IJP2023 B
IJP2023 B
net/publication/370953737
A two-surface contact model for DEM and its application to model fatigue
crack growth in cemented materials
CITATIONS
6 authors, including:
Jayantha Kodikara
Monash University (Australia)
372 PUBLICATIONS 5,858 CITATIONS
SEE PROFILE
Some of the authors of this publication are also working on these related projects:
Development of advanced deterioration models and test methods for the design of stabilised pavement bases View project
All content following this page was uploaded by Ha H. Bui on 23 May 2023.
PII: S0749-6419(23)00136-5
DOI: https://doi.org/10.1016/j.ijplas.2023.103650
Reference: INTPLA 103650
Please cite this article as: Vinh T. Le , Khoa M. Tran , Jayantha Kodikara , Didier Bodin ,
James Grenfell , Ha H. Bui , A two-surface contact model for DEM and its application to model
fatigue crack growth in cemented materials, International Journal of Plasticity (2023), doi:
https://doi.org/10.1016/j.ijplas.2023.103650
This is a PDF file of an article that has undergone enhancements after acceptance, such as the addition
of a cover page and metadata, and formatting for readability, but it is not yet the definitive version of
record. This version will undergo additional copyediting, typesetting and review before it is published
in its final form, but we are providing this version to give early visibility of the article. Please note that,
during the production process, errors may be discovered which could affect the content, and all legal
disclaimers that apply to the journal pertain.
1
A two-surface contact model for DEM and its application to model
fatigue crack growth in cemented materials
Vinh T. Lea, Khoa M. Trana, Jayantha Kodikaraa, Didier Bodinb, James
Grenfellb and Ha H. Buia,*
a
Department of Civil Engineering, Monash University, Australia
b
ARRB Group Ltd, Port Melbourne, VIC 3207, Australia
Abstract
This paper proposes a modelling approach that combines the discrete element method (DEM)
and a novel bonded contact model to characterise the fatigue response of cemented materials.
While DEM is commonly used to simulate bonded materials undergoing cracking, the
centrepiece of the present method is the development of the novel bonded fatigue model. This
new model couples damage mechanics and bounding surface plasticity theory to capture fatigue
crack growth in cement bridges between aggregates. Thanks to the incorporation of the bounding
surface plasticity, the proposed model provides a smooth transition from static to fatigue
damages and vice-versa in a unified manner, making it more flexible to capture damage
responses of cemented materials under different loading conditions (i.e. monotonic and cyclic
loadings). Moreover, the proposed approach automatically captures the hysteretic response in
cement bridges between aggregates under fatigue loadings without ad-hoc treatments. More
importantly, by removing the direct dependence of the fatigue damage variable on the number of
loading cycles, the modelling approach can be applied to simulate the fatigue behaviour of
cemented materials under cyclic variable load amplitudes. The proposed modelling approach is
evaluated against several strength tests to examine its predictive capability. Satisfactory
agreements with fatigue experiments are achieved for flexural modulus degradations, lifetimes
and sensitivity of stress levels under constant and variable amplitude cycles. This result suggests
that the proposed discrete modelling approach can be used to conduct numerical experiments for
insights into the fatigue behaviour of cemented materials.
Keywords: Discrete Element Method (DEM), Fatigue damage, Plasticity, Cohesive model,
Cemented materials, High-cycle fatigues.
1 Introduction
Fatigue in cemented materials caused by repeated loadings is a progressive process where micro-
structure changes gradually, leading to the growth of micro-cracks and eventually failure of
materials. This failure phenomenon has been considered a primary damage source in many
structures made of cementitious materials, such as bridges, pavements or foundations of offshore
wind farms. In general, fatigue response has three distinct phases: crack initiation, propagation
and structural failure (Ray & Kishen, 2010; Kachkouch et al., 2022; Song et al., 2022). Initially,
micro-cracks occur at localised discontinuous flaws in structures. Under repetitive loadings,
*
Corresponding author: ha.bui@monash.edu
2
these micro-cracks propagate at a stable rate, followed by the propagation of unstable cracks to
form macro-cracks. Finally, structural failures occur as the size of those cracks reaches a critical
value. One plausible hypothesis for fatigue responses in cemented materials is due to their
heterogeneous features (including aggregates and cement mortar) and material imperfections
(flaws), which are pre-existing micro-cracks at the interfaces between aggregates and cement
paste or the mortar matrix. This argument is then strongly supported by X-ray micro-CT images,
which show that microcracks propagate along the interfaces between the aggregates and mortar
binder in plain concrete materials under repeated loading (Vicente et al., 2018; Skarżyński et al.,
2019). Thus, one can interpret that the dissipated mechanism of fatigue behaviour is mainly
attributed to these interfaces' frictional and sliding behaviour (Desmorat et al., 2007; Baktheer et
al., 2021).
In addition to the fatigue crack growth, hysteresis is a typical nonlinear phenomenon in cemented
materials under cyclic/fatigue loadings (Reinhardt, 1984; Aslani & Jowkarmeimandi, 2012; Zhu
et al., 2017). Similar to the propagation of fatigue cracks, the hysteretic behaviour is induced by
the heterogeneity of material and structural flaws such as micro-cracks or voids (Chen et al.,
2017; Song et al., 2018; Zhang et al., 2022). In fact, during the loading-unloading process, the
roughness of cracked surfaces and aggregate interlock will generate frictional and sliding
behaviour of aggregate contacts. The hysteretic behaviour is observed due to the combining
effect of the micro behaviour on fracture planes and their surrounding elastic bulk materials.
From the damage mechanics point of view, the hysteresis represents the dissipative mechanisms
in materials, and the area of the hysteresis loops reflects the energy dissipation. The more the
dissipation energy, the greater the damage becomes. This dissipated mechanism was observed in
many experiments reported in the literature (Bazant & Xu, 1991; Isojeh et al., 2017; Keerthana
& Kishen, 2020; Cho et al., 2021), wherein the area of the hysteresis loops become larger with
the increasing number of loading cycles until tested specimens fail (Fig. 1). Therefore, capturing
realistic hysteresis of the material is one of the key features to developing a modelling model
capable of predicting fatigue crack propagation.
Modelling approaches for fatigue crack growth and hysteresis behaviour in cemented materials
can be generally classified into two categories: cohesive zone models (CZMs) and damage
mechanics-based models. The early research on cohesive models was motivated by the
mathematical theory of strip yield models (Dugdale, 1960; Barenblatt, 1962), which expressed
tractions in front of the crack tip as a function of the crack tip distance. Hillerborg et al. (1976)
were inspired by these models but proposed a different approach by formulating tractions based
on the crack opening. This traction-separation law was implemented into the finite element
method (FEM) for the first time to simulate the fracture process zone (FPZ) undergoing
softening damage ahead of a macro-crack in concrete. The approach was then further developed
to become cyclic cohesive zone models (CCZMs) by introducing fatigue laws. The CCZMs were
adopted to simulate fatigue crack growths in metals (Zheng et al., 2011; Xie et al., 2022),
composite materials (Moreo et al., 2007; Xu & Lu, 2013) and quasi-brittle materials (Toumi &
Bascoul, 2002; Xu & Yuan, 2009; Jin & Huang, 2014). However, the application of CCZMs is
restricted as crack paths are required to be predefined. Besides, cohesive laws in these models
were established in a one-dimensional space, making it challenging to capture the tensorial
behaviour of the FPZ in cemented materials where the fatigue behaviour stems from the
propagation of multi cracks. Additionally, a hysteretic cohesive law with different stiffnesses
depending on loading direction (unloading/reloading) was not considered (Xu & Yuan, 2009),
while a simple linear unloading path without irrecoverable separation/strain was assumed in the
3
tension regime (Nguyen et al., 2001; Eliáš & Le, 2012). In contrast to CCZMs, the damage
mechanics approach can effectively analyse multiple crack propagation (Liang et al., 2016;
Wang & Li, 2021). This method describes the material degradation by introducing
scalar/tensorial damage variables to represent the initiation, progressive growth and coalescence
of microcracks (Lemaitre, 1972, 2012). However, one of the challenges of fatigue problems is to
answer the question of what criteria are to determine damage initiation and damage crack
growth. To address this question, Marigo (1985) first developed a new criterion of
loading/unloading irreversibility, allowing damage propagation even for loading levels below the
yielding strength of materials. Inspired by this, Papa and Taliercio (1996); Alliche (2004); Zhang
et al. (2020) used the thermodynamic framework to develop anisotropic fatigue damage models
for concrete. In addition, the bounding surface plasticity theory is an effective alternative to
capture fatigue damage. The key idea of the method is to adopt two surfaces, including the inner
subloading surface and the outer bounding/limit surface (Dafalias & Popov, 1975; Krieg, 1975;
Fang et al., 2017; Zhao et al., 2020). While the former is used to characterise the damage
response of materials under repetitive loadings, the latter plays a role of a failure surface and
limits the geometrical boundary outside which the stress point and subloading surface are not
permitted (Prevost, 1982). The bounding surface theory has been widely applied to concrete
(Yang et al., 1985; Suaris et al., 1990; Pandolfi & Taliercio, 1998; Wu & Harvey, 2013).
