Next Article in Journal
High-Modulus Laminated SiC/AZ91 Material with Adjustable Microstructure and Mechanical Properties Based on the Adjustment of the Densities of the Ceramic Layers
Previous Article in Journal
A Circular Approach to Finished Tanned Leather: Regeneration by Cryogenic Technology
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Experimental–Computational Investigation of the Elastic Modulus of Mortar under Sulfate Attack

Key Laboratory of Roads and Railway Engineering Safety Control, Ministry of Education, Shijiazhuang Tiedao University, Shijiazhuang 050013, China
*
Author to whom correspondence should be addressed.
Materials 2023, 16(18), 6167; https://doi.org/10.3390/ma16186167
Submission received: 28 June 2023 / Revised: 24 August 2023 / Accepted: 8 September 2023 / Published: 12 September 2023

Abstract

:
External sulfate attack is an important factor causing a decrease in the mechanical properties of cement-based materials. In this paper, a computational prediction model of elastic modulus, considering the characteristics of sulfate corrosion from outside to inside and the influence of the interface transition zone (ITZ), was established to predict the elastic modulus of mortar under the external sulfate attack. Firstly, the backscattered electron (BSE) images of mortar and the algorithm of image threshold segmenting were used to determine a reasonable thickness of corroded ITZ. Secondly, the nanoindentation test was adopted to acquire the microscopic elastic parameters of phases (sand, cement, and ITZ) in corroded mortar. Moreover, the mortar mix proportion and Lu and Torquato’s model were adopted to calculate the volume fractions of phases. Finally, a computational prediction model of elastic modulus of mortar under sulfate attack was proposed with homogenization methods. The results indicate that the thickness of corroded ITZ is 20 mm, and the error values of elastic modulus between the theoretical prediction results and the experimental results are within 8%, indicating that the macroscopic elastic modulus of corroded mortar can be precisely predicted by the computational prediction model of elastic modulus.

1. Introduction

Elastic modulus has an important influence on the stress distribution and deformation of concrete structures, and, as such, it is a key parameter in the design of such structures [1]. Sulfate corrosion in environments, such as oceans, salt lakes, and groundwater, is an important factor causing durability issues and structural failure of cement materials [2]. The durability of cement-based material under sulfate attack has attracted widespread attention and research from scholars [3]. Mortar is between cement paste and concrete, and the elastic modulus of mortar is an important parameter connecting the elastic modulus of concrete and cement. Therefore, the study of the mortar’s elastic modulus under sulfate attack has a great significance.
As is known, the macroscopic mechanical behavior of material is largely dependent on the compositions, contents, and mechanical properties of phases at the nano/micro-scale [4]. The upscaling technique was accurately adopted to calculate the elastic parameters of material between the microscopic and macroscopic properties over the last several decades [5,6,7,8,9]. A multi-scale mechanical model was established with the Mori–Tanaka (M-T) method and self-consistent (SC) method to precisely predict the intrinsic elastic properties of concrete [10]. According to the results of the nanoindentation test, Li et al. divided the brine-eroded cement paste into two scales: C-S-H and cement paste [11]. The M-T method and SC method were used to predict the elastic modulus of C-S-H gel and cement paste after brine corrosion. Based on the cement hydration model, Huang et al. and Liang et al. used the multi-scale homogenization method to predict the elastic modulus of concrete from the C-S-H scale [1,7]. However, there are few studies on the elastic modulus of sulfate corrosive mortar with multi-scale homogenization methods.
The interface transition zone (ITZ) is between the matrix and the inclusions. The ITZ is the weakest phase in concrete material, with the characteristics of low elastic modulus, low strength, high porosity, and high permeability [12,13,14]. The macroscopic mechanical behavior of cement-based materials is largely dependent on the component content and mechanical properties of phases. Considering the influence of ITZ, Liang et al. divided the concrete into four scales and predicted the concrete’s elastic modulus with the homogenization methods [1]. The ITZ’s effective thickness is a key parameter for calculating the volume fraction of phases, and the macroscopic mechanical performances of concrete [15]. However, the thickness of the ITZ was not considered in Liang’s model. In order to evaluate the effective thickness of the ITZ, various microstructure of cement-based materials have been observed and researched, and the thickness of the ITZ from 15 μm to 50 μm is usually accepted [15]. Lutz et al. considered that the ITZ around the concrete aggregate was non-uniform, and it increased slowly with the increase in the distance to the aggregate’s center [16]. It is considered that the porosity of the ITZ is closer to the porosity of the matrix the farther away from the aggregate, and then the elastic modulus of concrete was studied by assuming the elastic modulus of the ITZ. Mondal et al. studied the elastic properties of the ITZ with the nanoindentation technology and found that the change in the elastic modulus of the indentation point in ITZ is regular, that is, the elastic modulus increases gradually as the indentation point approaches the center of the fine aggregate [17,18]. The average value of elastic modulus of ITZ was 18 GPa, about 85% of the value of the matrix. Chen et al. adopted the nearest surface distribution function of Lu and Torquato’s model to give a quantitative calculation formula for calculating the volume fraction of the ITZ between aggregate and matrix in cement-based composites, then used the random point sampling method to calculate the volume fraction of ITZ for verifying the effectiveness of the quantitative calculation formula [19,20]. Although the existing studies have calculated the influence of the ITZ on the overall performance of cement-based materials under different thicknesses with the help of formulas, they lack the experimental verification of the elastic performance of the ITZ and do not involve the influence law of the mortar ITZ on the overall elastic energy of mortar under a sulfate corrosion environment.
In this paper, a reasonable thickness of ITZ in corroded mortar was obtained using image processing software with the SEM-BSE image of mortar. Then, the elastic parameters of phases (sand, cement, and ITZ) in mortar were determined based on the nanoindentation test and theoretical calculation. In addition, the volume fraction of each phase in the mortar was determined by calculating the mortar mix ratio and applying Lu and Torquato’s model. Finally, the homogenization method was used to establish the elastic modulus prediction model of mortar under sulfate attack. The theoretical calculation results were compared with the experimental results to verify the effectiveness of the elastic modulus prediction model. The research route is shown in Figure 1.

