You are on page 1of 29

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/360697891

Assessment of fatigue resistance of concrete: S-N curves to the Paris’ law curves

Article  in  Construction and Building Materials · July 2022


DOI: 10.1016/j.conbuildmat.2022.127811

CITATIONS READS
0 159

4 authors:

Petr Miarka Stanislav Seitl


Brno University of Technology The Czech Academy of Sciences
53 PUBLICATIONS   172 CITATIONS    161 PUBLICATIONS   989 CITATIONS   

SEE PROFILE SEE PROFILE

Vlastimil Bilek Héctor Cifuentes


VŠB-Technical University of Ostrava Universidad de Sevilla
77 PUBLICATIONS   505 CITATIONS    76 PUBLICATIONS   874 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Fibre reinforced concrete fracture View project

Design of advanced cementitious materials based on target performance View project

All content following this page was uploaded by Petr Miarka on 13 June 2022.

The user has requested enhancement of the downloaded file.


Assessment of Fatigue Resistance of Concrete: S-N Curves to the Paris’ Law Curves
Authors Names and Affiliations
Petr Miarkaa,b*, Stanislav Seitla,b, Vlastimil Bílekc, Hector Cifuentesd
a
Institute of Physics of Materials, Czech Academy of Science, Žižkova 22, 616 00 Brno, Czech Republic
b
Brno University of Technology, Faculty of Civil Engineering, Veveří 331/95, 602 00 Brno, Czech
Republic
c
Faculty of Civil Engineering, Technical University of Ostrava, L. Podéště 1875/17, 708 33 Ostrava,
Czech Republic
d
School of Engineering, University of Seville, Camino de los Descubrimientos, S/N, Isla Cartuja, 41092
Sevilla, Spain.
*Corresponding Author
Petr Miarka (petr.miarka@vut.cz). Tel: +420 532 290 430
Graphical Abstract

Abstract
Fatigue behaviour of concrete materials is often investigated on un-notched specimens under the
compressive or bending loads. In this experimental study, a notched three-point bending (TPB)
specimen is used in high-cycle fatigue experiments to obtain the Wohler’s curve. Based on this
approach, a novel, yet relatively simple transition from the traditional Wohler’s curve to the Paris’ law
curve is proposed. Such a methodology allows one to obtain the Paris’ law material constants, which
are used to determine the fatigue failure of the structure or a component. The constants of the

1
experimentally determined material, measured in four different concrete mixtures, have been verified
by recalculating the number of cycles until the fatigue failure Nf by the integration of the Paris’ law
equation. The back-calculated number of cycles and the approximation of the S-N curve allowed for a
comparison with the experimental data. Furthermore, the initial notch tip was extended in this
approximation by the value of the critical distance. Such an extension allowed us to cover a wide range
of the experimental data and provided a better prediction of fatigue life. The proposed method was
verified on all four studied materials and showed satisfactory results.
Keywords: Fatigue; Concrete; Wohler curve; Paris’ law; HPC; HSC; AAC.
Abbreviations/nomenclature
LEFM linear elastic fracture mechanics
SIF stress intensity factor
FPZ fracture process zone
AAC alkali-activated concrete
HPC high-performance concrete
HSC high-strength concrete
FCGR fatigue crack growth rate
TPB three-point bending
PCC polycarboxylates-based
GBFS granulated blast furnace slag
 relative crack length (-)
 stress (MPa)
max maximum stress in the load cycle (MPa)
 stress range (MPa)
KI,max maximum value of the stress intensity factor (MPamm 1/2)
KI,C fracture toughness (MPamm1/2)
N cycle increment (cycle)
P force range (kN)
a crack length (mm)
a0 initial notch length (mm)
A, B material constants of the S-N field (-)
T specimen’s thickness (mm)
E Young’s modulus of elasticity (GPa)
k number of displacement values (-)
KI stress intensity factor for mode I (MPam1/2)
L specimen’s length (mm)
m, C Paris’law material constants
N fatigue load cycle (cycle)
Pmin, Pmax load, limit load (kN)
R load ratio Pmin/Pmax (-)
rC critical distance
W specimen’s width (mm)

2
1. Introduction
Concrete structures form an essential part of the local and global infrastructure, and their lifespan is
usually measured in decades. Most concrete structures, both currently built and already in service, are
made using reinforced and/or prestressed concrete building technology with steel bars or tendons with
a consistent material performance embedded in the concrete material [1, 2]. On the other hand, the
quality of the concrete mixture used and its treatment during the construction of load bearing structures
directly influence the service life [3] as it carries load and protects the steel reinforcement from
corrosion. Recent structural failures of a pedestrian bridge in Florida, USA [4], or the collapse of the
Polcevera viaduct in Genova, Italy [5, 6], showed that this catastrophic failure occurred due to the
combination of all unfavourable conditions to which the structure is exposed over its lifespan. During
the lifetime of the structure, there are various load actions and their combinations that can occur, i.e.
thermal effects, aggressive environment, concrete rheology, but mainly static and cyclic load actions.
Such a load combination can result in additional maintenance costs, or worse, the collapse of the
structure if the structure was poorly executed during the construction.
Material science has developed modern concrete mixtures to improve the mechanical performance,
durability, and reduce their carbon footprint [7, 8]. Such materials can be, for example, high-strength
concrete (HSC) [9], that allows for a material reduction in the cross-sectional dimensions of the
structure, high performance concrete (HPC) [10, 11], that exhibits a higher long-term performance and
durability of the structure, or alkali-activated concrete (AAC) [12, 13], that fully substitutes cement as
a binder in the mixture. Thus, the structures (e.g. bridges, beams, railway sleepers, wind turbines’
pylons) to be built from these modern concrete materials can benefit from a greater span length, a
shallow beam cross-section, and, most importantly, it can result in an extended service lifetime.
Besides reducing the carbon footprint, the durability can be considered as a major requirement of the
used concrete mixture as the concrete simultaneously covers and protects the steel reinforcement from
direct exposure to weather conditions. Such weather conditions are often aggressive (de-icing salts or a
marine environment) and often accelerate the corrosion of steel reinforcement [14, 15]. The enhanced
mixture’s performance and CO2 emission reduction are achieved by using mineral admixtures, e.g. silica
fume [16], ground granulated blast furnace slag [17], metakaolin or zeolite [18], etc. The use of these
admixtures results in a lower permeability of concrete as it fills in the pores and inner flaws present in
the concrete’s microstructure [19], and thus it results in an improved structure’s durability.
Despite the improved strength and performance, microcracks can form and increase in size throughout
the cover layer. Furthermore, most of the structures are exposed to cyclic load [20, 21], and such loading
can lead to a progressive long-term crack initiation and damage accumulation [22], which can contribute
to a longer defect initiation in the protecting cover layer [23, 24]. Thus, these unfavourable
circumstances have motivated the investigation of fatigue failure of concrete materials.
Fatigue is defined as a process of progressive, permanent changes in the internal material’s structure due
to the exposure to repeated/cyclic loading. Repeated load is mainly present on structures like bridges,
railway sleepers, machine foundations, highway/runway slabs, etc. For concrete materials, these
changes are mainly associated with the progressive growth of internal microcracks, resulting in a
significant increase in irrecoverable strain. At the macrolevel, this will manifest itself as a change in the
material mechanical properties. Different loading arrangements have been used in fatigue testing,
including compression, tension, and bending tests. The most common fatigue test method, by far, is via
flexural tests with induced cyclic load. In the literature, multiple studies are focused on this topic. A
study of flexural fatigue behaviour of plain concrete can be found in [25], of recycled concrete aggregate
in [26-28], and of fibre-reinforced concrete in [29-32]. The compressive fatigue behaviour of plain
concrete was studied in [33, 34] and studies investigating fibre-reinforced concrete can be found in [35-
37]. Moreover, a practical application of fatigue design for railway sleepers can be found in [38],
foundations of wind turbines in [39-41], a comprehensive analysis of reinforced beams under fatigue
load was studied in [42], and the combination of fatigue and chloride penetration is presented in [43].

3
Fatigue design of concrete structures is partly implemented in the standards for structural design,
Eurocode [44] or ACI [45]. These design approaches assume fatigue failure under compressive failure.
On the other hand, the recommendations of CEB-FIB Model Code (MC2010) [46] consider fatigue
failure under both compressive and tensile loads. Nonetheless, all these technical recommendations
assume a reduction of material strength to prevent fatigue failure by a certain value. In addition, MC
2010 provides S-N curves for fatigue design with the fatigue limit of 1  1028 cycles, which is a rather
unrealistic expectation. Furthermore, fatigue design of a structure is considered for the ultimate limit
state (ULS), while the serviceability limit state (SLS) is missing, i.e. the assessment of the crack width
wc, crack length l and deflection . This missing SLS fatigue consideration in standards may lead to an
unoptimized structure, which may eventually result in a visible damage of load-bearing structural
components.
Traditionally, for a specimen made of a metallic material, the fatigue resistance is assessed by the
Wöhler’s curve, usually called the S-N curve (evaluated by Basquin’s power law [47]), using smooth
specimens or by measuring the fatigue crack growth rate (FCGR) on a notched specimen based on the
linear elastic fracture mechanics (LEFM) as formulated by the Paris-Erdogan’s law [48]. While the
former approach studies the fatigue phenomenon at various stress levels and sets the material’s cyclic
endurance limit, the latter approach uses the notch on the specimen from which the crack can initiate,
and its increment during the load cycle da/dN is measured. In addition, FCGR is usually plotted over
the value of the stress intensity factor (SIF) KI; such a plot is sometimes referred to as the v-K curve.
This curve then provides complete information throughout the crack propagation process in the studied
material, i.e. crack initiation, propagation, and final brittle failure.
In this experimental study, we propose and assess a combination of the Wöhler’s curve and Paris-
Erdogan’s law evaluation on the measured experimental data. For this, we have performed fatigue
experiments on notched three-point bending (TPB) specimens made from various concrete materials
with similar mechanical properties as a direct result of a cooperation with a company producing railway
sleepers. The materials which have been studied in this experimental study are concrete C 50/60 [49],
HPC [50], HSC [51] and AAC [52]. The outcomes of this article are the S-N curves, which have been
obtained to characterize the fatigue resistance of the concrete mixture to the cyclic load and the Paris’
law material’s constants used in the prediction of the fatigue failure.

