A Review of The Interfacial Transition Zones in Concrete-Identification Physical Characteristics and Mechanical Properties

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 29

Engineering Fracture Mechanics 300 (2024) 109979

Contents lists available at ScienceDirect

Engineering Fracture Mechanics


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

A review of the interfacial transition zones in concrete:


Identification, physical characteristics, and mechanical properties
Qingqing Chen a, b, Jie Zhang a, b, c, *, Zhiyong Wang a, b, Tingting Zhao a, b,
Zhihua Wang a, b, *
a
Institute of Applied Mechanics, College of Mechanical and Vehicle Engineering, Taiyuan University of Technology, Taiyuan 030024, PR China
b
Shanxi Key Lab. of Material Strength & structural Impact, College of Mechanical and Vehicle Engineering, Taiyuan 030024, PR China
c
Department of Civil and Environmental Engineering, National University of Singapore, 1 Engineering Drive 2, 117576, Singapore

A R T I C L E I N F O A B S T R A C T

Keywords: Interfacial Transition Zones (ITZ) with high porosity are easier pathways for damage initiation
Interfacial transition zone (ITZ) and propagation. However, their mechanical performance has hardly been examined owing to the
Mechanical properties length scale of the microregions, as well as the heterogeneous and opaque constituent properties,
Affecting factors
which have been increasingly studied in recent years. This study presents a comprehensive review
Damage evolution
Physical characteristics
of the ITZ, determining strategies at the meso-micro level, quantitative characterization, and
physical features. It focuses on the effects of various casting and material composition factors on
the microstructural and mechanical properties of the ITZ, including porosity, thickness, volume,
and elastic modulus. The correlations between the microstructures and the essential mechanical
properties are discussed in detail. The weak bonding properties of the ITZ can be tailored by the
material composition, which changes its internal microstructure, enhances the adhesion of the
paste to aggregates, and fills micropores in cement. Several approaches can be employed to
achieve this, such as adding minerals, increasing the curing age, and adjusting the water-cement
ratio and sand fineness. The content of the weak ITZ can be reduced by optimizing the size
distribution and volume fraction of the aggregates. Furthermore, modelling the gradient distri­
bution of the ITZ thickness and elastic modulus is crucial in numerical simulations to enable an
accurate investigation of the damage accumulation process in concrete. This study also identifies
the challenges in the mechanical behavior of ITZ and serves as an engineering design guide for
more advanced fracture-resistant concrete structures.

1. Introduction

Owing to its low density, high porosity, and reduced stiffness and strength [1,2], the interfacial transition zone (ITZ) is often
considered the weakest region in concrete, the primary location for crack initiation and propagation, and the region prone to stress
concentrations. The damage induced by the propagated microcracks compromises the mechanical properties of concrete, potentially
leading to structural failure [3,4]. To understand the mechanisms of damage initiation and evolution in concrete, researchers have
focused on the determination and quantification of the physical and mechanical properties of the ITZ [1,5–7], as well as the

* Corresponding authors at: Institute of Applied Mechanics, College of Mechanical and Vehicle Engineering, Taiyuan University of Technology,
Taiyuan 030024, PR China.
E-mail addresses: zhangjie04@tyut.edu.cn (J. Zhang), wangzh077@163.com (Z. Wang).

https://doi.org/10.1016/j.engfracmech.2024.109979
Received 30 October 2023; Received in revised form 28 January 2024; Accepted 20 February 2024
Available online 24 February 2024
0013-7944/© 2024 Elsevier Ltd. All rights reserved.
Q. Chen et al. Engineering Fracture Mechanics 300 (2024) 109979

examination of various factors influencing these properties.


Farran first reported that the microstructures of the ITZ were significantly different from those of the cement matrix in concrete [8].
During the mixing and casting process of concrete, cement particles react with water to form various hydration products, and the
discontinuous pores of the cement particles are gradually filled with smaller crystals [9–11]. In contrast, owing to their water-
absorbing properties, the rough surfaces of the aggregates create a water-rich environment, leading to the formation of larger crys­
tals in the proximity of these aggregate particles [12,13]. Accordingly, the transition region between the cement and aggregates is
defined as the ITZ, reflecting the gradually changing distribution of voids, hydration products, and unreacted cement particles
[14–18]. The pronounced compositional differences between the ITZ and cement paste illustrate that the ITZ is a rather heterogeneous
and complex region with a thickness of a few tens of micrometers thick [19–22]. Distinguish the ITZ from aggregates and bulk paste is
challenging because of its unclear and irregular boundaries and difficulty in determining its physical and mechanical properties
[12,13,23]. Many scholars have conducted similar studies to identify and measure the ITZ in concrete using various testing methods
[9,14,24–28]. Recent advancements in imaging techniques have enabled researchers to investigate the nano- and microstructures of
the ITZ in more detail, providing visualization and characterization of its morphology, various physical features, and chemical
composition [23,25,29–31]. The coupled use of various techniques has allowed for quantitative descriptions of the physical charac­
teristics of the ITZ, and the establishment of correlations between micro- and mesoscopic structures and the macro-mechanical
behavior of concrete [12,16,25,29,31–35].
Despite the notable advancements in experimental studies on the ITZ, these methods are often restricted to a limited region of the
structure, making it challenging to capture the full range of ITZ properties and the internal damage process in the ITZ. Consequently,
numerical approaches have emerged as important and effective alternatives for exploring the effects of various factors on the me­
chanical properties of the ITZ in a comprehensive and detailed manner [36–38]. These models primarily focus on constructing the ITZ
microstructure and adopting appropriate modelling methods for the ITZ in mesoscopic numerical models. To overcome the limited
resolution of 3D imaging techniques, various numerical studies have focused on reconstructing complete 3D microstructures of ITZ
that exhibit porosity gradients [5,39–41]. These models provide a precise characterization of the degree of hydration within the ITZ to
accurately simulate its formation and evolution. Notably, codes including HYMOSTRUCT3D [27], µic [40], and CEMHYD3D [41] have

Fig. 1. The factors affecting the microstructure and mechanical properties of ITZ.

2
Q. Chen et al. Engineering Fracture Mechanics 300 (2024) 109979

implemented these models and accurately simulated critical microstructural features of ITZ. In addition, numerical meso-structural
models offer a practical way to study the damage and fracture of concrete with an explicit representation of microcrack initiation and
propagation [4,16,18]. There are various approaches to modelling the ITZ using finite and discrete element modelling. One such
approach involves ignoring the ITZ thickness and assuming an ideal bond between the aggregates and mortar matrix, which attributes
all inelastic and damage performance to mortar matrix [42,43]. Another method involves placing a zero-thickness cohesive element at
the interface between the cement and aggregate solids, which allows for the attribution of separate damage and failure behaviors to the
weak ITZ [34,44,45]. Additionally, ITZ can be represented using solid or finite-thickness elements that vary from 0.1 to 2 mm
[32,46,47], which are generally greater than the actual thickness to reduce meshing and computing costs. Furthermore, the discrete
element method allows for a more accurate representation of the heterogeneity of the ITZ [33,48,49] and is suitable for simulating the
post-peak behavior of concrete structures, particularly for modelling concrete under dynamic loading conditions. Compared with
experimental approaches, numerical simulations can provide real-time monitoring of the stress and strain distribution within the ITZ,
as well as the evolution of cracks and other damages, which is crucial for predicting the long-term mechanical performance of concrete
structures.
Various theoretical approaches have also been developed to describe the physical and chemical processes within the ITZ and
predict its behavior and properties [15,44,50–53]. These analytical expressions, which are derived from continuum [54], micro­
mechanical [55] and multi-scale models [50], mainly focus on the elastic modulus [15,19,44,56] and volume fraction of the ITZ
[51,56–58]. Although several formulas have been proposed to observe the volumes of the ITZ around 3D aggregates with various
geometrical morphologies [59–61], most neglect the overlaps between the ITZs without fully considering the interaction between the
three-phase constituents, which can be further addressed analytically based on the void exclusion probability [51,62]. In terms of the
elastic modulus of the ITZ, the current theoretical research can be summarized into three types. The first-generation models assumed
the ITZ to be a shell of constant thickness with a uniform elastic modulus throughout its region, neglecting the gradual transition of the
properties within the ITZ [15,17,56]. The second generation considers the ITZ region as a gradient material with varying properties
and approximates the elastic modulus of the ITZ as a function of its distance from the aggregate surface [55]. However, they can be
computationally intensive and require large amounts of data, and the initial cement gradient in the ITZ cannot be properly accounted.
The third generation utilized a multi-scale model to determine the elastic moduli, considering the local water-to-cement (w/c) ratio
within the ITZ [50,63]. An innovative n-layered spherical inclusion model was proposed to consider the porosity variation with the w/
c ratio and hydration degree within the ITZ, as well as the distance from the aggregate surface [58]. This model divided an inho­
mogeneous ITZ into a series of homogeneous and isotropic concentric shell elements, thereby providing a more accurate and detailed
representation of the ITZ. Despite these efforts, it remains challenging to develop a comprehensive model that accurately captures all
the physical and chemical features of the ITZ and incorporates the effects of various factors, owing to its variability and heterogeneity.
Previous research has demonstrated that the geometrical and mechanical properties of the ITZ can be affected by many factors,
including the grain size distribution of the cement [5,43,64–66], water/cement ratio [29,64,67–69], aggregate content [5,70–73],
aggregate roughness [26,74–76], aggregate size distribution [70–72,77], type of aggregate, and curing age [5,67,78–80], as presented
in Fig. 1. Recent research has focused on determining the influencing mechanism of each factor on the ITZ to improve the micro­
structure and enhance its bonding properties. Based on their impact on various features of the ITZ, the influencing factors can be
classified into direct and indirect factors: (1). Direct factors such as the w/c ratio, mineral addition, and curing age have a remarkable
impact on the microstructure formed in the ITZ [9,81–83]. These factors can significantly influence the hydration process in concrete
by increasing the amount of cement that participates in chemical reactions. This leads to stronger bonds with the aggregates and
reduces the shrinkage and drying shrinkage of the cement paste, resulting in a denser microstructure in the ITZ. (2). Indirect factors,
such as the size distribution, roughness, and maximum diameter of the aggregates, play a crucial role in determining the total volume
of the ITZ through their effects on the coating area and thickness, although they have a slight influence on the formed ITZ micro­
structure [67,70–72]. Based on the influencing mechanisms on the physical and mechanical properties of the ITZ, various achievable
measures have been taken to improve the microstructures and optimize its properties by adjusting these factors.
In this study, a comprehensive review of the ITZ in concrete is conducted at the micro- and mesoscopic levels based on extensive
experimental, numerical, and theoretical analysis, focusing on the determination of the physical features using various techniques,
characterization of the ITZ mechanical properties, and the evaluation of various influencing factors. The objective of this study is to
establish a deeper understanding of the correlation between the microstructural features and the physical and mechanical properties of
the ITZ and provide insights into the evaluation of the deformation and cracking behavior of the ITZ.

2. Physical and mechanical features of ITZ

Many investigations have found that the root causes of weakness in the ITZ are the presence of microcracks, excessive porosity, and
a high concentration of calcium hydroxide crystals [29,84]. This leads to extreme heterogeneity and poor bonding strength at the
aggregate-cement interfaces [5]. As a weak link in the concrete, the ITZ provides an easier path for micro-crack development
[33,77,85,86]. Initiated microcracks originated from the pores within the ITZ propagate into clear cracks and ultimately lead to the
final fracture of the concrete, thereby significantly affecting the global mechanical performance of concrete structures [7,25].
Consequently, quantitative descriptions of the physical characteristics of the ITZ are critical for examining the mechanical properties of
concrete. Presently, micro-mechanical experiments on the ITZ are performed by observing the changing tendency of the mechanical
properties in the ITZ region to determine its physical features, such as porosity, thickness, and elastic modulus [87]. Numerous re­
searchers have used electron microscopy and nanoindentation techniques combined with in-situ scanning probe microscopy imaging
to quantify the micro-structural gradients throughout the ITZ, providing crucial insights into its morphology and microstructure

3
Q. Chen et al. Engineering Fracture Mechanics 300 (2024) 109979

[10,77,88].

2.1. Methodologies for identifying and measuring ITZ

Over the years, numerous methods have been employed to quantitatively determine the morphological features of the ITZ,
including the volume fraction, thickness, and porosity. This section provides a comprehensive review of the methodologies employed
in identifying and quantifying the micromechanical, microstructural, and chemical characteristics of the ITZ [24,25,33,34], clarifying
the technical characteristics, advantages, and limitations of each technique.
The application of various identification techniques has greatly contributed to the evaluation of pore structures and the deter­
mination of microscopic features and mineral phase content within ITZ, through the processing of different physical signals. The
commonly used identification techniques include X-ray computational tomography (XCT) [2,34,89–92], nuclear magnetic resonance
(NMR) [93–95], alternating current impedance spectroscopy (ACIS) [78,79], mercury intrusion porosimeter (MIP) [18,96,97], and
scanning acoustic microscopy (SAM) [98,99]. Although these techniques offer a broader spatial range for analyzing the three-
dimensional microstructure of the ITZ, but their limited resolution and accuracy may prevent the capture of small-scale features
such as individual mineral phases or components. Detailed descriptions of these techniques are provided below.
(1) X-ray computed tomography (XCT)
The XCT scanning technique enables rapid capture of the three-dimensional microstructural features of the ITZ, which can be
distinguished from the cement paste according to its porosity gradients along the distance from the aggregate. In addition, XCT is
useful for observing hydration processes [7] and determining the type and content of each physical phase in the ITZ [2,34,89–92].
However, the limited spatial resolution of XCT makes it challenging to fully distinguish the compositional information of interfacial
microstructures in concrete, and the signal-to-noise ratio and observation scope require further improvement. Mazzucco et al. [2]
proposed a robust geometric meso-scale reconstruction method for concrete materials using XCT, which enabled the acquisition of the
overall inner geometry and distribution of aggregates and voids through differences in the material density of the components.
(2) Alternating current impedance spectroscopy (ACIS)
The ACIS technique indirectly obtains the concrete microstructure by measuring the complex impedance of the concrete samples at
different AC frequencies. Although the dynamic measurement of cracks and pores in the ITZ can be achieved through ACIS, the
required equipment is relatively expensive, and the testing procedures are more complicated because impedance spectroscopy is
dependent on the size of the concrete specimens. Nevertheless, some researchers have effectively applied ACIS to examine the ITZ
microstructures of concrete samples with different types and sizes of coarse aggregates at different curing ages [83,84]. Findings
demonstrate that this technique is a powerful tool for evaluating the ITZ microstructure.
(3) Mercury intrusion porosimeter (MIP)
The MIP technique is commonly used to measure and evaluate the pore characteristics within the ITZ regions [18,96,97]. The size
distribution of the pores could be indirectly obtained by testing the pressures. However, the testing process can damage pore structures
inside the concrete, and detailed information on the pores, including their shapes and positions, cannot be provided.
(4) Scanning acoustic microscope (SAM)
Ultrasonic waves undergo alterations in their phase characteristics and energy as they travel through various physical mediums.
The resulting data is acquired and evaluated to generate gray-scale images that reveal the internal structural characteristics of the
concrete samples. The SAM can accurately locate defects, especially in the ITZ, by imaging the surface or interior of concrete [98,99].
Despite the high-resolution and nondestructive detection capabilities of SAM, the technology has strict requirements for specimens.
With the development of digital image processing technology, numerous measuring techniques have been efficiently employed to
determine the variation of mechanical properties of ITZ [10,100–105]. These methods mainly rely on the analysis of 2D digital images
obtained from scanning electron microscopy with a backscattered electron detector (SEM-BSE) [23,25,29,30,88,101,106–111],
including tests such as nanoindentation [112,113], microhardness [114]. Depending on the specific requirements of the study, it may
be necessary to employ multi-scale identification and measuring techniques to accurately identify the microstructure characteristics
and measure the mechanical properties of the ITZ [24,25,33,34,90,115]. These studies have made significant contributions to the
understanding of the ITZ in concrete and provided more realistic inputs for micromechanical models.
(1) Backscattered electron (BSE)
SEM is an extensive technology for qualitatively describing the morphological features of the ITZ and has greatly contributed to the
rapid development of the study of the microstructure and constituent phases over the past decades [23,25,29,30,101]. SEM, equipped
with energy spectrum detection, can determine the element type and distribution of the hydration products in the ITZ. Various physical
signals sent by the SEM are used to present the surface microstructure of the concrete specimens, which are selectively examined in
representative regions and then magnified successively to obtain the microstructure of the tested areas. Many SEM-based studies have
been conducted to characterize the microstructure of the ITZ in concrete [7,22], including the analysis of lateral variation in the ITZ
surrounding aggregates and the variation in mean composition with distance from aggregate edges by Diamond [88], as well as the
examination of the characteristics of pores and hydration products in ITZ regions conducted by Liao et al. [106]. To analyze the ITZ
microstructure using SEM-BSE images, one useful method involves cutting narrow strips in the SEM-BSE image starting from the
aggregate surface and evaluating the porosity variation along each strip [67]. Gao et al. [5] applied BSE image analysis at two
magnifications (100 × and 500 × ) to analyze porosity profiles. A 100-pixel image provided a global view of the ITZ, and a typical
porous region of the ITZ can be detected near the aggregates. A 500-pixel image was used for the quantitative analysis of the unhy­
drated grains and hydration products, stacked irregularly with intersecting pores of various sizes and random shapes. With the
extensive application of SEM technology, the morphology and pore distribution of the ITZ have been observed [10,107–111].

4
Q. Chen et al. Engineering Fracture Mechanics 300 (2024) 109979

However, the limited size of the examined micro area may hinder an accurate representation of the entire microstructure, and the
mechanical properties of the ITZ are almost unavailable owing to the loadings of the SEM device at the microscopic level.
(2) Nanoindentation
Nanoindentation is a valuable tool for quantifying the local mechanical properties of cementitious materials, including the elastic
modulus, hardness, and fracture toughness at both the micro- and nanoscales [112,113,116]. A significant benefit of nanoindentation
over micro-indentation is its ability to analyze smaller indentation sizes, which enables the characterization of the intrinsic properties
of the microstructural gradients in the ITZ. Moreover, it allows a comparison of the mechanical characteristics of the cement matrix,
ITZ, and aggregates. The distributions of indentation modulus and hardness obtained using this method can also be used to determine
the thickness of the ITZ. This technique enables quantitative assessment of the nanomechanical properties of different phases
[21,107,117], and has been used in conjunction with SEM to characterize the microstructure of the ITZ [5,107]. For instance, Gao et al.
[118] employed a coupled nanoindentation and SEM approach to investigate the microstructures and elastic properties of the ITZ,
whereas Khedmati et al. [107] adopted nanoindentation coupled with SEM-EDX and laser scanning to examine the weakest zone in
concrete. By coupling nano-mechanical testing with a transmission electron microscope, it is possible to establish a one-to-one rela­
tionship between the load-depth characteristics and stress-induced microstructure evolution, allowing real-time monitoring of the
microstructure during testing [24]. This technique was developed to analyze the mechanical properties of the microstructural gra­
dients at the ITZ in concrete [119,120], although its application is restricted by the size of the nano-indent in comparison with the
thickness of the ITZ. Moreover, the preparation of the samples can be challenging and may involve various steps, including slicing the

Fig. 2. (a) SEM micrograph of ITZ in concrete [31]. (b) Backscattering image of the concrete sample and the element content distribution along the
line scan path [32,129]. (c) Schematic of porosity distribution in ITZ [39].