Nonetheless, most of these models were applied to conventional triaxial tests or uniaxial tests
under a low number of cyclic loadings.
Besides appropriate constitutive models, numerical methods also play a crucial role in predicting
fatigue behaviour at structural levels. Most existing constitutive fatigue models have been
developed and implemented into computational continuum approaches, such as FEM (Bodin et
al., 2004; Zhu et al., 2014; Kirane & Bažant, 2015; Zreid & Kaliske, 2016; Daneshyar &
Ghaemian, 2017). However, the continuum methods require ad-hoc treatments to cope with the
heterogeneous response and localised failures in cemented materials, leading to the complexity
of model formulations. Thus, there is a need to have an alternative modelling approach to
address these challenges effectively. Unlike the continuum approach, the Discrete Element
Method (DEM) pioneered by Cundall and Strack (1979) has great advantages of replicating the
microstructural features and naturally dealing with the localised failures in materials. In fact, a
cemented material domain in DEM is represented by an assemblage of particles featuring their
physical properties. These particles can move freely but are bonded together at their contacts.
Under external loadings, these contacts degrade and eventually break, forming cracks between
particles. This progress is similar to the propagation process of physical cracks in cemented
materials. Using this discrete nature, Nguyen et al. (2019) first developed a fatigue-bonded
contact model for DEM to investigate fatigue degradation of Cemented-Treated Bases (CTB)
materials. In addition, microstructural effects with changing particle arrangements on the fatigue
behaviour were also captured naturally in the study thanks to the capability of DEM, making it
advantageous over existing continuum approaches. Other straightforward linear bonded contact
models for modelling CTB, incorporating time-dependent fatigue damage, were later proposed
by Zhao et al. (2021a); Zhao et al. (2021b). However, the latter models are limited to opening
mode crack due to neglecting the bond decay caused by shear force, and their DEM simulations
require additional consideration of actual microstructure using digital image processing (DIP)
technology.
The review presented above highlights the main dissipated mechanisms of fatigue responses, i.e.
the gradual degradation of interfaces between cement matrix and aggregates and the crack
4
growth along these interfaces. The discrete nature of DEM and the previous fatigue models
developed for DEM simulations can capture such crack development; however, it is also
necessary to develop a bonded model to characterise both fatigue crack growth and hysteretic
behaviour simultaneously. Therefore, this paper aims to propose a DEM-based modelling
approach, balancing computational cost and an effective constitutive model, to describe the
fatigue behaviour in cemented materials. While DEM has a role in modelling homogeneity and
localised failures of materials, the centrepiece of the proposed approach is a novel fatigue
damage-plasticity constitutive model capable of naturally capturing the fatigue damage
mechanisms and hysteretic behaviour at cement binders. This model adopts the bounding surface
plasticity concept, incorporating a sub-loading surface, which can contract and translate in the
stress space to account for the progressive propagation of fatigue behaviour in cemented
materials. The failure caused by static and fatigue damages is also captured and linked through a
unified constitutive formulation thanks to the interpolation rule. The new fatigue model can
balance simplicity and versatility as it is formulated in two principal directions of the DEM
contact and can be applied to both monotonic and fatigue loading conditions.
Fig. 1. Widening hysteresis loops of the three-point bending test under fatigue loading (after
Bazant and Xu (1991)).
The organisation of the paper is as follows. A brief background of DEM modelling is given in
Section 2. Subsequently, Section 3 presents the fatigue model formulation, implementation
algorithm, and cycle jump technique. Next, the demonstration of model behaviour and
parametric studies at the constitutive level are given in Section 4. This is followed by the
validation and capability of the proposed modelling approach in predicting monotonic and
fatigue responses of cemented materials at structural scales in Section 5. Finally, a conclusion
drawn from this research is summarised in Section 6.
̈ (1)
5
̇ (2)
where and are the centroid displacement and rotational velocity of particle , respectively;
is particle mass, is the acceleration of gravity and is moment of inertia.
Fig. 2. (a) Forces and moments acting on a DEM particle; (b) Normal and shear stress vectors
acting on the bonded contact.
The vector force ( ) and moment ( ) acting on the particle are given by:
∑ (3)
∑ (4)
where and are the external force and moment vectors assigned to particle centre (Fig.
2a), respectively; is the vector connecting the centre of particle and the contacting point of
particle and ; is the number of particles in contact with the -th particle; is the force at
the contact between particles and , which can be calculated based on the contact stress ( )
and the contact representative area ( ) as in Eq. (5); and and are the global
damping force and moment computed using Eqs. (6), (7), and are used to damp out particle
accelerations (Tran et al., 2020; Tran et al., 2021):
(5)
̈ (6)
6
̇ (7)
where is the global damping coefficient and takes a value of suggested by Potyondy and
Cundall (2004); is the resultant force vector of all external, contact and gravity forces acting
on particle i; and is total moment acting on particle ; is the contact stress vector at the
contact between particle i and j; and is the contact area.
The contact stress vector relates to the relative displacement of bonded particles through a
bonded contact model, which represents the mechanical behaviour of cemented bridge between
two aggregates and will be detailed in the next section. After calculating all the necessary forces
and moments acting on each particle, the accelerations at a time step “ ” is then calculated by
using Eqs. (1), (2). The standard leapfrog algorithm is then used to integrate the equations of
motion. Eventually, the translational and rotational motions of particle at the next time step are
updated following:
̇ ̇ ̈ (8)
̇ (9)
̇ (10)
̇ (11)
where is the computational time step that must satisfy the following condition to ensure the
numerical stability of the above time integration:
√ (12)
with is the stability coefficient, and is the ratio of particle mass and spring stiffness.
The relative displacement in a bonded contact can be decomposed into elastic ( , plastic ( )
and fatigue ( ) components as follows:
(13)
The contact stress-displacement relationships in normal and shear directions can then be written
as:
* + [ ]* + (14)
7
where , are the normal and shear stresses in the bonded contact, respectively; is the
damage variable; , are the normal and shear elastic stiffness per initial unit area of the
contact; , are the total normal and shear displacements; , , , and represent the
plastic and fatigue normal, and shear displacements, respectively. The introduction of the
Heaviside step function in Eq. (14) ensures that the damaged part affects only the normal
stress in tension. In fact, when the normal stress changes from tension to compression, tensile
cracks tend to close, leading to the recovery of the compressive stiffness.
To regulate plastic and fatigue growths in bonded contacts, a new model is proposed based on
the below assumptions, which facilitate the calculation of inelastic displacements and stress
developments:
1. There exists a bounding surface representing the overall material strength. This
surface shrinks upon damage evolution to reflect the loss of the load-carrying
capacity of the bonded contact.
2. There exists a subloading surface controlling the growth of fatigue damage. The
subloading surface can translate and contract within the tension-shear domain and
below the bounding surface.
3. The softening modulus of materials varies along the stress trajectory, depending on
the distance between the current stress point on the sub-loading surface and the
projection of it onto the bounding surface.
4. The non-associated flow rule describes the plastic/fatigue flow and variation of the
softening modulus.
Bounding surface
The bounding surface takes the following standard form of the hyperbola:
[ ] * + (15)
√
where , , and , in which is the internal
friction coefficient, and and are the tensile strength and cohesion of the bonded contact,
respectively.
Vertex ( ) and centre ( ) points of the bounding surface are illustrated in the normal-shear
stress space (Fig. 3a). As these points lie on the normal-stress axis, their values can be derived
from the geometric property of hyperbola (i.e. , where ) as follows:
(16)
(17)
The bounding surface gradually shrinks in the stress space as the damage variable evolves from 0
to 1 (Fig. 3b). As the material is almost damaged ( ), the linear relationship between the
normal and shear stresses is obtained from Eq. (15): , which is the classical frictional
Mohr-Coulomb criterion. The shrinkage of the bounding surface is an effective way to
characterise the softening behaviour of cemented materials or soft rocks under mixed-mode
loading conditions (Le et al., 2017; Nguyen et al., 2017a; Nguyen et al., 2017b; Le et al., 2018;
Lu et al., 2022). This shrinkage also provides a smooth transition from the initial yield to the
final failure of the material.