2. Materials and Methods

2.1. Materials

P.I 52.5 Portland cement was selected in this study. The mineral composition of the cement is shown in Table 1. The natural river sand was selected as the fine aggregate, with a minimum size of 0.035 mm, a maximum size of 4.75 mm, a fineness modulus of 2.45, and a mud content (by mass) < 1.5%.
Sodium sulfate (Na2SO4) was selected to prepare the sulfate solution, and the molar concentration of sulfate ion in sulfate solution was 352 mol/m3 (mass concentration was 5%).

2.2. Mix Design and Samples Preparation

In this study, the mortar mixing ratio design is shown in Table 2.
The mixed mortar was poured into 20 mm × 20 mm × 40 mm molds, cured in a laboratory with a constant temperature (20 ± 5 °C) and humidity (65%) for 24 h, and then demolded. Subsequently, the mortar samples were cured under standard curing conditions (temperature of 20 °C and relative humidity of 95%) for 28 d. Finally, mortar samples were soaked in Na2SO4 solution at 20 ± 2 °C. In order to ensure a constant concentration of sulfate ions, the Na2SO4 solution was replaced every 3 d.
The uniaxial compression test, the nanoindentation test, and the SEM-BSE test were conducted, and mortar samples were selected with a processing period of 88 d (sulfate corrosion time of 60 d). For the uniaxial compression test, each mix proportion selected five mortar specimens to obtain the constitutive relation curves. For the nanoindentation test, specimens were selected from mortar specimens with dimensions of 5 mm × 5 mm × 10 mm. Owing to the high requirement of flatness for the nanoindentation test, nanoindentation specimens were pretreated [21,22]. Firstly, in order to prevent the further hydration of the specimen, the specimen was placed in an alcohol phenolphthalein solution, and then the specimen was taken out and placed in the mold, which was carried out with epoxy resin. Then, in the coarse polishing process, the specimens were polished on the grinding and polishing machine using abrasive paper of 180 meshes, 240 meshes, 400 meshes, 600 meshes, 800 meshes, and 1200 meshes, respectively. Then, the specimens were polished using canvas and silk cloth with a suspension of 0.25 μm diamond. It should be noted that the upper and lower planes of specimens must be parallel during the grinding process. Furthermore, atomic force microscopy (AFM) was used to measure the surface roughness of the polished specimens. The polishing process ended when the surface roughness was less than 100 nm. Finally, the specimens were washed in an ultrasonic bath with anhydrous ethanol to remove the suspensions and debris on the specimen surface, and then the specimens were dried naturally. The specimens for the nanoindentation test with a thickness of 4 mm are shown in Figure 2. The SEM-BSE test specimens were selected from the specimens after the nanoindentation test with additional treatment. After the nanoindentation test, specimens were sprayed with gold for 120 s. Then, a gold sprayed film with a thickness of 10 nm was formed on the surface of the specimen to achieve a better observation effect.

2.3. Experimental Methods

2.3.1. Uniaxial Compression Test

The uniaxial compression experiments were conducted by a servo-controlled setup (Zwick/Roell Z050) with a loading speed of 0.5 mm/min. Two linear variable differential transformers (LVDT1 and LVDT2) were symmetrically arranged 10 mm away from two sides of the specimen to measure the deformation under loading. Figure 3 shows the uniaxial compression test of the mortar specimen. In order to increase the accuracy of experimental results, the maximum and minimum elastic modulus obtained from uniaxial compression experiments were eliminated in each group with the other three left. Then, the three remaining test dates of each group were used to calculate the macroscopic static compressive elastic modulus through the method of taking the average value.

2.3.2. Nanoindentation Test

Instrumented nanoindentation (G200), produced by Agilent Company, Santa Clara, CA, USA, was adopted to acquire the microstructural mechanical properties of corroded mortar specimens. The elastic modulus and hardness of the specimen varying with the nanoindentation depth were obtained by a continuous stiffness method (CSM). The schematic of the nanoindentation test for the microscopic elastic modulus of the corroded mortar is shown in Figure 4.
During the loading process, elastic deformation firstly appeared in the specimen. As for the plastic deformation of the specimen, the loading curve was non-linear with the increase in loading. During unloading, only the elastic properties of the indentation point were restored. Figure 5 shows the load(P)–displacement(H) curve during the typical loading and unloading of the indenter. The curve was used to analysis the hardness and elastic modulus of indentation points. The microscopic elastic modulus E can be calculated by the Oliver–Pharr principle [23], as shown in the following Equation (1):
E = 1 ν 2 2 β S A π 1 ν i 2 E i
where E is the elastic modulus of tested material, ν is the Poisson’s ratio of tested material, β represents a correction factor (due to the Berkovich indenter, β = 1.034), S is the contact stiffness ( S = d P / d h h = h max , P is the maximum loading), A is the corresponding contact area at the maximum load, ν i , and E i denotes the parameters of indenter ( ν i = 0.07, E i = 1141 GPa).
The nanoindentation test was carried out on ITZ, cement, and sand of the mortar specimens. An 8 × 8 lattices of 64 points on the corroded mortar specimen was measured to obtain the microscopic elastic modulus.
In order to obtain a relatively stable elastic modulus value, the indentation depth needs to reach the micrometer scale [1]. Moreover, it can be more clearly observed by the scanning electron microscope (SEM) with the increase in the indentation depth. Then, the surface area of the indentation point will increase and the possibility of interaction between two adjacent points will increase. The ITZ’s thickness in concrete was 15 µm to 50 µm, and the ITZ’s thickness in mortar was smaller than that in concrete [24,25,26]. For the sake of the accurate elastic modulus of ITZ, the closer the distance between every two adjacent nanoindentation points, the better the test result, at least theoretically. However, there should be a sufficient distance between two adjacent indentation points to avoid mutual influence. Therefore, the depth of nanoindentation points and the spacing of nanoindentation points were set as 2000 nm and 15 µm, respectively.

2.3.3. SEM-BSE Test

The micromorphology of nanoindentation points and corroded mortar was observed by backscattered electron (SEM-BSE) images. The SEM-BSE images were obtained by Quanta 250FEG, produced by FEI Co., Hillsboro, OR, USA. The SEM-BSE images were acquired with a resolution of 3.5 nm under a high vacuum mode, and the distance between the sample and scanning electron microscope lens was 10 mm. Figure 6 shows the micromorphology of a lattice map of nanoindentation points and a single nanoindentation point of cement mortar (M1) observed by the SEM-BSE test. The region in the red box in Figure 6a is the lattice map of nanoindentation points, and the region in the blue box in Figure 6b is a single nanoindentation point.