2. Theoretical Background
Firstly, we introduce the traditional Wöhler’s curve used for the description of the material’s fatigue life
and the Paris-Erdogan’s law for FCGR measurement. Afterwards, the calculation of the SIF and fracture
toughness KIC are introduced. Lastly, we present the used methodology for the transition from the
Wöhler’s curve to Paris’ curve.

2.1 Fatigue Analysis


The S-N curve relates the stress ( or S) to the number of cycles N needed for the fatigue failure. The
fatigue lifetime is evaluated by using the Basquin’s power law, which is described as:
𝜎 = 𝐴𝑁 (1)
where  is the applied stress during the load cycle, N is the load number of cycles needed for the fatigue
failure, and A and B are the material’s constants characterizing the S-N field. A wide range of data of S-
N curves with various experimental set-ups measured on concrete materials can be found in the study of
Susmel [53].
However, the S-N curve’s approach provides only limited information of the material’s fatigue failure,
and in the most design application it is recommended to reduce the applied stress below the measured
curve to achieve the demanded structure’s fatigue endurance. In contrast, the Paris-Erdogan law [48] is

4
more beneficial from the material science viewpoint as the crack growth is related to the fracture
parameter SIF KI. This relationship of FCGR da/dN to KI is expressed by the power law as:
𝑑𝑎
= 𝐶𝐾 , (2)
𝑑𝑁
where C and m are the experimentally obtained material’s constants dependent on the asymmetry of the
load cycle R, loading frequency f, time t, environmental conditions, material properties, and the
specimen size (See [54]). The values of the constant m, i.e. the slope of the curve from Eq. (2), can have
values of 2 – 4 for metals, 10 for intermetallic materials, and can be higher than 20 for ceramics [55];
for rocks, it has a value of 22 [56]. This power law has found application in concrete (quasi-brittle)
materials as well [57, 58] with the values of the constant m equal to 11. However, the results in [57, 58]
are based on the maximum of 1103 load cycles.
The power law in Eq. (2) describes the stable crack growth in a material, which is described by a linear
function. This relation is accompanied by two asymptotic values of SIF: the threshold values KI,th and
the value of the material’s fracture toughness KIC. The threshold value KI,th implies the onset of fatigue
cracks, which are not growing, while the KIC values imply the unstable crack growth and the brittle
fracture.
This relationship has been verified in innumerable crack growth studies for the past 60 years, however,
some adjustments to the Paris’ law to cover different materials’ behaviours were done by Ritchie [59],
Ray [60], Kirane  Bažant [61], and Pugno et. al [62]. In his study, Ritchie covers the effects of both
KI,max and KI on the crack growth rates, Pugno et. al proposed a generalised version of Paris’ law, which
is related to the material’s S-N curve, while Ray and Kirane  Bažant focused their study on covering
the size-effect present in a concrete material. Regarding the size-effect [57], there have been studies
performed on concrete compressive specimens [37] and on TPB in [63] as well as for rocks in [56]. Le
[56] showed that the size-effect has a relatively minor influence on the Paris-Erdogan law constant m.
Nonetheless, these size-effect studies are investigating the fatigue damage with a maximum of 1  104
load cycles. The prediction of fatigue life by numerical approaches was recently proposed by Xi in [64]
and by Baktheer in [65], both focusing on the S-N curve simulation.
Generally speaking, fatigue is a linear phenomenon mainly dependent on the stress level  and on the
value of asymmetry of the load cycle R [66]. This effect is clearly observed on the measured Paris’
curves for metallic materials as presented by Klesnil and Lukáš in [67] and recently demonstrated in
[68]. The R ratio influences the S-N curves as well as the Paris’ curves. This effect on concrete was
studied in [69].
The asymmetry of the load cycle R is expressed as the ratio between the minimum force Pmin and the
maximum force Pmax applied per load cycle. Equally to the applied forces, the asymmetry of the load
cycle can be expressed in the applied stress . The asymmetry of the load cycle is usually expressed as:
𝑃 𝜎
𝑅= = , (3)
𝑃 𝜎
where Pmin and Pmax are the minimum and maximum forces per load cycle, and min and max are the
minimum and maximum stresses per loading cycle. Most of the structures are loaded with an
asymmetrical load with R = 0.1, which puts the present crack in the material’s internal structure under
permanent crack opening conditions, i.e. 10 % of the maximum load.
Another effect that influences the FCGR is a free-edge singularity or 3D effects of fracture, which result
in typical beachmarks or crack front curvature ahead of the fatigue crack tip. Such effects have been
studied by Bažant [70], Nakamura [71], Pook [72], de Matos [73], Hutař [74], and recently by Oplt [75].
These studies were mainly done on steel or other metallic alloys with plastic behaviour. Thus, the
formation of beachmarks in the fatigue crack propagation direction is able to develop due to the free-
edge singularity. However, these beachmarks or 3D effects are difficult to observe on concrete materials,

5
as there is no plastic behaviour, and multiple micro-cracks are developing ahead of the notch tip. Another
issue for beachmarks to be developed in concrete may be its heterogeneity and the wall-effect [76],
which places the cement paste on the edge of the specimen due to the setting of the concrete mixture in
the mould.

2.2 Stress Intensity Factor for Mode I


A TPB test specimen is often used due to its simple shape and a relatively easy experimental set-up.
This geometry is well acknowledged among the researchers to obtain the fracture and fatigue parameters
of quasi-brittle materials [77, 78]. Another reason to use TPB is that it provides an extensive use in the
measurement of fatigue parameters for various materials [79]. During the fatigue test, a SIF KI is used
[80], which is calculated by using Eq. (5) according to ASTM E647 [81].
3𝑃𝑆
𝐾 = 𝜋𝑎 𝑌 (𝛼), (4)
2𝑇𝑊
where  is the relative crack length expressed as a/W, a0 is the crack length, W is the specimen’s width,
T is the specimen’s thickness, S is the span between the supports, P is the applied load range, and YI is
the geometry shape function, which is expressed as a polynomial function:
𝑌 (𝛼) = +1.0312 − 1.1836(𝛼) + 3.1988(𝛼) + 0.8632(𝛼) , (5)
By substituting the maximum measured force for one cycle N = 1, i.e. static load, one can obtain the
value of the material’s fracture toughness KIC. The fracture toughness KIC serves as an asymptotic limit
in the Paris’ law and in the estimation of the material’s critical distance rC. Critical distance rC is the
distance from the crack tip in which the stress reaches its critical value or tensile strength and leads to
failure. It was originally derived by Irwin [82] for metals as an approximate size of the plastic zone
ahead of the crack tip. The literature recognizes two basic values of rC depending on the boundary
conditions:
1 𝐾
𝑟 = − 𝑝𝑙𝑎𝑛𝑒 𝑠𝑡𝑟𝑒𝑠𝑠, (6)
2𝜋 𝜎
1 𝐾
𝑟 = − 𝑝𝑙𝑎𝑛𝑒 𝑠𝑡𝑟𝑎𝑖𝑛, (7)
6𝜋 𝜎
Where t is the material’s tensile strength, usually referred to as fc,t in concrete. For concrete materials,
the critical distance finds application in the prediction of brittle fracture for mode I [83] as well as for
the mixed-mode I/II [49, 84]. The use of the critical distance is limited, since concrete is considered as
a quasi-brittle material, which does not evince a plastic zone ahead of the crack tip, but it has a much
bigger fracture process zone (FPZ) [85], in which multiple micro-cracks can occur and later collide into
one major crack.
In addition, the literature offers multiple ways to assess the FPZ length, i.e. the so-called characteristic
length lch proposed by Hillerborg [86], Ayatollahi and Akbardoost [83]. They have related the critical
distance rC to the use of higher-order terms of the Williams expansion. Bažant [87], Otsuka [88] and
Alanazi [89] referred the size of the FPZ to the maximum aggregate’s size dagg. Since the fatigue damage
is considered to be linear elastic, the use of Irwin’s original formulation based on LEFM of critical
distance rC found application in the high-cycle fatigue analysis.

2.3 Prediction of Fatigue Life


The Paris’ law can be used as a prediction of the material’s failure due to cyclic load, and thus finds a
practical application in structural design. If the material constants from the Paris’ law C and m are
known, then by using them in Eq. (2), the number of cycles needed for the fatigue failure Nf can be
estimated by the following integral equation:

6
1 𝑑𝑎
𝑁 = , (8)
𝐶 ∆𝜎√𝜋 𝑎 (𝑌 (𝛼))
where a0 is the initial crack length– usually notch length – and the aC is the critical crack length at which
brittle fracture occurs. The critical crack length aC is directly related to the material’s fracture toughness
KIC, i.e. the material’s failure occurs when a = aC. The critical length aC can be calculated using Eq. (4)
by substitution of the material’s fracture toughness KIC and applied stress ; this yields to following
equation:
1 𝐾
𝑎 = , (9)
𝜋 𝜎𝑌 (𝛼 )
in which the geometry shape function YI(αC) is dependent on the critical length ratio αC = aC/W and has
the same form as the one mentioned in Eq. (5).
Due to the presence of FPZ [90, 91] ahead of the crack tip, which is believed to extend the brittle failure
by the formation of multiple microcracks, the definite location of the fatigue crack front a0 as used in
Eq. (8) cannot be assessed. Based on this assumption, the initial notch length a should be extended to
reduce the distance in which the fatigue cracks can propagate. Furthermore, the fatigue failure is
considered linear elastic, which suggests extending the initial crack/notch length by the value of the
critical distance rC. This leads to a higher integration limit in Eq. (8) as the initial notch length has the
form of a0 = a + rC. This initial notch length extension is schematically explained in Figure 1.

