lothenbach, B.; Scrivener, K.L. On the relevance of volume increase for the length changes of mortar bars in sulfate solutions. Cem. Concr. Res. 2013, 46, 23–29. [Google Scholar] [CrossRef]
  28. Zhu, J.; Al-samawi, M.; Yang, Y.; Al-Shakhdha, N.A. An approach for simulating and evaluating the effect of sulfate attack on the durability of concrete bridges based on a three-dimensional cellular automata model. Eng. Struct. 2023, 291, 116451. [Google Scholar] [CrossRef]
  29. Ran, B.; Li, K.; Fen-Chong, T.; Omikrine-Metalssi, O.; Dangla, P. Spalling rate of concretes subject to combined leaching and external sulfate attack. Cem. Concr. Res. 2022, 162, 106951. [Google Scholar] [CrossRef]
  30. Melara, E.K.; Trentin, P.O.; Pereira, E.; Medeiros-Junior, R.A. Contribution to the service-life modeling of concrete exposed to sulfate attack by the inclusion of electrical resistivity data. Constr. Build. Mater. 2022, 322, 126490. [Google Scholar] [CrossRef]
  31. Qin, S.; Zhang, M.; Zou, D.; Liu, T. A failure thickness prediction model for concrete exposed to external sulfate attack. Constr. Build. Mater. 2024, 416, 135202. [Google Scholar] [CrossRef]
  32. Yu, Y.; Zhang, Y.X. Numerical modelling of mechanical deterioration of cement mortar under external sulfate attack. Constr. Build. Mater. 2018, 158, 490–502. [Google Scholar] [CrossRef]
  33. Yin, G.J.; Zuo, X.B.; Sun, X.H.; 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]
  34. Yin, G.J.; Zuo, X.B.; Tang, Y.J.; Ayinde, O.; Wang, J.L. 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]
  35. Sun, C.; Yuan, L.; Zhai, X.; Qu, F.; Li, Y.; Hou, B. Numerical and experimental study of moisture and chloride transport in unsaturated concrete. Constr. Build. Mater. 2018, 189, 1067–1075. [Google Scholar] [CrossRef]
  36. Whittaker, M.; Black, L. Current knowledge of external sulfate attack. Adv. Cem. Res. 2015, 27, 532–545. [Google Scholar] [CrossRef]
  37. Zhang, M.; Qin, S.; Lyu, H.; Chen, C.; Zou, D.; Zhou, A.; Li, Y.; Liu, T. A transport-chemical-physical-mechanical model for concrete subjected to external sulfate attack and drying-wetting cycles. Eng. Fract. Mech. 2023, 293, 109726. [Google Scholar] [CrossRef]
  38. Zou, D.; Zhang, M.; Qin, S.; Liu, T.; Tong, W.; Zhou, A.; Jivkov, A. Calcium leaching from cement hydrates exposed to sodium sulfate solutions. Constr. Build. Mater. 2022, 351, 128975. [Google Scholar] [CrossRef]
  39. Qin, S.; Zou, D.; Liu, T.; Jivkov, A. A chemo-transport-damage model for concrete under external sulfate attack. Cem. Concr. Res. 2020, 132, 106048. [Google Scholar] [CrossRef]
  40. Zou, D.; Qin, S.; Liu, T.; Jivkov, A. Experimental and numerical study of the effects of solution concentration and temperature on concrete under external sulfate attack. Cem. Concr. Res. 2021, 139, 106284. [Google Scholar] [CrossRef]
  41. Flatt, R.J.; Scherer, G.W. Thermodynamics of crystallization stresses in DEF. Cem. Concr. Res. 2008, 38, 325–336. [Google Scholar] [CrossRef]
  42. Scherer, G.W. Crystallization in pores. Cem. Concr. Res. 1999, 29, 1347–1358. [Google Scholar] [CrossRef]
  43. Scherer, G.W. Stress from crystallization of salt. Cem. Concr. Res. 2004, 34, 1613–1624. [Google Scholar] [CrossRef]
  44. Kunther, W.; Lothenbach, B.; Scrivener, K. Influence of bicarbonate ions on the deterioration of mortar bars in sulfate solutions. Cem. Concr. Res. 2013, 44, 77–86. [Google Scholar] [CrossRef]
  45. 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]
  46. Bary, B.; Leterrier, N.; Deville, E.; Le 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]