3. Experimental Results and Discussion

3.1. Macroscopic Elastic Modulus of Corroded Mortar

The stress–strain curve of the eroded mortar sample was obtained via a uniaxial compression test, so that the macroscopic elastic modulus (E) of the sample could be obtained. The macroscopic elastic moduli of corroded mortars with types of M1 and M2 were 29.88 GPa and 33.33 GPa, respectively.

3.2. Thickness of ITZ in Corroded Mortar

The BSE images of corroded mortar were processed by the algorithm of image threshold segmenting. Each of cubic volume elements (voxels) in the image was assigned a unique grayscale value, then the data were organized as a collection of cubic volume elements (voxels).
Next, the data of distinct material phases were segmented into several distinct categories [7]. The material phases, such as the pores and matrix in the mortar, were distinguished and quantified by the threshold intensity segmentation method [27].
Figure 7 shows the SEM-BSE image processing process of ITZ in mortar (M1) at 88 d. Figure 7a shows the SEM-BSE image of corroded mortar. The yellow line in the figure is the edge of sand. After selecting the region of 60 µm at the edge of sand and dividing it into 12 regions successively, the thickness of each region was 5 µm. Then, median filtering and image binarization technology were used to process the BSE image to distinguish the pores and matrix [28]. According to the previous research results, the gray segmentation threshold between matrix and pores was 72.1 [29]. Figure 7b shows the binarization results of BSE images. In each region, blue parts are micropore structures, and the left part is the matrix. According to the pixel number of the matrixes and micropore structures in each region, the porosity of regions at different distances from the sand edge can be calculated by Equation (2), as follows:
ϕ = n p n × 100 %
where ϕ is the porosity of the image processing region, n p is the pixel number of micropore structures, and n is the total number of pixels in the processing region.
Firstly, the porosity of each region was calculated, and then 10 regions with the same distance to the edge of the sand are selected, and the average porosity of these 10 regions was calculated as the porosity of the region. According to Equation (2), the variation in porosity with the increase in the distance to the edge of the fine aggregate can be obtained, as shown in Figure 8. The porosity first decreases rapidly and then gradually tends towards a stable value with the increase in the distance from the sand edge. According to the previous study of Lutz, the porosity of cement in mortar is a stable value [16]. The porosity of ITZ in mortar gradually decreases with the increase in the distance from the surface of the sand, and finally tends to the value of the cement porosity [16]. Therefore, the ITZ thickness of corroded mortar specimens is about 20 µm.

3.3. Nanoindentation Results of Corroded Mortar

A lattice of 8 × 8 nanoindentation points, including the sand, ITZ, and paste, was carried out around the fine aggregate in corroded mortar. Figure 9 shows the SEM-BSE image and elastic modulus results of nanoindentation points of mortar specimen (M1) at 88 d. The microscopic elastic modulus of each point is marked on Figure 9b, corresponding to the locations. It is seen that in the gray regions (sands) in Figure 9b, the minimum value of the micro-elastic modulus is about 55 GPa, the maximum value is about 70 Gpa, and the average value is 63.1 Gpa. According to the analysis in Section 3.2, the area within 20 µm at the edge of sands is ITZ. It is seen that the microscopic elastic modulus of nanoindentation points in yellow regions (ITZ) range from 11 Gpa to 28 Gpa with an average value of 14.8 Gpa. The microscopic elastic modulus of cement paste ranges from 12 Gpa to 28 Gpa, with an average value of 21.9 Gpa.
Figure 10 shows the SEM-BSE image and the elastic modulus results of nanoindentation points of mortar specimen (M2) at 88 d. The gray regions in Figure 10b represent sands, and the elastic modulus of sand is close to that of sand in mortar (M1). The minimum value of the micro-elastic modulus of sand is about 55 Gpa, the maximum value is about 70 Gpa, and the average value is 63.4 Gpa. It can be found that the microscopic elastic modulus of nanoindentation points, within 20 µm at the edge of sands on yellow regions (ITZ) in Figure 10b, range from 13 Gpa to 28 Gpa, with an average value of 18.4 Gpa. The microscopic elastic modulus of cement paste ranges from 10 Gpa to 34 Gpa, and the average value is 23.0 Gpa.

4. Evaluation of the Elastic Modulus of Corroded Mortar by Homogenization Method

As known, the macro-elastic modulus of mortar can be measured by the static compressive test, and the macro-mechanical properties are mainly dependent on micromechanical properties and micro-components. The homogenization methods can be used to calculate elastic modulus of mortar from multi-scale for studying the influence of sulfate corrosion on the mortar.