Figure 1: Notch tip extension by critical distance.

This notch tip extension can represent the inner structure’s flaw, i.e. shrinkage cracks, pores, or cracks
induced by the preparation of the notch. A similar approach was used in the assessment of fracture
toughness by Karihaloo [92] in his effective crack model.

2.4 Transition from the S-N Curve to the Paris’ Curves


Similarly to the S-N curve, one can plot a K-N (SIF vs load cycles) curve using Eq. (4), as the
experimental data was measured on the notched specimens. Assuming the same experimental data as in
Eq. (1), this brings the following equation:
𝐾 = 𝜓𝑁 , (10)
where  is the constant and β is the slope of the fitted curve. If the same experimental data is used in
the same manner as for Eq. (1), the parameter β in Eq.(10) is identical to the parameter B in Eq. (1), as
the KI in this case uses the same value of stress as in the S-N curve assessment.
Furthermore, this suggests plotting FCGR, da/dN against the values of KI, i.e. create an inverse function
to the Paris’ law as presented in Eq. (2). If Eq. (2) is plotted in double logarithmic coordinates, the power
law becomes a linear function, and after some mathematical manipulation, one can obtain the following
inverse function:

7
1 𝑑𝑎
𝐾 = − C from Eq. (1), (11)
𝑚 𝑑𝑁
where 1/m is the slope of the curve and C is the constant. Similarly, the load cycles N in Eq. (10) can be
replaced by FCGR da/dN, which gives the following form, again in double logarithmic coordinates:
𝑑𝑎
𝐾 =𝛽 + 𝜓 from Eq. (10), (12)
𝑑𝑁
where β is the slope, which is again identical to the parameter B from Eq. (1), and  is the constant,
which in this case has a different value from  in Eq. (10).
From both Eqs. (11) and (12), one can observe the similarity in the values of 1/m and β as both
parameters are obtained from the same experimental data. Thus, the slope of the S-N curve B is identical
to the parameter β in the inverse version of the Paris’ law. Consequently, the constant C can be evaluated
from Eq. (12) as follows:
1
𝐶 = −𝛽 ∙ 𝜓 𝑜𝑟 − ∙𝜓 (13)
𝑚

3. Materials
In this experimental study, we considered various concrete types extensively used in the fabrication of
precast concrete structural elements. They include C 50/60, high-performance concrete (HPC), high-
strength concrete (HSC), and alkali-activated concrete (AAC). The mixtures selected in this study show
a good early strength, and overall have a good long-term performance. Furthermore, they are currently
used for the fabrication of load-bearing highway bridge beams and railway sleepers. The used concrete
mixtures are introduced and complemented with their measured mechanical properties. The measured
mechanical properties were according to European standards.

3.1 Mixtures Composition


The C 50/60 mixture served as an initial reference mixture, from which the rest of the concrete types
were designed, as it is massively used in the production of precast concrete elements. The HPC and HSC
were composed to improve the mechanical performance of the mixtures by adding mineral admixtures
into the mixture. On the other hand, the AAC material was designed with the intent to fully replace the
cement in the mixture.
CEM I 42.5 R is used as a binder in C 50/60, the HPC and HSC mixtures together with natural sand 0/4
mm as a fine aggregate, and crushed aggregates from high-quality granite, which served as 4/8 mm and
8/16 mm aggregates. The polycarboxylates-based (PCC) superplasticizer Glenium 300 was used to reach
good workability in all these mixtures. The cement was supplemented with multiple mineral admixtures
[10] in the case of the HPC and HSC mixtures, i.e. granulated blast furnace slag (GBFS), limestone and
metakaolin Metaver was used. The concrete mixtures were mixed in a volume of 1 m 3. The studied
mixtures compositions per m3 are presented in Table 1.
Table 1: Composition of C 50/60, HPC, and HSC mixtures per m3.

Fine Crushed Crushed Super-


CEM I
Material GBFS Limestone Metakaoline aggregate aggregate aggregate Water plasticizer
42.5 R
0/4 4/8 8/16 (PCC)
C 50/60 450 - - - 690 215 845 180 4.5
HPC 575 40 30 80 590 185 725 165 20
HSC 450 60 15 75 400 600 - 165 17

As it can be seen from Table 1 the C 50/60 and HPC are using a similar amount of aggregates up to
8/16, which improves the compressive strength, while in the case of HSC this aggregate of 8/16 size
was fully replaced by 4/8 aggregates. Furthermore, both the HPC and HSC mixtures have a similar ratio

8
of water to the binder of 0.25, while for C 50/60, it is 0.4 –in this case, it is the traditional water to
cement ratio w/c.
In contrast to the above-mentioned traditional cement-based concrete types stands the AAC mixture,
which deserves more attention in the mixture description and composition [12, 93]. The studied AAC
mixture was designed based on the formerly performed mechanical tests, see [94, 95]. The traditional
binder Portland cement was fully substituted by the GBFS. The dry mass of the activator was 8 % and
the water to slag ratio was 0.45. The activator ratio of the mixture was composed according to [96].
Sodium water glass (Na-WG) and potassium hydroxide were combined to reduce efflorescence and to
provide an appropriate silicate modulus of the activator (Ms = 0.67 or mass ratio (K2O + Na2O) / SiO2 is
60/40). This activator’s content is convenient for both setting and strength. Naphthalene-based
plasticizer was also used to reach a better workability. The AAC’s material composition per m 3 is shown
in Table 2.
Table 2: Mixture composition of used AAC concrete per m3.
50 % sand crushed
Na-WG PSN- crushed
GBFS solution of water aggregates
Ms = 2.0 plasticizer 0/4 aggregates 8/16
KOH 4/8
[kg/m3] 450 45 34 159 10 855 385 400

In this case, a similar amount of binder was used as for the C 50/60 mixture, which helped to achieve a
similar water to binder ratio of 0.45. To increase the compressive strength and Young’s modulus, a
maximum size of aggregate 8/16 mm was used from the same high-quality granite mine as for the
traditional cement-based mixtures. Furthermore, a full substitution of cement in the AAC mixture has a
typically white colour of the matrix, while the rest of the mixtures have their matrix in the typical colour
of Portland cement dark grey. These studied materials and their cross-sections are presented in Figure
2.
All four studied mixtures were poured into the moulds immediately after mixing. Afterwards, the
specimens were carefully covered with PE-foil to prevent any moisture exchange with the environment.
The specimens were stored outside of the laboratory environment (precast concrete plant), with a
temperature ≈ 5 - 25oC for a period of 28 days.

(a) (b)

9
(c) (d)
Figure 2: The internal structure of used materials (a) – C 50/60, (b) – HPC, (c) – HSC, and (d) – AAC.

3.2 Mechanical Properties


The above-mentioned concrete mixtures were tested according to European standards to obtain their
mechanical properties. The tested mechanical properties, using standardized specimens size and shape,
were the compressive cube strength fc,cube, the compressive cylindrical strength fc,cyl, the flexural strength
fc,t, the indirect tensile strength ft, Young’s Modulus E and the bulk density . For the indirect tensile
strength, we used the Brazilian disc test with a radius of 150 mm and the thickness of approx. 30 mm.
These parameters are presented together with the standard deviation for all studied mixtures in Table 3.
Table 3: Measured mechanical properties according to European standards for all studied concrete mixtures.

Material C 50/601 HPC HSC AAC


fc,cube [MPa] 85.8 ± 2.9 100.5 ± 2.3 106.2 ± 2.5 62.0 ± 1.5
fc,cyl [MPa] 72.8 ± 2.5 84.7 ± 3.2 94.1 ± 2.8 48.0 ± 3.4
fc,t [MPa] 5.5 ± 0.31 6.3 ± 0.6 9.3 ± 0.9 3.153 ±0.2
E [GPa] 38.3 ± 0.3 40.6 ± 0.4 41.0 ± 0.6 26.3 ± 1.1
 [kg/m3] 2390 ± 27.32 2342 ± 25.54 2390 ± 26.18 2245.0 ± 14.58

From the measured mechanical properties as presented in Table 3 the following observation can be
made. The ACC mixture performed with the lowest mechanical properties for all the tested categories.
The HPC and HSC, if compared to C 50/60, showed an increase in the measured strength and Young’s
modulus. Thus, the use of proper mineral admixtures can significantly improve the mixture’s mechanical
performance, while keeping the same amount of cement. The mechanical properties of HPC and HSC
mixtures do not vary between each other as the use of 8/16 mm crushed aggregates in the HPC mixture
improves the measured compressive strength.

3.3 Fracture Mechanical Parameters


To illustrate the typical non-linear behaviour of the tested concrete mixtures, fracture tests were
performed on TPB specimens with a geometry height W  thickness T of 80  80 mm with a span S of
480 mm. The fracture tests were performed according to RILEM recommendations [77] on a servo-
hydraulic testing rig with a maximum capacity of 200 kN. The speed of the induced vertical
displacement δ was equal to 0.025 mm/s and it was measured by using a linear variable displacement
transducer (LVDT). The measured load-displacement (P - ) diagrams at the age of 7 days are presented
in Figure 3.

1
Please note that the C 50/60 mixture would fit the grade of C 70/85 based on the measured mechanical properties.
However, the precast plant is using the C 50/60 indication, so we will keep this grade in our study.

10
Figure 3: Experimentally measured load-displacement (P - ) diagrams for various concrete mixtures.

From Figure 3, a similar post-peak behaviour can be observed for all the studied concrete mixtures, also
the measured maximum loads are corresponding to the general expectations based on the measured
mechanical properties, i.e. the AAC mixture shows the lowest value of the maximum force, while HSC
showed the highest value. The measured P -  curves for the C 50/60 and HPC mixtures are somewhere
in between with similar values of the maximum force. In addition to this, HSC showed a significant
brittle behaviour during these fracture tests. To provide more information about the fracture mechanical
parameters, the measured value of the fracture toughness KIC and specific fracture energy Gf are
presented in Table 4.
Table 4: Measured fracture mechanical parameters of studied concrete mixtures.

