You are on page 1of 17

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

net/publication/353148446

Early age cracking risk in a massive concrete foundation slab: Comparison of


analytical and numerical prediction models with on-site measurements

Article  in  Construction and Building Materials · September 2021


DOI: 10.1016/j.conbuildmat.2021.124135

CITATIONS READS

12 671

4 authors:

Aneta Smolana Barbara Klemczak


Silesian University of Technology Silesian University of Technology
18 PUBLICATIONS   144 CITATIONS    57 PUBLICATIONS   621 CITATIONS   

SEE PROFILE SEE PROFILE

Miguel Azenha Dirk Schlicke


University of Minho Graz University of Technology
242 PUBLICATIONS   2,806 CITATIONS    74 PUBLICATIONS   405 CITATIONS   

SEE PROFILE SEE PROFILE

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

COST TU1404 STSM: Evaluation of cracking charactestics of cement-based materials and multiphysics simulations of early age concrete View project

COST TU1404 WG2 benchmark on numerical modelling of CBM at early age - stage 1 View project

All content following this page was uploaded by Barbara Klemczak on 15 July 2021.

The user has requested enhancement of the downloaded file.


Construction and Building Materials 301 (2021) 124135

Contents lists available at ScienceDirect

Construction and Building Materials


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

Early age cracking risk in a massive concrete foundation slab: Comparison


of analytical and numerical prediction models with on-site measurements
Aneta Smolana a, *, Barbara Klemczak a, Miguel Azenha b, Dirk Schlicke c
a
The Silesian University of Technology, Faculty of Civil Engineering, Akademicka 5, 44-100 Gliwice, Poland
b
ISISE, University of Minho, Campus de Azurem, 4800-058 Guimarães, Portugal
c
The Graz University of Technology, Institute of Structural Concrete, Lessingstraße 25/1, A-8010 Graz, Austria

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

Keywords: Mass foundation slabs represent unique structures because of the considerable thermal effects generated by the
Mass concrete exothermic reactions of cement hydration that occur during concrete curing. Arising temperature variations in
Early age cracking early age concrete generate tensile thermal stresses that may reach values resulting in cracking in foundation
Hydration
slabs. That is why a realistic estimation of early-age thermal loads and induced stresses is essential in engineering
Thermal stress
Foundation slabs
science and practice. In this work, several alternative methods that can be used in the assessment of the early age
Modelling cracking risk have been thoroughly reviewed and discussed. First, a brief review of analytic and numerical
methods has been performed to present the possible design-making paths. Next, a real mass foundation slab is
analysed using the described analytical and numerical methods, with simultaneous reference to the measure­
ments made during the construction process. Finally, the advantages and weak points of each method are
discussed.

1. Introduction bending stresses. Nevertheless, restraint stresses are routinely consid­


ered to be of lower importance in mass foundation slabs [6]. Self-
ACI 116R [1] describes mass concrete as: ‘any volume of concrete induced and restraint stresses might gain significant values matching
with dimensions large enough to require the measures be taken to cope the actual tensile strength of the concrete and consequently cause
with generation of heat from the hydration of the cement and attendant cracking. That is of relevant importance from the durability point of
volume change, to minimize cracking.’ In such structures, during the view, as early age cracking both can initiate corrosion of the rein­
concrete curing, the interior of the mass concrete element initially ex­ forcement and reduce the later load-bearing capacity of the structure
pands more than the external layers cooled due to the heat exchange [7].
with the environment. The reason for such volume changes is a tem­ In thick foundation slabs, self-induced stresses associated with the
perature increase as a result of the exothermic nature of the cement non-uniform volume changes due to considerable differences in tem­
hydration. Due to the low thermal conductivity of concrete, high- perature between the interior and surface layers of the element are of
temperature gradients may occur between the center and the surface predominant importance [6]. Considering the stress distribution over
of structural elements with thick cross-sections. The consequence of the slab thickness during concrete hardening, cracks can appear in two
these volume changes is the formation of stresses in the concrete stages of concrete curing. In the phase of the hydration temperature
element. Developing stresses might be described as self-induced stresses increase, cracks may appear at the external surfaces of the element,
and restrained stresses [1–5]. Self-induced stresses arise due to the in­ especially at the top surface. Later, the cracks may close when the
ternal restraints in the element, resulting from heterogeneous volu­ temperature of the whole concrete member equilibrates [8]. In the
metric changes within the element. External limitation of the cooling phase, the possibility of cracks formation is related to the
deformation freedom additionally causes restraint stresses. In founda­ inversion of the stresses’ sign at the element thickness [9,10]. At that
tion slabs, restraints occur in the contact between the bottom surface of time, cracks may be expected in the interior of the structure.
the slab and the subsoil. Besides, gradually distributed deformations Difficulties with the assessment of the thermal cracking risk in mass
across the thickness of the slab cause uplift which produces in turn concrete structures are related to many factors affecting the magnitude

* Corresponding author.
E-mail address: aneta.smolana@polsl.pl (A. Smolana).

https://doi.org/10.1016/j.conbuildmat.2021.124135
Received 8 March 2021; Received in revised form 15 June 2021; Accepted 28 June 2021
Available online 10 July 2021
0950-0618/© 2021 The Author(s). Published by Elsevier Ltd. This is an open access article under the CC BY-NC-ND license
(http://creativecommons.org/licenses/by-nc-nd/4.0/).
A. Smolana et al. Construction and Building Materials 301 (2021) 124135

Fig. 1. Basic dimensions of the analysed slab of the sluice Sülfeld-Süd [29,30].