4.1. Homogenization Method

The corroded mortar can be divided into two levels according to the multi-scale method, as shown in Figure 11.
On level I, the cement paste matrix, with a characteristic length range from 10−6 m to 10−4 m, is composed of unhydrated cement particles (C3S, C2S, C3A, and C4AF), hydrated products (LD C-S-H, HD C-S-H, CH, C4AH13, and C3 (A,F)H6), sulfate corrosion products (Aft), and pores, etc. The elastic modulus of cement paste can be obtained by the micromechanical model and homogenization method from the micro-elastic modulus of each phase in cement paste. As is known, when the reference medium in composites was hard to determine, the self-consistent (SC) method can be selected as an effective homogenization method [30,31,32]. In the SC method, the homogenized medium is considered as the reference medium. In this study, cement paste was regarded as a material containing 10 different phases, and the SC method was adopted as the homogenization method. The bulk modulus and shear modulus of the cement paste matrix can be calculated according to Equations (3) and (4), as follows [31,32]:
k S C hom = k 0 + r = 1 N c r ( k r k 0 ) ( 3 k S C hom + 4 μ S C hom ) 3 k r + 4 μ S C hom
μ S C hom = μ 0 + r = 1 N 5 c r μ S C hom ( μ r μ 0 ) ( 3 k S C hom + 4 μ S C hom ) 3 k S C hom ( 3 μ S C hom + 2 μ r ) + 4 μ S C hom ( 2 μ S C hom + 3 μ r )
where k S C hom and μ S C hom represent the homogenized bulk modulus and shear modulus, respectively.
On level II, two kinds of multi-level homogenization methods were considered, as shown in Figure 12, to predict the macro-effective elastic parameters of mortar under sulfate attack. Mortar was treated as a composite material consisting of sand, ITZ, and cement paste, as shown in Figure 12a. For the first step, the near-field ITZ and far-field cement paste were considered equivalent by the SC method to form an equivalent matrix, as shown in Figure 12b. For the second step, the mortar was composed of the equivalent matrix and sand, and the macro-effective elastic modulus of mortar can be estimated by the two-phase sphere model, as rendered in Figure 12c.
Using the elastic method and Eshelby equivalent medium method [32,33], Christensen [34,35] proposed the following formula for solving the effective bulk modulus in the two-phase sphere model:
k ¯ s = k E + c s ( k s k E ) ( 3 k E + 4 μ E ) 3 k E + 4 μ E + 3 ( 1 c s ) ( k s k E )
where k ¯ s is the bulk modulus of equivalent particles, k s is the bulk modulus of sand, k E is the bulk modulus of equivalent matrix, μ E is the shear modulus of equivalent matrix, and c s is the volume fraction of sand, which can be obtained from the following Equation (6):
c s = ( a b ) 3 = f s f s + f E
where f s and f E are the volume fractions of sand and equivalent matrix, respectively.
The effective shear modulus of equivalent particles can be obtained from the following Equations (7)–(14) [36]:
A ( μ ¯ s μ E ) 2 + 2 B ( μ ¯ s μ E ) + C = 0
where
A = 8 ( μ s μ E 1 ) ( 4 5 ν E ) η 1 c 10 / 3 2 [ 63 ( μ s μ E 1 ) η 2 + 2 η 1 η 3 ] 7 / 3 + 252 ( μ s μ E 1 ) η 2 c 5 / 3 50 ( μ s μ E 1 ) ( 7 12 ν E + 8 ν E 2 ) η 2 c + 4 ( 7 10 ν E ) η 2 η 3
B = 2 ( μ s μ E 1 ) ( 4 5 ν E ) η 1 c 10 / 3 2 [ 63 ( μ s μ E 1 ) η 2 + 2 η 1 η 3 ] 7 / 3 252 ( μ s μ E 1 ) η 2 c 5 / 3 + 75 ( μ s μ E 1 ) ( 3 ν E ) ν E η 2 c + 3 ( 15 ν E 7 ) η 2 η 3
C = 4 ( μ s μ E 1 ) ( 5 ν E 7 ) η 1 c 10 / 3 2 [ 63 ( μ s μ E 1 ) η 2 + 2 η 1 η 3 ] 7 / 3 + 252 ( μ s μ E 1 ) η 2 c 5 / 3 + 25 ( μ s μ E 1 ) ( ν E 2 7 ) η 2 c 7 ( 7 + 5 ν E ) η 2 η 3
η 1 = ( μ s μ E 1 ) ( 49 50 ν s ν E ) + 35 ( μ s μ E ) ( ν s 2 ν E ) + 35 ( 2 ν s ν E )
η 2 = 5 ( μ s μ E 8 ) + 7 ( μ s μ E + 4 )
η 3 = ( μ s μ E ) ( 8 10 ν E ) + ( 7 5 ν E )
μ ¯ s = B ± B 2 A C A μ E
where A, B, C, η 1 , η 2 , and η 3 are the coefficients of the shear modulus solving process, μ ¯ s is the shear modulus of the equivalent particle, and μ E is the shear modulus of the equivalent matrix.
According to Equations (5)–(14), the bulk modulus k hom and shear modulus μ hom of mortar can be obtained from the micro-elastic parameters of sand and equivalent matrix. Then, the macro-elastic modulus of mortar obtained by homogenization methods can be calculated by the following Equation (15):
E hom = 9 k hom μ hom 3 k hom + μ hom
where E hom represents the elastic modulus of mortar, and k hom and μ hom are the bulk modulus and shear modulus of mortar, respectively.

4.2. Mechanical Parameters of Phases

Under the corrosion of an external sulfate environment, sulfate ions enter the mortar through micropore structures and react with cement hydration products to produce expansive products, leading to the change in the mechanical properties of mortar [37,38,39,40].
It is generally believed that sulfate ions will not react with sand in mortar [41]. Therefore, it is assumed that the mechanical properties of sand do not change. It can be seen from the nanoindentation test results in Section 3.3 that the elastic modulus of sand in corroded mortar with different types has little difference. Then, the average elastic modulus of sand with the value of 63.25 Gpa was selected for calculating the macro-effective elastic parameters of the mortar. Furthermore, Poisson’s ratio of sand was 0.14 [41]. The shear modulus and bulk modulus of the sand can be calculated by Equations (16) and (17), and the values were 27.74 Gpa and 29.28 Gpa, respectively. Equations (16) and (17) are as follows:
μ = E 2 ( 1 + ν )
k = E 3 ( 1 2 ν )
In a sulfate environment, sulfate ions entering into mortar specimens and cause the corrosion of cement-based materials as part of a gradually decreasing process from outside to inside. Therefore, the corrosion degree of mortar was uneven inside and outside. Considering the influence of temporal and spatial distribution for the process of sulfate attack on mortar, the author of this paper coupled the effects of cement hydration and sulfate corrosion and established an X-CT-hydration–deterioration model [42]. On level I, the corroded cement paste matrix was regarded as a composite material composed of 10 phases on the micro-scale. Based on the X-CT-hydration–deterioration model, the compositions and volume fractions of microphases in cement-based material under sulfate attack were quantitatively predicted [42]. The elastic parameters of microphases in cement were obtained through a large number of nanoindentation experiments [29,43,44,45,46,47]. Then, through the SC method, the elastic parameters of cement paste were calculated with volume fractions and the elastic parameters of microphases [29]. The theoretical calculation results of elastic parameters of cement paste at 88 d are shown in Table 3 [29]. The theoretical calculation results were the output parameters of the elastic parameters of the cement paste on Level I, and also the input parameters of the elastic parameters of the cement paste on Level II.
According to the results of the nanoindentation tests in Section 3.3, the ratio of microscopic elastic modulus of ITZ to that of cement in the same corrosion layer of corroded mortar (M1) was about 0.7, and the value of that with type M2 was 0.8. Mondal et al. found that the elastic modulus ratio of ITZ and matrix was stable at a constant value [17,18]. Then, it was assumed that the microscopic elastic modulus ratio of ITZ and cement in corroded mortar (M1) remained at a constant value of 0.7, and the value of that with type M2 was 0.8. Therefore, according to the macro-elastic modulus of cement paste in Table 3, the macro-elastic modulus of ITZ in the corroded mortar (M1) was 15.2 Gpa, and the macro-elastic modulus of ITZ in the corroded mortar (M2) was 18.9 Gpa. Furthermore, Poisson’s ratio of ITZ was 0.24. Then, the bulk modulus and the shear modulus of ITZ in the corroded mortar (M1) were 9.74 Gpa and 6.13 Gpa, respectively. In the corrosion mortar (M2), the bulk modulus of ITZ is 12.12 Gpa, and the shear modulus is 7.62 Gpa.