Material C 50/60 HPC HSC AAC


KIC [MPamm1/2] 23.6 ± 2.4 24.2 ± 2.1 25.9 ± 1.9 19.4 ± 2.8
Gf [N/mm] 142.52 ± 5.4 108.15 ± 7.2 184.38 ± 6.2 83.51 ± 4.3

The presented values of fracture parameters in Table 4 are again in agreement with the trend observed
for the measured mechanical properties as well as for the observations made in Figure 3 on the measured
P- curves.
More information about the mechanical and fracture mechanical parameters of the studied materials can
be found in the previously published studies for C 50/60 in [49], for HSC in [97], for HPC [50, 84] and
for AAC in [52].

4. Fatigue Experimental Details


In this section, the specimen’s geometry for the fatigue tests is briefly described together with the fatigue
experimental set-up, and the measurement of the S-N curve is explained.

4.1 Specimen’s Geometry


In this experimental study, a standard TPB specimen was used. The TPB specimens were fabricated
from the above-mentioned concrete mixtures. The typical rectangular specimen had a length L of 200
mm with a square cross-section height W  thickness T of 80  80 mm. The initial notch a0 had the
length of 8 mm, which produces the relative crack length ratio α of 0.1 (see Figure 4). The notch was
created by an ordinary saw with a diamond blade.

11
(a) (b)
Figure 4: TPB specimen’s geometry (a) and its cross-section specifications (b).

4.2 Fatigue Experimental Test Set-up


In this subsection, the fatigue experimental set-up and the experimental methods are introduced. The
fatigue test was conducted on a TPB specimen using a Zwick/Roell Amsler HC25 servo-hydraulic
testing rig with a maximum load capacity of 25 kN. The working frequency of cyclic load was set to 10
Hz to cover all fatigue regions from low cycle to high cycle in a reasonable time, which was studied in
[98] and [99]. The applied asymmetry of the cycle R was set to 0.1, as it is most common for the load
presence on civil engineering structures.
A similar approach to fatigue testing was used for red ceramic waste aggregate concrete by Seitl et al.
[100] and for the NaOH-activated ACC material by Šimonová et al. [101]. The used fatigue
experimental set-up is presented in Figure 5.

Figure 5: Fatigue experimental set-up.

The fatigue experiments were carried out with a different number of cycles until the fatigue failure for
each tested specimen to provide the whole data range for the S-N curve: Nf = 1, to cover the static
strength and to have an initiation point for the construction of the S-N curve; cyclic load until fatigue
failure with Nf < 2 ×106, a linear part of the S-N curve, which was obtained by reducing the Pmax of the
previous specimen’s failure by 10%; and runouts with Nf > 2 ×106, which corresponds to the fatigue
limit set for concrete materials. The fatigue test was finished when the failure of the specimen occurred
or when the fatigue limit was reached. This fatigue limit was set to Nf = 2 ×106 cycles, due to the time
consumption of the experiments and due the recommendations proposed by Lee and Barr [21]. It should
be pointed out that this fatigue limit is not finite as the failure can occur with more load cycles.

5. Experimental Results and Discussion


In what follows, the experimentally measured S-N curves are presented together with the discussion of
the evaluated coefficients of the Basquin’s law. Subsequently, the FCGR according to the Paris’ law
and its materials’ constants are calculated from the same experimental data as the S-N curves.
Furthermore, a validation of the proposed method for the determination of the Paris’ law coefficients is
presented together with a parametric study of its constants.

12
5.1 The S-N Curves
As the experiments were performed on the notched specimens, the stress for the construction of the S-
N curves has to include the notch concentration and it cannot be calculated using the classic Euler-
Bernoulli’s beam theory. Thus, Eq. (4) is used to cover such stress conditions ahead of the crack tip.
Furthermore, depending on the used forces, various S-N curves can be constructed covering different
stress levels, i.e. using the maximum force Pmax, the force amplitude Pa = (Pmax – Pmin)/2 or the force
range P = Pmax – Pmin as seen in a load cycle.
Such experimentally measured S-N curves of each material with various stress levels as seen in the load
cycle are presented in Figure 6. Please note that the experimental data, i.e. the maximum force Pmax and
the number of cycles until the fatigue failure Nf, for all the studied mixtures is presented in Appendix B.

(a) (b)

(c) (d)
Figure 6: Experimentally measured S-N curves for various stress levels and for all the studied concrete materials – (a) C
50/60, (b) HPC, (c) HSC and (d) AAC.

As it can be observed from Figure 6, the fatigue experimental results confirm a general expectation and
agree with the measured mechanical properties of each tested material, i.e. the AAC material shows the
lowest fatigue performance due to the lowest flexural strength, while the HSC shows the best fatigue
performance and the best flexural strength. This observation confirms the evaluated Basquin’s law
coefficients. The calculated Basquin’s law coefficients A and B from the experiments together with the
coefficient of determination R2 are presented in Table 5.
Table 5: Measured material’s constants from Basquin’s law for various concrete materials.
A [MPa]
B [-] R2 [-]
Load level a  max
C 50/60 3.399 5.506 6.797 -0.0301 0.830
HSC 3.818 6.185 7.636 -0.0315 0.850

13
HPC 3.013 4.881 6.025 -0.0166 0.815
AAC 2.117 3.430 4.234 -0.0375 0.880

The values of the Basquin’s law coefficient A as presented in Table 5 corresponds/is in agreement with
the measured flexural strength fc,t as presented in Table 3 for each material and follows the same
observed trend, i.e. the AAC mixture has the lowest fatigue performance, and both the HSC and C 50/60
materials show a similar fatigue performance as well as in the mechanical tests. In contrast, the HPC
mixture shows a lower fatigue performance if compared to the HSC and C 50/60 materials, especially
in the cycle range from Nf = 1 cycle up to Nf = 2 ×104 cycles. This is due to the lower B coefficient of
the HPC material with the approx. value of -0.017, while the rest of the materials have the value of the
B coefficient equal to approx. -0.03. In addition, the lower value of the B coefficient results in a slower
fatigue degradation, which leads to a higher stress at the fatigue failure Nf.
This statement of a lower fatigue degradation can be documented by back-calculating the stress fat
values for a certain number of cycles Nf using the material’s constants A and B from Table 5 from
Basquin’s law. The comparison of the back-calculated stress levels are presented in Table 6 together
with the percentage difference to the static flexural strength fc,t.
Table 6: Comparison of calculated fatigue life prediction for all the studied concrete mixtures using evaluated Basquin’s law
coefficients.
Diff. to
A B
N fat C 50/60
Diff. to static
[cycles] [MPa] strength [%]
[%]
C 50/60 5.506 -0.0301 2×106 3.558 35 35
HSC 6.185 -0.0317 2×106 3.902 29 37
HPC 4.881 -0.0166 2×106 3.836 30 21
AAC 3.43 -0.0375 2×106 1.991 64 42

The back-calculated fatigue stress limits as presented in Table 6 used the fatigue failure at Nf of 2 ×106,
which leads to the following observation. One can see a reduction of the fatigue endurance limit for
both C 50/60 and HSC by 35%, by 21 % for HPC, and by 42 % for the ACC material, which is the
highest reduction. Furthermore, if the absolute values of the back-calculated fatigue stress levels fat are
compared, the HPC material’s stress is higher than in the C 50/60, which is in contradiction to the
measured static strengths.
Such statements are better observable from the comparison of the measured S-N curves for the values
of the stress range  for all materials in Figure 7.

Figure 7: Comparison of evaluated S-N curves for studied materials.

14
The comparison presented in Figure 7 shows a better fatigue resistance for the HPC mixture when the
load cycles exceed 1104 cycles. In addition, the HPC might surpass the HSC in the fatigue performance.
However, this might be the case for the load cycles Nf greater than 1107 cycles. Moreover, with the
current set-up and time consumption, it is impractical to perform experiments with load cycles greater
than 2106 cycles. Nevertheless, this presented fatigue result again confirms the necessity to perform
long-term fatigue tests of concrete materials as their service lifetime in the structure is scheduled to
decades.

5.2 Transition from the S-N Curves to the Paris’ Curves


Performing fatigue experiments on notched specimens allows for analysing the measured fatigue data
by using the Paris’ law. For this, the values of the SIF range KI and the crack growth rate da/dN have to
be assessed. The SIF KI is calculated by using Eq. (4) with the measured specimen’s geometry and the
maximum applied load per cycle Pmax. However, the crack growth rate da/dN is in question as the
experimental set-up is focused only on the measurement of load cycles until failure Nf for various stress
levels .
FCGR is usually measured on metallic materials, for which the crack increment per load cycle da/dN
can be accurately observed and measured on the specimen’s surface. This is hardly ever done for
concrete as the crack usually shifts from a straight path with the origin at the notch tip. On the other
hand, the heterogeneity of concrete and the application of the LEFM suggest various ways to set up the
proper crack growth rate without any measurement of the actual crack increment per load cycle.
Here we propose various options to establish the crack growth rate for which the brittle fracture will
occur after reaching the measured number of cycles until the fatigue failure Nf: a) the crack will
propagate throughout the specimen’s ligament da = 72 mm, b) the crack will propagate for da = 1 mm;
c) the crack will propagate for the length of the maximum aggregate size Dmax, i.e. da = 8 mm; and d)
the crack will propagate for the length equal to the critical distance rC. These assumptions are presented
in Figure 8.

Figure 8: Schematic representation of proposed fatigue crack propagation and specimen’s brittle failure.