and development of volume changes. Additionally, these changes occur accessible for the structural engineer.
with a simultaneous variation of mechanical properties. Reduction of Therefore, simple and realistic methods for predicting early age
the cracking risk may be provided by suitable prevention measures concrete cracking are still desired. Analytical methods developed both in
resulting from practical experiences [11–13]. However, although many scientific works [9,22,23] and thematic guidelines [24–28], fit into this
methods are used to improve the cracking resistance of early age mass demand. The advantage of analytical approaches over numerical
concrete [3,14,15], the prediction and control of hydration effects are methods is their ease using, but a simultaneous rough estimation of the
still indispensable. Simultaneously, the determination of the above­ cracking risk, which comes from many applied simplifications, must be
mentioned thermal effects and the potential cracking risk in hardening conscious.
concrete remains an extraordinarily complex task requiring: This paper presents the study on early age cracking risk in a mass
foundation slab, focusing on a comparison of alternative methods in
- determination of nonlinear and transient temperature fields crack prediction. The behaviour of a real mass foundation slab has been
including the maximum temperature due to self-heating and tem­ analysed using analytical and numerical methods. First, four analytical
perature differential between the interior and the surfaces of the slab, procedures proposed in British [25], American [24], Japanese [26], and
- identifying imposed thermal strains, German/Austrian [27,28] guidelines are investigated. Next, the FE
- identifying distribution and magnitude of self-induced and restraint model has been used for the prediction of the early age cracking risk.
stress, Finally, the results have been compared to the measurements made
- checking the cracking risk, during the construction process.
- calculation of the width of the crack.
2. Description of the analysed slab
Generally, quite many research works are devoted to modelling the
abovementioned issues. At the same time, it should be mentioned that The analysed slab is the first construction stage (stage 1) of the
full modelling of these effects requires interdisciplinary knowledge, massive foundation of the sluice Sülfeld-Süd in Germany [29]. The
covering hygro-thermal, chemical, and mechanical issues. In this considered slab with a thickness of 2 m and dimensions in a plan view
context, many works are devoted to the mathematical formulation of 26.5 m × 41.5 m is presented in Fig. 1. The following mix composition of
heat transfer and mechanical effects in early age concrete [16–20]. For the concrete was applied in the slab: cement CEM III/A 32.5 N (240 kg/
solving such complex mathematical equations, the finite element m3), water (150 kg/m3), gravel aggregate: 0/2 mm (703 kg/m3), 2/8
method (FEM) is frequently applied. A review of existing modelling mm (222 kg/m3), 8/16 mm (462 kg/m3), 16/32 mm (462 kg/m3), fly
approaches describing the early age behaviour of massive concrete ash (110 kg/m3), BV 15 (Melius) 1.5% (3.6 kg/m3). The measured 28-
structures shows a wide range of models’ complexity and modelling day properties of concrete were as follows: compressive strength –
techniques [21]. Undoubtedly, the finite element analysis (FEA) pro­ 34.2 MPa, the tensile strength – 1.83 MPa, the modulus of elasticity –
vides a comprehensive recognition of the early age effects considering 34.4 GPa. The registered adiabatic temperature rise was equal to 30 ◦ C.
all material and technological factors. However, it also brings problems The slab reinforcement consisted of bars ϕ25 mm with a spacing of 15
associated with introducing many input data to the FE model, often cm and concrete cover 6 cm. The reinforcement was located at the
difficult to define, as well as using specialized software not typically bottom, top, and lateral surfaces of the slab. The slab was cast

Table 1
The data of continuous casting and the initial temperature of each layer [29,30].

2
A. Smolana et al. Construction and Building Materials 301 (2021) 124135

The ambient temperature registered during the construction process


of the slab (Fig. 2) was mapped exactly in the numerical model. In this
regard, the calculation procedures of the analytical models only require
providing some average values, which are specified in Table 6.