5
Q. Chen et al. Engineering Fracture Mechanics 300 (2024) 109979

samples into tiny cuboids, impregnating them with epoxy resin under vacuum conditions, polishing them with diamond particles, and
cleaning them in an ethanol ultrasonic bath. The elastic modulus measured via nanoindentation may be affected by the presence of
pores in the sample, the area filled with epoxy resin, and the spacing between the indentation test points [29].
(3) Digital image correlation (DIC)
Digital image correlation (DIC) is a rapidly evolving technique for accurate full-filed deformation measurements that has been
widely employed in the analysis of concrete materials [121]. For instance, Chen et al. [122] employed DIC to examine the distribution
of shrinkage displacement and strain in concrete. DIC has been confirmed as a highly precise method for the damage identification and
failure measurement of concrete, as reported in numerous studies [123–126]. In addition, the coupling of DIC with SEM has been
utilized to quantitatively evaluate the mechanical properties and thickness of the ITZ, enabling precise determination of deformation
within the ITZ regions [29].

2.2. Porosity

The ITZ exhibits a significantly higher porosity, which is typically two to three times greater than that of the hardened cement
matrix away from the aggregate surfaces. This is evident in the SEM microstructure image of the concrete, which shows large
concentrated voids and cracks within the ITZ regions. The ITZ surrounding the aggregates can be defined as a region with excessive
porosity, which can be further divided into three layers based on the physical factors that influence its pore structure [20]. The first

Fig. 3. (a)The determination of ITZ thickness based on the porosity profiles [26,130,131]. (b) The variation of the porosity of different specimens at
various ratio of w/c [64,68]. (c) Measured porosity profiles based on the BSE image analysis of 0-50μ m and 0-15μ m for the mortars with different
casting factors [5].

6
Q. Chen et al. Engineering Fracture Mechanics 300 (2024) 109979

layer, a thin region of 0 μm–10 μm surrounding the aggregates, is primarily affected by the water film, resulting in a remarkably high
degree of porosity [107]. In the second layer, located 10 μm–30 μm away from the aggregate surface, the ITZ properties are more
influenced by the surface roughness of aggregates and the relative distance from the aggregate-cement boundary [23]. The third layer,
situated more than 40 μm from the aggregate surface, is determined by the surface morphology of aggregates, where the ITZ
microstructure is relaxed to the pore structure of cement.
A comprehensive understanding of the gradient performance of the porosity in the ITZ region is essential for studying its me­
chanical properties. Some models have been developed to introduce a gradient in the ITZ to characterize its microstructure
[50,55,127]. The gradient in the ITZ microstructure was determined by analyzing the BSE images of concrete, which are widely
adopted in models. To further quantify the ITZ porosity gradient, a quantitative relationship between the porosity and grey value of
BSE images was established [128]. While the maximum porosity of the ITZ is commonly reported to be approximately two to three
times that of cement paste [5,96,127], the exact value is difficult to determine, because it is affected by several factors, such as the
shape, type of aggregates, and w/c ratios. Although various investigations have provided a reasonable range for ITZ porosity,
determining its appropriate values remains the primary focus for the initiation and evolution of damage in the ITZ.
To evaluate the porosity of ITZ in concrete, some researchers have employed SEM-EDS and conventional optical microscopy
techniques [31]. Based on the grey-level analysis of the BSE images, it was observed in Fig. 2(a) that the porosity of the ITZ was higher
than that of cement, which is consistent with previous studies [67]. High resolution BSE images clearly reveal the distribution of
various physical phases, including mainly CH and ettringite crystals and a small quantity of C-S-H on the concrete surface, and provide
a clear representation of the microstructure of concrete [32,129]. Through the analysis of the BSE images using a grey-level approach,
the physical phases of sand, aggregates, pores, and reacted and unreacted cement were clearly determined, as shown in Fig. 2(b). Along
the scan path, the Ca and Si content distributions can be classified into various parts. The gradual increase in the Si content at the
aggregate boundary suggests the deposition of hydration products. Furthermore, a significant increase in the Ca content in the ITZ
region demonstrates that calcium hydroxide is the primary contributor to the formation of hydration products. Compared to the
cement paste, there were significant fluctuations in the levels of Ca and Si within the ITZ, revealing the existence of micropores in this
region. These findings provide supplementary evidence supporting the notion that the ITZ exhibits a higher degree of porosity than
other concrete components.
Considering the nonuniform distribution of pores within the ITZ, virtual 3D ITZ microstructures have been developed as an
alternative to the real microstructure of the ITZ [39], which reproduces the probabilistic descriptions of micropore microstructures and
the porosity distribution along the ITZ. The porosity distribution in the ITZ region obtained from the 2D SEM images is shown in Fig. 2
(c). Whereas the porosity along the bulk cement paste fluctuated around the mean value, the value increased as it approached the
aggregate surface.
Because the ITZ is a statistical concept, the average porosity distribution curves can be used to identify its regions. Many researchers
[9,26,130,131] have determined the ITZ thickness by examining the porosity profiles perpendicular to the interface within the
designated region of interest, as illustrated in Fig. 3(a)(b). A steep porosity gradient can be observed within a small distance (up to 10
μm) from the aggregate boundary, which is attributed to the formation of a thick water film surrounding the aggregates [106]. As the
position moved away from the aggregate boundary, the ITZ thickness was determined at the point where the porosity finally became
uniform.
To quantitatively evaluate the porosity, some researchers have conducted SEM-BSE image analysis to study the effects of different
casting factors on the porosity as a function of the distance from the aggregate-cement interface [68,132]. Many investigations have
shown that the w/c ratio is a critical factor that affects the quality of the ITZ and cement [6,23,67,88]. Laboratory experiments further
demonstrated that a decrease in w/c led to a significant reduction in the porosity of both the ITZ and cement (Fig. 3(b)) [64,68]. This
can be attributed to the densification of the microstructure and the filling of the pores by various hydration products [5,69]. A lower w/
c ratio in concrete enables more cement particles to be closely distributed around the aggregates, resulting in a reduction in the
porosity and thickness of the ITZ [31]. Based on the BSE image analysis and the Hymostruc3D model developed by Gao et al. [5], the
effects of the curing age, w/c ratio, and volume fraction of sand on the thickness and porosity of the ITZ are discussed, as presented in
Fig. 3(c). It is worth noting that the microstructure becomes denser with the increased curing age (28–90 days) and volume fraction of
sand particles (10–30 %), whereas it becomes looser with the increasing w/c ratio (0.3–0.5), especially within the 15–50 μm zones. A
longer curing age promoted a higher degree of hydration, and the filling action of the hydration products caused a significant reduction
in porosity within the ITZ [80]. These findings are consistent with previous studies on the effects of the w/c ratio on the porosity and
microstructure of ITZ. The porosity decreases consistently with the increase in volume fraction of sand particles, while the mortar
specimen (D3) with 30 % sand grain content shows a slight increase in porosity around 45 μm from the particle boundaries, indicating
that some variations in the homogeneity of the paste occur with the increasing sand grain content, as elaborated in Ref. [73]. To
investigate the influence of various casting factors on the porosity of the ITZ, researchers have employed an equivalent circuit model to
measure the resistance of the connected pores and the MIP technique to examine the ITZ microstructure [79]. The findings revealed
that incorporating aggregates into cement leads to a more porous ITZ, as evidenced by the smaller resistance of the connected pores in
concrete specimens with aggregates than in those without aggregates. Furthermore, the ITZ porosity index increased with increasing
aggregate size owing to the accumulation of a thicker water film around the larger aggregates, which caused a more porous micro­
structure. However, some studies found that the influence of aggregate size on the ITZ porosity was negligible, likely due to the small
differences between the aggregate sizes tested (500 μm, 1000 μm, and 2000 μm) [40]. By testing cement specimens with various mix
designs, researchers have also investigated the influence of cement fineness on the ITZ properties and specimen strength [64,133].
Increasing cement fineness reduces the porosity of concrete, resulting in a denser and more homogeneous microstructure with refined
pore structures.

7
Q. Chen et al. Engineering Fracture Mechanics 300 (2024) 109979

To date, few studies have focused on the influence of aggregate surface roughness on the formation and development of porosity
within the ITZ. The surface texture of the aggregates significantly influenced the variation in the ITZ characteristics. Based on the
quantitative analysis of SEM-BSE images and a K-means clustering algorithm, the uneven distribution of physical properties and in­
fluence of the aggregate surface texture on the ITZ microstructure have been investigated in some studies [134], as presented in Fig. 4.
The thickness and porosity of the ITZ can vary considerably around a single aggregate, depending on its location relative to the
aggregate boundary [135]. As shown in Fig. 4(a), the interfacial porosity above the upper surface of the aggregates was smaller than
that on the side and lower surfaces of the aggregates, indicating an uneven distribution of interfacial porosity around the same
aggregate. More porous ITZ microstructures can be examined below the aggregates because of the density differences of various
physical phases, which results in the formation of water-rich pockets below the aggregates, whose size depends on the size of the
aggregates [136–138]. The results presented in Fig. 4(b) indicate that the ITZ porosity gradients around rough surfaces were larger
than those around smooth surfaces, primarily because of the greater impact of rough surfaces on the packing of cement particles
compared to smooth surfaces. Furthermore, the ITZ microstructure surrounding the smoother aggregates was more compact.
In general, the porosity of the ITZ is affected by a range of factors, including the w/c ratio, curing age, cement fineness, aggregate
content, aggregate surface roughness, and aggregate size [5]. Specifically, the porosity of the ITZ decreased significantly with
increasing curing age and cement fineness, as well as with decreasing w/c ratios. Although the w/c ratio and aggregate content
influenced only the porosity of the ITZ, the curing age affected both the thickness and porosity of the ITZ.

2.3. Thickness

One widely accepted viewpoint is that the ITZ is most likely caused by a size discrepancy between the aggregate particles and
cement grains [9,41]. As cement grains are mixed with aggregates, the normal packing of cement particles can be disrupted because of
wall effects, with smaller cement particles are being packed near aggregates. Previous studies have suggested that the ITZ thickness
should be comparable to the median particle size of the cement in concrete; however, these studies only considered the cement phase
without accounting for the influence of aggregates [9]. Some researchers consider the ITZ to be as a slip layer without thickness
[115,139], whereas others believe that it has actual mechanical behavior and determine its thickness using backscattered electron
(BSE) techniques [140].

2.3.1. Thickness measurement


Measuring the thickness of the ITZ is challenging because of its small length and opaque properties. Several methods have been
proposed to determine the ITZ thickness [6,9,141]. One approach introduced by Hu et al. [142] defined the ITZ thickness as the

Fig. 4. The influences on the porosity distribution within ITZ of (a) three analyzed relative locations around aggregates and (b) surface roughness at
different locations around the aggregate [26].

8
Q. Chen et al. Engineering Fracture Mechanics 300 (2024) 109979

distance from the aggregate surface to the point at which the bulk porosity was obtained. In another criterion proposed by Bentz et al.
[143], the ITZ thickness is calculated as the distance from the aggregate boundary at which the initial pore volume is greater than 10 %
of the pore volume in the cement. However, this criterion is inaccurate because of the arbitrary 10 % choice. In addition, a different
criterion characterizes the ITZ thickness as the distance from the aggregate surface where the porosity reaches the average porosity of
the entire cement. Moreover, some studies have utilized the Ca/Si ratio to determine the ITZ thickness, which is defined as the distance
from the aggregate boundary where the Ca/Si values are stable [12].
Numerous studies have reported various estimates of the ITZ thickness range, which depend on the measuring techniques and
analysis models employed. Typically, a polished cross-section of concrete specimens imaged by BSE is utilized to obtain the
morphological features of concrete at the microscopic level [5,6,10]. Using the SEM-BSE images, different physical phases present in
the cement matrix, such as sand, pores, and hydration products, were identified according to their brightness order. Subsequently, the
ITZ thickness around the aggregates was determined using successive strips based on a concentric expansion technique [6,8,106,141].
As illustrated in Fig. 5, some studies have used the inflection points of the porosity profiles to define the boundaries between the
cement and ITZ regions and then calculated the ITZ thickness [26,29,31,89,144].
Table 1 lists ITZ thicknesses obtained using different measurement techniques in various studies. Basheer et al. [145] provided a
general evaluation of the ITZ thickness varying from 0 to 100 μm, while SEM-based [11] investigations suggested that the thickness
was more precise between 10 and 50 μm. In studies using large single aggregates, such as cylindrical cores or tubes, the measured ITZ
thickness was greater than 50 μm [106]. According to scanning electron microscope images, Scrivener et al. [9] discovered that the
variations of mechanical properties in ITZ were progressive and most noticeable in the initial 15 μm to 20 μm around aggregates. The
thickness and mechanical properties of the ITZ were also analyzed experimentally using a coupling strategy of SEM and the DIC
method [29]. The electron probe technique was used to determine the ITZ thickness around aggregates of varying sizes [32], and the
statistical analysis indicated that the thickness typically followed a normal distribution. Some studies have also indicated that the ITZ
thickness differs slightly for various sizes of aggregates and for different areas around the same aggregate [26].
According to experimental and numerical studies, the thickness and properties of the ITZ can be affected by different variables,
including the volume fraction of the aggregates, curing age, w/c, and sand grain size [5,31,63,65,114,131,148,152,153]. Although the
ITZ thickness has been found to depend on the w/c ratio and maximum size of cement particles in conventional concrete [66], it
appears to be independent of the aggregate diameter [154]. Consequently, earlier research on ITZ formation focused on the cement
matrix [42,155], paying less attention to the influence of aggregates. It has been shown that the ITZ thickness varies between several
micrometres and the mean size of cement grains [43] and is a function of paste thickness, w/c ratio, sand fineness, and curing age [65].

2.3.2. Thickness in numerical modellings


In numerical simulations, there are various approaches to characterizing the ITZ during the modelling process, as shown in Fig. 6.
The easiest approach is to ignore the ITZ thickness [42,43] and assume an ideal bond between the aggregates and mortar matrix.
Although these models attribute all inelastic and damage performance to mortar matrix [156–158] and are unable to predict the post-
peak softening stage of stress–strain curves, they are often used in some studies because of the ease of establishing the models. In
particular, for the dynamic behavior of large concrete structures subjected to high-intensity loads, such as penetration, high-speed
impacts, or explosions [159–161], it is necessary to compromise between the ITZ thickness and computational effort. The second
method characterizes the ITZ by placing a zero-thickness cohesive element at the interface between the cement and aggregate solids

Fig. 5. Determination of ITZ thickness around aggregates based on the BSE image of concrete [26,29,31,89,144].

9
Q. Chen et al. Engineering Fracture Mechanics 300 (2024) 109979

Table 1
The determination of ITZ thickness based on various techniques.
Reference Years Thickness of ITZ Measurement method

Asbridge et al. [146] 2002 20μ m-100μ m Microhardness testing


Yang et al. [147] 2003 Around 20μ m (10 % silica fume) Accelerated chloride migration test (ACMT)
Scrivener et al. [9,10] 2004 15μ m-50μ m Backscattered electron images
Rossignolo [12] 2009 55μ m SEM connected to an energy dispersion X-ray analyzer (EDX)
Xiao et al. [16] 2013 0.01 mm-0.05 mm Nanoindentation technique
Gao et al. [5,13] 2013;2014 15μ m BSE imaging
Skarżyński et al. [89] 2016 30μ m-50μ m Scanning electron microscope
Xu et al. [148] 2017 35μ m-40μ m Grid nanoindentation,
Branch et al. [149] 2018 10μ m-40μ m BSE (FEI Quanta 650 Scanning Electron Microscope)
Tian et al. [32] 2018 39.5μ m-42μ m Electron probe technique
Gao et al. [68] 2018 15μ m-30μ m SEM-BSE image analysis
Nežerka et al. [101] 2019 7.3μ m-23.46μ m (mineral admixtures) Electron microscopy (SEM) with a FEG-SEM Merlin Zeiss microscope
Li et al. [150] 2021 29.62μ m; 20.79μ m (10 % silica fume) Grid nanoindentation test
Li et al. [151] 2022 Around 50μ m SEM-BSE test

Fig. 6. Typical ITZ modelling methods in numerical simulations [32,36,37,49,167].