The above-proposed assumptions as presented in Figure 8 lead to various FCGR da/dN, which can
significantly reduce the specimen’s fatigue performance. Therefore, we put focus on selecting the proper
ligament length in which the fatigue crack can propagate. Afterwards, we concentrated on the influence
of FCGR on the values of Paris’ law coefficients. A comparison of various FCGR of each proposed case
(a)-(d) are presented in Table 7.
Table 7: Various FCGR da/dN depending on the crack length increment da.
Nf = 1 Nf = 1104 Nf = 2106
Case
[cycle] [cycle] [cycle]
(a) – da = 72 mm 72 7.210-3 3.610-5
(b) – da = 1 mm 1 110-4 510-7
(c) – da = 8 mm 8 810-4 410-6

15
(d) – da = rC ≈ 4 mm 4 410-4 210-6

Assuming various distances in which the crack can propagate per load cycle, this definitely leads to
numerous values of the material’s constant C, and mainly to different specimen’s final failures. In
addition to this, FCGR da/dN has an insignificant influence on the value of the coefficient m, because
the slope of the Paris’ curve does not change with different values of FCGR and because of the use of
the S-N curves in this experimental study, for which KI values are evaluated. Omitting certain
experimental data has a more significant influence on the Paris’ material constants, for instance, omitting
the values of KI for N = 1, which represents the material’s fracture toughness KIC (case I in fitting results),
or the values of N > 2106 cycles (case II in fitting results) as they are close to the threshold value KI,th,
and for which the specimen can withstand cyclic load and did not fail in the brittle manner. These two
illustrative fitting conditions together with the various FCGR for the proposed cases (a)-(d) are presented
in Figure 9 for the HPC material.

(a) – case I (b) – case II


Figure 9: Influence of data range on the values of Paris’ law constants for HPC material in double logarithmic coordinates.

From the fitted experimental data presented in Figure 9, a clear difference in the value of the constant m
is observable, i.e. for case I – m = 24.454 and for case II – m = 18.858, respectively. Hence, for case I,
the fatigue crack is propagating more rapidly than for case II. Furthermore, if the coefficients of
determination R2 are compared, excluding data leads to a worse R2 = 0.811 for case II, while the whole
experimental data set leads to R2 = 0.878. Similarly to the parameter m, the fitted experimental data
presented in Figure 9 shows different values of the constant C. This is again related to the different
FCGR (see Table 7) and to the different data sets used in the evaluation of this constant. Nonetheless,
this difference in the values of the constant C for cases I and II has a considerable influence on the
evaluation of the fatigue life of the analysed structure than the values of the constant C obtained for the
cases (a)-(d).
Based on these shortcomings in the evaluation of the Paris’ material constants and based on the
experimental set-up, we propose a combination of fitting experimental data measured on both the S-N
and Paris’ curves in order to provide the most accurate fitting. For this, Eq. (12) and Eq. (2) are used to
obtain the lowest difference in the 1/m or β (from Eq. (12)) parameter. Both R2 and the difference of 1/m
are controlled to achieve the most accurate data fit as well as the Paris’ material constants. For this, the
experimental data set had to be adjusted, i.e. the experimental data was omitted in the fitting process to
achieve both the highest R2 value and the lowest difference of the 1/m value. The results of various
fitting conditions are presented in Table 8.
Table 8: Various fitting option for both S-N and Paris’ curves for data set measured on HPC material.
S-N curve based fit Eq. (12) Paris’ law based fit Eq. (2) Diff.
R2 [-]
1/m C/m m C m C 1/m C/m 1/m
All data 0.878 0.036 1.484 27.854 -41.329 24.455 -36.495 0.041 1.492 0.0050
Excl. N = 1 0.873 0.045 1.509 22.419 -33.823 19.572 -29.835 0.051 1.524 0.0065

16
Excl. N >2 106 0.844 0.033 1.483 29.979 -44.465 25.306 -37.740 0.039 1.491 0.0062
Excl. N = 1 and
0.811 0.043 1.562 23.259 -35.033 18.859 -28.796 0.053 1.527 0.010
N > 2  106
Excl. KI = max 0.859 0.035 1.482 28.471 -42.182 24.447 -36.485 0.041 1.492 0.0058
Excl. KI,max =
0.896 0.041 1.497 24.505 -36.672 21.947 -33.044 0.045 1.506 0.0048
min for N = 1
Best fit 0.915 0.039 1.497 25.320 -37.903 23.158 -34.837 0.043 1.504 0.0037

The fitting results presented in Table 8 show different values of the obtained material’s constants if both
approaches are used. Furthermore, the increase in the value of the coefficient of determination R2 reduces
the difference of the parameter 1/m. A clear improvement of fitting is observed, however, the difference
in both used methods remains. This is due to the material’s heterogeneity, which is in correlation with
the value of the parameter R2. To achieve the best fitting conditions, only three pieces of experimental
data have been excluded in the used data set for fitting. The used data set together with the achieved best
value of the coefficient of determination R2 is shown in Figure 10.

(a) (b)
Figure 10: Comparison of two fitting approaches which were used in the evaluation of FCGR in double logarithmic
coordinates – a) S-N and b) – Paris’ law.

5.3 Fatigue Crack Growth Rates


This proposed fitting methodology has also been used for the other materials discussed in this study, i.e.
C 50/60, HSC and AAC, with the same goals – increasing the R2 value and reducing the difference of
the 1/m value. The evaluated material constants for all four studied materials are presented in Table 9.
Table 9: Comparison of obtained Paris’ law material constants assuming the best fit of the adjusted experimental data set.
S-N curve based fit Eq.(12) Paris’ law based fit Eq. (2) Diff.
R2 [-]
1/m C/m m C m C 1/m C/m 1/m
C 50/60 0.917 0.031 1.434 32.834 -47.095 30.119 -43.348 0.033 1.439 0.0027
HPC 0.915 0.039 1.497 25.320 -37.903 23.158 -34.837 0.043 1.504 0.0037
HSC 0.920 0.018 1.407 56.917 -80.095 52.360 -73.915 0.019 1.412 0.0015
AAC 0.977 0.034 1.207 29.165 -35.191 28.485 -34.427 0.035 1.209 8  10-4

The presented evaluated material’s constants in Table 9 show a relatively good agreement between the
used methods, S-N-based and Paris-based, and the evaluated material constants show similar values for
all the studied materials. In contrast, the HSC material shows the biggest difference in both evaluated
constants of S-N as well as for the Paris’ curve when these obtained values are compared to the rest of
the materials. This is due to the insufficient number of experimental data (only 11 tests were performed).
Furthermore, the fatigue experiments for this material were carried out in the high-cycle (N > 104)
region, while for the rest of the studied materials, the fatigue experiments were carried out to cover the
whole fatigue spectrum from N = 1 cycle to N > 2  106 cycles. Additionally, a similar difference of the
material constant m was shown by Le in his study [56], yet again, these results were related to the low-
cycle fatigue region, while in this study, we focused more on the high-cycle fatigue region.

17
Furthermore, the constant C is influenced by choosing the proper ligament length in which the fatigue
crack can propagate. However, the values of the constant C for case (a) from Figure 8 and from Table 7
were used as the experimental data was obtained from the S-N curve, which showed a brittle failure of
the specimen. The comparison of the evaluated material constants based on the S-N curves (Eq. (12))
is presented in Figure 11, while the Paris’ based fitting (Eq. (2)) is shown in Figure 12.

(a) (b)

(c) (d)
Figure 11: Comparison of evaluated S-N based constants for all the studied materials in double logarithmic coordinates – (a)
C 50/60, (b) – HPC, (c) – HSC and (d) – AAC.

(a) (b)

18
(c) (d)
Figure 12: Comparison of evaluated Paris’ law constants for all the studied materials in double logarithmic coordinates – (a)
C 50/60, (b) – HPC, (c) – HSC and (d) – AAC.

The excluded data in the fitting can be observed in both Figure 11 and Figure 12. In total, a maximum
of 3 experimental results were excluded to achieve the best value of R2 and the lowest difference of 1/m.
Furthermore, this excluded data had a more minor influence on the obtained material’s constant than if
the measured data for certain cycles were excluded, e.g., all data with the load cycles of Nf > 2 106.

(a) (b)
Figure 13: Comparison of obtained Paris’ curves for all the studied materials – (a) absolute value coordinates and (b) –
relative coordinates of KI/KIC.

To illustrate and compare the fatigue behaviour of the studied concrete mixtures, the obtained Paris’
curves were compared in absolute coordinates, i.e. da/dN – KI,max (see Figure 13(a)) and in relative
coordinates, i.e. da/dN – KI,max/KIC (Figure 13(b)), respectively. Such a comparison of FCGR presented
in relative coordinates (Figure 13(b)) shows that the C 50/60 and HSC mixtures have almost identical
curves. Furthermore, the AAC mixture shows higher FCGR for lower values of SIF (due to the lowest
mechanical performance) and the HPC mixture exhibits lower values of FCGR for higher values of SIF
(due to the highest mechanical performance). This observation results in a higher applied stress, which
is again in agreement with the observation found based on the measured mechanical properties.

5.4 Validation of the Proposed Methodology


In order to justify and verify the proposed methodology for the assessment of fatigue failure, the
evaluated/measured Paris’ law constants are used in the Eq. (10) from which the number of cycles to
fatigue failure Nf can be obtained. The use of Eq. (10) suggests compiling a simulated S-N curve of the
studied material by plotting the maximum stress max against the calculated number of cycles to the
fatigue failure Nf.