4. Preliminary assessment of the slab sensitivity to early age


cracking

The preliminary checking of potential proneness to thermal cracking


effects at pre-design stages (hence requiring more detailed studies) can
be made by resorting to simplified approaches such as the massivity
index (sometimes described as ‘surface modulus’, ms ) [35]. For the case
study herein presented, with a slab-on-ground with dimensions 41.5 ×
26.5 × 2 m3, the surface modulus can be easily calculated by the ratio of
the exposed area (disregarding the bottom surface in contact with the
ground) divided by the volume. In such conditions, ms becomes 0.62 /m,
Fig. 2. Registered variation of the ambient temperature [29,30].
which is under the 2.0 /m threshold, and hence the structure is
considered as massive with a predominant impact of thermal strains and
continuously for 12.5 h in August 2005. The initial temperature of each
close-to-adiabatic conditions in the core [36].
layer of the slab is depicted in Table 1 while ambient temperature
Very recent work has proposed improvements to the surface modulus
registered during the construction process of the slab is visible in Fig. 2.
approach by adjusting corrective factors regarding the cement quantity,
The top surface of the slab was only shielded from directly incident solar
the presence of supplementary cementitious material, or even the ex­
radiation, while the lateral surfaces were protected with the formwork
pected adiabatic temperature rise and the environmental/mixture
for the first 7 days of concrete curing.
temperatures [35]. The 0.62 /m massivity has been corrected by
dividing it by 3 factors:
3. Basic data for the comparative study
- factor kf to take into account the presence of supplementary mate­
Basic material properties were taken from the measurements
rials in the binder, which resulted in kf = 0.55, in correspondence to
[29,31]. Other parameters were adopted from [32,33] for the concrete,
the presence of ~ 30% fly ash in the binder,
and from [34] for the subsoil. The necessary values for the sluice slab,
- factor kb to take into account the actual binder content in the mix,
subbase, and subsoil are listed in Table 2. The basic properties of the
which was 350 kg/m3, resulting in kb = 1.167,
reinforcement bars are depicted in Table 3.
- factor kΔT which depends on the temperature of the fresh mixture
In the numerical approach, the analysed slab was reduced to ¼ of the
(Tfresh = 24 ◦ C), the average ambient temperature (Tambient = 18 ◦ C)
structure by using symmetry conditions. A 10 cm layer of lean concrete
and the adiabatic temperature rise expected (Tad,rise = 30 ◦ C, as
(subbase) and subsoil have been modelled, thus the actual support of the
available from experimental data [40]), thus resulting in kΔT = 0.55.
slab has been reproduced. No slip layer between the slab and subbase
has been applied. The dimensions of the analysed structure and the
The resulting corrected massivity factor Mcorr is 0.81 /m, which still
cooperating soil are depicted in Fig. 3a. The view of the FE mesh is
is far smaller than 2 /m, hence still dictating the need for a careful
presented in Fig. 3b.
thermal study of this structure, as performed in the present paper.
Two options of the casting process are applied in the performed FE
study, which requires clarification of the data. In the first case (later
denoted as ‘continuous’), the thickness of the slab was divided into 7 5. Assessment of the cracking risk with analytical approaches
layers (Table 1) to reproduce the process of continuous casting in the FE
model. Such analysis requires the transformation of the boundary con­ Two basic approaches for crack prediction may be distinguished in
ditions depending on the advancement of the casting (Fig. 4). Thus, the analytical models: the stress approach and strain approach [37,38].
values of the heat transfer coefficient for each type of boundary condi­ According to the stress-based method, the crack is formed when the
tion are listed in Table 4. In the second case, the slab was assumed to be tensile stress in the given point of the structure (σ t ) exceeds the actual
poured instantly, i.e., in one phase ‘1ph’. This case better corresponds to tensile strength (ft ). The condition takes the following form:
the analytical models, where the duration of the casting process is not σ t > ft (1)
considered. Necessary data for the simplified one-phase analysis is
presented in Table 5. The strain-based approach involves the determination of the tensile
strain (ε). In that case, reference is made to the strain capacity (εctu ),
which depends on the concrete class and the type of aggregate used in
Table 2 concrete. The condition for the cracking occurrence is formulated as:
Material properties [29,31–34].
ε > εctu (2)
Property Value

Sluice slab Subsoil Subbase Analytical models based on the outlined stress or strain method are
− 3 developed in theoretical works [9,24,25,39] but also are applied in
E-modulus (28-day), GPa 34.4 30 × 10 27
Poisson’s ratio 0.2 0.2 0.2
Density, kg/m3 2400 2070 2400 Table 3
Thermal expansion coefficient, 1/oC 1.2 × 10− 5 1 × 10− 5 1 × 10− 5
Properties of reinforcement bars [29,31,32].
Coefficient of thermal conductivity, W/ 3 (αH *=0); 1.4 1.7
(m oC) 2.1 (αH Property Value
*=1) Young’s modulus, GPa 210
Thermal capacity, J/(m3oC) 2.3 × 106 2.15 × 1.95 × Thermal expansion coefficient, 1/◦ C 1.2 × 10− 5
106 106 Plastic hardening (in the numerical model) No hardening
* Characteristic yield strength, MPa 500
αH – degree of heat development.

3
A. Smolana et al. Construction and Building Materials 301 (2021) 124135

Fig. 3. The dimensions of the analysed slab (a) and the FE model (b).

Fig. 4. Types of boundary conditions in the FE model.

Table 4 Table 6
Heat transfer coefficient [29,30]. Data for the analytical calculation.
Part Heat transfer coefficient, hcr Temperature Value*

Subsoil (SO)
(◦)
SO-1* (top surface): 30 W/ Cm2 , remaining surfaces: adiabatic Initial temperature of concrete, oC 24
Minimum ambient temperature, oC 11
conditions
(◦ ) Mean ambient temperature, oC 18
Subbase (SB) SB-1*: 30 W/ Cm2 Maximum ambient temperature, oC 28
(◦ )
Sluice slab ST-1* to ST-7* (top surface): 30 W/ Cm2 ; SL-1* to SL-7* (lateral *
(ST) (◦ ) (◦ ) based on registered values presented in Table 1 and Fig. 2.
surfaces): 5.2 W/ Cm (first 7 days after casting); 30 W/ Cm2
2

(after removal of formwork)


- stress-based approach, according to ACI Committee [24],
*
- notation to Fig. 4. - stress-based approach (related to the thermal cracking index), ac­
cording to JCI Guidelines for Control of Cracking of Mass Concrete
2016 [26],
Table 5 - stress-based approach, according to German/Austrian guidelines
Data for one-phase FE analysis.
[27,28].
Part Heat transfer coefficient, hcr Initial
temperature
Apart from the applied crack criterion, the above-mentioned guide­
lines also differ in the methods of determining the key values for the
(◦ )
Subsoil top surface (SO): 30 W/ Cm2 ; remaining 16 ◦ C
(SO) surfaces: adiabatic conditions thermal cracking risk of mass foundation slab. Hence, the next sections
(◦ )
Sluice slab top surface (ST): 30 W/ Cm2 ; lateral surfaces 24 ◦ C highlight the differences in the listed guidelines, concerning the
(ST) (◦ 2
)
(SL): 5.2 W/ Cm (first 7 days after casting); 30 following issues:
(◦ )
W/ Cm2 (after removal of formwork)
- the temperature difference between the interior and the surface of
the slab, which is crucial as the self-induced stresses, caused by the
thematic guidelines and standards. Due to the wide availability of internal restraints, dominating in mass foundation slabs,
guidelines, the following subsections discuss the methods proposed - the restraint factor describing the level of the internal restraints in
there. The following analytical methods enclosed in British, American, the mass slab,
Japanese, and German/Austrian guidelines are discussed: - the recommended time (concrete age) for the cracking risk
prediction,
- strain-based approach, according to CIRIA C766 [25], - crack width calculation.