8
Degradation in cemented materials can be caused by static or fatigue loadings, and thus the
damage variable in Eq. (15) should capture these conditions. It is assumed that this variable
can be decomposed into two components, the static damage variable ( ) and the fatigue damage
varibale ( ). When the stress state stays on the bounding surface, the static damage increment
( ̇ ) will occur, while the fatigue damage increment ( ̇ ) develops when the subloading/fatigue
criterion is satisfied. Accordingly, the damage variable is expressed as the sum of both damage
increments:
∑( ̇ ̇ ) (18)
Fig. 3. (a) The vertices and centres of the bounding and subloading surface in the stress space;
(b) The shrinkage of the bounding surface and the subloading surface upon damage evolution.
The plastic potential function for non-associated flow rules is defined as follows:
[ ] * + (19)
√
where is the parameter controlling the non-associativity.
With the defined plastic potential function, the non-associated flow rules defined the plastic
increments in the normal and shear directions as follows:
̇ ̇ ̇ ̇ (20)
̇ ̇
̇ (21)
where is defined in Section 3.1.1, and can be referred to as the plastic softening modulus.
9
Static damage variable takes the following incremental form:
̇ ̇ with ̇ √( ̇ ) ( ̇ ) (22)
[ ] * + (23)
√
where , , and , in which parameter decides
the initial position of the subloading surface in the stress space and evolves upon the growth of
damage variable, and is the parameter controlling the current position of .
It is intriguing to note that the parameter of the subloading surface is mathematically
equivalent to D of the bounding surface in the viewpoint of controlling the shape of these curves.
Based on this observation, the top subloading surface coincides with the bounding one only
when and . In other words, as the stress trajectory reaches the bounding surface,
the memory of past stress reversal events recorded by the subloading surface through is
erased. Similar to the bounding surface, the coordinates of the vertex and centre of the
subloading surface also stay on the axis (Fig. 3a). Therefore, and can be expressed as
(i.e., , where ):
(24)
(25)
It is noted that the subloading surface is a discontinuous domain limited by the top and bottom
subloading surfaces, in which is defined as:
) (26)
where subscripts “ ” and “ ” represent the top and bottom, respectively. The top subloading
surface ( ) is given by the same equation as Eq. (23), namely , whereas the bottom
subloading surface ( ) is obtained by taking the opposite sign of Eq. (23), that is, .
From the physical point of view, the less stress applied to the materials, the larger the fatigue life
becomes, and the effect of fatigue damage will gradually disappear (Sounthararajah et al., 2018;
Baktheer & Becks, 2021). Therefore, it is necessary to assume that there is a fatigue threshold,
10
and fatigue damage occurs when the stress level surpasses this limit. In our model, we assume
that the initial position of the top subloading surface ( ), by substituting initial into Eq. (23),
plays a role as a fatigue threshold. The fatigue damage is activated when the stress state meets
this verge; otherwise, the material behaviour is always elastic. As soon as the fatigue is activated,
the bottom subloadding surface will appear, forming a subdomain between the top and bottom
subloading surfaces. We call this domain the elastic domain of the subloading surface ( ), and
elastic behaviour occurs within this domain.
The hysteresis loops observed in the notched three-point bending beam experiments (Fig. 1)
indicate that the distance between two successive loops becomes larger and larger under further
cyclic loadings. These hysteresis loops also represent the degradation of stiffness modulus and
the progressive damage development in the material. To capture these experimental
observations, the parameter in Eq. (26) is assumed as:
(27)
where is the initial value of the elastic domain.
It is obvious that upon the development of damage variable , the elastic domain shrinks,
leading to the contraction of the subloading surface, which is considered the dissipated fatigue
process in materials. Referring to the stress path OSS’ in Fig. 3a, the behaviour is elastic for SS’
and becomes elastoplastic when the stress point S’ moves outside the subloading surface,
resulting in its translation and shrinkage. Such movement and contraction of the subloading
surface help the model capture hysteretic behaviour naturally, which will be demonstrated in
Section 4.2 From Eq. (27), when reaches unity, the elastic domain has vanished.
Consequently, the bounding and subloading surfaces will coincide and become one straight line,
corresponding to the final damage stage of the bonded contact (Fig. 3b).
The non-associated flow rule for fatigue growth is written in a form consistent with the
representation in Eqs. (20):
̇ ̇
̇ ̇ (28)
where the subscript “ ” indicates the top and bottom subloading surfaces when the stress point
lies on the top and bottom subloading surfaces, respectively; and is the fatigue potential
function corresponding to the subloading surface , that is, and with
being the general fatigue potential function of the subloading surface, and takes the following
form:
[ ] * + (29)
√
The value of in Eq. (28) is determined by using the interpolation rule:
(30)
( )
In the above equation, is the softening modulus evaluated on the bounding surface at the
image point. As shown in Fig. 3a, if the current stress point is at I, its image point (i.e. point K) is
11
defined as the intersection of the line OI and the bounding surface. In addition, IK is the current
distance between the stress and the image points, and is the parameter controlling the
development of fatigue deformation when the subloading surface translates. As the stress point
is approaching point (i.e. ), ; by contrast, while , there is ,
indicating no fatigue damage. As a result, the above interpolation rule helps capture the essential
physics of fatigue response in cemented materials. The higher the stress level applied to the
materials, the more inelastic or fatigue deformation occurs.
To describe the fatigue-induced damage development under cyclic loadings, the incremental
fatigue damage variable is provided as:
̇ ̇ with ̇ √( ̇ ) ( ̇ ) (31)
The fatigue damage rate is closely linked to static damage increment through the “image rule”.
Eq. (31) returns to Eq. (22) when the stress point I approaches K. This implies that the new
model successfully connects static/brittle and fatigue laws together based on the interpolation
rule. More importantly, unlike existing DEM contact models for fatigue, which directly link the
fatigue damage variable to the number of cycles (Nguyen et al., 2019), the model proposed in
this paper is independent of the number of cycles. As a result, our proposed model can be applied
to cyclic/fatigue loading conditions, wherein the cycles are non-periodic with variable
amplitudes and shapes. Considering the kinematics of the subloading surface with supports of
the geometric properties of hyperbolas (Fig. 3a), the following relations hold:
(32)
The above relation assumes that the position of the subloading surface cannot overtake the
bounding surface. Additionally, as the vertices and centres of bounding and subloading surfaces
stay on the normal stress axis, the translation rules may be formulated by postulating that the
relative motion of point with respect to is directed along . In other words, ̇ ̇ ̇
with ̇ being the scalar factor. As the bounding surface shrinks along the -axis, the subloading
surface also translates and contracts simultaneously along this axis, the following incremental
relations can be established for the change in position of and :
̇
̇ ̇ (33)
̇
̇ ̇ (34)
Substituting Eq. (32) into Eq. (35), the kinematics of the subloading surface can be described as:
̇ ̇
̇ ̇ ̇ (36)
12
The scalar parameter ̇ in Eq. (36) can be determined from the consistency condition ̇ :
̇
̇ (37)
where
(38)
̇ ̇ (39)
By substituting Eqs. (37), (38), (39) into Eq. (36), we have:
̇ ̇ ̇
̇ [ ̇ ] (40)
̇ ̇ ̇ (43)
̇ ̇ ̇ √( ) ( ) ̇ ̇ (44)
By substituting Eqs. (20) and (44) into Eq. (43), the scalar ̇ is determined as
̇
(45)
13
The increments of internal variables can then be computed easily following Eqs. (20) and (22).
Accordingly, the corrected stress is updated as:
̇ ̇ (46)
Similar to the algorithm used for the bounding surface, the same algorithm is used to update the
stress and calculate internal variables when the stress point lies on the subloading surface.
However, OK and IK in Eq. (30) should be computed before updating stress. As point I is now
in the trial step, its coordinates are , thereby the slope of the line OI is
Combined with Eq. (15), we obtain:
( ) (47)
(48)
√
where
The smaller solution is chosen as we only consider a quarter of bounding and subloading
surfaces in the stress space. Subsequently, the trial shear stress of point K is computed as
. The values of OK and IK can then be calculated as follows:
√( ) ( )
(49)
√( ) ( )
Once the stress is corrected based on the values of IK, OK and other internal variables, the
kinematic hardening is then updated using Eq. (41). The detailed stress return algorithm for
the proposed fatigue model is shown in Algorithm 1.