[34,44,45,162–164]. This allows for the attribution of inelastic and damage behavior to cement and damage and failure behavior to
the weak ITZ [45,165–167]. To reduce the meshing and computing costs, a third option is to adopt solid or finite-thickness elements to
model the ITZ [46,47], but with a thickness greater than the physical thickness. One example is the finite element model developed by
Tian et al. [32] which accurately represents the distribution of aggregates, cement, and the ITZ. In this model, ITZ is represented by
elements surrounding the aggregates with a thickness of 40.7 μm, and the elements near the aggregate-cement interface region are
gradually refined to reflect the geometrical characteristics of ITZ. Table 2 presents the various modelling methods for the ITZ and the
corresponding thicknesses of the ITZ obtained by different researchers. Some investigators employ a constant ITZ thickness sur­
rounding all aggregates [33,90,165,168–172], with values ranging from 0.2 to 2 mm [157,170,173,174], whereas others believe that
the ITZ thickness varies with aggregates size [175]. Some models also employ the discrete element method (DEM) with ITZ, repre­
sented by contact elements, connecting beams or springs, to study the mechanical performance of concrete [33,48,49]. Because of the
enormous computational burden, the values of the ITZ thickness adopted in the models presented above were greater than the actual
thickness of the ITZ. To account for the actual thickness of ITZ while reducing the computational cost, an aggregate expansion method
(AEM) is proposed [176], where ITZ is represented as a 50 μm thick layer of wedge elements surrounding the aggregate. This method

10
Q. Chen et al. Engineering Fracture Mechanics 300 (2024) 109979

Table 2
The modelling methods of ITZ in numerical simulations.
References Years ITZ thickness Modelling method Specimens size

Zheng et al. 2007 0.05 mm 3D mesoscopic concrete model by simulation algorithm with 25 × 25 × 25 mm3
[177] periodic boundary conditions
Sun et al. 2007 10μ m-50μ m The Virtual Cement and Concrete Testing Laboratory 50 × 75 × 230 mm3
[178] (VCCTL) module
Zhou et al. 2008 0.2 mm-0.8 mm Using the commercial software AUTODYN to develop two- ∅ 50 mm
[165,169] dimensional models of concrete
Kim et al. 2011 200μ m Random 2D circular shape aggregates model; 3D meso-scale 100 × 100 mm2 (2D);
[179] RVE (representative volume element) of concrete 50 × 50 × 50 mm3
(3D)
Song et al. 2012 0.5 mm The pseudo 3D mesoscale model with random polygon- 50 × 50 × 50 mm3
[170] shaped coarse aggregates
Pedersen et al. 2013 0.5 mm Two-dimensional mesoscopic finite element model with 100 × 74 mm2 (2D)
[174] random aggregate
Huang et al. 2015; 0.1 mm-0.4 mm Based on high-resolution micro-scale XCT images to 37.2 × 37.2 × 37.2
[90,171] 2016 construct the 3D model mm3
Keinde et al. 2015 50μ m; 100μ m 3D three-phase model with spherical aggregates using finite 33 × 33 × 33 mm3
[180] element methods
Skarżyński et al. 2016 0.025 mm X-ray micro-computed tomography to develop 2D 50 × 80 mm2
mesoscopic numerical model
[89]
Zhou et al. 2017 0.5 mm-1 mm Classical “take-and-place” procedure to generate the 3D 50 × 50 × 50 mm3
[181] meso-structure
Lv et al. [182] 2018 Determined by the element size (0.8 mm) Using the grid mapping method to construct the 3D meso- ∅ 120 × 100 mm
scale model
Zhang et al. 2018 Defined by the contact between cement Random discrete-element aggregate model using PFC2D 150 × 150 mm2(2D);
[38] and aggregate elements, without solid ∅ 150 mm(2D)
particles
Yu et al. [172] 2018 Equating the size of each element (0.5 mm) X-ray computerized tomography to construct 3D numerical 45 × 47.5 × 45.5
model; Digital image processing (DIP) techniques (edge mm3
detection)
Zhang et al. 2019 1.0 mm Based on Voronoi technology to generate 3D models by 25 × 25 × 25 mm3
[183] shrinking and extending processes
Wu et al. [184] 2020 35μ m-55μ m Based on COMSOL Multiphysics programs to construct 3D 100 × 100 × 100
the concrete finite element models with sphere aggregates mm3
Maleki et al. 2020 0.02 mm-1 mm Adopting take and place method to generated 3D meso- 25 × 25 × 25 mm3
[37] structure of concrete
Sharma et al. 2020 2μ m layer consisted by consecutive four Using a vector based modelling platform (μic) ____
[40] 0.5 mm thick layers
Nitka et al. 2020 0.75 mm, composed of 3 rows of spheres Two-dimensional mesoscopic models based on the micro-CT 80 × 50 × 40 mm3
[36] images

has only been validated for image-based models [162].

2.4. Volume

Determining the volume fraction of ITZ is essential for investigating the quantitative relationship between the microstructure and
macroscopic properties of concrete [15]. However, experimental measurement of the ITZ volume fraction is challenging due to the
overlapping of the neighboring ITZ layers with a high degree. Over the past decades, researchers have made theoretical and numerical
efforts to assess the ITZ volume fraction in concrete. In theoretical aspects, Garboczi and Bentz [51] proposed a theoretical formula to
observe the ITZ volume fraction around spherical aggregates in normal concrete, which has been extended to predict the elastic
modulus of concrete as a three-phase composite containing spherical aggregates, ITZs, and cement paste [59,60]. Researchers have
also developed a formula to calculate the ITZ volume fraction [57] with the aid of computer simulation technology to quantify the
influences of ITZ on the macro-properties of concrete. In numerical simulation, to simplify the modeling of ITZ, it is considered as a
uniform-thickness shell surrounding the aggregate [185]. In the study conducted by Aouissi et al. [186], the ITZ volume of VITZ was
regarded as a fraction of the volume of aggregates of Vagg . For Vagg below to 60 %, VITZ = 0.3Vagg ; and for Vagg higher than 60 %, VITZ =
0.5Vagg . Some researchers [56,177,178,187] have adopted an approximate method to calculate the ITZ volume fraction by multiplying
the surface area of aggregates by a uniform thickness of ITZ layers, where the ITZ thickness is primarily dependent on the median size
of the cement grains instead of the aggregate size. This method provides relatively accurate results, but it is time-consuming. Based on
the statistical geometry of composites, the nearest surface distribution function in physics presented by Lu and Torquato [62] has been
used to predict the ITZ volume fraction in concrete [57,151,188], as expressed in Eq. (1). The direct analytical solution for calculating
the ITZ volume fraction significantly contributes to the study of the macro mechanical behavior of concrete.
2
VITZ = 1 − Vagg − (1 − Vagg )exp(− πNv (etITZ + dtITZ 3
+ gtITZ )) (1)

11
Q. Chen et al. Engineering Fracture Mechanics 300 (2024) 109979

where
〈 〉
4 R2
e= (2)
1 − Vagg
〈 〉2
4〈R〉 8πNv R2
d= + (3)
1 − Vagg (1 − Vagg )2

〈 〉 〈 〉3
4〈R〉 16πNv R2 〈R〉 64Aπ2 Nv2 R2
g= + 2
+ (4)
3(1 − Vagg ) 3(1 − Vagg ) 27(1 − Vagg )3

where Vagg is the aggregate volume content; tITZ is the assumed ITZ thickness extend from the aggregate surface; Nv is the total number
of aggregate particles per unit volume; the parameters e, d, g are the mean number over the particle size distribution, which are
determined by sieve analysis. 〈R〉 is the average aggregate radius, and 〈R2 〉 is the average square aggregate radius over the entire size
distribution; A can be assigned with different values (0, 2, or 3) in relation to the analytical approximation [62].
To calculate the ITZ volume fraction, a multi-scale model [70] was proposed by investigators, which disregarded the effects of
aggregate size and mineralogical properties and neglected the overlap between ITZs during the assessment of the cement volume
fraction. Moreover, an n-layered spherical inclusion model was proposed by Zheng et al. [58] to predict the elastic moduli of concrete
with an inhomogeneous ITZ and to derive an analytical solution for its volume fraction for general aggregate gradation.
Many researchers have quantitatively analyzed the effects of various factors on ITZ volumes, including the aggregate volume
fraction, maximum aggregate diameter for Fuller gradation, and thickness. Some studies have indicated that an increase in aggregate
volume fraction corresponds to an increase in surface area, leading to a larger ITZ volume fraction [189]. Considering that the elastic
modulus of the matrix is much lower than that of the aggregates, an increase in the aggregate volume fraction indicates a decrease in
the matrix volume fraction. This results in an increase in the overall elastic modulus, which is the combined effect of the ITZ and the
aggregate volume fraction. The investigation conducted by Chen et al. [190] also obtained similar results, observing that with
increasing aggregate volume fraction, the ITZ volume fraction initially increased, peaked at a certain value of aggregate volume, and
then decreased with further addition of aggregate. Moreover, a general theoretical framework for the fraction of ITZs around 3D
arbitrary convex particles was developed by Xu et al. [61] by adopting the contraction factor (CF) as the shape descriptor of an
arbitrary nonconvex particle. The results indicated that the ITZ volume fraction first increased and then decreased with an increasing
aggregate volume fraction for a given CF. These findings suggest that the negative impact of the ITZ on the macro-properties of
concrete can be compensated by additional aggregates. These studies also indicated that, with a determined aggregate fraction, the ITZ
volume fraction monotonically increased as the maximum aggregate diameter decreased. This exploration is in line with the case of 3D
spheroidal particles introduced in the literature [70]. This is because the aggregate surface area per unit volume of concrete decreases
as the maximum aggregate diameter increases. Therefore, the size distribution of the aggregates has a pronounced impact on the ITZ
fraction. From the viewpoint of this interfacial property, a more suitable uniform distribution can be selected to replace the four-
gradation, three-gradation, or two-gradation to design an optimal fully-graded concrete.
However, assuming a uniform ITZ thickness around the aggregate leads to an overestimation the ITZ volume fraction and, sub­
sequently, its influence on the mechanical behavior of concrete. Samal et al. [191] developed an analytical approach to predict and
assign differential ITZ thickness around each aggregate based on Voronoi tessellation in the model. Furthermore, a theoretical
framework was introduced to estimate the ITZ volume fraction, which incorporates the concept of differential ITZ thickness, and
variations in response to different parameters, such as maximum and minimum aggregate sizes, aggregate volume fraction, and w/c
ratio, were systematically analyzed. The elastic modulus of concrete was calculated using a four-phase generalized self-consistent
scheme to validate the proposed approach with the experimental results. It was demonstrated that the estimates of the elastic
modulus of concrete with the assumption of differential ITZ thickness are substantially better than the estimates assuming uniform ITZ
thickness for low modulus values of the ITZ and the midrange of aggregate fractions.
As mentioned previously, the total volume of the ITZ was determined based on the influence of multiple factors on the coating area
and thickness. The coating regions of the ITZ were predominantly dependent on the volume fraction and size distribution of the
aggregates, while exhibiting a modest correlation with the maximum diameter and surface roughness of the aggregates.

2.5. Mechanical properties

2.5.1. Elastic modulus


The elastic modulus of the ITZ plays a significant role in studying the mechanical properties of the ITZ, which helps to better
develop the relationship between the microscale mechanical behavior and the structural response of concrete [14–17]. However,
estimating the real elastic modulus of the ITZ presents significant challenges because of its opaque and inhomogeneous components.
Recent studies have conducted expensive and time-consuming microscale experiments to investigate the elastic modulus of ITZ
[14,20,116,192,193].
The mechanical parameters of the ITZ can be determined based on the performance of the cement in concrete. Revilla et al. [194]
presented a novel method for estimating the elastic stiffness of aggregate-ITZ systems in concrete, focusing on the porosity and volume
of the components, and validated it with the limestone-ITZ and electric arc furnace slag-ITZ systems. In previous works [17,56], the

12
Q. Chen et al. Engineering Fracture Mechanics 300 (2024) 109979

elastic modulus of the ITZ was found to be uniformly distributed and smaller than that of bulk paste by a constant factor represented by
Eq. (5). Lee et al. [195] conducted a comparative study of the elastic modulus of concrete with experimental results, assuming a
difference of 30 %–70 % between the elastic moduli of the ITZ and cement. Various theoretical models have been proposed to predict
the elastic modulus of concrete with an inhomogeneous ITZ, as described below. The three-phase model [196] was extended to a four-
phase model by Ramesh et al. [15], where ITZ was regarded as a homogeneous phase in concrete. The results indicated that EEITZ CP
= 0.5
was appropriate for predicting the experimental results with an ITZ volume of approximately 10 %. Using a four-phase model, some
studies concluded that the bulk modulus of the ITZ was 0.7 times that of cement, and the shear and elastic moduli of the ITZ were
nearly half those of the cement phase [44]. In a developed double-inclusion model [56], it was reported that the value of EEITZ
cp
varied
from 0.2 to 0.4 for 20 μm thick of ITZ and 0.5 to 0.7 for 40 μm thick of ITZ, where spherical aggregates coated with a homogeneous ITZ
layer were considered as a double inclusion incorporated in the cement [197]. Based on the developed n-layered spherical inclusion
model, the effects of maximum aggregate diameter, aggregate gradation, and ITZ thickness on the elastic modulus of concrete were
evaluated quantitatively [74]. Moreover, differential effective medium theory (D-EMT) could also be adopted to compute the elastic
properties of concrete by assigning elastic moduli to the corresponding bulk paste matrix, ITZ, and aggregates [178,198].
Eitz = αEcp (5)

where α is a constant with a proposed value between 0.2 and 0.8 [19,56,150], and Eitz and Ecp are the elastic moduli of the ITZ and
cement paste, respectively.
To investigate the variation in the elastic modulus of the ITZ, a coupled strategy between SEM and DIC techniques has been
proposed [29], which not only simplifies the sample preparation but also overcomes the challenge of measuring the nominal
compressive elastic modulus (ENC ) of the ITZ. The results indicated that the ENC of the ITZ area was less than that of the aggregate and
mortar. As the areas in the ITZ gradually approached those of the aggregate and mortar, the ENC values of these areas approximated the
values of the two materials. The ENC of the ITZ area and the distance from the aggregate-mortar boundary are related by a quadratic
function, as shown in Fig. 7. With increasing distance from the aggregate-mortar boundary, the ENC of the ITZ area first decreased and
then increased, and the average ENC of the ITZ area was approximately 33 % that of the mortar. These findings are consistent with those
of Rao et al. [74] and Gu et al. [199], who reported the ITZ strength to be approximately one-third to half that of the mortar. To account
for the inhomogeneity of the ITZ, some studies have assumed a power-law distribution gradient of porosity in the ITZ [56]. The same
power-law function can be used to predict the elastic modulus variation in the ITZ because the elastic modulus is considered to depend
only on porosity.
Previous studies evaluated the distribution of the elastic modulus within the ITZ and reached varying conclusions through different
theoretical models and experimental means, as illustrated in Fig. 8. The simplest and most commonly adopted assumption is the
uniform distribution of the elastic modulus throughout the ITZ (see Fig. 8(a)). Some researchers have suggested that the elastic
modulus of the ITZ is the weakest in the region close to the aggregate and increases with the distance from the aggregate until it reaches
the elastic modulus of the cement paste (see Fig. 8(b)). However, experimental studies have shown that the elastic modulus of the ITZ
follows a quadratic gradient distribution with distance from the aggregates (see Fig. 8(c)). At present, determining the elastic modulus
of the ITZ in numerical models has become a topic of great interest to researchers.

2.5.2. Macroscopic strengths


Over the past few decades, various experimental methods have been proposed to measure the ITZ properties. Malachanne et al.
[200] conducted a numerical study on the paste-aggregate composite, using a cohesive zone model with adhesion, friction, and
unilateral contact in a finite element model to test the experimental interface response, and demonstrated that this method effectively

Fig. 7. The quadratic function relation between the average ENC and the distance from the aggregate boundary [29].

13
Q. Chen et al. Engineering Fracture Mechanics 300 (2024) 109979

Fig. 8. Different distribution of elastic modulus of ITZ in concrete.

describes the mechanical behavior of the ITZ up to rupture. Subsequently, Nesrine et al. [201] performed numerical modelling of a
local shear test on a cement paste-aggregate composite, employing a cohesive zone model and 3D finite element analysis to identify the
behavior of the interface, with the damage characterized by the cohesive law at this interface. The results revealed that the most fragile
region in concrete is the ITZ, characterized by its thinness and higher porosity, which significantly influenced the mechanical behavior,
particularly in the initiation and propagation of cracks within the less resistant zones. Some researchers have conducted tensile, shear
and bending tests to construct a strength relationship between the bulk paste and ITZ. Mielniczuk et al. [202] performed direct tensile
tests on samples comprising two grains bound by cement paste, which were analyzed experimentally at a local scale (cement-grain
interface) during mechanical testing. This research characterized the behavior and fracture process of the cement interface at a local
scale, focusing on the evolution of force and stiffness during mechanical testing and the progression of cracking and rupture in
cemented grains. Such local characteristics provide strong support for the numerical modelling of concrete at the macroscopic scale.
According to the research findings of Rao et al. [74], the tensile or shear bonding strength of the interface depends on the roughness of
the aggregates when using the same mortar and aggregate type. The tensile bond strength was approximately 50 % of the tensile
strength of the mortar, and this bond strength decreased as the w/c ratio increased, although not as fast as the compressive strength of
the mortar. Under Mode I failure, the tensile bond strength at the interface with the rough concrete surface was approximately one-
fourth to one-half of the tensile strength of the cement mortar. On average, the tensile bond strength of the interface was approximately
one-third that of the mortar. To conduct the numerical simulation of the Brazilian splitting test and the uniaxial compression test,
Zhang et al. [38] regarded the ITZ strength as 0.4 times the strength of cement mortar, which yielded simulation results closest to those
of the experiments [74,106]. Additionally, the mesoscopic numerical model of the uniaxial compressive and tensile tests revealed that
concrete could be regarded as a two-phase medium comprising aggregate and cement mortar when the strength of the ITZ was at least
70 % that of the cement mortar [38], and the ITZ could no longer be considered the weakest region.
Testing the fracture properties of the ITZ is significantly more challenging than testing aggregates and cement mortar. Various
approaches have been proposed for measuring the strength properties of ITZ. Some researchers [75] designed notched rock-mortar
specimens by casting fresh mortar on the rock surface and found that the tensile strength of the interface between the rock and
mortar was in the range of 1/16–1/3.5 times that of the mortar. Jebli et al. [203] presented a numerical identification strategy for the
tangential cohesive response (in mode II). Experimental tests on parallelepipedic samples at the classical aggregate scale were con­
ducted at various stages of hydration using specific devices for direct tensile and shear testing of the composites. The results indicated
that the tensile strength at the cement-aggregate interface was approximately 30 % lower than that of the cement paste; similarly, the
shear strength of the interface was lower than that of the cement paste. Similarly, many bonding strength tests on the ITZ were
conducted in previous investigations [76,199], and it was found that the bonding strength of the ITZ was approximately 50 % of the
tensile strength of the mortar. In addition, artificial neural networks (ANN) can be utilized [189] to predict the ITZ tensile strength and
fracture energy. The results showed that the ratio of ITZ tensile strength to mortar tensile strength was about 0.46, and the fracture
energy of ITZ was about 0.4 times that of the mortar.

3. Factors affecting on the physical and mechanical properties of ITZ

The mechanical properties and microstructure of the ITZ can be influenced by a variety of factors, such as the type, surface
roughness, size distribution of aggregates, size of sand particles, cement fineness, mineral additions, curing conditions, and w/c ratio
[9,11,19,31,67,114,204,205]. Understanding the variations in the ITZ properties resulting from these factors is crucial for investi­
gating the mechanical behavior of concrete. This section systematically summarises and analyses various factors that affect the
physical features and properties of the ITZ and identifies the influencing mechanisms. These findings are helpful for identifying
methods for reducing the formation of the ITZ and compacting the ITZ structure, which is essential for improving the mechanical
properties of concrete structures.

3.1. Mineral addition

The efficacy of cement in concrete has been found to be limited, with only 60–70 % of the cement effectively utilized in engineering

14
Q. Chen et al. Engineering Fracture Mechanics 300 (2024) 109979

Fig. 9. The nominal compressive elastic modulus (ENC ) and the morphology of samples of ITZ with limestone and granite aggregates with water-
cement of 0.23, 0.35 and 0.53 [29].