4
A. Smolana et al. Construction and Building Materials 301 (2021) 124135

Further, the procedures and calculation examples presented in the based on the aggregate type (Table 7). It should be mentioned that the
next sections refer to the most common cracks in mass foundation slabs. guidelines extended the recommendations in this regard by a significant
Hence, Sections 5.1-5.4 are focused on the determination of possible factor influencing the discussed coefficient, which is the type of sand
cracking on the top surface of the slab, induced by the internal restraints used in concrete. At the same time, Eurocode 2 [32] recommends that
during the heating phase. unless more reliable information is available, the coefficient of thermal
expansion should be assumed of 10 µε/◦ C. While this is a representative
value for some aggregates, there are materials, as flint gravels, with
5.1. CIRIA C766 [28]
higher values of 12–14 µε/◦ C. This 20–40% difference in αT may be the
difference between compliant and non-compliant crack widths. There­
CIRIA C766 guidelines are complementary to Eurocode 2 [32,33],
fore, if no data are at hand, the guidelines recommend using the value of
which to a limited extent describes early age volume changes in con­
thermal expansion αT equal to 12 µε/◦ C.
crete. Therefore, the discussed guidelines can be perceived as a British
The guideline’s method for estimating the autogenous shrinkage εca
commentary to European standards. The risk of cracking is appraised by
and the drying shrinkage followed the procedures from Eurocode 2 [32].
comparing the crack-inducing strain εr with the tensile strain capacity of
At the same time, in worked examples presented in the guidelines the
the concrete εctu . Then, the condition for the crack occurrence is
shrinkage strains εca and εcd in Eq. (4) are neglected, assuming that in
formulated as follows:
mass foundation slabs the principal reason for the possible cracking is
εr > εctu (3) the thermal gradient and the internal restraints.
Due to the dominating effect of the internal restraints in mass slabs,
The condition is recommended to be verified after 3 days of concrete
the guidelines recommend the value of the factor R equal to 0.42.
curing. Early age crack-inducing strain εr may be calculated using the
Although the external restraints of mass slabs cast on the ground are
expression:
considered as less important, some recommendations are also provided,
εr = K1 R(αT ΔT + εca + εcd ) (4) as listed in Table 8. Nevertheless, no recommendations for a specified
length/thickness ratio or the type of foundation are given.
with:
To assess the cracking risk, the maximum tensile strain capacity, εctu ,
K1 , the coefficient considering creep under sustained loading; the
should be known. The following general formula is suggested to calculate
recommended value is K1 = 0.65,
this value for early age concrete:
R, the restraint factor describing the degree of deformation freedom,
αT , the coefficient of thermal expansion, με/◦ C, εctu = fctm (3days)/Ecm (3days) (6)
ΔT, the temperature difference, for mass slabs is taken as the dif­
The values of εctu obtained from Eq. (6) are valid under conditions of
ference between the temperature of the interior and the surface of the
short-term loading, whereas stresses induced to the early age thermal
element, ◦ C,
loads are sustained. It has two effects on εctu . The first represents an
εca , the autogenous shrinkage strains, με,
increase of the strain capacity by a factor 1/K1 , representing the relax­
εcd , the drying shrinkage strains, με.
ation of stress due to creep. Thus, for the supposed 35% reduction in
The calculation of the temperature difference ΔT remains the most
stress, K1 = 0.65 is introduced. The second is a reduction of the failure
substantial difficulty and requires the determination of thermal fields
stress under sustained loading. The recommended value of the coeffi­
variable in time and the slab volume. The guidelines provide an iteration
cient for sustained loading is K2 = 0.7. Finally, the value of εctu is
method for predicting the temperature rise and temperature differen­
increased due to the net effect by the factor K2 /K1 which is equal to to
tials using the adiabatic temperature data. The proposed approach is
1.08 according to CIRIA C766. Estimated values of εctu are listed in
based on the standard heat diffusion theory and relies on spreadsheet
Table 9. The provided values are calculated for concrete class C30/37,
calculations. Principally, it assumes that the temperature within a cell,
and other classes should be modified according to the formula:
at a particular time increment, is calculated as the mean of the tem­
[ ( )]
peratures in the adjacent cells in the previous increment added to the εctu = εctuC30/37 0.63 + fck,cube /100 (7)
increment of adiabatic temperature within the time increment, accord­
ing to the formula: If a crack is detected based on Eq. (3), the crack width is calculated
( ) from the formula:
Tt,j = 0.5 Tt− Δt,j− 1 + Tt− Δt,j+1 + ΔTad (5) [ ]
k1 ϕ
w = sr,max εcr = εcr 3.4cc + 0.425 (8)
with: ρeff
Tt,j , the temperature within a cell at a particular time increment t, in
the cell j, with:
Tt− Δt,j− 1 , Tt− Δt,j+1 , the temperature within the previous time incre­ sr,max , the maximum crack spacing, m,
cc , the concrete cover, m,
ment t − Δt, in the adjacent cells j − 1, j + 1,
k1 , the coefficient considering the bond properties of the reinforce­
ΔTad , the adiabatic temperature rise within the time increment t.
ment (Eurocode 2 [32] recommends the value of 0.8 for high bond bars
The coefficient of thermal expansion of the concrete αT may be taken
and 0.7 for typical bars, however, [25] recommends the higher value
k1 = 1.14),
Table 7
ϕ, the bar diameter, m,
The thermal expansion coefficient based on CIRIA C766 [25].
ρeff , the reinforcement ratio, calculated as ρeff = As /Ac,eff ,
Aggregate αT , με/∘ C
As , the reinforcement area, m2,
Sand from the same rock Siliceous sand Ac,eff , the effective area of concrete in tension around the reinforce
Basalt 10.0 10.5
Flint gravel 12.0 12.0
Glacial gravel 13.0 13.0 Table 8
Quartzite 14.0 14.0 External restraint R for mass slabs [25].
Granite 10.0 10.5
Restraint conditions R
Limestone 9.0 9.5
Siliceous limestone 10.5 11.0 Massive pour cast onto blinding 0.1–0.2
Sandstone 12.5 12.5 Base of massive pour cast onto existing mass concrete 0.3–0.4