  47. Basista, M.; Weglewski, W. Micromechanical modeling of sulphate corrosion in concrete: Influence of ettringite forming reaction. Theor. Appl. Mech. 2008, 35, 29–52. [Google Scholar] [CrossRef]
  48. Pel, L. Moisture transport in porous building materials. Heron J. 1995, 41, 95–105. [Google Scholar]
  49. Zhang, Y. Mechanics of Chloride Ions Transport in Concrete. Ph.D. Thesis, Zhejiang University, Hangzhou, China, 2008. (In Chinese). [Google Scholar]
  50. Li, C. Study on Water and Ionic Transport Processes in Cover Concrete under Drying-wetting Cycles. Ph.D. Thesis, Tsinghua University, Beijing, China, 2009. (In Chinese). [Google Scholar]
  51. Brunauer, S.; Emmett, P.H.; Teller, E. Adsorption of gases in multimolecular layers. J. Am. Chem. Soc. 1938, 60, 309–319. [Google Scholar] [CrossRef]
  52. Chaube, R.; Kishi, T.; Maekawa, K. Modelling of Concrete Performance: Hydration, Microstructure and Mass Transport, 1st ed.; CRC Press: London, UK, 1999. [Google Scholar]
  53. Wu, J. Modelling and Simulation of Chloride Ions Transmission in Concrete. Master’s Thesis, Harbin Institute of Technology, Harbin, China, 2012. (In Chinese). [Google Scholar]
  54. Li, C.; Li, K.; Chen, Z. Numerical analysis of moisture influential depth in concrete during drying-wetting cycles. Tsinghua Sci. Technol. 2008, 13, 696–701. [Google Scholar] [CrossRef]
  55. Guan, B.W.; Wu, J.Y.; Yang, T.; Xu, A.H.; Sheng, Y.P.; Chen, H.X. Developing a model for chloride ions transport in cement concrete under dynamic flexural loading and dry-wet cycles. Math. Probl. Eng. 2017, 2017, 5760512. [Google Scholar] [CrossRef]
  56. Gummerson, R.J.; Hall, C.; Hoff, W.D.; Hawkes, R.; Holland, G.N.; Moore, W.S. Unsaturated water flow within porous materials observed by NMR imaging. Nature 1979, 281, 56–57. [Google Scholar] [CrossRef]
  57. Yin, G.J.; Li, L.B.; Wen, X.D.; Miao, L.; Wang, S.S.; Zuo, X.B. Numerical modelling on moisture and sulfate ion transport in unsaturated concrete slab under dry-wet cycles. J. Build. Eng. 2024, 89, 109296. [Google Scholar] [CrossRef]
  58. Wong, S.F.; Wee, T.H.; Swaddiwudhipong, S.; Lee, S.L. Study of water movement in concrete. Mag. Concr. Res. 2001, 53, 205–220. [Google Scholar] [CrossRef]
  59. Liu, P. Study on the Moisture Transport Law of Concrete under Dry-wet Cycles and Sulfate Attack. Master’s Thesis, Nanjing University of Science and Technology, Nanjing, China, 2013. (In Chinese). [Google Scholar]
  60. Zheng, S.; He, R.; Chen, H.; Wang, Z.; Huang, X.; Liu, S. Three-dimensional reconstruction and sulfate ions transportation of interfacial transition zone in concrete under dry-wet cycles. Constr. Build. Mater. 2021, 291, 123370. [Google Scholar] [CrossRef]
  61. Zhang, J.; Sun, M.; Hou, D.; Li, Z. External sulfate attack to reinforced concrete under drying-wetting cycles and loading condition: Numerical simulation and experimental validation by ultrasonic array method. Constr. Build. Mater. 2017, 139, 365–373. [Google Scholar] [CrossRef]
  62. Li, J.; Xie, F.; Zhao, G.; 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]
  63. Shan, Z.Q.; Yin, G.J.; Wen, X.D.; Miao, L.; Wang, S.S.; Zuo, X.B. Numerical simulation on transport-crystallization-mechanical behavior in concrete structure under external sulfate attack and wetting–drying cycles. Mater. Des. 2024, 241, 112908. [Google Scholar] [CrossRef]
  64. Müllauer, W.; Beddoe, R.E.; Heinz, D. Sulfate attack expansion mechanisms. Cem. Concr. Res. 2013, 52, 208–215. [Google Scholar] [CrossRef]