15
Q. Chen et al. Engineering Fracture Mechanics 300 (2024) 109979

experiments. The remaining cement acts as a filler and does not contribute to the strength of concrete, resulting in the wastage of raw
materials. The addition of active mineral admixtures is a practical and effective method to improve the mechanical properties of
concrete and achieve greater economic efficiency. These admixtures can also reduce the friction resistance between the mixing ma­
terials, resulting in more uniform concrete components. The presence of mineral additives in concrete reduces its overall porosity and
improves its structural density and strength. The microstructure of the ITZ can be strengthened by optimizing the particle size dis­
tribution of cementitious materials, resulting in reduced porosity and a lower w/c ratio. This is typically achieved through the use of
mineral additions such as natural pulverized limestone and metakaolin, as well as industrial materials such as granulated blaster
furnace slag, fly ash, and silica fume. These particles are much finer than cement particles and contribute to particle compaction,
leading to the densification of the ITZ [18,206–208]. This approach is commonly employed for high-strength concrete [9,81].
According to published researches, the addition of silica fume is an effective method for enhancing the strength of concrete by
improving the ITZ [12,209]. Some previous studies have shown that adding 10 % silica fume reduces the ITZ thickness from 50 μm to
10 μm in comparison to concrete without mineral addition, which leads to a 100 % increase in interfacial bond strength and fracture
energy [12]. Moreover, pozzolan is widely accepted as a partial replacement for Portland cement in mortar and concrete production
[210]. The experimental results showed that the addition of mineral admixtures resulted in a denser ITZ and an optimized pore
structure, especially at later curing stages. Some workers [211] employed digital image analysis to investigate the fracture behavior of
concrete with fly ash as an admixture under shear conditions and concluded that the addition of fly ash significantly enhanced the
crack resistance of the concrete.

3.2. Aggregates

Many studies have suggested that aggregates play a significant role in ITZ formation. Physical interactions between the aggregate
grains and mortar can have a pronounced effect on the development of the ITZ microstructure. For instance, the thickness of the water
films that form surrounding coarse aggregates can be reduced owing to their higher water absorption, leading to an improved ITZ
microstructure.

3.2.1. Type of aggregates


The strength of concrete at the macroscopic level mainly depends on the quality of the cement mortar and the surface charac­
teristics of the aggregates. However, as the quality of the cement mortar improves, the influence of the type and surface texture of the
aggregates on the concrete strength becomes more prominent. Previous studies have examined the effects of the aggregate type on the
macroscopic mechanical properties of concrete, indicating that the elastic modulus of concrete increased with the aggregate stiffness
and strength [15]. Studies have also demonstrated that the bonding strength of the ITZ is closely related to the type and surface
roughness of the aggregates [19,212], and that the ITZ properties are significantly influenced by both the size and surface topography
of the aggregates [14,67].
Extensive research has been conducted to determine the effects of aggregates with varying shapes and surface textures on the
physical features of the ITZ [11,19,26,213]. The physical interactions between mortar and aggregates are primarily determined by the
surface texture and water absorption of the aggregates. Irregular and rough aggregates tend to impede crack initiation and propagation
more strongly than smooth and rounded aggregates [19]. The influence of the aggregate surface texture on the ITZ properties is similar
to that of the wall effects, where anhydrous cement particles cannot pack as tightly near the smooth areas of aggregates with convex
and concave surfaces, leading to a higher porosity of the ITZ microstructure. As described in Section 2.2, the most significant variation
in porosity is observed at a distance of 10–40 μm from the aggregate surface, and the surface texture obviously influences the ITZ
properties at distances greater than 10 μm. Furthermore, research has suggested that the mineral composition and surface texture of
aggregates influence the nucleation and growth of hydration products, resulting in variations in the mechanical properties of the ITZ
[57].
The effect of aggregate features on the formation and properties of the ITZ was investigated using the AC impedance technique with
three types of coarse aggregates: limestone, granite, and basalt [79]. This research reveals that the surface texture and water ab­
sorption of aggregates play crucial roles in the formation of ITZ microstructure. Specifically, the ITZ microstructure around the
limestone aggregates at an early age was denser, owing to its rougher surface and higher water absorption. By contrast, the ITZ
microstructure around the granite and basalt aggregates became denser because of the higher chemical activity between the aggregates
and mortar.
Several studies have investigated the mechanical properties and morphology of the ITZ by analyzing the effects of two different
types of aggregates (limestone and granite) and three w/c ratios [29]. As shown in Fig. 9, the rougher surface of the limestone ag­
gregates enhanced the bonding strength between the aggregates and the mortar. Conversely, the smoother and flatter surface of the
granite aggregates reduced the constraint on mortar shrinkage deformation, resulting in a higher probability of crack generation.
Additionally, at various w/c ratios, the ITZ thickness around the limestone aggregates was found to be relatively greater, and the
nominal compressive elastic modulus was lower than that around the granite aggregates. It can be concluded that the ITZ formed by
limestone aggregates exhibited a greater tendency towards deformation than that formed by granite aggregates. Moreover, some
scholars have investigated the effects of the ITZ mechanical properties on the macro-mechanical behavior and impact response of
concrete using gravel aggregates and copper slag [63]. The authors observed significant differences in the surface roughness and
stiffness between the two types of aggregates, resulting in distinct local stress concentrations within the ITZ and varying impact re­
sponses, as shown in Fig. 10. Specifically, the ITZ with rougher aggregates exhibited greater energy dissipation inside the mixture,
leading to a higher bearing capacity of the concrete and stronger impact resistance at the macro level. These findings demonstrate the

16
Q. Chen et al. Engineering Fracture Mechanics 300 (2024) 109979

important role of the ITZ properties in determining the impact response of concrete, which suggests that the mechanical characteristics
of the ITZ can be optimized to enhance the overall performance of concrete under impact loading.
By improving the physical interactions between the close contact points, the increased surface roughness of the aggregates can
improve the interfacial strength against frictional resistance and increase the required area at fracture within the ITZ regions.
Consequently, concrete with rougher aggregates has a greater load-bearing capacity and requires more energy for the final failure.

3.2.2. Diameter and size distribution of aggregates


The influence of the aggregate size on the ITZ thickness is illustrated in Section 2.3.1. Earlier studies [70,71] suggested that the ITZ
thickness is primarily dominated by the median diameter of the cement particles rather than by the aggregate size [134]. Through
experimental examination using the electron probe technique, a statistical analysis indicated that the ITZ thickness around aggregates
of varying sizes typically followed a normal distribution. In addition, thicker ITZ layers surrounded larger aggregates, and thinner ITZ
surrounded smaller aggregates. Some studies have also indicated that the ITZ thickness differs slightly for different areas around the
same aggregate [26].
The influence of the maximum size of the aggregate on the ITZ fraction has also been explored in previous studies [72], which was
evaluated using the maximum equivalent diameter of the aggregates. Fig. 11(a) illustrates that the ITZ volumes decrease mono­
tonically with an increase in the maximum equivalent diameter of the aggregates for a given aggregates volume. This finding is also
consistent with the findings of previous studies on spheroidal particles presented in the reference [70]. This can be explained by the
fact that an increase in the maximum equivalent diameter of the aggregates decreased the fineness of the particles with a certain
aggregate content, resulting in a reduction in the aggregate surface area per unit volume in concrete.
Additionally, Fig. 11(b) shows the influence of the aggregate gradation on the ITZ volume fractions. For a certain aggregate
content, the ITZ volumes increase in the following order: uniform gradation < four-gradation < three-gradation < two-gradation <
one-gradation, where the uniform gradation is characterized by the Fuller distribution. The ITZ volumes are significantly influenced by
the aggregate gradation, and some studies have obtained similar conclusions through a sensitivity analysis of various affecting factors
[70,71]. Fig. 11(c) presents the influence of the aggregate volume fraction of on the ITZ volumes with varying maximum equivalent
diameters of aggregates Dmax for the Fuller gradation and equal volume fraction (EVF) gradation from references [70,71], where the
ITZ thickness of tITZ and the minimum aggregate size of Dmin are assumed as tITZ = 30mm, Dmin = 0.15mm[71] and tITZ = 0.03mm,
Dmin = 0.15mm[70], respectively. The findings suggest that the ITZ volumes of VITZ increase within a specific range as the aggregate
fraction of fagg increases, but decrease when fagg exceeds a critical value, as demonstrated in [71]. This can be attributed to the dense
stacking of aggregates caused by the excessive addition, which squeezes the space between the aggregates and leads to a high overlap
of the ITZ regions of adjacent aggregates. For the Fuller gradation, VITZ decreased significantly with increasing Dmax for a certain ITZ

Fig. 10. Load-time histories of the specimens with two types of aggregates during the first drop of impact [63].

17
Q. Chen et al. Engineering Fracture Mechanics 300 (2024) 109979

Fig. 11. The fraction of ITZs versus (a) the maximum size [72], (b) the gradation of aggregates [72] and (c) the fraction of aggregates [70,71].

thickness [71]. Because of the large number of small aggregates contained in the EVF gradation, the ITZ volumes for the EVF gradation
were always greater than those for the Fuller gradation with Vagg , Dmax , and tITZ in concrete. This conclusion is consistent with previous
research [71]. This highlights the significant influence of aggregate gradation on ITZ volumes. Therefore, optimizing the gradation of
aggregates in engineering structures to approximate the Fuller distribution should be prioritized to ensure an ideal distribution of
aggregates.
Comparing different studies [70,71], it is also noted that the VITZ of Sun et al. [71] is higher than that of Zheng et al. [70] for
concrete with an approximate aggregate size for a given Vagg . This could be attributed to the different values of the ITZ thickness, and
tITZ was set to 30μ m and 0.03 mm in Refs. [70,71], which directly causes a gap of the ITZ volume. According to the analysis presented
above, the ITZ thickness is the most important factor influencing the fraction of the ITZ for Fuller gradation, followed by the volume
fraction of aggregates. The maximum aggregate diameter had the least effect.
The influence of the aggregate size on the failure performance of concrete can also be understood by evaluating its effects on the
mechanical properties of the ITZ, which were performed using a single spherical aggregate in previous studies [77]. The effects of the
aggregate size on the compressive and tensile strengths of concrete are shown in Fig. 12. The experimental results indicated that the
compressive strength of concrete increased gradually with increasing single aggregate size, particularly in medium- and high-strength
concrete, owing to the comparatively greater compatibility between the aggregates and cement. Conversely, the tensile strength of
concrete decreased significantly with an increase in aggregate size, particularly in high-strength composites. This could be explained
by the fact that increasing the aggregate size strengthened the significant difference between the elastic moduli of these two physical
phases, resulting in an increased stress concentration and more microcracks in the ITZ. Consequently, the bonding strength of the ITZ
was found to be the primary determinant of the tensile strength of concrete, with negligible effects on the compressive strength.

3.3. Water-cement ratio

The water-cement ratio (w/c) is a key parameter that significantly influences the quality of the cement matrix and ITZ, which is
crucial for determining the strength and durability of concrete. Königsberger et al. [54] used continuum micromechanics to downscale

Fig. 12. Effects of aggregate size on (a) compressive strength and (b) tensile strength of low strength, medium-strength and high-strength concrete
specimens [77].

18
Q. Chen et al. Engineering Fracture Mechanics 300 (2024) 109979

concrete-level loading to maximum stresses in the ITZ and micron-sized hydrates, applying a Drucker-Prager failure criterion for model
validation. The results revealed that the concrete strength magnitude was governed by the ITZ stress concentration, and that the w/c
ratio was the dominant mixture design parameter. Numerous studies have explored the influence of w/c ratio on the thickness,
microstructure, and mechanical characteristics of the ITZ. It is widely acknowledged that increasing the w/c ratio leads to an increased
porosity in both the cement and ITZ phases [69]. Conversely, reducing the w/c ratio is an effective way to reduce the porosity of the
cement and ITZ phases, thereby improving the ITZ microstructure [69]. In contrast to the pore structure in mortar, the higher water
content and lower filling density of cement particles in the ITZ result in a higher porosity, which significantly influences the overall
properties of the ITZ [11]. As discussed in Section 2.2, it is widely accepted that decreasing the w/c ratio can reduce the ITZ thickness,
thereby enhancing the concrete strength and promoting denser microstructures.
As shown in Fig. 9 and described in Section 3.2.1, the influence of various w/c ratios on the concrete morphology was investigated.
As the w/c increased, the mortar surface became smoother, with fewer pores and more visible microcracks. Moreover, the increase in
w/c contributes to more microcracks and a reduction in the average ENC of the ITZ, which is caused by the shrinkage deformation of the
mortar owing to the higher water content [29]. The physical features could be compared at various w/c ratios with the same aggregate
type, and the correlation between ENC and the distance from the aggregates is illustrated in Fig. 13. It has been noticed that with the
increasing w/c, the ITZ thickness with granite aggregates decreases gradually, but the ITZ thickness with limestone aggregates first
increases and then decreases. The distribution of ENC is consistent with that of previous studies [14,16]. As the ITZ area approached the
aggregates and mortar, the ENC of the ITZ gradually approached the elastic moduli of these two materials.
To study the physical features and evolution of the ITZ, a multi-aggregate method was proposed, considering the effects of w/c
ratios and other practical factors [131]. The results indicated that the porosity decreased more rapidly with increasing distance from
the aggregate surface for lower w/c ratios, resulting in a reduced ITZ thickness. A linear function with a reasonably good correlation
coefficient can be employed to establish a quantitative relationship between the ITZ thickness and logarithmic w/c for a specific paste
thickness. A similar qualitative description of the relationship between the w/c and ITZ can be found in Ref. [67].
Because of the difference in the initial water content, the cement grains within the ITZ regions and bulk cement experience different
w/c ratios, necessitating consideration of the significant influence of the local w/c on the ITZ microstructure [13]. Considering the
influence of the w/c on the physical features of the ITZ, researchers have proposed two equations to fit the ITZ thickness and porosity
[5]. As illustrated in Section 2.2 (Fig. 2), the porosity of the ITZ increased gradually, and the ITZ microstructure appeared to loosen
with increasing w/c ratio. A higher w/c ratio could lead to a highly hydrated binder system and greater initial porosity, both of which
cause a deviation in the linear relationship between the porosity and w/c ratio. Nevertheless, there is no significant deviation from the
linear relationship between porosity and w/c ratio in the region where the distance from the aggregate surface is less than 15 μm.

3.4. Environmental factors

Environmental factors, such as temperature and humidity, have a significant impact on the mechanical behavior and integrity of
the Interfacial Transition Zone (ITZ) in concrete. Exposure to high temperatures causes the concrete material to degrade, and
microcracks initiate and grow owing to moisture loss, which provokes different physicochemical processes [214,215], such as the
disintegration of hydrated cement, shrinkage of cement, and thermal incompatibilities between the mortar and the aggregate. Many
studies indicated that the cement and ITZ are contractive between 230 ◦ C and 300 ◦ C, corresponding to the dehydration of these
phases, while the aggregates are always expansive for a temperature range of 20 ◦ C to 800 ◦ C [214,215]. Extensive studies [216–218]
have indicated that thermal incompatibility stress is generated at the ITZ between the mortar and aggregate owing to the difference in

Fig. 13. Distribution of the nominal compressive elastic modulus (ENC ) of concrete specimens at various w/c ratios with (a) Limestone aggregates
and (b) Granite aggregates [29].

19
Q. Chen et al. Engineering Fracture Mechanics 300 (2024) 109979

their thermal expansion coefficients between aggregate and mortar. This means that coarse aggregates with lower coefficients of
thermal expansion and high moduli, which typically have higher coefficients of thermal expansion, restrain the volume changes of the
ITZ and mortar under identical temperature variations, which typically have higher coefficients of thermal expansion [219]. This
restraint induces a tensile stress in the ITZ tangential to the coarse aggregate surface, resulting in microcracks in the ITZ. When the
microcracks in the ITZ accumulate to a certain degree, they may extend to the mortar and aggregate.
During heating, the ITZ is weakened owing to the dehydration of the cement paste and the decomposition of portlandite, coupled
with an increase in thermal strain due to aggregate expansion [220]. And due to the increasing thermal expansion difference and the
deterioration of the cement mortar with the increasing temperature, the damage to the concrete increased sharply after 500 ◦ C [221].
As revealed by the studies conducted by Shen et al [222], during the heating process, C-S-H exhibits shrinkage due to moisture loss,
which is an irreversible process; that is, it does not regain its original volume upon cooling, whereas aggregates expand during the
heating process and may occupy the space created by the shrinkage of the cement paste. Although the thermal expansion of aggregates
is typically reversible, their contraction during cooling is not accompanied by the cement paste regaining its original volume, leading
to a gap at the interface.
Nadeem et al. [223] examined the microstructure of the ITZ in concrete composed of crushed granite aggregates after exposure to
normal and elevated temperatures by quantitative analysis of SEM images. It is observed that the pore area fraction in the ITZ in
concrete continuously increased with the increasing temperature (27 ◦ C–800 ◦ C), which means that the degradation of the micro­
structure resulted in compressive strength and durability losses at elevated temperatures. Nadeem et al. [224] described the physical
condition of ITZ at three temperature ranges, namely: low-range temperatures (27 ◦ C–200 ◦ C), medium range temperatures
(200 ◦ C–400 ◦ C) and high-range temperatures (400 ◦ C–800 ◦ C). Different types of flaws were identified in the ITZ, including texture,
orientation, and local flaws, with a notable increase in the fractions of orientation and local flaws with increasing temperature. With
increasing temperatures, the physical characteristics of the ITZ gradually changed from a discrete or discontinuous flaw zone at normal
or mildly elevated temperatures to a continuous and highly porous flaw zone at elevated temperatures, indicating that this shift was
responsible for the loss of the mechanical and durability properties of concrete.
Extensive studies have shown that the degradation of mechanical performance is generally related to ITZ microcracks under
conditions of sharp temperature variations. Kanellopoulos et al. [225] proposed that the mechanical performance of high-performance
concrete decreased mainly due to the development and propagation of ITZ cracks after 90 thermal cycles (20 ◦ C–90 ◦ C). Huang et al.
[226] selected samples from three different environmental thermal fatigue cycles (0, 150, and 330) for BSE analysis to verify the effect
of environmental thermal fatigue on the ITZ microstructure. It can be seen that microcracks were mainly located in the ITZ during
environmental thermal fatigue cycling, and the area percentage of ITZ microcracks was over 70 %. This is attributed to the ITZ having a
higher porosity and lower elastic modulus than the matrix. Ye et al. [227] conducted the BSE image analysis of the mortar sample after
the thermal cycle in a temperature range of 20 ◦ C–60 ◦ C. The microstructure observed by SEM indicated that the ITZ and hardened
cement paste were the weak regions in the mortar sample and were susceptible to temperature variations. An increase in the maximum
cycling temperature induces deterioration of the ITZ and results in an expansion of the cracks located in the ITZ. On the other hand,
more microcracks will be generated on the hardened cement paste surface as the maximum cycling temperature increases, which may
finally connect together and form a relatively large crack. The deterioration of the ITZ becomes more significant after the thermal cycle
in a temperature range of 20 ◦ C–70 ◦ C. In addition to cracks located around the ITZ, microcracks were easily observed on the hardened
cement paste surface.
However, different concrete deterioration mechanisms were observed under low- and high-temperature conditions. Ziaei-Nia et al.
[228] found that freeze–thaw cycles under normal conditions can cause concrete damage owing to relatively large differences between
the aggregate and cement paste stresses and ITZ stresses. Yang et al. [229,230] investigated the evolution process of concrete ITZ under
the coupling function of fatigue load and freeze–thaw and evaluated the microstructure characteristics of concrete in different tem­
perature areas, including low temperature (from –22 ◦ C to − 10 ◦ C) and high temperature (32 ◦ C–40 ◦ C). The results indicated that the
ITZ structure deteriorated considerably under the coupling effect of fatigue load and freeze–thaw, such that the micro-hardness was
reduced significantly and the width of the ITZ increased. Remarkably, the deterioration phenomenon at low temperatures was more
significant than that at high temperatures.
In addition to being affected by factors such as temperature and freeze–thaw cycles, the mechanical properties of ITZ are also
affected by the combined effects of humidity and temperature, which are typically reflected in different curing methods [231,232].
Under standard curing, the ITZ of the concrete is closely connected to the matrix without evident micro-cracks. In contrast, the ITZ in
concrete under autoclaved curing is clearly different from that observed under standard curing and steam curing, primarily due to a
notable alteration in the composition of the concrete hydration products induced by autoclaved curing. The microstructure of ITZ
under autoclaved curing is denser than that under standard curing [233]. Zhang et al. [234] found that autoclaved curing facilitated
the consumption of CH crystals in the ITZ by secondary hydration, resulting in the formation of relatively dense tobermorite and thus a
more densified the ITZ. Both steam curing and autoclaved curing enhance the microhardness of ITZ compared to standard curing.