5
­
A. Smolana et al. Construction and Building Materials 301 (2021) 124135

Table 9 Table 11
Estimated values of εctu for strength class C30/37 under sustained loading [25]. Evaluation of cracking risk based on CIRIA C766 [25].
Aggregate type The tensile strain capacity εctu , με Calculated Source/assumption Value
value
Early age (3 Long-term (28
days) days) ΔT, ◦ C based on Table 10 28.7
αT , με/◦ C based on Table 7 12
Basalt 55 103
Flint gravel 60 112 R recommended value 0.42
Quartzite 66 123 K1 recommended value 0.65
Granite, gabbro 66 123 εr Eq. (4), with εca = 0, εcd = 0 94
Limestone, dolomite 74 137
εctu , με 64
Sandstone 83 154 Table 9, t = 3 days, Eq. (7) with fck,cube = 34.
Lightweight aggregate (sintered fly ash) with 95 176 2 MPa
natural sand
Cracking risk Eq. (3); εr > εctu YES
εcr , με Eq. (10) 62

ment to a depth of hc,eff , m2, calculated from the expression: k1 recommended value 1.14

{ } Bar diameter ϕ, data - Section 2, 3 0.25


H m
hc,eff = min ; 2.5(cc + ϕ/2) (9)
2 Bars spacing, m data - Section 2, 3 0.15
Bars cover, c, m data - Section 2, 3 0.06
It should be pointed that Eq. (8) is valid if the direction of the tensile
As , m2 – 32.72 ×
stress and reinforcement coincides as well as bonded reinforcement is
10− 4
fixed at reasonably close centres within the tension zone. This second
Ac,eff , m2 Eq. (9), hc,eff = 2.5(cc + ϕ/2) 1812.5 ×
condition results in the need to verify the maximum spacing of the 10− 4
reinforcement, which is defined as 5(cc + ϕ/2). ρeff ρeff = As /Ac,eff 0.01806
The crack inducing strain εcr is calculated from the expression: sr,max , m Eq. (8), max allowable reinforcement spacing: 0.875
150 mm < 362.5 mm
εcr = εr − 0.5εctu (10)
w, m Eq. (8) 0.054
The evaluation of the cracking risk in the analysed slab using this
analytical approach is given in Tables 10 and 11. First, the temperature
caused by the cement hydration was estimated using the method pro­ of the element,
vided by CIRIA C766 (Table 10). The predicted adiabatic temperature Kf , the coefficient of the degree of the restraint,
rise was 29.7 ◦ C, which is almost identical to the measured value of αT , the coefficient of thermal expansion, µε/◦ C,
30 ◦ C. The cracking risk was checked at the top surface of the slab in the ΔT, the temperature difference between the temperature of the
heating phase, thus, the maximum temperature difference between the interior and the surface of the slab, ◦ C,
center and the top was taken in Table 11. Such a maximum difference Ecm,eff (t), sustained modulus of elasticity of the 7-day concrete, MPa.
occurs in the intersection of the slab’s vertical planes of symmetry. The proportional relation between the stress and the strain occurs at
Following the guidelines procedure, the cracking risk was evaluated any point of the uncracked concrete member. Following the variation in
after 3 days of concrete curing. The cracking is detected in the examined the degree of restraint, the horizontal tensile stress varies throughout the
slab, but the crack width is relatively low (Table 11). Further comments, member continuously restrained at its base. This rule is reflected in the
as well as a reference to the on-site measurement considering the tem­ structural shape restraint factor KR , which distribution is dependent on
perature development, stress level, and cracking, are discussed later in the length to height ratio (L/H). So, the factor KR maybe calculated from
section 7. the following equations:

- for L/H equal to or greater than 2.5:


5.2. ACI Committee 207 [24] [( )/( ) ]h/H
L L
KR = − 2 +1 (13)
According to [24], the cracking will occur when the actual stress due H H
to restrained volume change, σt (t), reaches the actual concrete tensile
strength, ft (t). Therefore, the assessment of the early age cracking risk
comes down to the following condition:
- for L/H less than 2.5:
σ t (t) ≥ ft (t) (11) [( )/( ) ]h/H
L L
The recommended time for controlling the above condition is 7 days KR = − 1 + 10 (14)
H H
after concrete placing. The maximum tensile stress can be calculated
from the equation: where h is the height above the restraint body.
σ (t) = KR Kf αT ΔTEcm,eff (t) (12) The restraint factor Kf represents the degree of restraint in the con­
tact layer between the concrete element and the restraining body. Its
with: value is based on the stiffness ratio as follows:
KR , the coefficient of the degree of restraint distribution at the height
1
Kf = (15)
1 + AAcr EEcr
Table 10
The calculated temperature based on CIRIA C766 [25]. with:
Calculated value Value Ac , the area of the considered concrete element cross-section, m2,
Maximum temperature, ◦ C 47.7 Ar , the area of the element restraining the considered element, m2,
Maximum differential (center - top), ΔT, ◦ C 28.7 Ec , modulus of elasticity of the concrete, MPa,
Maximum differential (center - bottom), ◦ C 13.4 Er , modulus of elasticity of the restraining element, MPa.