  65. Lothenbach, B.; Bary, B.; Le Bescop, P.; Schmidt, T.; Leterrier, N. Sulfate ingress in Portland cement. Cem. Concr. Res. 2010, 40, 1211–1225. [Google Scholar] [CrossRef]
  66. Scherer, G.W. Factors affecting crystallization pressure. In Proceedings of the International RILEM Workshop on International Sulfate Attack and Delayed Ettringite Formation, Villars, Switzerland, 4–6 September 2002. [Google Scholar]
  67. Scrivener, K.L.; Taylor, H.F.W. Delayed ettringite formation: A microstructural and microanalytical study. Adv. Cem. Res. 1993, 5, 139–146. [Google Scholar] [CrossRef]
  68. Chen, H. Strengths Prediction of Concrete in Marine Environments Based on Dry-Wet Cycles and Salt Crystallization Translated into Stress. Ph.D. Thesis, Southeast University, Nanjing, China, 2018. (In Chinese). [Google Scholar]
  69. Yang, F. Damage Modeling of Concrete subjected to Coupled Action of Crystallization of Sulfate and Drying-Wetting Cycles. Master’s Thesis, Southeast University, Nanjing, China, 2012. (In Chinese). [Google Scholar]
  70. Ren, J.; Lai, Y.; Bai, R.; Qin, Y. The damage mechanism and failure prediction of concrete under wetting–drying cycles with sodium sulfate solution. Constr. Build. Mater. 2020, 264, 120525. [Google Scholar] [CrossRef]
  71. Yin, G.J.; Zuo, X.B.; Sun, X.H.; Tang, Y.J. Numerical investigation of the external sulfate attack induced expansion response of cement paste by using crystallization pressure. Model. Simul. Mater. Sci. Eng. 2019, 27, 025006. [Google Scholar] [CrossRef]
  72. De Maio, U.; Greco, F.; Lonetti, P.; Pranno, A. A combined ALE-cohesive fracture approach for the arbitrary crack growth analysis. Eng. Fract. Mech. 2024, 301, 109996. [Google Scholar] [CrossRef]
  73. Rimkus, A.; Cervenka, V.; Gribniak, V.; Cervenka, J. Uncertainty of the smeared crack model applied to RC beams. Eng. Fract. Mech. 2020, 233, 107088. [Google Scholar] [CrossRef]
Figure 1. The relationship between saturation and relative humidity in isothermal adsorption.
Figure 1. The relationship between saturation and relative humidity in isothermal adsorption.
Materials 17 03334 g001
Figure 2. Schematic diagram of modified critical pore radius (rs is the ideal critical pore radius; ta is thickness of the adsorbed water film).
Figure 2. Schematic diagram of modified critical pore radius (rs is the ideal critical pore radius; ta is thickness of the adsorbed water film).
Materials 17 03334 g002
Table 1. Empirical models for moisture transport coefficient.
Table 1. Empirical models for moisture transport coefficient.
StageFunction FormExpressionReference
WettingExponential D m = D m 0 e n s Li [54], Guan [55], Gummerson [56]
Power function D m = D m 0 s n Yin [57]
DryingS-shaped D m = D m 0 α 1 + 1 α 1 1 + ( 1 s ) / 1 s c N Li [54], Guan [55], Wong [58]
Constant D m = D 1 Li [50]
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

Qin, S.; Chen, C.; Zhang, M. Modeling of Concrete Deterioration under External Sulfate Attack and Drying–Wetting Cycles: A Review. Materials 2024, 17, 3334. https://doi.org/10.3390/ma17133334

AMA Style

Qin S, Chen C, Zhang M. Modeling of Concrete Deterioration under External Sulfate Attack and Drying–Wetting Cycles: A Review. Materials. 2024; 17(13):3334. https://doi.org/10.3390/ma17133334

Chicago/Turabian Style

Qin, Shanshan, Chuyu Chen, and Ming Zhang. 2024. "Modeling of Concrete Deterioration under External Sulfate Attack and Drying–Wetting Cycles: A Review" Materials 17, no. 13: 3334. https://doi.org/10.3390/ma17133334

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