4.3. Calculation of Volume Fractions of Microphases

According to the mix ratio of mortar, the volume fraction of sand in corroded mortar can be obtained by the following Equation (18):
f s = V s V w 0 + V c 0 + V s = m s / ρ s m w / ρ w + m c / ρ c + m s / ρ s
Where f s is the volume fraction of sand in the mortar, V s is the volume of sand in the mortar, V w 0 and V c 0 are the volume of initial water and cement in the mortar, respectively, m s , m w , and m c are the mass content of sand, water, and cement, respectively, ρ s , ρ w and ρ c the density of sand, water and cement, respectively.
The volume fractions of cement paste matrix (not including ITZ) and ITZ in the mortar were calculated by Lu and Torquato’s model, as shown in Figure 13 [48,49]. The model assumed the following. Firstly, fine aggregates were circular particles with different scales, randomly distributed in matrix. Secondly, the ITZ was a shell wrapping around the circular fine aggregate with a thickness of t, and the overlapping of the ITZ between adjacent aggregates was considered. Thirdly, the remaining part in the mortar was the cement paste matrix.
Based on the above assumptions, the volume fraction of the cement paste matrix was ev(t). If the thickness of ITZ is infinitely large, t → ∞, then the area of cement paste matrix tends to zero, ev(t) → 0. Therefore, the volume fraction of ITZ in corroded mortar can be expressed as follows:
f I T Z = 1 f s e v t
where f I T Z and f s are the volume fraction of the ITZ and the sand, respectively.
When the thickness t of ITZ and the particle size distribution of sand are known, ev(t) can be obtained, as follows:
e v t = ( 1 f s ) exp 2 f s S a 0 t D N ¯ 3 + a 1 t D N ¯ 2 + a 2 t D N ¯ , t 0
a 0 = 4 B 1 f s 1 f s + 3 f s S + 4 A f s 2 S 2 / 1 f s 3
a 1 = 6 B 1 f s + 9 f s S / 1 f s 2
a 2 = 3 / 1 f s
B = D N ¯ 2 / D N 2 ¯
S = D N 2 ¯ × D N ¯ / D N 3 ¯
where t is the thickness of ITZ in corroded mortar, A is a constant value (when the sand aggregate is assumed as circular, A = 0), D N ¯ is the average diameter of fine aggregates, D N 2 ¯ is the second order origin moment of average diameter of fine aggregates, and D N 3 ¯ is the third-order origin moment of average diameter of fine aggregates. Equation (26) is as follows:
f c = 1 f s f I T Z
Based on Lu and Torquato’s model and the mix proportion of mortar, the volume fractions of phases in the corroded mortar can be calculated as follows: (1) Based on the mix ratio of the mortar and the densities of sand ( ρ s = 2.6 g/cm3), cement ( ρ c = 3.10 g/cm3), and water ( ρ w = 1 g/cm3), the volume fraction of sand aggregate was calculated by Equation (18). (2) Substituting the particle size distribution of sand and the thickness of the ITZ (20 µm) into Equation (19), the volume fraction of ITZ was obtained by Lu and Torquato’s model. (3) Finally, the volume fraction of cement paste matrix (not including ITZ) was calculated by Equation (26). The theoretical calculation results of volume fractions the of phases in corroded mortar are shown in Table 4.

5. Verification of Elastic Modulus Prediction Model of Mortar under Sulfate Corrosion

5.1. Theoretical Calculation of Macroscopic Elastic Modulus of Corroded Mortar

Based on the mechanical parameters of the phases in Section 4.2 and the volume fractions of the phases in Section 4.3, the theoretical calculation of the macro-elastic modulus of corroded mortar were obtained by the two-scale homogenization method in Section 4.1. The prediction results of the elastic modulus of corroded mortar are listed in Table 5 and Table 6, respectively.

5.2. Comparison between Experimental Results and Theoretical Prediction Results of the Macro-Elastic Modulus of Corroded Mortar

The error value between the theoretical and experimental values of the macro-elastic modulus of corroded mortar can be calculated by the following Equation (27):
δ = E t ¯ E c E c × 100 %
where δ is the error value between the theoretical value and the experimental value of the elastic modulus, E t ¯ is the average value of experimental elastic modulus, and Ec is the theoretical elastic modulus. The error values of mortar (M1 and M2) were 7.7% and 1.9%, respectively, and the error value of the elastic modulus predicted by the theoretical model is less than 8% compared with the actual elastic modulus obtained by the test. The data show that the theoretical calculation model proposed in this paper can well predict the elastic modulus of corroded mortar.

6. Conclusions

In this paper, the elastic parameters and the volume fractions of the phases (sand, cement, and ITZ) in mortar were obtained by tests and theoretical calculation. Then, a computational prediction model of the elastic modulus was established with homogenization methods to predict the elastic modulus of mortar under external sulfate attack. The conclusions are shown as follows:
(1)
The BSE images of corroded mortar and the algorithm of image threshold segmenting can be used to acquire the thickness of ITZ, and the thickness of ITZ in corroded mortar at 88 d is 20 µm.
(2)
Combined with the results of nanoindentation test and the characteristics of corroded mortar, the macro-elastic modulus of ITZ in the corroded mortars can be accurately and effectively evaluated. The values of ITZ in corroded mortars (M1 and M2) are 15.2 GPa and 18.9 GPa, respectively.
(3)
Based on the mortar mix proportion and the thickness of ITZ, the volume fractions of sand, cement paste, and ITZ in corroded mortar can be well evaluated by Lu and Torquato’s model.
(4)
The computational elastic modulus prediction model, considering the influence of ITZ, can well predict the elastic modulus of mortar under sulfate attack.