6
A. Smolana et al. Construction and Building Materials 301 (2021) 124135

The mass foundation slabs are elements of the dominant influence of 2.0(cc + ϕ/2), m2.
internal restraints resulting in self-induced stresses. According to the At the same time, the guidelines indicate that Eq. (20) is based on
guidelines, the internal restraint may be treated as the continuous edge applied mechanical loads, without consideration of the volume change.
restraint, assuming the location of the effective restraining plane on the Because volume changes can cause the increase of the actual crack width
surface of zero stress in the element cross-section, which is dependent on over that estimated by the above formula, the guidelines leave the lib­
the actual temperature distribution in the concrete (Fig. 5). Therefore, erty for the designer to choose a more conservative expression, if
the factor Kf in massive foundation slabs can be calculated assuming necessary. The average cracks spacing is determined from the expres­
that the restraining element is the part of the slab’s section where sion:
compressive stresses occur. Additionally, it can be simplified that the wk
temperature at the thickness of the slab takes the parabolic distribution sr = (21)
KR KF αT ΔT − ft (t)/Ecm,eff (t)
and the depth of the balance line, ds, is approximately equal to 20% of
the slab thickness [24,25,36]. Based on the ACI procedure, the cracking risk in the considered slab
The guidelines recommend Schmidt’s method for the determination has been evaluated. The maximum difference in temperature between
of temperature difference ΔT. The description of the method and ex­ the center and the top of the slab was taken from the previous section, as
amples demonstrating its practical use are also enclosed in [24]. Schmidt’s method recommended in the ACI guidelines essentially
The modulus of elasticity, Ecm (t) is recommended to be determined complies with the iterative procedure from CIRIA C766. Following the
experimentally, but the analytical formula is also provided [24,39]: ACI guidelines procedure, the cracking risk was evaluated after 7 days of
√̅̅̅̅̅̅̅̅̅ concrete curing. The estimated crack width based on the ACI procedure
Ecm (t) = 0.043ρ1.5 fc (t) (16) is given in Table 12. Similarly, as in the previous section, the crack width
with: is relatively low (0.065 mm) and similar to the value obtained from
ρ, the volume density of the concrete, kg/m3, CIRIA guidelines, despite the different calculation methods. More
fc (t), the compressive strength of the concrete at age t, MPa. comments related to the on-site measurements are given in section 7.
Creep effects are considered by using the effective modulus of elas­
ticity, Ecm,eff (t), instead of the modulus of elasticity, Ecm (t) [40]: 5.3. JCI guidelines for control of cracking of mass concrete 2016
[26]
Ecm (t)
Ecm,eff (t) = (17)
1 + ϕ(t, t’ ) The JCI guidelines recommend the implementation of 3D finite
element analysis (FEA) for checking the possible thermal cracking.
The creep coefficient ϕ(t, t’ ) for typical curing conditions is derived
However, due to practicality and convenience, an analytical method for
from [40]:
simple evaluation is also provided. The method is based on the estima­
(t − t’ )0.6 tion equations of the thermal cracking index for the structure, obtained
ϕ(t, t’ ) = 2.35 (18)
10 + (t − t’ )0.6 from 3D-FEA application to 4849 model structures, in which three
impact factors closely related to thermal cracking are selected: shape
where t ’ is the loading time. and stiffness, materials and mixture proportions, and curing methods. In
The tensile strength used in the cracking condition is based on the the detection of cracking, the thermal cracking index is defined as:
compressive strength. For concrete slab with surfaces subjected to dry­
ing, the tensile strength can be taken as [24,39]: Icr =
ft (teq )
(22)
σ t (teq )
ft (t) = 0.0069[ρfc (t) ]0.5 (19)
where:
Finally, the maximum crack width at the surface may be calculated
ft (teq ), designed splitting tensile strength at temperature adjusted age
from the equation:
teq ,
√̅̅̅̅̅̅̅̅̅̅̅̅̅̅
wk = 0.00145σs 3 a1 Ac,eff (20) σ t (teq ), principal tensile stress at temperature adjusted age teq .

where: Table 12
σ s , the stress in reinforcing steel, MPa, Evaluation of cracking risk based on ACI [24].
a1 , the distance from the surface to the center of the reinforcing bars, Calculated value Source/assumption Value
m,
ΔT, C ◦
based on Table 10 28.7
Ac,eff , the effective area of concrete in tension to a depth of hc,eff =
KR Eq. (13), L/H = L/ds, L = 41.5 m, ds = 0.2H = 0.97
0.4 m
Kf Eq. (15), Ac = 16.6 m2, Ar = 49.8 m2, Ec = Er 0.75
Ecm (t), MPa Based on the measurements; t = 7 days 28.853
Ecm,eff (t), MPa Eq. (17), t = 7 days, ϕ(t, t0 ) = 0.572 18.359
ft (t), MPa t = 7 days, ft (28) = 1.83 MPa 1.29
σt (t), MPa Eq. (12) 4.60
Cracking risk Eq. (11), σt (t) ≥ ft (t) YES
Bar diameter ϕ, data - Section 2, 3 0.25
m
Bars spacing, m data - Section 2, 3 0.15
Bars cover, c, m data - Section 2, 3 0.06
As , m2 – 32.72 ×
10− 4
Ac,eff , m2 with hc,eff = 2.0(cc + ϕ/2) 1450 × 10− 4

ρeff ρeff = As /Ac,eff 0.02256


sr,max , m Eq. (21) 0.404
Fig. 5. Internal restraint diagram and the temperature change in the cross- w, mm Eq. (20) 0.065
section [24].