14
3.1.2 Cycle jump technique
To correctly predict the fatigue degradation process, the simulation should track the whole
trajectory of cycle-by-cycle damage states. This cycle-tracking process often requires a
substantial computational cost when involved in high-cycle fatigue (Han et al., 2020), but it
would become unnecessary for applications related to constantly repeated loads. For this reason,
the cycle jump technique has been proposed to save computational costs for these applications
(Van Paepegem & Degrieck, 2001; Turon et al., 2007). In this study, the cycle jump technique is
also adopted to obtain computational efficiency for applications related to constantly repeated
loads. Concerning our proposed contact model, this technique means that if the damage
development in the bonded contacts is stable in a certain set of loading cycles, the simulation can
jump from the first to the final load cycles of the set, and damage degradation of these loading
cycles is determined by extrapolation strategy (Rezazadeh & Carvelli, 2018; Nguyen et al.,
2019). The main idea of the technique can be illustrated in Fig. 4a, in which the damage after
cycles is extrapolated for the damage after cycles. If the damage evolution of a bonded
contact as a function of the number of cycles is known, the gradient of this function ( )
can be expressed as:
(50)
The number of cycles that can be jumped is then determined using the following equation:
(51)
, -
15
Fig. 4. (a) D-N relationship; (b) Schematic representation of cycle jump method (Van Paepegem
& Degrieck, 2001)
To complete the cyclic jump technique, the damage rate with respect to the number of loading
cycles is calculated as follows (Jimenez & Duddu, 2016):
( ) (52)
∑( ) ̇ ̇ (53)
where n is the number of simulation steps up to the end of the cycle jump.
16
4 Model behaviour
The stress-return algorithm presented in Algorithm 1 is implemented into the Particle Flow Code
(PFC 6.0) (Itasca, 2017), and it is used in all simulations presented in this study. To demonstrate
the constitutive performance of the proposed fatigue model, a sample created by bonding two
particles with a diameter of 0.1 m is simulated under monotonic loading of mode I/II and mixed-
mode cyclic loading conditions. For simulations under monotonic loads, a constant velocity is
applied to the particles in corresponding directions until reaching certain displacement values. It
is noticed that the simulation under the mode II loading is conducted under zero normal stress.
The behaviour of the proposed model under monotonic loading is presented in Section 4.1.
Meanwhile, the loading condition and the model performance under cyclic load are described in
Section 4.2. The contact model parameters of simulations in this section are presented in Table 1.
Table 1 The model parameters used in 2-particle tests
Model parameters Selected values
Elasto-plastic parameters
(Pa/m) 1012
(Pa/m) 1012
(Pa) 3.6106
(Pa) 40106
0.84
0.65
0.70
2
Fatigue parameters
0.45
0.13
1.0
(m) 210-5
17
Fig. 5. Stress-displacement curves: (a) Mode-I test; (b) Mode-II test.
Fig. 6. (a) Cyclic stress loading path (BC); (b) Bounding and subloading surfaces
18
Fig. 7 shows the results of the cyclic mixed-mode test when the cyclic normal load is applied.
The shear stress-shear displacement relation is a horizontal line due to the imposed constant
shear stress condition (Fig. 7a). Under this constraint and due to the dilation effect, shear
displacement grows. Meanwhile, the normal stress-displacement curve first undergoes a linear
elastic stage, followed by the elastoplastic behaviour when the load-unloading process is started.
As the stress point moves along with the subloading surface (during the load-unloading process),
the accumulated inelastic displacements are observed to increase with the rising number of
cycles (Fig. 7b). This growth contributes to fatigue damage development (Fig. 7c). In addition,
the ratio (referred to Fig. 6b) decreases because the bounding and subloading surfaces
shrink simultaneously under the development of the damage variable. The rate of shrinkage can
be reflected by the reduction in the elastic domain ( - Fig. 7c), which controls the size of the
subloading surface. As a result, the contraction of the elastic domain considerably affects
displacement-stress behaviour, which is shown by much wider hysteresis loops for the last cycles
(Fig. 7a). These larger loops also demonstrate the evolution of the damage variable, representing
the gradual reduction of loading/unloading paths.
One of the key features of the proposed fatigue model is the kinematic hardening parameter ( ),
which plays a vital role in controlling the kinematics of the subloading surface. Fig. 7d illustrates
the variation of with the number of simulation steps or loading cycles. At first, is positive
and moves in a narrow range. However, as the number of simulation steps increases, switches
its sign and locates in the growing band. To explain this variation, and , which are the
intersections between the top/bottom subloading surface and the normal-stress axis, respectively
(Fig. 6b), are calculated based on Eq. (24) as follows: ( ) ,
( ) . It is noted that the maximum value of and the minimum value of are
almost constant during constant cyclic loading tests. Fig. 7c shows that the elastic domain
between the top and bottom part of the sub-loading surface reduces as the damage variable
grows. Moreover, because the subloading surface size contracts, the value of tends to rise
(Eq. (26)), resulting in a decline of ( ) . Consequently, to keep the maximum value of
constant, the value of needs to go up. By contrast, the reduction of contributes to the
decrease of when the stress point is on the bottom subloading surface. Accordingly, the -
value band becomes larger, whereas the range of is the same (Fig. 7d). It is interesting that as
soon as the top subloading surface coincides with the bounding surface, the value is
automatically reset to be 0 (mentioned in Section 3.1). The memory subloading surface via is
thus eliminated, and the material response is controlled by the bounding surface, showing a
softening behaviour in Fig. 7a.
19
Fig. 7. Model behaviour: (a) Stress-displacement relation; (b) Accumulated fatigue
displacements; (c) Parameter values; (d) Kinematic hardening parameters.
20
Fig. 8. Influences of fatigue parameters on the damage evolution in the mix-mode cyclic tests:
(a) Effect of ; (b) Effect of .
Fig. 8a shows the effect of on the development of damage variable. The smaller initial elastic
domain contributes to the faster development rate of the damage variable. For example, to reach
the damage value of 0.4, the test with needs 38 loading-unloading cycles, while other tests
with and require 48 and 71 cycles, correspondingly. This is because the smaller initial
elastic domain allows the stress point to have more room to move in the stress space (Fig. 6b).
As the stress point translates with the sub-loading surface, the fatigue deformation (calculated by
using Eq.(28) would develop, thereby increasing the development of the fatigue damage variable
(Eq. (31)). At the same time, the shrinkage of the elastic domain would also accelerate the
growth of fatigue deformation (via Eq. (26)), and this process keeps continuing until the bonded
contact failures. A similar trend is also observed for the effect of on damage development
illustrated in Fig. 8b. The damage growth rate is inversely proportional to the value of , which
is consistent with Eqs. (28), (30). The larger value of is, the smaller value of fatigue
deformation generates, hence the lower damage evolution rate.
21
particle size and distribution change will require the recalibration of constitutive parameters
before using the model for parametric studies.
5.1 The general process of calibrating model parameters
Parameters used in the proposed DEM simulations are mesoscale constitutive parameters (i.e.,
the scale of contact bond), which cannot be directly measured from the standard laboratory
experiments (i.e. macroscale). As a result, the mesoscale constitutive parameters used in DEM
simulations are usually obtained by matching the numerical results to the experimental
counterparts based on a trial-and-error process. The proposed fatigue constitutive model requires
two sets of parameters: elastoplastic and fatigue parameters. The first set of parameters can be
subdivided into four sub-groups: bonded stiffness ( , ) associated with elastic behaviour;
tensile strength ( , cohesion ( ); softening parameters ( , , ) associated with fracture
energies deciding post-peak responses; bond friction ( ) and dilatancy parameter ( ) associated
with the frictional and dilatant behaviour of the bonded contact, respectively. All the parameters
in this set are first calibrated with data from the strength test. These parameters are then used to
calibrate the fatigue parameters in the simulation of a fatigue test. The fatigue threshold ( ),
controlling the number of contacts subjected to fatigue damage, is first calibrated by fitting the
numerical results to the deformation of experimental samples at the first stage of the modulus
degradation curve or crack tip displacement histories. Parameters and are calibrated to
match the deformation of specimens at the second and final stages. Finally, parameter is
obtained by fitting the fatigue life. For the cycle jump technique, the fracture energy is the
area of the load-displacement curve obtained from the strength test, while and are
calibrated to match the fatigue life.
In addition, the deformability method (Potyondy & Cundall, 2004) is adopted in the calibration
process. The normal and shear contact stiffnesses are now calculated from contact effective
modulus, , and normal/shear stiffness ratio ( ) as follows:
(54)
(55)
where and are the radius of two intact particles, is Poison’s ratio, is Young’s modulus
and is the shear modulus. The values of and can be the same for all bonded contacts
and are obtained through the calibration process.