3.5. Other factors

According to various investigations of ITZ based on both experimental and numerical methods, the physical features and me­
chanical properties of ITZ could also be affected by other variables, including the curing age and sand grain size, except for the
affecting factors presented above.
As presented in Sections 2.2, and 2.3, it was demonstrated that the curing age has a significant effect on both the thickness and
porosity of the ITZ. In the initial stage of hydration, cement particles exposed to water start the hydration process, resulting in

20
Q. Chen et al. Engineering Fracture Mechanics 300 (2024) 109979

formation of initial hydration products. Due to the uneven contact between cement particles and the aggregate surface, the micro­
structure with greater porosity is formed within ITZ. As hydration progresses to around 28 days, the continuous accumulation of
hydration products incrementally fills the pores within ITZ, resulting in a progressively denser structure. At this stage, the micro­
structure and mechanical properties of ITZ gradually approach those of the surrounding cement paste. Nonetheless, the ITZ remains,
for an extended duration, the most vulnerable segment within the concrete, exhibiting relatively elevated porosity and permeability.
However, the ITZ remains the weakest area in concrete, with a relatively higher porosity and permeability. Subsequently, the concrete
enters a stable hydration phase, which lasts for several months or even years. In this prolonged stage, the hydration reaction becomes
very slow, the concrete continues to gain strength, but at a much slower rate. A longer curing age led to a higher degree of hydration,
which significantly reduced the porosity within the ITZ, owing to the filling action of the hydration products.
There have been several long-term studies on concrete [23,106,130,210,235], investigating properties such as strength, durability,
and permeability over periods ranging from months to multiple decades, often focusing on the characteristics of ITZ in relation to
aging. Some researchers developed advanced computational models and simulations to predict the evolution of the ITZ. These models
can simulate the hydration processes and microstructural development over extended periods. For example, Gao et al. [13] employed a
continuum integrated model HYMOSTRUC to simulate the composite microstructure of aggregate-ITZ-bulk paste in concrete, focusing
on both the initial state and after 56 days of curing. Some studies [80] regarded the model structure at 100-hour hydration as
representative of mature concrete, and the computer simulation revealed that gradient structures were qualitatively maintained during
maturation. To simulate the acting process of cement hydration, the modified status-oriented computer model by Ma et al. [131]
indicated that the differences in porosities between ITZ and paste gradually decreased with the hydration process and nearly remained
constant at mature ages, which was supported by experimental findings [9]. Furthermore, the obtained simulation results also indi­
cated that the thickness of the ITZ was a function of the curing age, and the thickness would decreased or even disappeared with
increasing curing age, significantly decreasing from 7d to 28d [43]. Additionally, some experimental studies have been conducted to
reveal the correlation between the aging of concrete and physical characteristics of ITZ. Considering that the hydration products of the
concrete changed at different ages, Liao et al. [106] selected various ages of 1, 7, 14, 28 and 56 days to analyze the interfacial
characteristics of paste and aggregate and discuss the relationship between macroscopic properties and microstructures based on the
SEM microstructural analysis. Through the investigation into the development of hydration products within transition zone for
different ages, it was found that the density of the hydration products increased as the curing age increased, and the density of hy­
dration products increased as the distance from the aggregate edge increased at any curing age.
The influence of the cement particle size on the ITZ thickness in conventional concrete has been illustrated in detail in previous
studies [66], as mentioned in Section 2.3.1. Using a vector-based microstructural modelling platform, Sharma et al. [40] conducted a
parametric study to investigate the effects of various factors, including the w/c and particle size of cement and aggregates, on the
thickness and microstructure of the ITZ. This study also concluded that the ITZ thickness was mainly determined by the size of the
cement grains rather than that of aggregates, which is consistent with the results obtained in other studies [155]. The porosity dis­
tributions along the direction away from the aggregate boundary for different sand grades of mortar were evaluated [42], as shown in
Fig. 14. The results revealed that finer sand resulted in a thinner ITZ.
Research on the ITZ is crucial for understanding its characteristics such as porosity, thickness, volume, and chemical composition.
These factors are essential designing concrete structures that can withstand harsh environments like marine settings, chemical
exposure, or extreme temperatures. For instance, in areas with high acidity or alkalinity, such as near marine settings or areas where
deicing salts are frequently used, a denser ITZ with lower porosity is beneficial. This structure impedes the penetration of acidic or
alkaline solutions and significantly slows down the penetration of chloride. Thereby, optimizing the porosity of ITZ plays a pivotal role
in protecting both the concrete and its steel reinforcements from corrosive damage. Additionally, in colder regions, the ITZ in concrete
is more vulnerable to damage during freeze–thaw cycles due to its higher porosity and weaker microstructure. To counter this, various

Fig. 14. Comparison of porosity distributions for all mortars made with various sand sizes along the direction from aggregate surface to paste [42].

21
Q. Chen et al. Engineering Fracture Mechanics 300 (2024) 109979

mineral admixtures are commonly employed to enhance the pore structure of ITZ by chemically reacting with calcium hydroxide
released during cement hydration, resulting in the formation of additional cementitious compounds. Such enhancements result in a
denser ITZ microstructure with smaller, more evenly distributed pores, which robustly resists freeze–thaw cycles, thus markedly
reducing the risk of frost damage. In short, research on the ITZ of concrete is pivotal in enhancing the performance and durability of
concrete structures, particularly in environments that pose specific challenges like extreme temperatures or chemical exposure.

4. Conclusions and perspectives

This paper provides a comprehensive overview of the current state of knowledge on the physical features and mechanical prop­
erties of the ITZ, such as the porosity, thickness, volume, and elastic modulus, which are quantitatively characterized based on a
summary and comparison of extensive literature using theoretical, experimental, and numerical methods. The factors influencing the
physical and mechanical properties of the ITZ are evaluated, as illustrated in Fig. 15. These factors include the mineral addition, grain
size distribution of cement, w/c ratio, aggregate content, diameter and size distribution of aggregates, type of aggregate, and curing
age. These can be classified into direct and indirect factors according to their influence on the microstructure and total volume of the
ITZ, as described in Section 1. Based on a summary and comparison of large quantities of work, the following conclusions can be
drawn.

(1) Accurate characterization of the physical features of the ITZ is the premise for implementing reliable numerical simulations,
especially with regard to its thickness and elastic modulus. These properties provide valuable insights for refining ITZ modelling
in concrete for subsequent investigations. However, in most existing studies, these two types of physical features of the ITZ are
simply determined as constant parameters, which may make it difficult to fully reflect the influence of the variability and
heterogeneity of the ITZ on the macro-mechanical behavior and failure mechanism of concrete. Therefore, introducing more
advanced numerical models that incorporate the gradient distribution of the physical features of the ITZ to simulate the damage
accumulation process in concrete accurately may be a meaningful direction in the future. Specifically, ITZ thickness should be
determined based on the aggregate particle size, with larger particles resulting in a proportionately thicker ITZ that statistically
follows a normal distribution pattern. Correspondingly, the elastic modulus of the ITZ should follow a gradient distribution as a
quadratic function of distance from the aggregate boundary.
(2) Although analytical models of concrete considering the existence of the ITZ have been extensively studied, most have focused on
estimating its elastic modulus and volume fraction by evaluating the elastic modulus of the cement matrix and the aggregate
volume fraction. However, the uneven distribution of stress field and strain gradient within the ITZ region, owing to its ir­
regularity and heterogeneity, is still not well explained. Therefore, the development of an analytical model that considers
uneven stress fields within the ITZ to predict the varying bonding strength remains an important topic for future research.
(3) The microstructure and physical features of the ITZ are significantly influenced by multiple factors. Among these factors, the w/
c ratio, curing age, cement fineness, surface texture of aggregates, and mineral addition directly affect the morphological
features and microstructure of the ITZ by influencing the formation and distribution of various unhydrated and hydrated
products during the hydration process. The size distribution and content of the aggregates do not significantly affect the for­
mation and microstructure of the ITZ, but they do have a significant influence on the ITZ volume. The ITZ volume fraction is
primarily determined by the ITZ thickness, with the aggregate content being the secondary factor. For a given volume of ag­
gregates and ITZ thickness, the ITZ volumes decrease monotonically with increasing maximum equivalent diameter of the
aggregates.
(4) Improving the ITZ microstructure is a feasible strategy for enhancing the bonding strength and macro-mechanical performance
of concrete. Improvements in the properties of the ITZ can be achieved through the addition of mineral admixtures, increasing
the curing age, and adjusting the w/c ratio and sand fineness. These approaches aim to enhance the properties of the ITZ,
particularly its porosity and elastic modulus, by increasing the adhesion between the paste and aggregates and filling the
micropores within the cement matrix, thereby improving its structural density and mechanical properties. Although rougher
aggregates may lead to higher ITZ porosity, increased surface roughness can improve the interfacial strength against frictional
resistance and increase the required fracture area within the ITZ regions. Moreover, the actual size distribution of aggregates
should be optimized to approach the Fuller distribution as soon as possible in the design of concrete engineering.

Additionally, considering the booming advancements in deep learning and data mining, tremendous progress has been made in
recent decades in selecting constituent contents as key features for model inputs to predict a given set of target properties of concrete.
Through the quick extraction of valuable information, neural network models exhibit significant advantages in establishing quanti­
tative relationships between microstructures and the macro-physical and mechanical behaviors of the ITZ. Future research could be
conducted using a neural network model. (1) Quickly identify the ITZ phase in concrete and determine its width by analyzing the ITZ
microstructure using SEM-BSE image. (2) Establish a functional relationship between the mechanical properties of the ITZ and its
components. (3) Based on a comprehensive database of experiments and meso-scale simulations of concrete, the macro mechanical
properties and fracture performance of the ITZ can be accurately predicted by processing a set of initial inputs including the mix
compositions and curing conditions.

22
Q. Chen et al. Engineering Fracture Mechanics 300 (2024) 109979

Fig. 15. The relationship between various affecting factors and the multiscale characterization of ITZ.

CRediT authorship contribution statement

Qingqing Chen: Data curation, Writingoriginal draft, Investigation, Resources. Jie Zhang: Conceptualization, Funding acquisi­
tion, Writing-review & editing, Supervision. Zhiyong Wang: Conceptualization, Writing-review & editing. Tingting Zhao: Writing-
review & editing, Resources. Zhihua Wang: Writing-review & editing, Supervision.

Declaration of competing interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to
influence the work reported in this paper.

Data availability

Data will be made available on request.

Acknowledgements

This work is supported by the National Natural Science Foundation of China (Grant Nos. 12102292, 12102294, 12272257), the
special fund for Science and Technology Innovation Teams of Shanxi Province (Nos. 202204051002006). The financial contributions
are gratefully acknowledged.

References

[1] Xi X, Yin ZQ, Yang ST, Li CQ. Using artificial neural network to predict the fracture properties of the interfacial transition zone of concrete at the meso-scale.
Engng Fract Mech 2021;242:107488. https://doi.org/10.1016/j.engfracmech.2020.107488.
[2] Mazzucco G, Pomaro B, Xotta G, Garbin E, Majorana CE, Marchi NDe, Concheri G.. Meso-scale XCT-based modeling of ordinary concrete. Constr Build Mater
2021;286:122850. https://doi.org/10.1016/j.conbuildmat.2021.122850.
[3] Sicat E, Gong FY, Ueda T, Zhang DW. Experimental investigation of the deformational behavior of the interfacial transition zone (ITZ) in concrete during
freezing and thawing cycles. Constr Build Mater 2014;65:122–31. https://doi.org/10.1016/j.conbuildmat.2014.04.035.
[4] Liu YF, Zhang J, Zhao TT, Wang ZY, Wang ZH. Reconstruction of the meso-scale concrete model using a deep convolutional generative adversarial network
(DCGAN). Constr Build Mater 2023;370:130704. https://doi.org/10.1016/j.conbuildmat.2023.130704.
[5] Gao Y, De Schutter G, Ye G, Tan ZJ, Wu K. The ITZ microstructure, thickness and porosity in blended cementitious composite: effects of curing age, water to
binder ratio and aggregate content. Compos B Engng 2014;60:1–13. https://doi.org/10.1016/j.compositesb.2013.12.021.
[6] Wu K, Shi HS, Xu LL, Ye G, De Schutter G. Microstructural characterization of ITZ in blended cement concretes and its relation to transport properties. Cem
Concr Res 2016;79:243–56. https://doi.org/10.1016/j.cemconres.2015.09.018.
[7] Shi Y, Lv XD, Zhou SH, Liu ZA, Yang MC, Liu CF, et al. Mechanical properties, durability, and ITZ characteristics of full-grade dam concrete prepared by
aggregates with surface rust stains. Constr Build Mater 2021;305:124798. https://doi.org/10.1016/j.conbuildmat.2021.124798.