7
A. Smolana et al. Construction and Building Materials 301 (2021) 124135

The thermal cracking index for the whole concrete volume of the It should be pointed that the JCI guidelines suggest using Eq. (28) for
structure (1 lift) or with the consideration of the construction breaks the prediction of the crack width of a wall-type structure since it was
(more lifts) can be estimated from the equation: derived based on walls subjected to the predominant external restraint
at the bottom. Nevertheless, the calculation for the analysed slab has
Icr = (α1m η α2m ζ α3m ξ )β Icrc0 − Ib (23) been also performed. Contrary to the methods provided by CIRIA C766
and ACI, the maximum temperature difference between the center and
where:
the top of the slab is not determined in this procedure. Similarly, the
α1m , the impact factor of shape and stiffness of considered placing
time of the crack checking is also indefinite. The obtained cracking
lift,
probability is very low (5.95%) and slightly exceeds the limit value of
α2m , the impact factor of materials and mixture proportions of
5%. This results in a small crack width equal to 0.067 mm (Table 13),
considered placing lift,
which is almost the same value as obtained based on the ACI approach. A
α3m , the impact factor of curing method of considered placing lift,
wider comparison with measurements and other approaches is given in
Icrc0 , the basic cracking index equal to 0.77,
section 7.
Ib , the safety factor equal to 0.3,
η, ζ, ξ, β, constants representing the influence of the type of cement
on the thermal cracking index.
5.4. German and Austrian guidelines [27,28]
The above impact factors are determined using the following equa­
tions:
Both the German guideline for control of restraint-induced cracking
( )b1m1
H
( )b1m2
Bi0
( )b1m3
N0 in hydraulic structures [27], as well as the Austrian guideline for
α1m = a1m0 + a1m1 0 + a1m2 + a1m3 + analytical assessment of cracking risk in watertight structures [28]
H Bi N
( )b1m4 ( )b1m5 ( )b1m6 (24) provide analytical approaches for the assessment of imposed de­
+a1m4
L0
+ a1m5
Int0
+ a1m6
Ec /Er formations in RC members. Both guidelines are composed in a way to
L Int Ec0 /Er0 enable practicable but efficient design rather than predicting the phys­
( ) ical member behaviour in a scientifically accurate manner. Therefore,
Ta ( ) ( ) both guidelines focus on restraint forces and macrocracking (bending
Q∞ rAT
and separating cracks, respectively). Self-equilibrated stresses due to
Ta0
α2m = a2m0 + a2m1 e + a2m2 + a2m3 +
Q∞0 rAT0 (25) non-linear distributed deformations within the member, however, are
( )0.45 ( )
fc sAT hereby deliberately neglected. The reason is that these self-equilibrated
+a2m4 + a2m5
fc0 sAT0 stresses are usually mitigating the risk of macrocracking, as outlined in
[22] or [41], and thus, they are not regarded in the calculation. It shall
( )b ( )b3m2 ( )b3m3
Ta 3m1 hcr tf be noted that undesirable effects of self-equilibrated stresses are limited
α3m = a3m0 + a3m1 loge + a3m2 + a3m3 +
Ta0 h0 tf 0 by quality control measures for casting within the guideline.
( )b3m4 ( )b (26) In principle, the solutions of [27] and [28] start with the quantifi­
Int0 3m5
cation of the stress-effective deformation to be imposed with regard to
Tat +ΔTa
+a3m4 e Tat + a3m5
Int the concrete used, type and thickness of the viewed element and climatic
conditions. Following, resulting restraint stresses are calculated with
where:
regard to typical member behaviour and respective restraining condi­
a1m0 to a1m6 , b1m1 to b1m6 - constants of shape and stiffness depend­
tions. Depending on the guideline, these restraint stresses are then used
ing on the type of cement and the number of lift; a2m0 to a2m5 - constants
of materials and mixture proportions, depending on the type of cement
and number of lift; a3m0 to a3m5 , b3m1 to b3m5 - constants of curing Table 13
Evaluation of cracking risk based on JCI [26].
method, depending on the type of cement and number of lifts; H - the lift
(slab) height; L - the slab length; Bi0 - the slab width; N-the lift number; Calculated Source/assumption Value
value
Int - the placing interval; Ec /Er - Young’s modulus ratio between
restrained and restraining bodies; Ta - the placing temperature; Tat - the Icr0 recommended value 0.77
ambient temperature; Q∞ - the ultimate adiabatic temperature rise; rAT , Ib recommended value 0.3
sAT - constants related to the temperature rise; hcr - the heat transfer Q∞ based on tables given in [29], for low heat 45.0
coefficient; tf - the time until formwork removal. cement
α1m based on tables given in [29], for low heat 1.27
The applicable ranges of parameters listed above, as well as their cement
reference values (with "0" index), are given in corresponding tables or α2m based on tables given in [29], for low heat 24.231
detailed formulas which can be found in [26]. cement
When the thermal cracking index fulfills the condition: Icr ≥ 1.85, α3m based on tables given in [29], for low heat 0.91
cement
the probability of the crack is very low and does not exceed 5%. In other
β based on tables given in [29], for low heat 0.28
cases, the probability of cracking may be estimated from the formula: cement
[ { ( )− 4.29 } ] η based on tables given in [29], for low heat 1.6
Icr cement
P(Icr ) = 100 1 − exp − (27)
0.92 ζ based on tables given in [29], for low heat 1.0
cement
The maximum width of the thermal crack is calculated from the ξ based on tables given in [29], for low heat 0.6
cement
equation:
Icr Eq. (23) 1.76
( )
− 0.141 Cracking risk detected if Icr < 1.85 YES
wk = γ a + 0.0938 (Icr − 1.965) (28)
ρ P(Icr ) Eq. (27) 5.95%
As , m2 – 32.72 ×
where: ρ – reinforcement ratio related to the cross section of the slab, γa – 10− 4
safety factor within the range of 1.0–1.7, depending on the performance ρ ρ = As /Ac 0.0033
requirements. wk , mm Eq. (28) 0.067