22
Subsequently, these particles are allowed to rearrange until they reach their equilibrium locations
and are bonded together. The wall elements are then deleted, and four new rigid wall elements
with a width of 4 mm are generated to simulate the loading and supporting rollers (Fig. 9b),
which is referred to as Sample 1. Note that the proposed model is employed in all particle-
particle contacts, while the linear model with frictionless interactions (Potyondy & Cundall,
2004) is assigned to particle-wall contacts. More importantly, only a limited number of particles
is required for the simulations, which significantly reduces the computational cost for high-
fatigue tests thanks to the phenological contact model at the mesoscale.
As the experimental samples were tested using the stress-control mode, a servo-control technique
was adopted to replicate the experimental loading pattern in the simulations. For the monotonic
test (strength test), the specimen is loaded with a seating force of 50 N to bring the loading
rollers into contact with the specimen. The load is then increased at an equivalent rate of 833
N/min until the specimen ruptures. The applied load ( ) and mid-span deflection ( ) are
recorded during the loading processes. The flexural stress, flexural strain and flexural modulus
are then calculated using the elastic theory of beam as follows (Austroads, 2014):
(56)
(57)
(58)
where is the flexural strain, is the span length, is the beam height, and B is the beam
width, is the flexural stress, is the flexural modulus, is the maximum magnitude of
applied cyclic load, and is the deflection at the centre of the sample corresponding to .
Fig. 9. Numerical set-up for DEM simulations with different levels of randomness: (a) Specimen
geometry and boundary conditions for the four-point bending test; (b) Numerical Sample 1; (c)
Numerical Sample 2; (c) Numerical Sample 3.
23
Like the monotonic test, the flexural fatigue test was also conducted using the stress-control
program with a seating load of 50 N. Subsequently, the test involves applying cyclic haversine
load pulses of 9.1 Hz frequency. The maximum cyclic load is 900 N, corresponding to 77% of
the maximum flexural stress obtained in the strength test ( ), whereas the
minimum cyclic load is equal to the seating load, 50 N.
The calibrated parameters for simulations in this section are summarised in Table 2. The cycle
jump technique is only adopted for fatigue tests using the following parameters: ,
, .
24
extends toward the top surface until the beam fails. Therefore, the proposed model demonstrates
its capability to capture the damage response of cemented materials under monotonic loads.
Fig. 10. Strength test of four-point bending beam: (a) Flexural stress-strain curves; (b) Cracking
path in the experiment and DEM simulation.
25
stage of the modulus degradation curve experiences a stable drop. It should be noticed that there
are blips in the curve of fatigue damage contacts at cycles 100 and 1200 and the final stage of the
fatigue test. The reason behind these blips is a significant rearrangement of particles and bonded
contacts in the beam test. In fact, when the beam experiences considerable rearrangement, the
stress redistribution occurs within the specimen, resulting in a smaller number of contacts
involved in the fatigue damage process. However, as particles and contacts gradually rearrange
to a more stable condition, more fatigue contact numbers take part in the fatigue damage
procedure, showing a recovery in the number of fatigue contacts after a temporary drop.
It is also noticed that during the second phase, within which the fatigue damage contact numbers
reduce, there is a gradual increase in the number of static damage contacts, indicating a damage
transition process from fatigue damage to static damage. The transition can be explained by the
shrinkage of the bounding surface, representing the loss of the material strengths of bonded
contacts. The shrinkage of the bounding surface not only contributes to the increase in the
number of contacts involved in static damage but also results in less room for fatigue damage to
grow under reversal loadings. Consequently, fewer fatigue contacts are observed under further
fatigue loadings. In the final phase, when the microcracks propagate and coalesce to form a
major crack, the modulus curve declines rapidly, showing the sudden rupture of the beam under
stress-controlled loading. This sudden failure is caused by increased static damage contacts (i.e.
32 contacts) participating in the damage process, as shown in Fig. 12. Besides, the blips are seen
clearly at the last stage, wherein the specimen undergoes significant rearrangements before
failure.
Fig. 11. (a) The flexural stress-strain responses in the simulation of the monotonic and fatigue
tests obtained from Sample 1; (b) Comparison of the modulus reduction between different
numerical samples and experimental data in fatigue testing.
26
Fig. 12. Evolutions of fatigue damage contacts (i.e. contacts undergoing fatigue damage) and
static damage contacts (i.e. contacts undergoing static damage) during the fatigue test of Sample
1.
To further demonstrate the capability of the proposed model to capture the fatigue behaviour of
CTB, Sample 1 is repeated with different stress amplitudes, i.e., 0.8, 0.82, 0.85, 0.91 and 0.96 of
. Fig. 13a illustrates the flexural modulus degradation curves of the virtual samples under
different stress amplitudes. When the applied stress level increases, the modulus degradation rate
rises substantially, and fatigue failure occurs soon in the samples. The modulus reduction is the
result of the development of permanent deformation. As a higher stress ratio is applied to the
specimen, the stress at contacts is higher, causing more inelastic deformation as the stress
reaches the subloading surface quicker, and thus the modulus degradation decreases faster in the
specimen. This behaviour was also reported in experimental studies (Sounthararajah et al., 2018;
Zhao et al., 2021a; Zhao et al., 2021b).
The fatigue damage evolution in the most damaged bonded contact corresponding to each
numerical test is shown in Fig. 13b. It can be seen that the larger the stress amplitudes, the less
fatigue damage before the rupture of samples. All fatigue damages are less than one at 100%
fatigue life. This means that the specimen failure is determined by the material strength bounding
surface criterion. Higher stress magnitudes in bonded contacts will satisfy this surface criterion
sooner, causing the sample to fail earlier, whereas the fatigue damage growth remains small. In
other words, the specimen rupture is more brittle when the stress level is higher. This illustrates
the essence of linking the fatigue damage rate (Eq. (31)) to the static damage rate (Eq. (22))
through the interpolation rule. Fig. 13c reveals that the proposed model can capture the fatigue
life obtained from experimental results. The sample fatigue life decreases with an increase in the
maximum cyclic stress level. There is a small difference between experimental and numerical
results, but this difference becomes negligible as the maximum cyclic stress level decreases. This
variation may be due to the variation in the internal structure of experimental and numerical
samples. Fig. 13d shows the modulus degradation curves, in which the number of cycles is
normalised with the maximum number of cycles of each curve. Upon the increase of stress
levels, the normalised lines tend to shift up, but they fall into a narrow band, except for the stress
27
levels of 0.91 and 0.96 . This small band suggests that within a certain range of stress
amplitudes, the normalised S-N curve can be considered as a macro material property for model
calibration. Thanks to this, one can take advantage of the normalised curve to validate the model
with a suitable high-stress level, for example, 0.85 in this research. This is because the
higher stress magnitude resulting in lower fatigue life will save a significant amount of
computational time to complete the fatigue tests. The simulation results confirm the
recommendation discussed in Austroads (2014), in which the magnitude of the applied load
pulses in the fatigue tests is recommended not to be larger than 90% range of the breaking load.
Fig. 13. (a) The modulus degradation curves obtained in the numerical fatigue tests with different
stress levels; (b) The fatigue damage value of the most degradable contact at the bottom surface
of specimens corresponding to six stress ratios; (c) The S-N curves; (d) The normalised modulus
curves.
28
capability of the proposed model to capture fatigue behaviour under variable levels of cyclic
loads by simulating experiments conducted by Baktheer et al. (2021).
5.3.1 Simulation setup
The sample geometry and loading conditions of the three-point bending test are illustrated in Fig.
14. The numerical sample is reproduced using 3420 DEM particles with radii varying uniformly
from 2.5 to 4 mm. This specimen is loaded by a wall element placed in the middle of the top
surface and supported by two other wall elements. During the loading process, the magnitude of
the applied load and crack tip opening displacement (CTOD) of the notch are recorded. For the
strength test, the beam is loaded under strain-controlled mode at a rate of 1mm/min, while the
fatigue test is conducted under variable loading amplitude with a sinusoidal waveform of 20 Hz
frequency. The lower load level is held constant at Smin = 0.1Peak, where Peak is the highest
load the sample can carry in the strength test. Meanwhile, the upper load level begins with S max =
0.5Peak and is then gradually increased after 10 cycles with the rise by S = 0.05, shown in Fig.
15. The calibrated parameters are summarised in Table 2.
Fig. 14. The three-point bending test: (a) Geometry and experiment setup; (b) Numerical
specimen.
Fig. 15. The loading amplitude used in the three-point fatigue test.