Author Contributions

Conceptualization, Z.G.; methodology, Z.G. and H.Z.; validation, Y.L. and Y.W.; formal analysis, Z.G.; investigation, Z.G. and H.Z.; resources, Y.W.; data curation, Z.G. and Y.G.; writing—original draft Z.G. and Y.G.; writing—review and editing, Z.G. and Y.W.; visualization, Y.L., P.S. and Y.W.; supervision, Z.G. and P.S.; project administration, Y.G. and L.W.; funding acquisition, L.W. and Y.W. All authors have read and agreed to the published version of the manuscript.

Funding

This study was funded by the National Natural Science Foundation of China (No. 52308272, 52378171), the Natural Science Foundation of Hebei Province (No. E2022210005, E2021210088, E2020210017), and the Shijiazhuang Railway University Youth Research and Innovation Project (Natural) (ZQK202302).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The authors gratefully acknowledge the support received from the Key Laboratory of Roads and Railway Engineering Safety Control (Shijiazhuang Tiedao University) and the Key Laboratory of Urban Security and Disaster Engineering of the Ministry of Education (Beijing University of Technology).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Liang, S.; Wei, Y.; Wu, Z. Multiscale modeling elastic properties of cement-based materials considering imperfect interface effect. Constr. Build. Mater. 2017, 154, 567–579. [Google Scholar] [CrossRef]
  2. 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]
  3. Oliver, W.C.; Pharr, G.M. Measurement of hardness and elastic modulus by instrumented indentation: Advances in under-standing and refinements to methodology. J. Mater. Res. 2004, 19, 3–20. [Google Scholar] [CrossRef]
  4. Hu, C.; Li, Z. Micromechanical investigation of Portland cement paste. Constr. Build. Mater. 2014, 71, 44–52. [Google Scholar] [CrossRef]
  5. Buljak, V.; Cocchetti, G.; Cornaggia, A.; Garbowski, T.; Maier, G.; Novati, G. Materials mechanical characterizations and structural diagnoses by inverse analyses 20. In Handbook of Damage Mechanics; Springer: New York, NY, USA, 2015; pp. 619–642. [Google Scholar]
  6. Huang, J.; Krabbenhoft, K.; Lyamin, A.V. Statistical homogenization of elastic properties of cement paste based on X-ray microtomography images. Int. J. Solids Struct. 2013, 50, 699–709. [Google Scholar] [CrossRef]
  7. Zhang, Y.; Zhang, S.; Jiang, X.; Zhao, W.; Wang, Y.; Zhu, P.; Yan, Z.; Zhu, H. Uniaxial tensile properties of multi-scale fiber reinforced rubberized concrete after exposure to elevated temperatures. J. Clean. Prod. 2023, 389, 136068. [Google Scholar] [CrossRef]
  8. Li, Y.; Hao, J.; Wang, Z.; Guan, Z.; Wang, R.; Chen, H.; Jin, C. Experimental-Computational Investigation of Elastic Modulus of Ultra-High-Rise Pumping Concrete. J. Adv. Concr. Technol. 2020, 18, 39–53. [Google Scholar] [CrossRef]
  9. Zhang, Y.; Zhang, S.; Zhao, W.; Jiang, X.; Chen, Y.; Hou, J.; Wang, Y.; Yan, Z.; Zhu, H. Influence of multi-scale fiber on residual compressive properties of a novel rubberized concrete subjected to elevated temperatures. J. Build. Eng. 2023, 65, 105750. [Google Scholar] [CrossRef]
  10. Bernard, O.; Ulm, F.-J.; Lemarchand, E. A multiscale micromechanics-hydration model for the early-age elastic properties of cement-based materials. Cem. Concr. Res. 2003, 33, 1293–1309. [Google Scholar] [CrossRef]
  11. Li, Y.; Wang, P.; Wang, Z. Evaluation of elastic modulus of cement paste corroded in bring solution with advanced ho-mogenization method. Constr. Build. Mater. 2017, 157, 600–609. [Google Scholar] [CrossRef]
  12. Chen, H.S.; Stroeven, P.; Ye, G.; Stroeven, M. Influence of Boundary Conditions on Pore Percolation in Model Cement Paste. Key Eng. Mater. 2006, 302–303, 486–492. [Google Scholar] [CrossRef]
  13. Chen, H.; Sun, W.; Stroeven, P. Interfacial transition zone between aggregate and paste in cementitious composites (I): Ex-perimental techniques. J. Chin. Ceram. Soc. 2004, 32, 63–69. [Google Scholar]
  14. Chen, H.; Stroven, P. Interfacial transition zone between aggregate and paste in cementifious composties (II): Mechanism of formation and degradation of interfacial transition zone microstructure and its influence factors. J. Chin. Ceram. Soc. 2004, 32, 70–79. [Google Scholar]
  15. Machoviè, V.; Kopecký, L.; Němeček, J.; Kolář, F.; Svítilová, J.; Bittnar, Z.; Andertová, J. Raman micro-spectroscopy mapping and microstructural and micromechanical study of interfacial transition zone in concrete reinforced by poly(ethylene tereph-thalate) fibres. Ceram. Silik. 2008, 52, 54–60. [Google Scholar]
  16. Lutz, M.P.; Monteiro, P.J.; Zimmerman, R.W. Inhomogeneous Interfacial Transition Zone Model for the Bulk Modulus of Mortar. Cem. Concr. Res. 1997, 27, 1113–1122. [Google Scholar] [CrossRef]
  17. Mondai, P.; Shah, S.R.; Marks, L.D. Nanoscale characterization of cementitious materials. Aci Mater. J. 2008, 105, 174–179. [Google Scholar]
  18. Mondal, P.; Shah, S.P.; Marks, L.D. Nanomechanical Properties of Interfacial Transition Zone in Concrete. In Nanotechnology in Construction 3: Proceedings of the NICOM3; Springer: Berlin/Heidelberg, Germany, 2009; pp. 315–320. [Google Scholar]
  19. Chen, H.S. Quantitative calculation of volume fraction of interface transition zone in cement based composites. J. Harbin Inst. Technol. 2006, 23, 133–142. [Google Scholar]
  20. Chen, H.; Sun, W.; Stroeven, P.; Sluys, L.J. Analytical solution of the nearest surface spacing between neighboring aggregate grains in cementitious composites. J. Chin. Ceram. Soc. 2005, 33, 859–863. [Google Scholar]