23
Q. Chen et al. Engineering Fracture Mechanics 300 (2024) 109979

[8] Farran J. Introduction the transition zone- discovery and development. In: Maso JC, editor. Interfacial Transition Zone in Concrete, RILEM Report 11. London,
UK: E&FN Spon; 1996. xiii-xvi.
[9] Scrivener KL, Crumbie AK, Laugesen P. The interfacial transition zone (ITZ) between cement paste and aggregate in concrete. Interface Sci 2004;12(4):411–21.
https://doi.org/10.1023/B:INTS.0000042339.92990.4c.
[10] Scrivener KL. Backscattered electron imaging of cementitious microstructures: understanding and quantification. Cem Concr Compos 2004;26:935–45.
https://doi.org/10.1016/j.cemconcomp.2004.02.029.
[11] Scrivener KL, Nemati KM. The percolation of pore space in the cement paste aggregate interfacial zone of concrete. Cem Concr Res 1996;26(1):35–40. https://
doi.org/10.1016/0008-8846(95)00185-9.
[12] Rossignolo JA. Interfacial interactions in concretes with silica fume and SBR latex. Constr Build Mater 2009;23:817–21. https://doi.org/10.1016/j.
conbuildmat.2008.03.005.
[13] Gao Y, De Schutter G, Ye G, Huang HL, Tan ZJ, Wu K. Characterization of ITZ in ternary blended cementitious composites: experiment and simulation. Constr
Build Mater 2013;41:742–50. https://doi.org/10.1016/j.conbuildmat.2012.12.051.
[14] Xiao J, Li W, Sun Z, Lange DA, Shah SP. Properties of interfacial transition zones in recycled aggregate concrete tested by nanoindentation. Cem Concr Compos
2013;37:276–92. https://doi.org/10.1016/j.cemconcomp.2013.01.006.
[15] Ramesh G, Sotelino ED, Chen WF. Effect of transition zone on elastic moduli of concrete materials. Cem Concr Res 1996;26:611–22. https://doi.org/10.1016/
0008-8846(96)00016-6.
[16] Xiao JZ, Li WG, Corr DJ, Shah SP. Effects of interfacial transition zones on the stress-strain behavior of modeled recycled aggregate concrete. Cem Concr Res
2013;52:82–99. https://doi.org/10.1016/j.cemconres.2013.05.004.
[17] Li GQ, Zhao Y, Pang SS. Four-phase sphere modeling of effective bulk modulus of concrete. Cem Concr Res 1999;29:839–45. https://doi.org/10.1016/S0008-
8846(99)00040-X.
[18] Winslow DN, Cohen MD, Bentz DP, Snyder KA, Garboczi EJ. Percolation and pore structure in mortars and concrete. Cem Concr Res 1994;24(1):25–37.
https://doi.org/10.1016/0008-8846(94)90079-5.
[19] Prokopski G, Halbiniak J. Interfacial transition zone in cementitious materials. Cem Concr Res 2000;30:579–83. https://doi.org/10.1016/S0008-8846(00)
00210-6.
[20] Sorelli L, Constantinides G, Ulm FJ, Toutlemonde F. The nano-mechanical signature of ultra high performance concrete by statistical nanoindentation
techniques. Cem Concr Res 2008;38:1447–56. https://doi.org/10.1016/j.cemconres.2008.09.002.
[21] Wang XH, Jacobsen S, He JY, Zhang ZL, Lee SF, Lein HL. Application of nanoindentation testing to study of the interfacial transition zone in steel fiber
reinforced mortar. Cem Concr Res 2009;39:701–15. https://doi.org/10.1016/j.cemconres.2009.05.002.
[22] Barnes BD, Diamond S, Dolch WL. Hollow shell hydration of cement particles in bulk cement paste. Cem Concr Res 1978;8(3):263–72. https://doi.org/
10.1016/0008-8846(78)90095-9.
[23] Diamond S, Huang JD. The ITZ in concrete-a different view based on image analysis and SEM observations. Cem Concr Compos 2001;23(2–3):179–88. https://
doi.org/10.1016/S0958-9465(00)00065-2.
[24] Zhu XY, Yuan Y, Li LH, Du YC, Li F. Identification of interfacial transition zone in asphalt concrete based on nano-scale metrology techniques. Mater Des 2017;
129:91–102. https://doi.org/10.1016/j.matdes.2017.05.015.
[25] Lyu K, She W, Miao CW, Chang HL, Gu Y. Quantitative characterization of pore morphology in hardened cement paste via SEM-BSE image analysis. Constr
Build Mater 2019;202:589–602. https://doi.org/10.1016/j.conbuildmat.2019.01.055.
[26] Lyu K, Garboczi EJ, She W, Miao CW. The effect of rough vs. smooth aggregate surfaces on the characteristics of the interfacial transition zone. Cem Concr
Compos 2019;99:49–61. https://doi.org/10.1016/j.cemconcomp.2019.03.001.
[27] Ramaniraka M, Rakotonarivo S, Payan C, Garnier V. Effect of interfacial transition zone on diffuse ultrasound in thermally damaged concrete. Cem Concr Res
2022;152:106680. https://doi.org/10.1016/j.cemconres.2021.106680.
[28] Farnam Y, Geiker MR, Bentz D, Weiss J. Acoustic emission waveform characterization of crack origin and mode in fractured and ASR damaged concrete. Cem
Concr Compos 2015;60:135–45. https://doi.org/10.1016/j.cemconcomp.2015.04.008.
[29] He JT, Lei D, Xu WX. In-situ measurement of nominal compressive elastic modulus of interfacial transition zone in concrete by SEM-DIC coupled method. Cem
Concr Compos 2020;114:103779. https://doi.org/10.1016/j.cemconcomp.2020.103779.
[30] He JT, Lei D, Luzio GD, Bai ZuFP, Px.. Mechanical properties measurement and micro-damage characterization of ITZ in concrete by SEM-DIC method. Opt
Lasers Engng 2022;155:107064. https://doi.org/10.1016/j.optlaseng.2022.107064.
[31] Vargas P, Restrepo-Baena O, Tobón JI. Microstructural analysis of interfacial transition zone (ITZ) and its impact on the compressive strength of lightweight
concretes. Constr Build Mater 2017;137:381–9. https://doi.org/10.1016/j.conbuildmat.2017.01.101.
[32] Tian Y, Tian ZS, Jin NG, Jin XY, Yu W. A multiphase numerical simulation of chloride ions diffusion in concrete using electron microprobe analysis for
characterizing properties of ITZ. Constr Build Mater 2018;178:432–44. https://doi.org/10.1016/j.conbuildmat.2018.05.047.
[33] Skarzynski L, Nitka M, Tejchman J. Modelling of concrete fracture at aggregate level using FEM and DEM based on X-ray μCT images of internal structure.
Engng Fract Mech 2015;145:13–35. https://doi.org/10.1016/j.engfracmech.2015.08.010.
[34] Trawiński W, Tejchman J, Bobiński J. A three-dimensional meso-scale modelling of concrete fracture, based on cohesive elements and X-ray μCT images.
Engng Fract Mech 2018;189:27–50. https://doi.org/10.1016/j.engfracmech.2017.10.003.
[35] Xie YT, Corr DJ, Jin F, Zhou H, Shah SP. Experimental study of the interfacial transition zone (ITZ) of model rock-filled concrete (RFC). Cem Concr Compos
2015;55:223–31. https://doi.org/10.1016/j.cemconcomp.2014.09.002.
[36] Nitka M, Tejchman J. Meso-mechanical modelling of damage in concrete using discrete element method with porous ITZs of defined width around aggregates.
Engng Fract Mech 2020;231:107029. https://doi.org/10.1016/j.engfracmech.2020.107029.
[37] Maleki M, Rasoolan I, Khajehdezfuly A, Jivkov AP. On the effect of ITZ thickness in meso-scale models of concrete. Constr Build Mater 2020;258:119639.
https://doi.org/10.1016/j.conbuildmat.2020.119639.
[38] Zhang SL, Zhang CS, Liao L, Wang CL. Numerical study of the effect of ITZ on the failure behaviour of concrete by using particle element modelling. Constr
Build Mater 2018;170:776–89. https://doi.org/10.1016/j.conbuildmat.2018.03.040.
[39] Kim SY, Kim JS, Kang JW, Han TS. Construction of virtual interfacial transition zone (ITZ) samples of hydrated cement paste using extended stochastic
optimization. Cem Concr Compos 2019;102:84–93. https://doi.org/10.1016/j.cemconcomp.2019.04.012.
[40] Sharma M, Bishnoi S. Influence of properties of interfacial transition zone on elastic modulus of concrete: evidence from micromechanical modelling. Constr
Build Mater 2020;246:118381. https://doi.org/10.1016/j.conbuildmat.2020.118381.
[41] Carrara P, De Lorenzis L. Consistent identification of the interfacial transition zone in simulated cement microstructures. Cem Concr Compos 2017;80:224–34.
https://doi.org/10.1016/j.cemconcomp.2017.03.008.
[42] Lyu K, She W, Chang HL, Gu Y. Effect of fine aggregate size on the overlapping of interfacial transition zone (ITZ) in mortars. Constr Build Mater 2020;248:
118559. https://doi.org/10.1016/j.conbuildmat.2020.118559.
[43] Grondin F, Matallah M. How to consider the interfacial transition zones in the finite element modelling of concrete? Cem Concr Res 2014;58:67–75. https://
doi.org/10.1016/j.cemconres.2014.01.009.
[44] Hashin Z, Monteiro PJM. An inverse method to determine the elastic properties of the interphase between the aggregate and the cement paste. Cem Concr Res
2002;32(8):1291–300. https://doi.org/10.1016/S0008-8846(02)00792-5.
[45] Wang XF, Zhang MZ, Jivkov AP. Computational technology for analysis of 3D meso-structure effects on damage and failure of concrete. International Journal
of Solid and structures 2016;80:310–33. https://doi.org/10.1016/j.ijsolstr.2015.11.018.
[46] Chen QQ, Zhang YH, Zhao TT, Wang ZY, Wang ZH. Mesoscale modelling of concretes subjected to triaxial loadings: mechanical properties and fracture
behaviour. Materials 2021;14:1099. https://doi.org/10.3390/ma14051099.

24
Q. Chen et al. Engineering Fracture Mechanics 300 (2024) 109979

[47] Wang ZH, Chen QQ, Wang ZY, Xiong J. The investigation into the failure criteria of concrete based on the BP neural network. Engng Fract Mech 2022;275:
108835. https://doi.org/10.1016/j.engfracmech.2022.108835.
[48] Nagai K, Sato Y, Ueda T. Mesoscopic simulation of failure of mortar and concrete by 3D RBSM. J Adv Concr Technol 2005;3(3):385–402. https://doi.org/
10.3151/jact.3.385.
[49] Nitka M, Tejchman J. A three-dimensional meso-scale approach to concrete fracture based on combined DEM with X-ray μCT images. Cem Concr Res 2018;
107:11–29. https://doi.org/10.1016/j.cemconres.2018.02.006.
[50] Nadeau JC. A multiscale model for effective moduli of concrete incorporating ITZ water-cement ratio gradients, aggregate size distributions, and entrapped
voids. Cem Concr Res 2003;33(1):103–13. https://doi.org/10.1016/S0008-8846(02)00931-6.
[51] Garboczi EJ, Bentz DP. Analytical formulas for interfacial transition zone properties. Adv Cem Bas Mat 1997;6(3):99–108. https://doi.org/10.1016/S1065-
7355(97)81592-1.
[52] Aquino MJ, Li Z, Shah SP. Mechanical properties of the aggregate and cement interface. Adv Cem Bas Mat 1995;2:211–23. https://doi.org/10.1016/1065-
7355(95)90040-3.
[53] Königsberger M, Pichler B, Hellmich C, Olek J. Micromechanics of ITZ-aggregate interaction in concrete part I: stress concentration. Journal of American
Ceramic Society 2014;97(2):535–42. https://doi.org/10.1111/jace.12591.
[54] Königsberger M, Hlobil M, Delsaute B, Staquet S, Hellmich C, Pichler B. Hydrate failure in ITZ governs concrete strength: a micro-to-macro validated
engineering mechanics model. Cem Concr Res 2018;103:77–94. https://doi.org/10.1016/j.cemconres.2017.10.002.
[55] Lutz MP, Zimmerman RW. Effect of an inhomogeneous interphase zone on the bulk modulus and conductivity of a particulate composite. International Journal
of Solid and Structures 2005;42(2):429–37. https://doi.org/10.1016/j.ijsolstr.2004.06.046.
[56] Yang CC. Effect of the transition zone on the elastic moduli of mortar. Cem Concr Res 1998;28(5):727–36. https://doi.org/10.1016/S0008-8846(98)00035-0.
[57] Garboczi EJ, Bentz DP. Multiscale analytical/numerical theory of the diffusivity of concrete. Adv Cem Bas Mat 1998;8(2):77–88. https://doi.org/10.1016/
S1065-7355(98)00010-8.
[58] Zheng JJ, Zhou XZ, Jin XY. An n-layered spherical inclusion model for predicting the elastic moduli of concrete with inhomogeneous ITZ. Cem Concr Compos
2012;34(5):716–23. https://doi.org/10.1016/j.cemconcomp.2012.01.011.
[59] Dridi M. Analysis of effective diffusivity of cement based materials by multi-scale modelling. Mater Struct 2013;46(1–2):313–26. https://doi.org/10.1617/
s11527-012-9903-5.
[60] Zouaoui R, Miled K, Limam O, Beddy A. Analytical prediction of aggregates’ effects on the ITZ volume fraction and young’s modulus of concrete. Int J Numer
Anal Meth Geomech 2017;41:976–93. https://doi.org/10.1002/nag.2660.
[61] Xu WX, Duan QL, Ma HF, Chen W, Chen HS. Interfacial effect on physical properties of composite media: interfacial volume fraction with non-spherical hard-
core-soft-shell structured particles. Sci Rep 2015;5:16003. https://doi.org/10.1038/srep16003.
[62] Lu BL, Torquato S. Nearest-surface distribution-functions for polydispersed particle-systems. Phys Rev A 1992;45(8):5530–44. https://doi.org/10.1103/
PhysRevA.45.5530.
[63] Erdem S, Dawson AR, Thom NH. Influence of the micro- and nanoscale local mechanical properties of the interfacial transition zone on impact behavior of
concrete made with different aggregates. Cem Concr Res 2012;42(2):447–58. https://doi.org/10.1016/j.cemconres.2011.11.015.
[64] Korouzhdeh T, Eskandari-Naddaf H, Kazemi R. The ITZ microstructure, thickness, porosity and its relation with compressive and flexural strength of cement
mortar; influence of cement fineness and water/cement ratio. Front Struct Civ Engng 2022;16(2):191–201. https://doi.org/10.1007/s11709-021-0792-y.
[65] Rangaraju PR, Olek J, Diamond S. An investigation into the influence of inter-aggregate spacing and the extent of the ITZ on properties of Portland cement
concretes. Cem Concr Res 2010;40(11):1601–8. https://doi.org/10.1016/j.cemconres.2010.07.002.
[66] Zheng JJ, Li CQ, Zhou XZ. Characterization of microstructure of interfacial transition zone in concrete. ACI Mater J 2005;102(4):265–71.
[67] Elsharief A, Cohen MD, Olek J. Influence of aggregate size, water cement ratio and age on the microstructure of the interfacial transition zone. Cem Concr Res
2003;33(11):1837–49. https://doi.org/10.1016/S0008-8846(03)00205-9.
[68] Gao YY, Hu CL, Zhang YM, Li ZJ, Pan JL. Characterisation of the interfacial transition zone in mortars by nanoindentation and scanning electron microscope.
Mag Concr Res 2018;70(18):965–72. https://doi.org/10.1680/jmacr.17.00161.
[69] Prokopsi G, Langier B. Effect of w/c ratio and silica fume addition on the fracture toughness and morphology of fractured surfaces of gravel concretes. Cem
Concr Res 2000;30(9):1427–33. https://doi.org/10.1016/S0008-8846(00)00332-X.
[70] Zheng JJ, Guo ZQ, Pan XD, Stroeven P, Sluys LJ. ITZ volume fraction in concrete with spheroidal aggregate particles and application: part I. Numerical
algorithm Magazine of Concrete Research 2012;63(7):473–82. https://doi.org/10.1680/macr.2011.63.7.473.
[71] Sun GW, Sun W, Zhang YS, Liu ZY. Numerical calculation and influencing factors of the volume fraction of interfacial transition zone in concrete. Science
China-Technological Sciences 2012;55(6):1515–22. https://doi.org/10.1007/s11431-011-4737-x.
[72] Xu WX, Han ZM, Tao L, Ding QH, Ma HF. Random non-convex particle model for the fraction of interfacial transition zones (ITZs) in fully-graded concrete.
Powder Technol 2018;323:301–9. https://doi.org/10.1016/j.powtec.2017.10.009.
[73] Gao Y, De Schutter G, Ye G. Micro- and meso-scale pore structure in mortar in relation to aggregate content. Cem Concr Res 2013;52:149–60. https://doi.org/
10.1016/j.cemconres.2013.05.011.
[74] Rao GA, Prasad BKR. Influence of the roughness of aggregate surface on the interface bond strength. Cem Concr Res 2002;32:253–7. https://doi.org/10.1016/
S0008-8846(01)00668-8.
[75] Rao GA, Raghu Prasad BK. Influence of type of aggregate and surface roughness on the interface fracture properties. Mater Struct 2004;37(5):328–34. https://
doi.org/10.1007/BF02481679.
[76] Hong L, Gu XL, Lin F. Influence of aggregate surface roughness on mechanical properties of interface and concrete. Constr Build Mater 2014;65:338–49.
https://doi.org/10.1016/j.conbuildmat.2014.04.131.
[77] Akçaoğlu R, Tokyay M, Celik T. Assessing the ITZ microcracking via scanning electron microscope and its effect on the failure behavior of concrete. Cem Concr
Res 2005;35(2):358–63. https://doi.org/10.1016/j.cemconres.2004.05.042.
[78] Kong L, Hou L, Bao X. Application of AC impedance technique in study of lightweight aggregate-paste interface. Constr Build Mater 2015;82:332–40. https://
doi.org/10.1016/j.conbuildmat.2015.02.079.
[79] Kong LJ, Hou LR, Wang YH, Sun GW. Investigation of the interfacial transition zone between aggregate cement paste by AC impedance spectroscopy. Journal
of Wuhan University of Technology (Materials Science) 2016;31(4):865–71. https://doi.org/10.1007/s11595-016-1460-2.
[80] Stroeven P, Hu J. ITZ’ s structural evolution during hydration in model concrete. Mag Concr Res 2009;61(5):371–7. https://doi.org/10.1680/
macr.2008.00095.
[81] Golewski GL. The influence of microcrack width on the mechanical parameters in concrete with the addition of fly ash: consideration of technological and
ecological benefits. Constr Build Mater 2019;197:849–61. https://doi.org/10.1016/j.conbuildmat.2018.08.157.
[82] Golewski GL. Determination of fracture toughness in concretes containing siliceous fly ash during mode III loading. Struct Engng Mech 2017;62:1–9. https://
doi.org/10.12989/sem.2017.62.1.001.
[83] Golewski GL. Effect of fly ash addition on the fracture toughness of plain concrete at third model of fracture. J Civ Engng Manag 2017;23:613–20. https://doi.
org/10.3846/13923730.2016.1217923.
[84] Stroeven P, Stroeven M. Reconstructions by SPACE of the interfacial transition zone. Cem Concr Compos 2001;23(2–3):189–200. https://doi.org/10.1016/
s0958-9465(00)00076-7.
[85] Zhao HY, Wu ZH, Liu A, Zhang LC. Numerical insights into the effect of ITZ and aggregate strength on concrete properties. Theor Appl Fract Mech 2022;120:
103415. https://doi.org/10.1016/j.tafmec.2022.103415.
[86] Chang HL, Feng P, Lyu K, Liu J. A novel method for assessing C-S-H chloride adsorption in cement pastes. Constr Build Mater 2019;225:324–31. https://doi.
org/10.1016/j.conbuildmat.2019.07.212.