29
5.3.2 Model predictions
Fig. 16a compares the experimental and numerical results for the applied load versus CTOD of
the specimen under the static load. An excellent agreement between the predictive response of
the proposed modelling approach and the experiments is achieved. The load-CTOD curves of all
samples in this test undergo three distinct stages. The predictive curve begins with linear
behaviour until reaching the elastic limit force of 14 kN, followed by the hardening behaviour up
to the peak load of 18.7 kN. In the final stage, the softening behaviour is observed, represented
by the significant drop in load after the peak. The load then gradually decreases until the rupture
of the beam. The softening observation is attributed to the strain-controlled mode adopted in the
monotonic test of the three-point bending beam. This is opposite to the residual phase after
reaching the flexural strength, as the four-point bending specimen is conducted under stress-
controlled loading in Section 5.2.2. The hardening and softening behaviours are the consequence
of the combined effects of plastic and damage growth at multiple contacts. Indeed, as the sample
is loaded, contacts experience different stress levels and yield at different times, thereby
observing the hardening behaviour. Meanwhile, the softening behaviour is recorded when most
contacts have yielded. Fig. 16b shows the cracking path developed in the numerical sample,
which was not reported in the experimental study. The major crack initiates from the notch tip,
and then propagates toward the top surface during the softening behaviour of the loading curve.
Compared to the sudden failure of the four-point bending beam in the stress-controlled strength
test, the three-point bending beam fails gradually due to the application of the strain-controlled
program.
Fig. 16. Three-point bending strength test: (a) Load-CTOD curves; (b) Cracking pattern at the
end of the test.
In the cyclic simulation, as the maximum number of loading cycles in this fatigue test is only
100, the cycle jump technique was not adopted. Instead, the numerical fatigue test is loaded
cycle by cycle, as tracing the complete path of successive damage stages is the main priority.
Fig. 17a compares the load-CTOD response obtained in the cyclic test to that of the monotonic
test. The proposed model can predict the inelastic deformation development under repeated loads
with the maximum cyclic loading amplitudes much lower than the peak load in the strength test.
Compared to the load-CTOD curve obtained from the experiments presented in Fig. 17b, the
proposed model can generally capture the gradual increase in CTOD for the first 50 cycles, then
a rapid rise for several last cycles, which can be attributed to the accumulated fatigue damage
30
due to the increase in loading amplitude, facilitating the shrinkage of the bounding surface and
thus the sample rapture. This agreement with the experiment demonstrates the capability of the
proposed model in capturing cyclic tests subjected to variable loads.
The fatigue damage ( ) and total damage ( ) variables at three contacts shown in Fig. 17c are
plotted in Fig. 17d. These are the three most degraded contacts along the cracking path in the
fatigue test. It can be seen that the fatigue and total damages increase continuously during cyclic
loading. The damage variable of contact 1 starts at the first cycle, while it is delayed for other
contacts. For example, the damage variable of contact 2 and 3 develop only when the number of
cycles is 10 and 40, respectively. This observation is consistent with the propagation of the
cracking path, which starts at the crack tip (i.e., contact 1) and then develops and propagates
upwards to contacts 2 and 3 under cyclic loadings. For the first 50 cycles, the total damage is
mainly contributed by the fatigue damage. However, after cycle 50, the total damages increase
dramatically while the accumulated fatigue damages grow considerably until the sample
ruptures. At this stage, the total damage value is nearly 1, whereas the maximum value of the
fatigue damage in three contacts is less than 0.6. This implies that the static damage ( ̇ ̇
̇ ), controlled by the bounding surface which stands for material strength, contributes
enormously to the rupture of the fatigue test.
One intriguing thing is that the total damage (Fig. 17b) rises quickly at the cycle of increasing
applied load, corresponding to the new CTOD peak, as shown in Fig. 17a and b. The jump in the
new CTOD peak is called the effect of sudden overloads when the fatigue test is conducted
under variable amplitudes (Ray & Chandra Kishen, 2012; Kirane & Bažant, 2015). This
influence is seen in Fig. 17b, where the proposed model can capture the sudden increase in the
maximum experimental CTOD for the first 40 cycles. Although predicted CTOD matches
somewhat with their counterparts in the experiments for the remaining cycles, the proposed
model shows its ability to capture the impact of overloads through rapid growth in CTOD until
the sample failure at cycle 84, which is close to the fatigue life of 86 cycles in the experiment.
Based on this outcome, the applicability of bounding surface plasticity theory and independent-
cycle fatigue damage law is demonstrated in capturing fatigue response under variable loadings.
31
Fig. 17. Results of three-point bending test under cyclic loading: (a) Load-CTOD responses from
simulation; (b) CTOD-N relations; (c) The most three degradable contacts near the notch; (d)
Damage variable-N responses of three bonded contacts.
32
6 Conclusion
This work proposes a discrete-based approach to model the fatigue behaviour of cemented
materials, wherein DEM is employed to replicate the heterogeneity and crack development in
cemented materials. The centrepiece of the paper is the development of a novel bonded cohesive-
frictional fatigue model at the grain scale, describing the degrading response of interfaces
between aggregates and cement binders subjected to cyclic loadings. This new model couples
damage mechanics and bounding surface plasticity theory, thus giving rise to a more natural
approach to describe the inelastic and hysteresis responses of cemented materials. Moreover,
instead of treating static and fatigue damage evolutions separately, the proposed model
successfully connects both damages through the interpolation rule, enabling a smooth transition
from static damage to fatigue.
A fair agreement between numerical and experimental results demonstrates the abilities of the
proposed modelling approach to predict the fracture and fatigue responses of cemented materials
under both monotonic and cyclic loadings. For the strength tests, the model captures well the
softening and hardening behaviour of materials corresponding to strain-controlled and stress-
controlled modes. Furthermore, the proposed method can reproduce the flexural modulus
degradation curves and the fatigue lives of four-point bending beams with different stress ratios.
The influence of stress levels on the macro fatigue responses is naturally captured thanks to the
variation of softening modulus depending on the stress magnitude at the constitutive level.
Besides, the modelling approach can predict the histories of crack tip displacements under cyclic
variable load amplitudes thanks to the fatigue damage variable which is independent of the
number of loading cycles. This distinctive feature shows the potential of the model to apply to
reality, wherein the practical structures can be subjected to non-periodic cyclic loadings with
variable amplitudes, shapes or even sudden overloads. In addition, the effectiveness of the novel
fatigue constitutive model in characterising dissipative processes within materials under cyclic
loadings can open more potentials for better interpreting of fatigue failure mechanism.
33
Appendix A
Algorithm 1. Stress return algorithm of the proposed fatigue model
Stress return algorithm
Input value: ̇ ; Output values: ̇
Initial value: , , threshold = 0, bottom = 0. Threshold is fatigue limit depending on the initial value of
Constant: , , , , , , , , ,
{[ ] [ ]} { [ √ ]}
{[ ] [ ]} { [ √ ]}
{[ ] [ ]} { [ √ ]}
If then
threshold =1
End
If ( ) then (On bounding surface)
̇
Else
If (( ) and (threshold==1)) then (On top subloading surface)
Compute internal variables: IK, OK, SK, SI and then ̇
Compute increment of internal variables: ̇ , ̇
Compute correct stress: ̇ ̇
Update internal variables: ̇ , ̇ ,
Compute and update kinematic variables: , ̇
Else
34
Appendix B
In this section, we demonstrate the capability of the proposed modelling approach to predict
material behaviour in several standard tests.
B1. Direct tensile test
This test is conducted under both monotonic and fatigue loading conditions. The monotonic test
is conducted under a strain-controlled program, while the fatigue test is carried out under a
stress-controlled program. Both tests adopted the same material constitutive parameters as the
four-point bending test in Section 5.2 and presented in Table 2. The particle size ranges from
1mm to 2mm. The geometry, loading, and boundary conditions for the direct tensile test are
illustrated in Fig. 18a. During the loading processes, the displacement at the top and the reaction
force at the bottom of the specimen are recorded and are displayed in Fig. 18b through the load-
displacement response. Since the monotonic test was conducted under a strain-control program,
the load-displacement curve exhibits rapid softening behaviour after reaching the peak load,
followed by a gradual decrease in the final stage of failure.
(a) (b)
Fig. 18. (a) Geometry, loading and boundary conditions of the direct tensile test; (b) The
predictive load-displacement curve of the monotonic test and damage evolution of bonded
contacts obtained at the end of the test.
Next, the above test is repeated for the cyclic load (i.e., fatigue test) using a stress-controlled
program with the maximum and minimum loads set at 80% and 10% of the peak load
determined in the monotonic test. Fig. 19a illustrates the predicted load-displacement response
of the fatigue test. Under cyclic loading, the permanent displacement gradually increases, as
shown by the displacement moving away from the origin at the end of the unloading path. The
permanent displacement initially increases dramatically, followed by a stable stage before
sharply rising towards failure, as depicted in Fig. 19b.