  21. Han, J.; Pan, G.; Sun, W. Elastic modulus change investigation of cement paste before and after carbonation using nanoindentation technique. Procedia Eng. 2012, 27, 341–347. [Google Scholar] [CrossRef]
  22. Li, Y.; Zhang, G.; Wang, Z.; Wang, P.; Guan, Z. Integrated experimental-computational approach for evaluating elastic modulus of cement paste corroded in brine solution on microscale. Constr. Build. Mater. 2018, 162, 459–469. [Google Scholar] [CrossRef]
  23. Oliver, W.C.; Pharr, G.M. An improved technique for determining hardness and elastic modulus using load and displacement sensing indentation experiments. J. Mater. Res. 1992, 7, 1564–1583. [Google Scholar] [CrossRef]
  24. Scrivener, K.L.; Crumbie, A.K.; Laugesen, P. The Interfacial Transition Zone (ITZ) Between Cement Paste and Aggregate in Concrete. Interface Sci. 2004, 12, 411–421. [Google Scholar] [CrossRef]
  25. Bernard, F.; Kamali-Bernard, S. Numerical study of ITZ contribution on mechanical behavior and diffusivity of mortars. Comput. Mater. Sci. 2015, 102, 250–257. [Google Scholar] [CrossRef]
  26. Yun, G.; Schutter, G.D.; Ye, G.; Huang, H.; Tan, Z.; Kai, W. Porosity characterization of ITZ in cementitious composites: Concentric expansion and overflow criterion. Constr. Build. Mater. 2013, 38, 1051–1057. [Google Scholar]
  27. Sun, X.; Dai, Q.; Ng, K. Computational investigation of pore permeability and connectivity from transmission X-ray microscope images of a cement paste specimen. Constr. Build. Mater. 2014, 68, 240–251. [Google Scholar] [CrossRef]
  28. Lafhaj, Z.; Goueygou, M.; Djerbi, A.; Kaczmarek, M. Correlation between porosity, permeability and ultrasonic parameters of mortar with variable water/cement ratio and water content. Cem. Concr. Res. 2006, 36, 625–633. [Google Scholar] [CrossRef]
  29. Zou, D.; Cheng, H.; Liu, T.; Qin, S.; Yi, T.-H. Monitoring of concrete structure damage caused by sulfate attack with the use of embedded piezoelectric transducers. Smart Mater. Struct. 2019, 28, 105039. [Google Scholar] [CrossRef]
  30. Hill, R. Theory of mechanical properties of fibre-strengthened materials—III Self-consistent model. J. Mech. Phys. Solids 1965, 13, 189–198. [Google Scholar] [CrossRef]
  31. Budiansky, B. On the elastic moduli of some heterogeneous materials. J. Mech. Phys. Solids 1965, 13, 223–227. [Google Scholar] [CrossRef]
  32. Eshelby, J.D. The determination of the elastic field of an ellipsoidal inclusion, and related problems. Proc. R. Soc. Lond. 1957, 241, 376–396. [Google Scholar]
  33. Timoshenko, S.; Goodier, J. Theory of Elasticity, 3rd ed.; Mc Graw Hill: New York, NY, USA, 1970. [Google Scholar]
  34. Christensen, R.M. Mechanics of Composite Materials; John Wiley & Sons, Inc.: New York, NY, USA, 1979. [Google Scholar]
  35. Christensen, R.; Lo, K. Solutions for effective shear properties in three phase sphere and cylinder models. J. Mech. Phys. Solids 1979, 27, 315–330. [Google Scholar] [CrossRef]
  36. Zhu, H.; Chen, Q.; Yan, Z.; Ju, J.W.; Zhou, S. Micromechanical models for saturated concrete repaired by the electrochemical deposition method. Mater. Struct. 2013, 47, 1067–1082. [Google Scholar] [CrossRef]
  37. Sun, C.; Chen, J.; Zhu, J.; Zhang, M.; Ye, J. A new diffusion model of sulfate ions in concrete. Constr. Build. Mater. 2013, 39, 39–45. [Google Scholar] [CrossRef]
  38. Rajasekaran, G. Sulphate attack and ettringite formation in the lime and cement stabilized marine clays. Ocean Eng. 2005, 32, 1133–1159. [Google Scholar] [CrossRef]
  39. Chen, J.-K.; Jiang, M.-Q. Long-term evolution of delayed ettringite and gypsum in Portland cement mortars under sulfate erosion. Constr. Build. Mater. 2009, 23, 812–816. [Google Scholar] [CrossRef]
  40. Bonakdar, A.; Mobasher, B. Multi-parameter study of external sulfate attack in blended cement materials. Constr. Build. Mater. 2010, 24, 61–70. [Google Scholar] [CrossRef]
  41. Li, Y.; Li, Y.; Wang, P.; Wang, Z.; Guan, Z. Methodology of Evaluating Elastic Modulus of Cement Mortar Corroded in Brine. J. Adv. Concr. Technol. 2019, 17, 434–448. [Google Scholar] [CrossRef]
  42. Li, Y.; Guan, Z.; Wang, Z.; Zhang, G.; Ding, Q. Evaluation of Elastic Modulus of Cement Paste in Sodium Sulfate Solution by an Advanced X-CT-Hydration-Deterioration Model with Self-Consistent Method. Adv. Cem. Res. 2019, 32, 1–41. [Google Scholar] [CrossRef]
  43. Tennis, P.D.; Jennings, H.M. A model for two types of calcium silicate hydrate in the microstructure of Portland cement pastes. Cem. Concr. Res. 2000, 30, 855–863. [Google Scholar] [CrossRef]
  44. Jennings, H.M. Refinements to colloid model of C-S-H in cement: CM-II. Cem. Concr. Res. 2008, 38, 275–289. [Google Scholar] [CrossRef]
  45. Wu, L.; Song, G.; Lei, B. Prediction of elastic modulus of hardend cement pastes. J. Build. Mater. 2011, 14, 819–823. [Google Scholar]
  46. Lv, J. Prediction of the Elastic Moduli of Cement-Based Materials; Zhejiang University of Technology: Zhejiang, China, 2007. [Google Scholar]
  47. Sun, G.; Sun, W.; Zhang, Y.; Liu, Z. Quantitative calculation on volume fraction of hydrated products in Portland cement. J. Southeast Univ. Nat. Sci. Ed. 2011, 41, 606–610. [Google Scholar]
  48. Chen, H.; Sun, W.; Zhou, Y.; Piet, S.; Johannes, S.L. Analytical solution of the nearest surface spacing between neighboring binder grains in fresh cement paste. Acta Mater. Compos. Sin. 2007, 24, 127–133. [Google Scholar]