25
Q. Chen et al. Engineering Fracture Mechanics 300 (2024) 109979

[87] Xu WX, Chen HS, Chen W, Zhu ZG. Theoretical estimation for the volume fraction of interfacial layers around convex particles in multiphase materials. Powder
Technol 2013;249(3):513–5. https://doi.org/10.1016/j.powtec.2013.09.010.
[88] Diamond S. Considerations in image analysis as applied to investigations of the ITZ in concrete. Cem Concr Compos 2001;23(2–3):171–8. https://doi.org/
10.1016/S0958-9465(00)00085-8.
[89] Skarżyński Ł, Tejchman J. Experimental investigations of fracture process in concrete by means of X-ray micro-computed tomography. Strain 2016;52:26–45.
https://doi.org/10.1111/str.12168.
[90] Huang YJ, Yang ZJ, Ren WY, Liu GH, Zhang CZ. 3D meso-scale fracture modelling and validation of concrete based on in-situ X-ray computed tomography
images using damage plasticity model. International Journal of Solid and Structures 2015;67:340–52. https://doi.org/10.1016/j.ijsolstr.2015.05.002.
[91] Yang Z, Ren W, Sharma R, McDonald S, Mostafavi M, Vertyagina Y, et al. In-situ X-ray computed tomography characterisation of 3D fracture evolution and
image-based numerical homogenisation of concrete. Cem Concr Compos 2017;75:74–83. https://doi.org/10.1016/j.cemconcomp.2016.10.001.
[92] Ren WY, Yang ZJ, Sharma R, Zhang C, Withers PJ. Two-dimensional X-ray CT image based meso-scale fracture modelling of concrete. Engng Fract Mech 2015;
133:24–39. https://doi.org/10.1016/j.engfracmech.2014.10.016.
[93] Li Y, Zhang H, Liu G, Hu D, Ma X. Multi-scale study on mechanical property and strength prediction of aeolian sand concrete. Constr Build Mater 2020;247:
118538. https://doi.org/10.1016/j.conbuildmat.2020.118538.
[94] Wang C, Zhang M, Wang Q, Dai J, Luo T, Pei W, et al. Research on the influencing mechanism of nano-silica on concrete performances based on multi-scale
experiments and micro-scale numerical simulation. Constr Build Mater 2022;318:125873. https://doi.org/10.1016/j.conbuildmat.2021.125873.
[95] Qiu J, Zhu M, Zhou Y, Guan X. Effect and mechanism of coal gangue concrete modification by fly ash. Constr Build Mater 2021;294:123563. https://doi.org/
10.1016/j.conbuildmat.2021.123563.
[96] Bourdette B, Ringot E, Ollivier JP. Modelling of the transition zone porosity. Cem Concr Res 1995;25(4):741–51. https://doi.org/10.1016/0008-8846(95)
00064-J.
[97] Ye G. Percolation of capillary pores in hardening cement pastes. Cem Concr Res 2005;35(1):167–76. https://doi.org/10.1016/j.cemconres.2004.07.033.
[98] Hussin A, Poole C. Petrography evidence of the interfacial transition zone (ITZ) in the normal strength concrete containing granitic and limestone aggregates.
Constr Build Mater 2011;25(5):2298–303. https://doi.org/10.1016/j.conbuildmat.2010.11.023.
[99] Prasad M, Manghnani MH, Wang Y, Zinin P. Acoustic microscopy of Portland cement mortar aggregate/paste interfaces. J Mater Sci 2000;35:3607–13. https://
doi.org/10.1023/a:1004873815763.
[100] Zhang Q, Ji T, Yang Z, Wang C, Wu H. Influence of different activators on microstructure and strength of alkali-activated nickel slag cementitious materials.
Constr Build Mater 2020;235:117449. https://doi.org/10.1016/j.conbuildmat.2019.117449.
[101] Nežerka V, Bílý P, Hrbek V, Fládr J. Impact of silica fume, fly ash, and metakaolin on the thickness and strength of the ITZ in concrete. Cem Concr Compos
2019;103:252–62. https://doi.org/10.1016/j.cemconcomp.2019.05.012.
[102] Peled A, Castro J, Weiss WJ. Atomic force and lateral force microscopy (AFM and LFM) examination of cement and cement hydration products. Cem Concr
Compos 2013;36(1):48–55. https://doi.org/10.1016/j.cemconcomp.2012.08.021.
[103] Peled A, Weiss J. Hydrated cement paste constituents observed with atomic force and lateral force microscopy. Constr Build Mater 2011;25:4299–302. https://
doi.org/10.1016/j.conbuildmat.2011.04.066.
[104] Shen P, Lu L, He Y, Wang F, Hu S. The effect of curing regimes on the mechanical properties, nano-mechanical properties and microstructure of ultra-high
performance concrete. Cem Concr Res 2019;118:1–13. https://doi.org/10.1016/j.cemconres.2019.01.004.
[105] Chaurasia L, Bisht V, Singh LP, Gupta S. A novel approach of biomineralization for improving micro and macro-properties of concrete. Constr Build Mater
2019;195:340–51. https://doi.org/10.1016/j.conbuildmat.2018.11.031.
[106] Liao K, Chang P, Peng Y, Yang C. A study on characteristics of interfacial transition zone in concrete. Cem Concr Res 2004;34(6):977–89. https://doi.org/
10.1016/j.cemconres.2003.11.019.
[107] Khedmati M, Kim YR, Turner JA, Alanazi H, Nguyen C. An integrated microstructural-nanomechanical-chemical approach to examine material-specific
characteristics of cementitious interphase regions. Mater Charact 2018;138:154–64. https://doi.org/10.1016/j.matchar.2018.01.045.
[108] Dobiszewska M, Schindler AK, Pichor W. Mechanical properties and interfacial transition zone microstructure of concrete with waste basalt powder addition.
Constr Build Mater 2018;177:222–9. https://doi.org/10.1016/j.conbuildmat.2018.05.133.
[109] Golewski GL. Evaluation of morphology and size of cracks of the interfacial transition zone (ITZ) in concrete containing fly ash (FA). J Hazard Mater 2018;357:
298–304. https://doi.org/10.1016/j.jhazmat.2018.06.016.
[110] Wu K, Wang K, Xu LL, Sun DD, Wang FZ, De Schutter G. Damage evolution of blended cement concrete under sodium sulfate attack in relation to ITZ volume
content. Constr Build Mater 2018;190:452–65. https://doi.org/10.1016/j.conbuildmat.2018.09.013.
[111] Nadesan MS, Dinakar P. Micro-structural behavior of interfacial transition zone of the porous sintered fly ash aggregate. Journal of Building Engineering 2018;
16:31–8. https://doi.org/10.1016/j.jobe.2017.12.007.
[112] Constantinides G, Ulm FJ. The nanogranular nature of C-S-H. Journal of Mechanics and Physics of Solids 2007;55(1):64–90. https://doi.org/10.1016/j.
jmps.2006.06.003.
[113] Sarris E, Constantinides G. Finite element modeling of nanoindentation on C-S-H: effect of pile-up and contact friction. Cem Concr Compos 2013;36:78–84.
https://doi.org/10.1016/j.cemconcomp.2012.10.010.
[114] Gao JM, Qian CX, Liu HF, Wang B, Li L. ITZ microstructure of concrete containing GGBS. Cem Concr Res 2005;35(7):1299–304. https://doi.org/10.1016/j.
cemconres.2004.06.042.
[115] Brown L, Allison PG, Sanchez F. Use of nanoindentation phase characterization and homogenization to estimate the elastic modulus of heterogeneously
decalcified cement pastes. Mater Des 2018;142:308–18. https://doi.org/10.1016/j.matdes.2018.01.030.
[116] Luo ZY, Li WG, Wang KJ, Shah SP. Research progress in advanced nanomechanical characterization of cement-based materials. Cem Concr Compos 2018;94:
277–95. https://doi.org/10.1016/j.cemconcomp.2018.09.016.
[117] Khedmati M, Alanazi H, Kim YR, Nsengiyumva G, Moussavi S. Effects of Na2O/SiO2 molar ratio on properties of aggregate-paste interphase in fly ash-based
geopolymer mixtures through multiscale measurements. Constr Build Mater 2018;191:564–74. https://doi.org/10.1016/j.conbuildmat.2018.10.024.
[118] Gao YY, Hu CL, Zhang YM, Li ZJ, Pan JL. Investigation on microstructure and microstructural elastic properties of mortar incorporating fly ash. Cem Concr Res
2018;86:315–21. https://doi.org/10.1016/j.cemconcomp.2017.09.008.
[119] Alanazi H. Study of the interfacial transition zone characteristics of geopolymer and conventional concretes. Gels 2022;8(2):105. https://doi.org/10.3390/
gels8020105.
[120] Zhu W, Bartos PJM. Assessment of interfacial microstructure and bond properties in aged GRC using a novel microindentation method. Cem Concr Res 1997;27
(11):1701–11. https://doi.org/10.1016/S0008-8846(97)00155-5.
[121] Lei D, Huang JF, Xu WX, Wang WC, Zhang P. Deformation analysis of shear band in granular materials via a robust plane shear test and numerical simulation.
Powder Technol 2018;323:385–92. https://doi.org/10.1016/j.powtec.2017.10.027.
[122] Chen Y, Wei JX, Huang HL, Jin W, Yu QJ. Application of 3D-DIC to characterize the effect of aggregate size and volume on non-uniform shrinkage strain
distribution in concrete. Cem Concr Compos 2018;86:178–89. https://doi.org/10.1016/j.cemconcomp.2017.11.005.
[123] Zhao BN, Lei D, Fu JJ, Yang LQ, Xu WX. Experimental study on micro-damage identification in reinforced concrete beam with wavelet packet and DIC method.
Constr Build Mater 2019;210:338–46. https://doi.org/10.1016/j.conbuildmat.2019.03.175.
[124] Lei D, Huang ZT, Bai PX, Zhu FP. Experimental research on impact damage of xiaowan arch dam model by digital image correlation. Constr Build Mater 2017;
147:168–73. https://doi.org/10.1016/j.conbuildmat.2017.04.143.
[125] Lei D, Yang LQ, Xu WX, Zhang P, Huang ZT. Experimental study on alarming of concrete micro-crack initiation based on wavelet packet analysis. Constr Build
Mater 2017;149:716–23. https://doi.org/10.1016/j.conbuildmat.2017.05.159.
[126] Li WG, Xiao JZ, Sun ZH, Shah PS. Failure processes of modeled recycled aggregate concrete under uniaxial compression. Cem Concr Compos 2012;34:1149–58.
https://doi.org/10.1016/j.cemconcomp.2012.06.017.

26
Q. Chen et al. Engineering Fracture Mechanics 300 (2024) 109979

[127] Herve E, Care S, Seguin JP. Influence of the porosity gradient in cement paste matrix on the mechanical behavior of mortar. Cem Concr Res 2010;40(7):
1060–71. https://doi.org/10.1016/j.cemconres.2010.02.010.
[128] Jiang ZL, Huang QH, Xi YP, Gu XL, Zhang WP. Experimental study of diffusivity of the interfacial transition zone between cement paste and aggregate. J Mater
Civ Engng 2016;28(10):04016109. https://doi.org/10.1061/(ASCE)MT.1943-5533.0001637.
[129] Tran DL, Mouret M, Cassagnabère F. Impact of the porosity and moisture state of coarse aggregates, and binder nature on the structure of the paste-aggregate
interface: elementary model study. Constr Build Mater 2022;319:126112. https://doi.org/10.1016/j.conbuildmat.2021.126112.
[130] Zhang HZ, Gan YD, Xu YD, Zhang SZ, Schlangen E, Šavija B. Experimentally informed fracture modelling of interfacial transition zone at micro-scale. Cem
Concr Compos 2019;104:103383. https://doi.org/10.1016/j.cemconcomp.2019.103383.
[131] Ma HY, Li ZJ. Multi-aggregate approach for modelling interfacial transition zone in concrete. ACI Mater J 2014;111(2):189–200. https://doi.org/10.14359/
51686501.
[132] Wong HS, Head MK, Buenfeld NR. Pore segmentation of cement-based materials from backscattered electron images. Cem Concr Res 2006;36(6):1083–90.
https://doi.org/10.1016/j.cemconres.2005.10.006.
[133] Hu J, Ge Z, Wang KJ. Influence of cement fineness and water-to-cement ratio on mortar early-age heat of hydration and set times. Constr Build Mater 2014;50:
657–63. https://doi.org/10.1016/j.conbuildmat.2013.10.011.
[134] Lyu K, She W. Determination of aggregate surface morphology at the interfacial transition zone (ITZ). J Vis Exp 2019;154:e60245.
[135] Zhu ZG, Chen HS. Aggregate shape effect on the overestimation of interface thickness for spheroidal particles. Powder Technol 2017;313:218–30. https://doi.
org/10.1016/j.powtec.2017.03.014.
[136] Josserand L, Coussy O, Larrard FD. Bleeding of concrete as an ageing consolidation process. Cem Concr Res 2006;36(9):1603–8. https://doi.org/10.1016/j.
cemconres.2004.10.006.
[137] Basheer PAM, Nolan É. Near-surface moisture gradients and in situ permeation tests. Constr Build Mater 2001;15(2–3):105–14. https://doi.org/10.1016/
S0950-0618(00)00059-3.
[138] Abdul Razak H, Chai HK, Wong HS. Near surface characteristics of concrete containing supplementary cementing materials. Cem Concr Res 2004;26(7):883–9.
https://doi.org/10.1016/j.cemconcomp.2003.10.001.
[139] Liang SM, Wei Y, Wu ZH. Multiscale modeling elastic properties of cement-based materials considering imperfect interface effect. Constr Build Mater 2017;
154:567–79. https://doi.org/10.1016/j.conbuildmat.2017.07.196.
[140] Jia ZJ, Han YG, Zhang YM, Qiu C, Hu CL, Li ZJ. Quantitative characterization and elastic properties of interfacial transition zone around coarse aggregate in
concrete. Journal of Wuhan University of Technology (Materials Science) 2017;32(4):838–44. https://doi.org/10.1007/s11595-017-1677-8.
[141] Brough AR, Atkinson A. Automated identification of the aggregate-paste interfacial transition zone in mortars of silica sand with Portland or alkali-activated
slag cement paste. Cem Concr Res 2020;30(6):849–54. https://doi.org/10.1016/S0008-8846(00)00254-4.
[142] Hu J, Stroeven P. Properties of the interfacial transition zone in model concrete. Interface Sci 2004;12(4):389–97. https://doi.org/10.1023/B:
INTS.0000042337.39900.fb.
[143] Bentz DP, Garboczi EJ. Computer modelling of interfacial transition zone microstructure and properties. In: Alexander MG, Arliguie G, Ballivy G, Bentur A,
Marchand J, editors. Engineering and Transport Properties of the Interfacial Transition Zone in Cementitious Composites. RILEM Publications S.A.R.L; 1999.
p. 349–85.
[144] Zhu ZG, Provis JL, Chen HS. Quantification of the influences of aggregate shape and sampling method on the overestimation of ITZ thickness in cementitious
materials. Powder Technol 2018;326:168–80. https://doi.org/10.1016/j.powtec.2017.12.008.
[145] Basheer L, Kropp J, Cleland DJ. Assessment of the durability of concrete from its permeation properties: a review. Constr Build Mater 2001;15(2):93–103.
https://doi.org/10.1016/S0950-0618(00)00058-1.
[146] Asbridge AH, Page CL, Page MM. Effects of metakaolin, water/binder ratio and interfacial transition zones on the microhardness of cement mortars. Cem Concr
Res 2002;32(9):1365–9. https://doi.org/10.1016/s0008-8846(02)00798-6.
[147] Yang CC. Effect of the interfacial transition zone on the transport and the elastic properties of mortar. Mag Concr Res 2003;55(4):305–12. https://doi.org/
10.1680/macr.2003.55.4.305.
[148] Xu J, Wang BB, Zuo JQ. Modification effects of nanosilica on the interfacial transition zone in concrete: a multiscale approach. Cem Concr Compos 2017;81:
1–10. https://doi.org/10.1016/j.cemconcomp.2017.04.003.
[149] Branch JL, Epps R, Kosson DS. The impact of carbonation on bulk and ITZ porosity in microconcrete materials with fly ash replacement. Cem Concr Res 2018;
103:170–8. https://doi.org/10.1016/j.cemconres.2017.10.012.
[150] Li Y, Liu YZ, Wang R. Evaluation of the elastic modulus of concrete based on indentation test and multi-scale homogenization method. Journal of Building
Engineering 2021;43:102758. https://doi.org/10.1016/j.jobe.2021.102758.
[151] Li Y, Liu Y, Jin C, Mu J, Li H, Liu J. Multi-scale creep analysis of river sand and manufactured sand concrete considering the influence of ITZ. Constr Build
Mater 2022;344:128175. https://doi.org/10.1016/j.conbuildmat.2022.128175.
[152] Xu J, Corr DJ, Shah SP. Nanomechanical investigation of the effects of nanoSiO2 on C-S-H gel/cement grain interfaces. Cem Concr Compos 2015;61:7–17.
https://doi.org/10.1016/j.cemconcomp.2015.04.011.
[153] Heo YS, Sanjayan JG, Han CG, Han MC. Relationship between inter-aggregate spacing and the optimum fiber length for spalling protection of concrete in fire.
Cem Concr Res 2012;42(3):549–57. https://doi.org/10.1016/j.cemconres.2011.12.002.
[154] Zheng JJ, Xiong FF, Wu ZM, Jin WL. A numerical algorithm for the ITZ area fraction in concrete with elliptical aggregate particles. Mag Concr Res 2009;61(2):
109–17. https://doi.org/10.1680/macr.2007.00123.
[155] Chen HS, Zhu ZG, Liu L, Sun W, Miao CW. Aggregate shape effect on the overestimation of ITZ thickness: quantitative analysis of platonic particles. Powder
Technol 2016;289:1–17. https://doi.org/10.1016/j.powtec.2015.11.036.
[156] Wriggers P, Moftah SO. Mesoscale models for concrete: homogenisation and damage behaviour. Finite Elements in Analysis Design 2006;42(7):623–36.
https://doi.org/10.1016/j.finel.2005.11.008.
[157] Du XL, Jin L, Ma GW. Numerical simulation of dynamic tensile-failure of concrete at meso-scale. Int J Impact Eng 2014;66:5–17. https://doi.org/10.1016/j.
ijimpeng.2013.12.005.
[158] Park SW, Xia Q, Zhou M. Dynamic behavior of concrete at high strain rates and pressures: II. numerical simulation. Int J Impact Eng 2001;25:887–910. https://
doi.org/10.1016/s0734-743x(01)00021-5.
[159] Zhang J, Wang ZY, Yang HW, Wang ZH, Shu XF. 3D meso-scale modeling of reinforcement concrete with high volume fraction of randomly distributed
aggregates. Constr Build Mater 2018;164:350–61. https://doi.org/10.1016/j.conbuildmat.2017.12.229.
[160] Zhang J, Chen WS, Hao H, Wang ZY, Wang ZH, Shu XF. Performance of concrete targets mixed with coarse aggregates against rigid projectile impact. Int J
Impact Eng 2020;141:103565. https://doi.org/10.1016/j.ijimpeng.2020.103565.
[161] Zhang J, Wang ZY, Wang ZH, Shu XF, Li QM. Penetration trajectory of rigid projectile in the heterogeneous meso-scale concrete target. Adv Struct Engng 2022;
25(7):1469–82. https://doi.org/10.1177/13694332221087342.
[162] Trawiński W, Bobiński J, Tejchman J. Two-dimensional simulations of concrete fracture at aggregate level with cohesive elements based on X-ray μCT images.
Engng Fract Mech 2016;168:204–26. https://doi.org/10.1016/j.engfracmech.2016.09.012.
[163] Zhou RX, Lu Y. A mesoscale interface approach to modelling fractures in concrete for material investigation. Constr Build Mater 2018;165:608–20. https://doi.
org/10.1016/j.conbuildmat.2018.01.040.
[164] Yilmaz O, Molinari JF. A mesoscale fracture model for concrete. Cem Concr Res 2017;97:84–94. https://doi.org/10.1016/j.cemconres.2017.03.014.
[165] Zhou XQ, Hao H. Mesoscale modelling of concrete tensile failure mechanism at high strain rates. Comput Struct 2008;86(21–22):2013–26. https://doi.org/
10.1016/j.compstruc.2008.04.013.
[166] Yang ZJ, Su XT, Chen JF, Liu GH. Monte Carlo simulation of complex cohesive fracture in random heterogeneous quasi-brittle materials. International Journal
of Solid and Structures 2009;46(17):3222–34. https://doi.org/10.1016/j.ijsolstr.2009.04.013.