35
(a) (b)
Fig. 19. (a) Load-displacement response for the direct tensile test under fatigue loading; (b)
Displacement-number of cycles (N) relationship under fatigue loading
36
(a) (b)
Fig. 20. (a) The high-low two-step loading amplitude used in the three-point bending fatigue
test; (b) Load-crack tip opening displacement response obtained from simulation.
37
Acknowledgement
This research work is part of a research project (Project No IH18.02.4) sponsored by the SPARC
Hub (https://sparchub.org.au) at the Department of Civil Engineering, Monash University,
funded by the Australian Research Council (ARC) Industrial Transformation Research Hub
(ITRH) Scheme (Project ID: IH180100010). The financial support from Monash International
Scholarship (Le), and the Australian Research Council via Future Fellow Project FT200100884
(Bui) are also gratefully acknowledged. Part of this research was undertaken with the assistance
of resources and services from the National Computational Infrastructure (NCMAS-2023-101).
38
References
Alliche, A. (2004). Damage model for fatigue loading of concrete. International Journal of
Fatigue, 26(9), 915-921.
Aslani, F., & Jowkarmeimandi, R. (2012). Stress–strain model for concrete under cyclic loading.
Magazine of concrete research, 64(8), 673-685.
Austroads. (2014). Cemented Materials Characterisation: Final Report. Austroads technical
report, Austroads, Sydney, Australia AP-R462-14.
Baktheer, A., Aguilar, M., & Chudoba, R. (2021). Microplane fatigue model MS1 for plain
concrete under compression with damage evolution driven by cumulative inelastic shear
strain. International Journal of Plasticity, 102950.
Baktheer, A., & Becks, H. (2021). Fracture mechanics based interpretation of the load sequence
effect in the flexural fatigue behavior of concrete using digital image correlation.
Construction and Building Materials, 307, 124817.
Barenblatt, G. I. (1962). The mathematical theory of equilibrium cracks in brittle fracture.
Advances in applied mechanics, 7, 55-129.
Bazant, Z. P., & Xu, K. (1991). Size effect in fatigue fracture of concrete. ACI Materials
Journal, 88(4), 390-399.
Bodin, D., Pijaudier-Cabot, G., de La Roche, C., Piau, J.-M., & Chabot, A. (2004). Continuum
damage approach to asphalt concrete fatigue modelling. Journal of Engineering
Mechanics-ASCE, 130(6), 700-708.
Bui, H. H., & Nguyen, G. D. (2021). Smoothed particle hydrodynamics (SPH) and its
applications in geomechanics: From solid fracture to granular behaviour and multiphase
flows in porous media. Computers and Geotechnics, 138, 104315.
Chen, X., Huang, Y., Chen, C., Lu, J., & Fan, X. (2017). Experimental study and analytical
modeling on hysteresis behavior of plain concrete in uniaxial cyclic tension.
International Journal of Fatigue, 96, 261-269.
Cho, B. H., Nam, B. H., & Khawaji, M. (2021). Flexural fatigue behaviors and damage evolution
analysis of edge-oxidized graphene oxide (EOGO) reinforced concrete composites.
Cement and Concrete Composites, 104082.
Cundall, P. A., & Strack, O. D. (1979). A discrete numerical model for granular assemblies.
geotechnique, 29(1), 47-65.
Dafalias, Y., & Popov, E. (1975). A model of nonlinearly hardening materials for complex
loading. Acta mechanica, 21(3), 173-192.
Daneshyar, A., & Ghaemian, M. (2017). Coupling microplane-based damage and continuum
plasticity models for analysis of damage-induced anisotropy in plain concrete.
International Journal of Plasticity, 95, 216-250.
Desmorat, R., Ragueneau, F., & Pham, H. (2007). Continuum damage mechanics for hysteresis
and fatigue of quasi‐ brittle materials and structures. International journal for numerical
and analytical methods in geomechanics, 31(2), 307-329.
Dugdale, D. S. (1960). Yielding of steel sheets containing slits. Journal of the Mechanics and
Physics of Solids, 8(2), 100-104.
Eliáš, J., & Le, J.-L. (2012). Modeling of mode-I fatigue crack growth in quasibrittle structures
under cyclic compression. Engineering Fracture Mechanics, 96, 26-36.
Fang, H., Zheng, H., & Zheng, J. (2017). Micromechanics-based multimechanism bounding
surface model for sands. International Journal of Plasticity, 90, 242-266.
Han, S., Yang, X., Shi, D., Miao, G., Huang, J., & Li, R. (2020). Microstructure-sensitive
modeling of competing failure mode between surface and internal nucleation in high
cycle fatigue. International Journal of Plasticity, 126, 102622.
39
Hillerborg, A., Modéer, M., & Petersson, P.-E. (1976). Analysis of crack formation and crack
growth in concrete by means of fracture mechanics and finite elements. Cement and
concrete research, 6(6), 773-781.
Isojeh, B., El-Zeghayar, M., & Vecchio, F. J. (2017). High-cycle fatigue life prediction of
reinforced concrete deep beams. Engineering Structures, 150, 12-24.
Jimenez, S., & Duddu, R. (2016). On the parametric sensitivity of cohesive zone models for
high-cycle fatigue delamination of composites. International Journal of Solids and
Structures, 82, 111-124.
Jin, G., & Huang, X. (2014). Predicting Fatigue Damage of Asphalt Concrete Using a Cohesive
Zone Model. International Journal of Pavement Research & Technology, 7(5).
Kachkouch, F. Z., Noberto, C. C., Babadopulos, L. F. d. A. L., Melo, A. R. S., Machado, A. M.
L., Sebaibi, N., Boukhelf, F., & El Mendili, Y. (2022). Fatigue behavior of concrete: A
literature review on the main relevant parameters. Construction and Building Materials,
338, 127510.
Keerthana, K., & Kishen, J. C. (2018). An experimental and analytical study on fatigue damage
in concrete under variable amplitude loading. International Journal of Fatigue, 111, 278-
288.
Keerthana, K., & Kishen, J. C. (2020). Micromechanics of fracture and failure in concrete under
monotonic and fatigue loadings. Mechanics of Materials, 148, 103490.
Kirane, K., & Bažant, Z. P. (2015). Microplane damage model for fatigue of quasibrittle
materials: Sub-critical crack growth, lifetime and residual strength. International Journal
of Fatigue, 70, 93-105.
Krieg, R. (1975). A practical two surface plasticity theory.
Le, L. A., Nguyen, G. D., Bui, H. H., Sheikh, A. H., & Kotousov, A. (2018). Localised failure
mechanism as the basis for constitutive modelling of geomaterials. International Journal
of Engineering Science, 133, 284-310.
Le, L. A., Nguyen, G. D., Bui, H. H., Sheikh, A. H., Kotousov, A., & Khanna, A. (2017).
Modelling jointed rock mass as a continuum with an embedded cohesive-frictional
model. Engineering Geology, 228, 107-120.
Lemaitre, J. (1972). Evaluation of dissipation and damage in metals submitted to dynamic
loading. Journal of the Mechanical Behavior of Materials.
Lemaitre, J. (2012). A course on damage mechanics. Springer Science & Business Media.
Liang, J., Ren, X., & Li, J. (2016). A competitive mechanism driven damage-plasticity model for
fatigue behavior of concrete. International Journal of Damage Mechanics, 25(3), 377-
399.
Lu, D. X., Nguyen, N. H., & Bui, H. H. (2022). A cohesive viscoelastic-elastoplastic-damage
model for DEM and its applications to predict the rate-and time-dependent behaviour of
asphalt concretes. International Journal of Plasticity, 157, 103391.
Marigo, J. (1985). Modelling of brittle and fatigue damage for elastic material by growth of
microvoids. Engineering Fracture Mechanics, 21(4), 861-874.
Moreo, P., Garcia-Aznar, J., & Doblare, M. (2007). A coupled viscoplastic rate-dependent
damage model for the simulation of fatigue failure of cement–bone interfaces.
International Journal of Plasticity, 23(12), 2058-2084.
Nguyen, G. D., Nguyen, C. T., Nguyen, V. P., Bui, H. H., & Shen, L. (2016b). A size-dependent
constitutive modelling framework for localised failure analysis. Computational
Mechanics, 58(2), 257-280.
Nguyen, N. H., Bui, H. H., Kodikara, J., Arooran, S., & Darve, F. (2019). A discrete element
modelling approach for fatigue damage growth in cemented materials. International
Journal of Plasticity, 112, 68-88.
40
Nguyen, N. H., Bui, H. H., Nguyen, G. D., & Kodikara, J. (2017b). A cohesive damage-
plasticity model for DEM and its application for numerical investigation of soft rock
fracture properties. International Journal of Plasticity, 98, 175-196.