  49. Garboczi, E.J.; Bentz, D.P. Analytical formulas for interfacial transition zone properties. Adv. Cem. Based Mater. 1997, 6, 99–108. [Google Scholar] [CrossRef]
Figure 1. Roadmap of this paper.
Figure 1. Roadmap of this paper.
Materials 16 06167 g001
Figure 2. Specimens for nanoindentation test.
Figure 2. Specimens for nanoindentation test.
Materials 16 06167 g002
Figure 3. Uniaxial compression test of the mortar specimen.
Figure 3. Uniaxial compression test of the mortar specimen.
Materials 16 06167 g003
Figure 4. Schematic of nanoindentation test for the corroded mortar.
Figure 4. Schematic of nanoindentation test for the corroded mortar.
Materials 16 06167 g004
Figure 5. The load–displacement curve of a typical indentation point.
Figure 5. The load–displacement curve of a typical indentation point.
Materials 16 06167 g005
Figure 6. SEM-BSE of nanoindentation points. (a) Lattice map of nanoindentation points; (b) single nanoindentation point.
Figure 6. SEM-BSE of nanoindentation points. (a) Lattice map of nanoindentation points; (b) single nanoindentation point.
Materials 16 06167 g006
Figure 7. Image processing of ITZ. (a) SEM-BSE image; (b) image after threshold segmentation.
Figure 7. Image processing of ITZ. (a) SEM-BSE image; (b) image after threshold segmentation.
Materials 16 06167 g007
Figure 8. Variation in porosity with the increase in the distance from the sand edge.
Figure 8. Variation in porosity with the increase in the distance from the sand edge.
Materials 16 06167 g008
Figure 9. The schematic diagram of the nanoindentation results of corroded mortar (M1). (a) SEM-BSE image of the nanoindentation region; (b) microscopic elastic modulus of 8 × 8 nanoindentation points.
Figure 9. The schematic diagram of the nanoindentation results of corroded mortar (M1). (a) SEM-BSE image of the nanoindentation region; (b) microscopic elastic modulus of 8 × 8 nanoindentation points.
Materials 16 06167 g009
Figure 10. The schematic diagram of the nanoindentation results of corroded mortar (M2). (a) SEM image of nano indentation region; (b) microscopic elastic modulus of 8 × 8 nanoindentation points.
Figure 10. The schematic diagram of the nanoindentation results of corroded mortar (M2). (a) SEM image of nano indentation region; (b) microscopic elastic modulus of 8 × 8 nanoindentation points.
Materials 16 06167 g010
Figure 11. Multiscale microstructure of corroded mortar.
Figure 11. Multiscale microstructure of corroded mortar.
Materials 16 06167 g011
Figure 12. Two-step homogenization method. (a) Mortar; (b) equivalent result of the first step; (c) equivalent result of the second step.
Figure 12. Two-step homogenization method. (a) Mortar; (b) equivalent result of the first step; (c) equivalent result of the second step.
Materials 16 06167 g012
Figure 13. Schematic of Lu and Torquato’s model.
Figure 13. Schematic of Lu and Torquato’s model.
Materials 16 06167 g013
Table 1. Mineral composition of cement.
Table 1. Mineral composition of cement.
CompositionC3AC2SC3SC4AF
Content (%)7.4116.7963.9411.86
Table 2. Mortar mixing ratio design.
Table 2. Mortar mixing ratio design.
TypeWater (g)Cement (g)Sand (g)w/c
M140075015000.53
M2312.290013500.35
Table 3. Theoretical calculation results of the elastic parameters of cement paste at 88 d (Gpa).
Table 3. Theoretical calculation results of the elastic parameters of cement paste at 88 d (Gpa).
w/cBulk ModulusShear ModulusElastic Modulus
0.5314.038.6321.48
0.3515.139.3323.22
Table 4. The theoretical calculation results of volume fractions of phases in mortar (%).
Table 4. The theoretical calculation results of volume fractions of phases in mortar (%).
TypeSandCement Paste ITZ
M147.041.911.1
M246.043.110.9
Table 5. Prediction results of elastic modulus of corroded mortar (M1).
Table 5. Prediction results of elastic modulus of corroded mortar (M1).
Homogenization ProcessInput ParametersOutput Results
Elastic ParametersVolume FractionsElastic Parameters
k (GPa) μ (GPa) C (%) k (GPa) μ (GPa)E (GPa)
Level I: Equivalent matrix
(SC method)
ITZ9.746.1320.9412.988.0419.99
Cement14.038.6379.06
Level II: mortar
(two-phase ball model)
Equivalent matrix12.988.045317.9013.5132.38
Sand29.2827.7447
Table 6. Prediction results of the elastic modulus of corroded mortar (M2).
Table 6. Prediction results of the elastic modulus of corroded mortar (M2).
Homogenization ProcessOutput ResultsOutput Results
Elastic ParametersVolume FractionsElastic Parameters
k (GPa) μ (GPa) C (%) k (GPa) μ (GPa)E (GPa)
Level I: Equivalent matrix
(SC method)
ITZ12.127.6220.5514.458.9522.25
Cement15.139.3379.45
Level II: mortar
(two -phase ball model)
Equivalent matrix14.458.955418.8414.1633.97
Sand29.2827.7447
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

Guan, Z.; Zhang, H.; Gao, Y.; Song, P.; Li, Y.; Wu, L.; Wang, Y. Experimental–Computational Investigation of the Elastic Modulus of Mortar under Sulfate Attack. Materials 2023, 16, 6167. https://doi.org/10.3390/ma16186167

AMA Style

Guan Z, Zhang H, Gao Y, Song P, Li Y, Wu L, Wang Y. Experimental–Computational Investigation of the Elastic Modulus of Mortar under Sulfate Attack. Materials. 2023; 16(18):6167. https://doi.org/10.3390/ma16186167

Chicago/Turabian Style

Guan, Zhongzheng, Haochang Zhang, Yan Gao, Pengfei Song, Yong Li, Lipeng Wu, and Yichao Wang. 2023. "Experimental–Computational Investigation of the Elastic Modulus of Mortar under Sulfate Attack" Materials 16, no. 18: 6167. https://doi.org/10.3390/ma16186167

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