27
Q. Chen et al. Engineering Fracture Mechanics 300 (2024) 109979

[167] Tang LW, Zhou W, Liu XH, Ma G, Chen MX. Three-dimensional mesoscopic simulation of the dynamic tensile fracture of concrete. Engng Fract Mech 2019;211:
269–81. https://doi.org/10.1016/j.engfracmech.2019.02.015.
[168] Lilliu G, van Mier JGM. 3D lattice type fracture model for concrete. Engng Fract Mech 2003;70:927–41. https://doi.org/10.1016/S0013-7944(02)00158-3.
[169] Zhou XQ, Hao H. Modelling of compressive behaviour of concrete-like materials at high strain rate. International Journal of Solid and Structures 2008;45:
4648–61. https://doi.org/10.1016/j.ijsolstr.2008.04.002.
[170] Song ZH, Lu Y. Mesoscopic analysis of concrete under excessively high strain rate compression and implications on interpretation of test data. Int J Impact Eng
2012;46:41–55. https://doi.org/10.1016/j.ijimpeng.2012.01.010.
[171] Huang YJ, Yang ZJ, Chen XW, Liu GH. Monte Carlo simulations of meso-scale dynamic compressive behavior of concrete based on X-ray computed tomography
images. Int J Impact Eng 2016;97:102–15. https://doi.org/10.1016/j.ijimpeng.2016.06.009.
[172] Yu QL, Liu HY, Yang TH, Liu HL. 3D numerical study on fracture process of concrete with different ITZ properties using X-ray computerized tomography.
International Journal of Solid and Structures 2018;147:204–22. https://doi.org/10.1016/j.ijsolstr.2018.05.026.
[173] Jin L, Yu WX, Du XL, Yang WX. Mesoscopic numerical simulation of dynamic size effect on the splitting-tensile strength of concrete. Engng Fract Mech 2019;
209:317–32. https://doi.org/10.1016/j.engfracmech.2019.01.035.
[174] Pedersen RR, Simone A, Sluys LJ. Mesoscopic modeling and simulation of the dynamic tensile behavior of concrete. Cem Concr Res 2013;50:74–87. https://
doi.org/10.1016/j.cemconres.2013.03.021.
[175] Chen HB, Xu B, Mo YL, Zhou TM. Behavior of meso-scale heterogeneous concrete under uniaxial tensile and compressive loadings. Constr Build Mater 2018;
178:418–31. https://doi.org/10.1016/j.conbuildmat.2018.05.052.
[176] Li SG, Li QB. Method of meshing ITZ structure in 3D meso-level finite element analysis for concrete. Finite Elem Anal Des 2015;93:96–106. https://doi.org/
10.1016/j.finel.2014.09.006.
[177] Zheng JJ, Zhou XZ. Percolation of ITZs in concrete and effects of attributing factors. J Mater Civ Engng 2007;19(9):784–90. https://doi.org/10.1061/(ASCE)
0899-1561(2007)19:9(784).
[178] Sun ZH, Garboczi EJ, Shah SP. Modeling the elastic properties of concrete composites: experiment, differential effective medium theory, and numerical
simulation. Cem Concr Compos 2007;29:22–38. https://doi.org/10.1016/j.cemconcomp.2006.07.020.
[179] Kim SM, Abu Al-Rub RK. Meso-scale computational modeling of the plastic-damage response of cementitious composites. Cem Concr Res 2011;41(3):339–58.
https://doi.org/10.1016/j.cemconres.2010.12.002.
[180] Keinde D, Kamali-Bernard S, Bernard F, Cisse I. Effect of the interfacial transition zone and the nature of the matrix-aggregate interface on the overall elastic
and inelastic behaviour of concrete under compression: a 3D numerical study. European Journal of Environmental and Civil Engineering 2014; Published
online March 17, 2014:1-10. 10.1080/19648189.2014.896757.
[181] Zhou RX, Song ZH, Lu Y. 3D mesoscale finite element modelling of concrete. Comput Struct 2017;192:96–113. https://doi.org/10.1016/j.
compstruc.2017.07.009.
[182] Lv TH, Chen XW, Chen G. The 3D meso-scale model and numerical tests of split hopkinson pressure bar of concrete specimen. Constr Build Mater 2018;160:
744–64. https://doi.org/10.1016/j.conbuildmat.2017.11.094.
[183] Zhang YH, Chen QQ, Wang ZY, Zhang J, Wang ZH, Li ZQ. 3D mesoscale fracture analysis of concrete under complex loading. Engng Fract Mech 2019;220:
106646. https://doi.org/10.1016/j.engfracmech.2019.106646.
[184] Wu LJ, Ju XL, Liu MW, Guan L, Li MYF, Ml.. Influences of multiple factors on the chloride diffusivity of the interfacial transition zone in concrete composites.
Compos B 2020;199:108236. https://doi.org/10.1016/j.compositesb.2020.108236.
[185] Garboczi EJ, Bullard JW. Contact function, uniform-thickness shell volume, and convexity measure for 3D star-shaped random particles. Powder Technol 2013;
237:191–201. https://doi.org/10.1016/j.powtec.2013.01.019.
[186] Aouissi F, Yang CC, Brahma A, Bederina M. Study of the influence of the interfacial transition zone on the elastic modulus of concrete using a triphasic model.
J Adhes Sci Technol 2016;30(9):994–1005. https://doi.org/10.1080/01694243.2015.1136135.
[187] Garboczi EJ, Berryman JG. Elastic moduli of a material containing composite inclusions: effective medium theory and finite element computations. Mech
Mater 2001;33(8):455–70. https://doi.org/10.1016/S0167-6636(01)00067-9.
[188] Wu K, Han H, Li HX, Dong BQ, Liu TJ, De Schutter G. Experimental study on concurrent factors influencing the ITZ effect on mass transport in concrete. Cem
Concr Compos 2021;123:104215. https://doi.org/10.1016/j.cemconcomp.2021.104215.
[189] Sheng PY, Chen ZT, Cui Y, Zhang JZ, Zhang H, Ji Z. Numerical modelling of interfacial transition zone influence on elastic modulus of concrete. Mag Concr Res
2020;72(16):811–9. https://doi.org/10.1680/jmacr.18.00564.
[190] Chen HS, Sun W, Stroeven P, Sluys LJ. Overestimation of the interface thickness around convex-shaped grain by sectional analysis. Acta Mater 2007;55:
3943–9. https://doi.org/10.1016/j.actamat.2007.03.009.
[191] Samal DK, Ray S, Thiyagarajan H. Influence of differential ITZ thickness around the aggregate on properties of concrete. J Mater Civ Eng 2022;34(7):
04022140. https://doi.org/10.1061/(ASCE)MT.1943-5533.0004280.
[192] Li WG, Xiao JZ, Sun ZH, Kawashima S, Shah SP. Interfacial transition zones in recycled aggregate concrete with different mixing approaches. Constr Build
Mater 2020;35:1045–55. https://doi.org/10.1016/j.conbuildmat.2012.06.022.
[193] Sáez del Bosquea IF, Zhu W, Howind T, Matíasa A, Sánchez de Rojasc MI, Medina C. Properties of interfacial transition zones (ITZs) in concrete containing
recycled mixed aggregates. Cem Concr Compos 2017;81:25–34. https://doi.org/10.1016/j.cemconcomp.2017.04.011.
[194] Revilla-Cuesta V, Skaf M, Santamaría A, Romera JM, Ortega-López V. Elastic stiffness estimation of AGGREGATE–ITZ system of concrete through matrix
porosity and volumetric considerations: explanation and exemplification. ArchivCivMechEng 2022;22(2):59. https://doi.org/10.1007/s43452-022-00382-z.
[195] Lee KM, Park JH. A numerical model for elastic modulus of concrete considering interfacial transition zone. Cem Concr Res 2008;38:396–402. https://doi.org/
10.1016/j.cemconres.2007.09.019.
[196] Christensen RM, Lo KH. Solutions for effective shear properties in three phase sphere and cylinder models. Journal of the Mechanics and Physics Solids 1979;
27(4):315–30. https://doi.org/10.1016/0022-5096(79)90032-2.
[197] Duplan F, Abou-chakra A, Turatsinze A, Escadeillas G, Brule S, Masse F. Prediction of modulus of elasticity based on micromechanics theory and application to
low-strength mortars. Constr Build Mater 2014;50:437–47. https://doi.org/10.1016/j.conbuildmat.2013.09.051.
[198] Haecker CJ, Garboczi EJ, Bullard JW, Bohn RB, Sun Z, Shah SP, et al. Modeling the linear elastic properties of Portland cement paste. Cem Concr Res 2005;35
(10):1948–60. https://doi.org/10.1016/j.cemconres.2005.05.001.
[199] Gu XL, Hong L, Wang ZL, Lin F. Experimental study and application of mechanical properties for the interface between cobblestone aggregate and mortar in
concrete. Constr Build Mater 2013;46:156–66. https://doi.org/10.1016/j.conbuildmat.2013.04.028.
[200] Malachanne E, Jebli M, Jamin F, Garcia-Diaz E, El Youssoufi MS. A cohesive zone model for the characterization of adhesion between cement paste and
aggregates. Constr Build Mater 2018;193:64–71. https://doi.org/10.1016/j.conbuildmat.2018.10.188.
[201] Nesrine S, Mouad J, Etienne M, et al. Identification of a cohesive zone model for cement paste-aggregate interface in a shear test. Eur J Environ Civ Engng
2023;27(6):2288–302. https://doi.org/10.1080/19648189.2019.1623082.
[202] Mielniczuk B, Jebli M, Jamin F, El Youssoufi MS, Pelissou C, Monerie Y. Characterization of behavior and cracking of a cement paste confined between
spherical aggregate particles. Cem Concr Res 2016;79:235–42. https://doi.org/10.1016/j.cemconres.2015.09.016.
[203] Jebli M, Jamin F, Malachanne E, Garcia-Diaz E, El Youssoufi MS. Experimental characterization of mechanical properties of the cement-aggregate interface in
concrete. Constr Build Mater 2018;161:16–25. https://doi.org/10.1016/j.conbuildmat.2017.11.100.
[204] Zhang HR, Ji T, Liu H. Performance evolution of the interfacial transition zone (ITZ) in recycled aggregate concrete under external sulfate attacks and dry-wet
cycling. Constr Build Mater 2019;229:116938. https://doi.org/10.1016/j.conbuildmat.2019.116938.
[205] Yue GB, Ma ZM, Liu M, Liang CF, Ba GZ. Damage behavior of the multiple ITZs in recycled aggregate concrete subjected to aggressive ion environment. Constr
Build Mater 2020;245:118419. https://doi.org/10.1016/j.conbuildmat.2020.118419.

28
Q. Chen et al. Engineering Fracture Mechanics 300 (2024) 109979

[206] Xu G, Beaudoin JJ, Jolicoeur C, Page M. The effect of a polynaphthalene sulfonate superplasticizer on the contribution of the interfacial transition zone to the
electrical resistivity of mortars containing silica and limestone fine aggregate. Cem Concr Res 2000;30(5):683–91. https://doi.org/10.1016/S0008-8846(00)
00222-2.
[207] Duan P, Shui ZH, Chen W, Shen CH. Effects of metakaolin, silica fume and slag on pore structure, interfacial transition zone and compressive strength of
concrete. Constr Build Mater 2013;44:1–6. https://doi.org/10.1016/j.conbuildmat.2013.02.075.
[208] Kong DY, Lei T, Zheng JJ, Ma CC, Jiang J, Jiang J. Effect and mechanism of surface-coating pozzalanics materials around aggregate on properties and ITZ
microstructure of recycled aggregate concrete. Constr Build Mater 2010;24(5):701–8. https://doi.org/10.1016/j.conbuildmat.2009.10.038.
[209] Xuan DX, Shui ZH, Wu SP. Influence of silica fume on the interfacial bond between aggregate and matrix in near-surface layer of concrete. Constr Build Mater
2009;23(7):2631–5. https://doi.org/10.1016/j.conbuildmat.2009.01.006.
[210] Duan P, Shui ZH, Chen W, Shen CH. Efficiency of mineral admixtures in concrete: microstructure, compressive strength and stability of hydrate phases. Appl
Clay Sci 2013;84–83:115–21. https://doi.org/10.1016/j.clay.2013.08.021.
[211] Golewski GL. Validation of the favorable quantity of FA in concrete and analysis of crack propagation and its length-using the crack tip tracking (CTT) method-
in the fracture toughness examinations under mode II, through digital image correlation. Constr Build Mater 2021;296:122362. https://doi.org/10.1016/j.
conbuildmat.2021.122362.
[212] Naderi S, Tu WL, Zhang MZ. Meso-scale modelling of compressive fracture in concrete with irregularly shaped aggregates. Cem Concr Res 2021;140:106317.
https://doi.org/10.1016/j.cemconres.2020.106317.
[213] Zhang D, Huang XM, Zhao YL. Investigation of the shape, size, angularity and surface texture properties of coarse aggregates. Constr Build Mater 2012;34:
330–6. https://doi.org/10.1016/j.conbuildmat.2012.02.096.
[214] Pham DT, Vu MN, Trieu HT, Bui TS, Nguyen-Thoi T. A thermo-mechanical meso-scale lattice model to describe the transient thermal strain and to predict the
attenuation of thermo-mechanical properties at elevated temperature up to 800 ◦ C of concrete. Fire Saf J 2020;114:103011. https://doi.org/10.1016/j.
firesaf.2020.103011.
[215] Pham DT, Nguyen TD, Vu MN, Chinkulkijniwat A. Mesoscale approach to numerical modelling of thermo-mechanical behaviour of concrete at high
temperature. Eur J Environ Civ Engng 2021;25(7):1329–48. https://doi.org/10.1080/19648189.2019.1577762.
[216] Peng Z, Wang Q, Zhou W, Chang X, Yue Q, Huang C. Meso-scale simulation of thermal fracture in concrete based on the coupled thermal–mechanical phase-
field model. Constr Build Mater 2023;403:133095. https://doi.org/10.1016/j.conbuildmat.2023.133095.
[217] Jin L, Zhang R, Du X. Characterisation of temperature-dependent heat conduction in heterogeneous concrete. Mag Concr Res 2018;70(7):325–39. https://doi.
org/10.1680/jmacr.17.00174.
[218] Jin L, Liu K, Zhang R, Yu W, Du X, Deng X. Combined effects of cryogenic temperature and strain rates on compressive behavior of concrete. Int J Damage
Mech 2022;31(9):1396–419. https://doi.org/10.1177/10567895221105590.
[219] Choi S, Na BU, Won MC. Mesoscale analysis of continuously reinforced concrete pavement behavior subjected to environmental loading. Constr Build Mater
2016;112:447–56. https://doi.org/10.1016/j.conbuildmat.2016.02.222.
[220] Roufael G, Beaucour AL, Eslami J, Hoxha D, Noumowé A. Influence of lightweight aggregates on the physical and mechanical residual properties of concrete
subjected to high temperatures. Constr Build Mater 2021;268:121221. https://doi.org/10.1016/j.conbuildmat.2020.121221.
[221] Gao Z, Wang L, Zhang H. Simulation of thermal expansion mismatch at high temperature. Underground Space 2023;8:210–28. https://doi.org/10.1016/j.
undsp.2022.03.007.
[222] Shen Z, Zhou H, Brooks A, Hanna D. Evolution of elastic and thermal properties of cementitious composites containing micro-size lightweight fillers after
exposure to elevated temperature. Cem Concr Compos 2021;118:103931. https://doi.org/10.1016/j.cemconcomp.2021.103931.
[223] Nadeem A, Memon SA, Lo TY. The performance of Fly ash and metakaolin concrete at elevated temperatures. Constr Build Mater 2014;62:67–76. https://doi.
org/10.1016/j.conbuildmat.2014.02.073.
[224] Nadeem A, Memon SA, Lo TY. Qualitative and quantitative analysis and identification of flaws in the microstructure of fly ash and metakaolin blended high
performance concrete after exposure to elevated temperatures. Constr Build Mater 2013;38:731–41. https://doi.org/10.1016/j.conbuildmat.2012.09.062.
[225] Kanellopoulos A, Farhat FA, Nicolaides D, Karihaloo BL. Mechanical and fracture properties of cement-based bi-materials after thermal cycling. Cem Concr Res
2009;39(11):1087–94. https://doi.org/10.1016/j.cemconres.2009.07.008.
[226] Huang H, An M, Wang Y, Yu Z, Ji W. Effect of environmental thermal fatigue on concrete performance based on mesostructural and microstructural analyses.
Constr Build Mater 2019;207:450–62. https://doi.org/10.1016/j.conbuildmat.2019.02.072.
[227] Ye B, Cheng Z, Ni X. Effects of multiple heating-cooling cycles on the permeability and microstructure of a mortar. Constr Build Mater 2018;176:156–64.
https://doi.org/10.1016/j.conbuildmat.2018.05.009.
[228] Ziaei-Nia A, Tadayonfar GR, Eskandari-Naddaf H. Effect of air entraining admixture on concrete under temperature changes in freeze and thaw cycles. Mater
Today: Proc 2018;5(2):6208–16. https://doi.org/10.1016/j.matpr.2017.12.229.
[229] Yang X, Shen A, Guo Y, Zhou S, He T. Deterioration mechanism of interface transition zone of concrete pavement under fatigue load and freeze-thaw coupling
in cold climatic areas. Constr Build Mater 2018;160:588–97. https://doi.org/10.1016/j.conbuildmat.2017.11.031.
[230] Yang X, Jiang Y, Liu G. Study of microstructure of concrete pavement exposed to alternate load and temperature effect. Iran J Sci Technol Trans Civ Eng 2022;
46(1):473–82. https://doi.org/10.1007/s40996-021-00648-1.
[231] Shi J, Liu B, Wu X, Qin J, Jiang J, He Z. Evolution of mechanical properties and permeability of concrete during steam curing process. Journal of Building
Engineering 2020;32:101796. https://doi.org/10.1016/j.jobe.2020.101796.
[232] Vandamme M, Ulm FJ, Fonollosa P. Nanogranular packing of C-S–H at substochiometric conditions. Cem Concr Res 2010;40(1):14–26. https://doi.org/
10.1016/j.cemconres.2009.09.017.
[233] Zhang H, Ji T, Zeng X, Yang Z, Lin X, Liang Y. Mechanical behavior of ultra-high performance concrete (UHPC) using recycled fine aggregate cured under
different conditions and the mechanism based on integrated microstructural parameters. Constr Build Mater 2018;192:489–507. https://doi.org/10.1016/j.
conbuildmat.2018.10.117.
[234] Zhang H, Ji T, He B, He L. Performance of ultra-high performance concrete (UHPC) with cement partially replaced by ground granite powder (GGP) under
different curing conditions. Constr Build Mater 2019;213:469–82. https://doi.org/10.1016/j.conbuildmat.2019.04.058.
[235] Wang Y, Wu L, Wang Y, Liu C, Li Q. Effects of coarse aggregates on chloride diffusion coefficients of concrete and interfacial transition zone under experimental
drying-wetting cycles. Constr Build Mater 2018;185:230–45. https://doi.org/10.1016/j.conbuildmat.2018.07.049.

29

You might also like