Nguyen, N. H., Bui, H. H., Nguyen, G. D., Kodikara, J., Arooran, S., & Jitsangiam, P. (2017a).
A thermodynamics-based cohesive model for discrete element modelling of fracture in
cemented materials. International Journal of Solids and Structures, 117, 159-176.
Nguyen, O., Repetto, E., Ortiz, M., & Radovitzky, R. (2001). A cohesive model of fatigue crack
growth. International Journal of Fracture, 110(4), 351-369.
Pandolfi, A., & Taliercio, A. (1998). Bounding surface models applied to fatigue of plain
concrete. Journal of engineering Mechanics, 124(5), 556-564.
Papa, E., & Taliercio, A. (1996). Anisotropic damage model for the multiaxial static and fatigue
behaviour of plain concrete. Engineering Fracture Mechanics, 55(2), 163-179.
Potyondy, D. O., & Cundall, P. (2004). A bonded-particle model for rock. International journal
of rock mechanics and mining sciences, 41(8), 1329-1364.
Prevost, J. H. (1982). Two‐ surface versus multi‐ surface plasticity theories: A critical
assessment. International Journal for Numerical and Analytical Methods in
Geomechanics, 6(3), 323-338.
Ray, S., & Chandra Kishen, J. (2012). Fatigue crack growth due to overloads in plain concrete
using scaling laws. Sadhana, 37(1), 107-124.
Ray, S., & Kishen, J. C. (2010). Fatigue crack propagation model for plain concrete–An analogy
with population growth. Engineering fracture mechanics, 77(17), 3418-3433.
Reinhardt, H. W. (1984). Fracture mechanics of an elastic softening material like concrete.
HERON, 29 (2), 1984.
Rezazadeh, M., & Carvelli, V. (2018). A damage model for high-cycle fatigue behavior of bond
between FRP bar and concrete. International Journal of Fatigue, 111, 101-111.
Skarżyński, Ł., Marzec, I., & Tejchman, J. (2019). Fracture evolution in concrete compressive
fatigue experiments based on X-ray micro-CT images. International Journal of Fatigue,
122, 256-272.
Song, K., Wang, K., Zhang, L., Zhao, L., Xu, L., Han, Y., & Hao, K. (2022). Insights on low
cycle fatigue crack formation and propagation mechanism: A microstructurally-sensitive
modeling. International Journal of Plasticity, 154, 103295.
Song, Z., Konietzky, H., & Frühwirt, T. (2018). Hysteresis energy-based failure indicators for
concrete and brittle rocks under the condition of fatigue loading. International Journal of
fatigue, 114, 298-310.
Sounthararajah, A., Bui, H. H., Nguyen, N., Jitsangiam, P., & Kodikara, J. (2018). Early-age
fatigue damage assessment of cement-treated bases under repetitive heavy traffic loading.
Journal of Materials in Civil Engineering, 30(6), 04018079.
Suaris, W., Ouyang, C., & Fernando, V. M. (1990). Damage model for cyclic loading of
concrete. Journal of engineering mechanics, 116(5), 1020-1035.
Toumi, A., & Bascoul, A. (2002). Mode I crack propagation in concrete under fatigue:
microscopic observations and modelling. International journal for numerical and
analytical methods in geomechanics, 26(13), 1299-1312.
Tran, K. M., Bui, H. H., & Nguyen, G. D. (2021). A hybrid discrete-continuum approach to
model hydro-mechanical behaviour of soil during desiccation. arXiv preprint
arXiv:2106.04676.
Tran, K. M., Bui, H. H., Sánchez, M., & Kodikara, J. (2020). A DEM approach to study
desiccation processes in slurry soils. Computers and Geotechnics, 120, 103448.
Turon, A., Costa, J., Camanho, P., & Dávila, C. (2007). Simulation of delamination in
composites under high-cycle fatigue. Composites Part A: applied science and
manufacturing, 38(11), 2270-2282.
41
Van Paepegem, W., & Degrieck, J. (2001). Fatigue degradation modelling of plain woven
glass/epoxy composites. Composites Part A: Applied Science and Manufacturing, 32(10),
1433-1441.
Vicente, M. A., Ruiz, G., Gonzalez, D. C., Minguez, J., Tarifa, M., & Zhang, X. (2018). CT-
Scan study of crack patterns of fiber-reinforced concrete loaded monotonically and under
low-cycle fatigue. International Journal of Fatigue, 114, 138-147.
Wang, Y., & Li, J. (2021). A two-scale stochastic damage model for concrete under fatigue
loading. International Journal of Fatigue, 153, 106508.
Wei, X.-D., Nguyen, N. H., Bui, H. H., & Zhao, G.-F. (2021). A modified cohesive damage-
plasticity model for distinct lattice spring model on rock fracturing. Computers and
Geotechnics, 135, 104152.
Wu, R., & Harvey, J. (2013). A J2‐ plasticity model based on bounding surface concept.
International Journal for Numerical and Analytical Methods in Geomechanics, 37(7),
744-757.
Xie, D., Zhang, W., Lyu, Z., Liaw, P. K., Tran, H., Chew, H. B., Wei, Y., Ren, Y., & Gao, Y.
(2022). Plastic anisotropy and twin distributions near the fatigue crack tip of textured Mg
alloys from in situ synchrotron X-ray diffraction measurements and multiscale mechanics
modeling. Journal of the Mechanics and Physics of Solids, 165, 104936.
Xu, Q., & Lu, Z. (2013). An elastic–plastic cohesive zone model for metal–ceramic interfaces at
finite deformations. International Journal of Plasticity, 41, 147-164.
Xu, Y., & Yuan, H. (2009). Computational analysis of mixed-mode fatigue crack growth in
quasi-brittle materials using extended finite element methods. Engineering Fracture
Mechanics, 76(2), 165-181.
Yang, B.-L., Dafalias, Y. F., & Herrmann, L. R. (1985). A bounding surface plasticity model for
concrete. Journal of Engineering Mechanics, 111(3), 359-380.
Zhang, J., Liu, W., Zhu, Q., & Shao, J. (2022). A novel elastic–plastic damage model for rock
materials considering micro-structural degradation due to cyclic fatigue. International
Journal of Plasticity, 103496.
Zhang, L., Gao, Z., Haynes, R. A., Henry, T. C., & Yu, W. (2020). An anisotropic continuum
damage model for high-cycle fatigue. Mechanics of Advanced Materials and Structures,
1-15.
Zhao, X., Dong, Q., Chen, X., Han, H., & Zhang, T. (2021b). Evaluation of fatigue performance
of cement-treated composites based on residual strength through discrete element
method. Construction and Building Materials, 306, 124904.
Zhao, X., Dong, Q., Chen, X., Xiao, Y., & Zheng, D. (2021a). Fatigue damage numerical
simulation of cement-treated base materials by discrete element method. Construction
and Building Materials, 276, 122142.
Zhao, Y., Lai, Y., Pei, W., & Yu, F. (2020). An anisotropic bounding surface elastoplastic
constitutive model for frozen sulfate saline silty clay under cyclic loading. International
Journal of Plasticity, 129, 102668.
Zheng, L., Gao, Y., Lee, S., Barabash, R., Lee, J., & Liaw, P. K. (2011). Intergranular strain
evolution near fatigue crack tips in polycrystalline metals. Journal of the Mechanics and
Physics of Solids, 59(11), 2307-2322.
Zhu, Y., Kang, G., Kan, Q., & Bruhns, O. T. (2014). Logarithmic stress rate based constitutive
model for cyclic loading in finite plasticity. International Journal of Plasticity, 54, 34-55.
Zhu, Y., Kang, G., & Yu, C. (2017). A finite cyclic elasto-plastic constitutive model to improve
the description of cyclic stress-strain hysteresis loops. International Journal of Plasticity,
95, 191-215.
Zreid, I., & Kaliske, M. (2016). An implicit gradient formulation for microplane Drucker-Prager
plasticity. International Journal of Plasticity, 83, 252-272.
42
Vinh T. Le: Methodology, Software, Formal Analysis, Investigation, Writing- Original
Draft, Visualization. Khoa M. Tran: Software, Investigation, Writing - Review & Editing.
Jayantha Kodikara: Investigation, Writing - Review & Editing, Funding acquisition.
Didier Bodin: Investigation, Writing - Review & Editing. James Grenfell: Investigation,
Writing - Review & Editing. Ha H. Bui: Conceptualization, Methodology, Supervision,
Software, Investigation, Writing - Review & Editing Funding acquisition, Project
administration.
Declaration of interests
☒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.
☐The authors declare the following financial interests/personal relationships which may be considered
as potential competing interests:
43