19
Since concrete is a quasi-brittle material and its behaviour is characterized by the presence of FPZ, the
initial crack length a0 in Eq. (10) was extended based on the assumption made in Figure 1, i.e. the fatigue
crack propagates from the notch tip in the distance of rC. Afterwards, the stress has reached its critical
value or tensile strength, which resulted in the brittle fracture/failure of the specimen. Eq. (6) and Eq.
(7) were used to obtain the critical distance rC. The calculated values of the critical distance rC are based
on the measured values of the fracture toughness KIC for N = 1 and on the measured flexural strength fc,t.
The calculated values of the critical distances rC for all the investigated concrete mixtures are presented
in Table 10.
Table 10: Comparison of measured flexural strength fc,t and fracture toughness KIC with equivalent critical distances rC.
rC – plane rC – plane
Material fc,t [MPa] KIC [MPamm1/2]
stress [mm] strain [mm]
C 50/60 5.52 31.1 5.079 1.693
HPC 6.30 34.1 4.664 1.555
HSC 9.30 27.2 1.364 0.455
AAC 3.15 19.4 6.045 2.015

The presented values of critical distances are again in agreement with the general expectation based on
the measured material’s mechanical properties, i.e. the AAC material shows the highest values, while
the HSC shows the lowest ones. This can be related to the highest measured materials’ strengths of all
the tested materials, which is usually accompanied by a brittle behaviour. Moreover, the difference in
the values of KIC presented in Table 4 and in Table 10 is due to the different age of the concrete
specimens, i.e. the values in Table 10 were measured at the age higher than 90 days to be able to relate
it to fatigue tests as accurately as possible.
The back-calculated S-N curves were compiled for all four studied concrete mixtures by using Eq. (10)
with the Paris’ constants C and m from Table 9. Both cases of these constants, evaluated based on the
S-N and Paris’-based, were considered in the prediction of cycles until the fatigue failure Nf.
Subsequently, the back-calculated S-N curves considered no extension of the initial notch and both
values of critical distances rC to show its importance in the prediction of the fatigue life. The compiled
simulated S-N curves are presented in Figure 14.

(a) (b)

20
(c) (d)
Figure 14: Back-calculated S-N curves for the studied mixtures (a) – C 50/60, (b) – HPC, (c) – HSC and (d) – AAC.

The back-calculated S-N curves as presented in Figure 14 were compared with the experimentally
measured experimental data for the maximum stress max. A clear influence of the extension of the initial
notch a0 can be observed for all four investigated mixtures. Furthermore, the use of the constants m and
1/m obtained by both S-N and Paris’ curves ensures the same slope of the back-calculated curves.
Moreover, such compiled curves show a relatively good agreement with the experimental data in the
high-cycle fatigue region. This is related to the linear nature of the fatigue failure and to the undeveloped
FPZ in the tested specimen. Hence, this validates the use of LEFM for the assessment of the fatigue
failure of concrete materials.
Another interesting observation can be made in Figure 14. The notch tip a0 extension by the critical
distance rC clearly improves the prediction of the fatigue life, and it can represent higher and lower limits
of the fatigue curves. This fact is definitely supported by the fatigue results of the HPC mixture, for
which the discrepancy from the experimental data is relatively low. In contradiction to this is the
experimental data of the HSC mixture accompanied by a higher error in the predicted fatigue life. The
C 50/60 and AAC mixtures have more dispersed simulated S-N curves with the extent of the critical
distance for the plane stress condition from the experimental data. Nonetheless, these curves still belong
to the range of the measured experimental data.
Both simulated and experimentally evaluated fatigue curves are highly influenced by the typical
material’s heterogeneity and by the total number of the performed fatigue experimental tests, which are
not able to fully cover the whole testing range. This is clearly related to the number of the performed
fatigue experimental tests, e.g. in the case of the HSC mixture, for which only 11 specimens have been
tested, which definitely influenced the obtained material’s constants, and it resulted in a high error of
the simulated S-N curves. Therefore, we recommend using at least 15, preferably 24 specimens for cyclic
tests to reduce the error present in material constants, which constitutes in the same error in the back-
calculated S-N curves. This statement is mainly valid for the HPC mixture for which 24 specimens were
tested and the experimental data shows a reasonable scatter.
Although the proposed methodology for the transition from the S-N curve to the Paris’ curve provides
similar values of the material’s constant m, it still remains a great challenge to experimentally measure
the crack increment over a load cycle da/dN on a set-up for an S-N curve measurement. This, if
successful, may improve the accuracy of the value of the constant m as well as the whole proposed
method. Similarly, size-effect fatigue experiments are worthy of special attention on their own to see if
the size effect applies to the cases with more than 1  105 load cycles applied on the specimen, or if the
fatigue failure remains in the linear/brittle manner.
Lastly, these experimental results can eventually lead to adjustment of the calculation of the thickness
of the concrete protective cover layer of the steel reinforcement. Undetected crack in the protective cover

21
layer can start to propagate due to cyclic load, and eventually it can reach steel reinforcement. This can
significantly reduce structural service life due to the possible exposure of the reinforcement to air
humidity or to an aggressive environment, both phenomena leading to reinforcement corrosion.
Conclusions
The presented experimental study focused on the comprehensive fatigue analysis of various concrete
mixtures, namely C 50/60, HPC, HSC, and AAC, and proposed a novel, yet simple method for the
transition from the S-N field to the Paris’ curves.
The fatigue experimental tests were performed on a three-point bending (TPB) test set-up with an initial
notch. These specimens were loaded with a cyclic load with an asymmetry of cycle R = 0.1. Such
specimens were tested mainly in the high-cycle fatigue region with the number of load cycles higher
than 2  104 cycles. The fatigue experimental results are presented in the form of an S-N curve, from
which a material’s constants of Basquin’s power law were determined.
Furthermore, the use of notched specimens suggested to evaluate the fatigue experimental data by the
Paris’ law used in the assessment of the fatigue crack growth rate. For this, a novel transition method
from the S-N field to the Paris’ curves was proposed based on a simple function inversion. The evaluated
material’s constants m and C are related to the constants obtained from Basquin’s law. The applicability
of this method was assessed by a back calculation of the load cycles until the fatigue failure Nf using
material constants from both approaches.
The back calculation and the subsequent compilation of the simulated S-N curves employed the initial
notch tip a0 extension by the value of critical distance rC. Such results of the simulated S-N curves
verified the applicably of this proposed transition method and provided the upper and lower limits for
the fatigue life assessment. The discrepancy between the back-calculated S-N curve and the experimental
results is mainly related to the significant material’s heterogeneity and to the insufficient number of
tested specimens. Based on the experimental results presented in this study, a minimum of 15 (preferably
24) specimens are recommended to be tested to cover the material’s heterogeneity, which has a negative
and major influence on the evaluated values of the material’s constants.
Lastly, this experimental study proved the importance of fatigue testing of concrete materials in the
high-cycle fatigue region together with the assessment of the remaining fatigue life. This newly
proposed transition from the S-N curve to the Paris’s law represents a simple extension of the
experimental data. This method may lead to a wider use in experimental research as well as in applied
structural design due to its relatively simple experimental set-up and due to the adoption of one set of
experimental data for both approaches to obtain the structure’s fatigue life assessment. Nevertheless, the
fatigue failure of concrete materials is an interesting open topic deserving of further investigation.
Acknowledgements
Financial support provided by the Czech Science Foundation under project no. 21-08772S, as well as
by the “Ministerio de Ciencia, Innovación y Universidades” of the Spanish Government under project
no. PID2019-110928RB-C33 is gratefully acknowledged.
The first author acknowledges financial support of a project: “International mobility of researchers at
the Brno University of Technology II” funded by the MŠMT of the Czech Republic with registration
no. CZ.02.2.69/0.0/0.0/18_053/0016962.
Appendix A
Inversion of the Paris’ power law by transformation to double logarithmic coordinates of Eq. (2):

log = log(𝐶) + 𝑚 ∙ log (𝐾)

−𝐶 = 𝑚∙𝐾

22
− 𝐶=𝐾

Appendix B
The measured fatigue experimental data for all the studied mixtures.
Table 11: C 50/60 material

Specimen
Pmax [KN] N [cycle] Note
nmr.
1 10.5 1 Static test
2 10.5 1 Static test
3 9.6 154
4 8.5 166
5 8.6 1 370
6 7.6 11 739
7 7.2 15 731
8 7 16 040
9 8 79 960
10 7.4 97 065
11 7 150 410
12 8 402 258
13 6.8 2 000 000 Runout
14 6.8 2 565 308 Runout
15 7.8 3 431 548 Runout

Table 12: HPC

Specimen
Pmax [KN] N [cycle] Note
nmr.
1 12.5 1 Static test
2 11.6 1 Static test
3 11.4 1 Static test
4 10.6 1 Static test
5 10.3 50
6 10.3 80
7 9.4 1 094
8 10 1 524
9 10 1 840
10 9 2 686
11 9.8 3 387
12 9.6 4 168
13 9 10 181
14 9.2 11 098
15 8.6 27 796
16 7.8 36536
17 8.5 40 093
18 7.2 65 811
19 8.2 102 173
20 7.6 425 431
21 6.8 840 512
22 6.6 2 000 000 Runout
23 6.6 2 161 575 Runout
24 7.2 2 544 755 Runout

Table 13: HSC

Specimen
Pmax [KN] N [cycle] Note
nmr.
1 9.2 1 Static test
2 8 202 317
3 7.4 280 562
4 7.9 12 924

23
5 8.2 5 983
6 7.7 225 131
7 7.5 1 099 035
8 7.2 311 902
9 7 325 060
10 7.2 2 000 000 Runout
11 7 2 000 000 Runout

Table 14: ACC material

Specimen
Pmax [KN] N [cycle] Note
nmr.
1 6.2 1 Static test
2 7.1 1 Static test
3 6.4 1 Static test
4 5.5 121
5 5 1 106
6 4.8 2 106
7 4 10 137
8 4.2 49 926
9 4 96 204
10 4.1 503 895
11 4.5 1 643 729
12 3.8 2 000 000 Runout
13 3.8 2 000 000 Runout
14 3.8 2 478 717 Runout
15 3.8 2 550 911 Runout
16 4 3 000 000 Runout

References
[1] A.E. Naaman, Prestressed concrete analysis and design: Fundamentals, McGraw-Hill New York,
1982.
[2] E.G. Nawy, Prestressed Concrete: A Fundamental Approach, : Pearson College Div, 2005.
[3] A. Surahyo, Concrete Construction, Springer, Cham, Switzerland, 2019.
[4] X. Zhou, J. Di, X. Tu, Investigation of collapse of Florida International University (FIU) pedestrian
bridge, Engineering Structures, 200 (2019) 109733.
[5] M. Morgese, F. Ansari, M. Domaneschi, G.P. Cimellaro, Post-collapse analysis of Morandi’s
Polcevera viaduct in Genoa Italy, Journal of Civil Structural Health Monitoring, 10 (2020) 69-85.
[6] M. Domaneschi, C. Pellecchia, E. De Iuliis, G.P. Cimellaro, M. Morgese, A.A. Khalil, F. Ansari,
Collapse analysis of the Polcevera viaduct by the applied element method, Engineering Structures, 214
(2020) 110659.
[7] M. Schneider, The cement industry on the way to a low-carbon future, Cement and Concrete
Research, 124 (2019) 105792.
[8] S.A. Miller, V.M. John, S.A. Pacca, A. Horvath, Carbon dioxide reduction potential in the global
cement industry by 2050, Cement and Concrete Research, 114 (2018) 115-124.
[9] M.A. Caldarone, High-Strength Concrete: A Practical Guide, CRC Press, 2019.
[10] P.-C. Aïtcin, R.J. Flatt, Science and technology of concrete admixtures, Woodhead publishing,
2015.
[11] E.G. Nawy, Fundamentals of High-Performance Concrete, Wiley, 2001.
[12] J.L. Provis, J.S.J. van Deventer, Alkali Activated Materials: State-of-the-Art Report, RILEM TC
224-AAM, Springer Netherlands, 2013.
[13] J.L. Provis, Alkali-activated materials, Cement and Concrete Research, 114 (2018) 40-48.
[14] A. Neville, Chloride attack of reinforced concrete: an overview, Materials and Structures, 28 (1995)
63-70.
[15] M. Kušter Marić, J. Ožbolt, G. Balabanić, Reinforced concrete bridge exposed to extreme maritime
environmental conditions and mechanical damage: Measurements and numerical simulation,
Engineering Structures, 205 (2020) 110078.

24
[16] K.E. Hassan, J.G. Cabrera, R.S. Maliehe, The effect of mineral admixtures on the properties of
high-performance concrete, Cement and Concrete Composites, 22 (2000) 267-271.
[17] R. Yu, P. Spiesz, H.J.H. Brouwers, Development of an eco-friendly Ultra-High Performance
Concrete (UHPC) with efficient cement and mineral admixtures uses, Cement and Concrete
Composites, 55 (2015) 383-394.
[18] E. Vejmelková, D. Koňáková, T. Kulovaná, M. Keppert, J. Žumár, P. Rovnaníková, Z. Keršner, M.
Sedlmajer, R. Černý, Engineering properties of concrete containing natural zeolite as supplementary
cementitious material: Strength, toughness, durability, and hygrothermal performance, Cement and
Concrete Composites, 55 (2015) 259-267.
[19] C.S. Poon, S.C. Kou, L. Lam, Compressive strength, chloride diffusivity and pore structure of high
performance metakaolin and silica fume concrete, Construction and Building Materials, 20 (2006) 858-
865.
[20] E.O.L. Lantsoght, C. van der Veen, A. de Boer, Proposal for the fatigue strength of concrete under
cycles of compression, Construction and Building Materials, 107 (2016) 138-156.
[21] M.K. Lee, B.I.G. Barr, An overview of the fatigue behaviour of plain and fibre reinforced concrete,
Cement and Concrete Composites, 26 (2004) 299-305.
[22] S. Korte, V. Boel, W. De Corte, G. De Schutter, Behaviour of fatigue loaded self-compacting
concrete compared to vibrated concrete, Structural Concrete, 15 (2014) 575-589.
[23] R. François, S. Laurens, F. Deby, Corrosion and its Consequences for Reinforced Concrete
Structures, ISTE Press - Elsevier, 2018.
[24] B. Teplý, D. Vorechovská, Reinforcement Corrosion: Limit States, Reliability and Modelling,
Journal of Advanced Concrete Technology, 10 (2012) 353-362.
[25] J. Zhang, V.C. Li, H. Stang, Size Effect on Fatigue in Bending of Concrete, Journal of Materials in
Civil Engineering, 13 (2001) 446-453.
[26] B.S. Saini, S.P. Singh, Flexural fatigue life analysis of self compacting concrete containing 100%
coarse recycled concrete aggregates, Construction and Building Materials, 253 (2020) 119176.
[27] B.S. Saini, S.P. Singh, Flexural fatigue strength prediction of self compacting concrete made with
recycled concrete aggregates and blended cements, Construction and Building Materials, 264 (2020)
120233.
[28] S. Arora, S.P. Singh, Analysis of flexural fatigue failure of concrete made with 100% Coarse
Recycled Concrete Aggregates, Construction and Building Materials, 102 (2016) 782-791.
[29] S. Goel, S.P. Singh, Fatigue performance of plain and steel fibre reinforced self compacting
concrete using S–N relationship, Engineering Structures, 74 (2014) 65-73.
[30] J. Qiu, E.-H. Yang, Micromechanics-based investigation of fatigue deterioration of engineered
cementitious composite (ECC), Cement and Concrete Research, 95 (2017) 65-74.
[31] J. Qiu, X.-N. Lim, E.-H. Yang, Fatigue-induced deterioration of the interface between micro-
polyvinyl alcohol (PVA) fiber and cement matrix, Cement and Concrete Research, 90 (2016) 127-136.
[32] J.D. Ríos, H. Cifuentes, R.C. Yu, G. Ruiz, Probabilistic Flexural Fatigue in Plain and Fiber-
Reinforced Concrete, Materials, 10 (2017) 767.
[33] M.A. Farooq, Y. Sato, T. Ayano, K. Niitani, Experimental and numerical investigation of static and
fatigue behavior of mortar with blast furnace slag sand as fine aggregates in air and water, Construction
and Building Materials, 143 (2017) 429-443.
[34] T. Miura, K. Sato, H. Nakamura, Influence of primary cracks on static and fatigue compressive
behavior of concrete under water, Construction and Building Materials, 305 (2021) 124755.
[35] E. Poveda, G. Ruiz, H. Cifuentes, R.C. Yu, X. Zhang, Influence of the fiber content on the
compressive low-cycle fatigue behavior of self-compacting SFRC, International Journal of Fatigue, 101
(2017) 9-17.
[36] S. Blasón, E. Poveda, G. Ruiz, H. Cifuentes, A. Fernández Canteli, Twofold normalization of the
cyclic creep curve of plain and steel-fiber reinforced concrete and its application to predict fatigue
failure, International Journal of Fatigue, 120 (2019) 215-227.
[37] J.J. Ortega, G. Ruiz, E. Poveda, D.C. González, M. Tarifa, X.X. Zhang, R.C. Yu, M.Á. Vicente, Á.
de la Rosa, L. Garijo, Size effect on the compressive fatigue of fibre-reinforced concrete, Construction
and Building Materials, 322 (2022) 126238.
[38] E. Poveda, R.C. Yu, J.C. Lancha, G. Ruiz, A numerical study on the fatigue life design of concrete
slabs for railway tracks, Engineering Structures, 100 (2015) 455-467.

25
[39] I.D. Unobe, A.D. Sorensen, Multi-hazard analysis of a wind turbine concrete foundation under
wind fatigue and seismic loadings, Structural Safety, 57 (2015) 26-34.
[40] X. Bai, M. He, R. Ma, D. Huang, J. Chen, Modelling fatigue degradation of the compressive zone
of concrete in onshore wind turbine foundations, Construction and Building Materials, 132 (2017) 425-
437.
[41] J.A. de Lana, P.A.A.M. Júnior, C.A. Magalhães, A.L.M.A. Magalhães, A.C. de Andrade Junior,
M.S. de Barros Ribeiro, Behavior study of prestressed concrete wind-turbine tower in circular cross-
section, Engineering Structures, 227 (2021) 111403.
[42] P. Ganesh, A. Ramachandra Murthy, Static and fatigue responses of retrofitted RC beams with
GGBS based UHPC strips, Engineering Structures, 240 (2021) 112332.
[43] J. Wu, J. Xu, B. Diao, L. Jin, X. Du, Fatigue life prediction for the reinforced concrete (RC) beams
under the actions of chloride attack and fatigue, Engineering Structures, 242 (2021) 112543.
[44] European Committee for Standardization, Eurocode 2: Design of concrete structures—Part 1-1:
General rules and rules for buildings, Brussels, Belgium, (2004).
[45] ACI PRC-215-92: Considerations for Design of Concrete Structures Subjected to Fatigue Loading
(1997).
[46] CEB-FIB, Model Code 2010, First complete draft, Bulletin, 55 (2010).
[47] O.H. Basquin, The exponential law of endurance tests, in, pp. 625-630.
[48] P. Paris, F. Erdogan, A Critical Analysis of Crack Propagation Laws, Journal of Basic Engineering,
85 (1963) 528-533.
[49] S. Seitl, P. Miarka, V. Bílek, The mixed-mode fracture resistance of C 50/60 and its suitability for
use in precast elements as determined by the Brazilian disc test and three-point bending specimens,
Theoretical and Applied Fracture Mechanics, 97 (2018) 108-119.
[50] P. Miarka, S. Seitl, V. Bílek, Mixed-mode fracture analysis in high-performance concrete using a
Brazilian disc test, Materiali in Tehnologije, 53 (2019) 233-238.
[51] P. Miarka, R. Janssen, S. Seitl, W. de Corte, A Comparison of the Fracture Behaviour of Various
Concrete Grades under Mixed Mode I/II Loading, Trans Tech Publications Ltd, 2020.
[52] P. Miarka, L. Pan, V. Bílek, S. Seitl, H. Cifuentés, Influence of the chevron notch type on the values
of fracture energy evaluated on alkali-activated concrete, Engineering Fracture Mechanics, 236 (2020)
107209.
[53] L. Susmel, A unifying methodology to design un-notched plain and short-fibre/particle reinforced
concretes against fatigue, International Journal of Fatigue, 61 (2014) 226-243.
[54] R.O. Ritchie, Incomplete self-similarity and fatigue-crack growth, International Journal of Fracture,
132 (2005) 197-203.
[55] R.H. Dauskardt, M.R. James, J.R. Porter, R.O. Ritchie, Cyclic Fatigue-Crack Growth in a SiC-
Whisker-Reinforced Alumina Ceramic Composite: Long- and Small-Crack Behavior, Journal of the
American Ceramic Society, 75 (1992) 759-771.
[56] J.-L. Le, J. Manning, J.F. Labuz, Scaling of fatigue crack growth in rock, International Journal of
Rock Mechanics and Mining Sciences, 72 (2014) 71-79.
[57] Z.P. Bazant, K. Xu, Size Effect in Fatigue Fracture of Concrete, Materials Journal, 88 (1991) 390-
399.
[58] Z.P. Bazant, W.F. Schell, Fatigue Fracture of High-Strength Concrete and Size Effect, Materials
Journal, 90 (1993) 472-478.
[59] R.O. Ritchie, Mechanisms of fatigue-crack propagation in ductile and brittle solids, International
Journal of Fracture, 100 (1999) 55-83.
[60] S. Ray, J.M. Chandra Kishen, Fatigue crack propagation model and size effect in concrete using
dimensional analysis, Mechanics of Materials, 43 (2011) 75-86.
[61] K. Kirane, Z.P. Bažant, Microplane damage model for fatigue of quasibrittle materials: Sub-critical
crack growth, lifetime and residual strength, International Journal of Fatigue, 70 (2015) 93-105.
[62] N. Pugno, M. Ciavarella, P. Cornetti, A. Carpinteri, A generalized Paris’ law for fatigue crack
growth, Journal of the Mechanics and Physics of Solids, 54 (2006) 1333-1349.
[63] K. Kirane, Z.P. Bažant, Size effect in Paris law and fatigue lifetimes for quasibrittle materials:
Modified theory, experiments and micro-modeling, International Journal of Fatigue, 83 (2016) 209-220.

26
[64] L. Xu, M. Ma, L. Li, Y. Xiong, W. Liu, Continuum-based approach for modelling the flexural
behaviour of plain concrete beam under high-cycle fatigue loads, Engineering Structures, 241 (2021)
112442.
[65] A. Baktheer, J. Hegger, R. Chudoba, Enhanced assessment rule for concrete fatigue under
compression considering the nonlinear effect of loading sequence, International Journal of Fatigue, 126
(2019) 130-142.
[66] M. Klesnil, P. Lukáš, Effect of stress cycle asymmetry on fatigue crack growth, Materials Science
and Engineering, 9 (1972) 231-240.
[67] M. Klesnil, P. Lukác, Fatigue of Metallic Materials, Volume 71, Elsevier Science, Walthm, MA,
USA, 1992.
[68] T. Vojtek, P. Pokorný, T. Oplt, M. Jambor, L. Náhlík, D. Herrero, P. Hutař, Classically determined
effective ΔK fails to quantify crack growth rates, Theoretical and Applied Fracture Mechanics, 108
(2020) 102608.
[69] L. Saucedo, R.C. Yu, A. Medeiros, X. Zhang, G. Ruiz, A probabilistic fatigue model based on the
initial distribution to consider frequency effect in plain and fiber reinforced concrete, International
Journal of Fatigue, 48 (2013) 308-318.
[70] Z.P. Bažant, L.F. Estenssoro, Surface singularity and crack propagation, International Journal of
Solids and Structures, 15 (1979) 405-426.
[71] T. Nakamura, D.M. Parks, Antisymmetrical 3-D stress field near the crack front of a thin elastic
plate, International Journal of Solids and Structures, 25 (1989) 1411-1426.
[72] L.P. Pook, Some implications of corner point singularities, Engineering Fracture Mechanics, 48
(1994) 367-378.
[73] P.F.P. de Matos, D. Nowell, The influence of the Poisson’s ratio and corner point singularities in
three-dimensional plasticity-induced fatigue crack closure: A numerical study, International Journal of
Fatigue, 30 (2008) 1930-1943.
[74] P. Hutař, L. Náhlík, Z. Knésl, The effect of a free surface on fatigue crack behaviour, International
Journal of Fatigue, 32 (2010) 1265-1269.
[75] T. Oplt, P. Hutar, P. Pokorný, L. Náhlík, Z. Chlup, F. Berto, Effect of the free surface on the fatigue
crack front curvature at high stress asymmetry, International Journal of Fatigue, 118 (2019) 249-261.
[76] K.L. Scrivener, A.K. Crumbie, P. Laugesen, The Interfacial Transition Zone (ITZ) Between
Cement Paste and Aggregate in Concrete, Interface Science, 12 (2004) 411-421.
[77] RILEM TCS Recommendations, Determination of the fracture energy of mortar and concrete by
means of three-point bend tests on notched beams, Materials and structures, 18 (1985) 285-290.
[78] B.L. Karihaloo, Fracture Mechanics and Structural Concrete (Concrete Design and Construction
Series), Ed. Longman Scientific & Technical. United States, (1995).
[79] A. Carpinteri, Handbook of Fatigue Crack Propagation in Metallic Structures, Elsevier Science,
1994.
[80] M.L. Williams, On the Stress Distribution at the Base of a Stationary Crack, Journal of Applied
Mechanics, 24 (1956) 6.
[81] ASTM E647-15e1, Standard Test Method for Measurement of Fatigue Crack Growth Rates, in,
ASTM International, West Conshohocken, PA, 2015.
[82] G.R. Irwin, Analysis of stresses and strains near the end of a crack traversing a plate, Journal of
Applied Mechanics, 24 (1957) 361-364.
[83] M.R. Ayatollahi, J. Akbardoost, Size effects on fracture toughness of quasi-brittle materials – A
new approach, Engineering Fracture Mechanics, 92 (2012) 89-100.
[84] P. Miarka, S. Seitl, M. Horňáková, P. Lehner, P. Konečný, O. Sucharda, V. Bílek, Influence of
chlorides on the fracture toughness and fracture resistance under the mixed mode I/II of high-
performance concrete, Theoretical and Applied Fracture Mechanics, 110 (2020) 102812.
[85] S. Muralidhara, B.K.R. Prasad, H. Eskandari, B.L. Karihaloo, Fracture process zone size and true
fracture energy of concrete using acoustic emission, Construction and Building Materials, 24 (2010)
479-486.
[86] A. Hillerborg, M. Modéer, P.E. Petersson, Analysis of crack formation and crack growth in concrete
by means of fracture mechanics and finite elements, Cement and Concrete Research, 6 (1976) 773-781.

27
View publication stats

[87] Z.P. Bažant, Mechanics of fracture and progressive cracking in concrete structures, in: Fracture
mechanics of concrete: Structural application and numerical calculation, Springer, Dordrecht, The
Netherlands, 1985, pp. 1-94.
[88] K. Otsuka, H. Date, Fracture process zone in concrete tension specimen, Engineering Fracture
Mechanics, 65 (2000) 111-131.
[89] N. Alanazi, L. Susmel, Theory of Critical Distances and static/dynamic fracture behaviour of un-
reinforced concrete: length scale parameters vs. material meso-structural features, Engineering Fracture
Mechanics, 261 (2022) 108220.
[90] X.Z. Hu, F.H. Wittmann, Fracture energy and fracture process zone, Materials and Structures, 25
(1992) 319-326.
[91] X. Hu, K. Duan, Influence of fracture process zone height on fracture energy of concrete, Cement
and Concrete Research, 34 (2004) 1321-1330.
[92] B.L. Karihaloo, P. Nallathambi, Effective crack model for the determination of fracture toughness
(KIce) of concrete, Engineering Fracture Mechanics, 35 (1990) 637-645.
[93] F. Pacheco-Torgal, J.A. Labrincha, C. Leonelli, A. Palomo, P. Chindaprasirt, Handbook of Alkali-
Activated Cements, Mortars and Concretes, Woodhead Publishing, 2015.
[94] V. Bilek, T. Opravil, F. Soukal, Searching for practically applicable alkali-activated concretes, in:
C.S.a.X. Shen (Ed.) First International Conference on Advances in Chemically-Activated Materials
CAM'2010, 2010, pp. 28-35.
[95] V. Bilek, J. Hurta, P. Done, L. Zidek, Development of alkali-activated concrete for structures –
Mechanical properties and durability, Perspectives in Science, 7 (2016) 190-194.
[96] H. Szklorzova, V. Bilek, Influence of alkali ions in the activator on the performance of alkali-
activated mortars, in: B.a. Kersner (Ed.) 3rd International symposium Nontraditional cement and
concrete composites, Brno, 2003, pp. 777-784.
[97] P. Miarka, A.S. Cruces, S. Seitl, L. Malíková, P. Lopez-Crespo, Evaluation of the SIF and T-stress
values of the Brazilian disc with a central notch by hybrid method, International Journal of Fatigue, 135
(2020) 105562.
[98] A. Medeiros, X. Zhang, G. Ruiz, R.C. Yu, M.d.S.L. Velasco, Effect of the loading frequency on
the compressive fatigue behavior of plain and fiber reinforced concrete, International Journal of Fatigue,
70 (2015) 342-350.
[99] B. Zhang, D.V. Phillips, K. Wu, Effects of loading frequency and stress reversal on fatigue life of
plain concrete, Magazine of Concrete Research, (2015).
[100] S. Seitl, P. Miarka, H. Šimonová, P. Frantík, Z. Keršner, J. Domski, J. Katzer, Change of Fatigue
and Mechanical Fracture Properties of a Cement Composite due to Partial Replacement of Aggregate
by Red Ceramic Waste, Periodica Polytechnica Civil Engineering, 63 (2019) 152-159.
[101] H. Šimonová, B. Kucharczyková, V. Bílek, L. Malíková, P. Miarka, M. Lipowczan, Mechanical
Fracture and Fatigue Characteristics of Fine-Grained Composite Based on Sodium Hydroxide-Activated
Slag Cured under High Relative Humidity, Applied Sciences, 11 (2020) 259.

28

You might also like