8
A. Smolana et al. Construction and Building Materials 301 (2021) 124135

to either determine the required reinforcement for limiting the width of [ ( )]


macrocracks [27] or to even exclude macrocracking by application of an fct,eff = fctm 0.5 + 0.25 1 −
1
(31)
optimized concrete with strict quality measures [28]. (0.8 + hPl )2
The application of the analytical approach behind [27] and [28] is
demonstrated in the following for the example of the 2.0 thick slab. With where fctm is the concrete tensile strength at 28 days, MPa.
regard to the boundary conditions in this example, in particular appli­ On basis of these results the risk of macrocracking is excluded as long
cation of normal strength concrete and determination of the crack as it holds:
width, the determination equations in [27] are applied. σt
The imposed deformation due to hydration is determined concerning ≤ 1.0 (32)
fct,eff
the decisive situation for the crack occurrence of the viewed member.
Thick ground slabs are hereby assumed to be significantly affected by In the present case, the risk of cracking cannot be excluded. Direct
the heat storage in the subsoil. The resulting temperature gradients calculation of the crack width for a given reinforcement is not foreseen
cause bending stresses with the inversion of signs in the course of time. in [27] as it was created for practical application the other way around,
As outlined in [22], this entails two decisive points in time for verifi­ namely determination of required reinforcement for limitation of crack
cation: (i) occurrence of maximum bending tensile stresses at the top width. For the scientific purpose of the comparative study of this paper,
surface after the first days when the bottom surface reaches its however, one can change the calculation apparatus of [27]. One option
maximum temperature due to heat of hydration and (ii) superposition of is the direct conversion as shown in [42], which enables a direct
maximum bending tensile stresses at the bottom surface with maximum calculation but still gives safe side results as the conservative assumption
axial tensile stresses at the time of temperature equilibrium with the that concrete stresses at the end of the transfer length are equal to the
surrounding. The latter is usually negligible in the case of thick ground tensile strength is taken over. The more realistic calculation of crack
slabs on conventional soil due to a very small degree of axial restraint. As width, however, requires an appropriate assumption on the remaining
it also applies to the present ground slab, the analytical model is sub­ restraint stresses after cracking which can only be found by an iterative
sequently illustrated using the restraint stresses and crack width at the solution. The governing condition is the compatibility of deformations.
top of the slab (verification time (i)). The stress-effective temperature For thick concrete members, where cracking due to imposed de­
gradient, ΔTeff , for this verification time amounts: formations consists of a geometrically set primary crack plus secondary
( ) cracks in its vicinity, this can be formulated in general by:
ΔTeff = 0.6 k0 kFK kJZ ΔTad,7d + ΔTnom (29)
σ IIt ( )
with: wrest = wP (1 + 0.9n) + lcr − lslip (33)
Ecm
k0 , the basic factor for stress-effective temperature deformation
with:
depending on member thickness and concerning the effect of stiffness
wrest , the restrained deformation which is to be absorbed by the crack
evolution, calculated as: k0 = 0.14 + 0.2hPl ≤ 0.74,
system around one primary crack,
hPl , the slab thickness, m,
kFK , the factor regarding the concrete class, equal to 1.0 for C20/25; wP , the opening of primary crack,
1.05 for C25/30; 1.1 for C30/37, n, the number of pairs of secondary cracks next to the primary crack
kJZ , the factor regarding the casting season, equal to 1.0 for casting (see explanation below),
outside winter and 0.6 for winter casting, σ IIt , the remaining concrete stress in the area between geometrically
ΔTad,7d , the adiabatic temperature rise after seven days, oC, set primary cracks after cracking,
ΔTnom , the additional temperature increment concerning the differ­ lcr , the geometrically set distance between primary cracks,
ence between fresh concrete temperature and mean daily temperature of lslip , the length of the crack system where concrete slips on rein­
the ambient air, the recommended value is equal to 5 ◦ C. forcement.
The resulting restraint tensile stresses at the top surface due to the The deformation to be absorbed by the crack system around one
temperature gradient are usually determined assuming complete rota­ primary crack can be determined from the deformation according to the
tional restraint. It reads: restraint stresses in the uncracked state over the distance between pri­
mary cracks. For foundation slabs under a rotational restraint it can be
σ t = αT
ΔTeff
Ecm (30) assumed with:
2 √̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅
σt /
with: wrest,slab,top = lcr with lcr = fct,eff hPl 3γc (34)
Ecm
Ecm , the elastic modulus at 28 days (the effect of stiffness evolution is
already included in Eq. (29)), ( )
where γc is the density of reinforced concrete γ c = 25 kN/m3 .
αT , the coefficient of thermal deformability, 1/ oC. The number of pairs of secondary cracks (n) is usually calculated as
On basis of the determined stresses according to Eq. (30) the risk of the required number of cracks to absorb the imposed deformation while
macrocracking at the top surface could be assessed. However, the maintaining a crack width criterion (wk ) in the primary crack. This
regulation in [27] does not aim such a conclusion and determines condition is not adequate when using the apparatus in opposite direction
directly the required reinforcement for absorbing the herewith released and thus, the number n must be iteratively increased until the steel
deformation in the event of cracking while keeping the crack width stresses in the last secondary crack fall below the cracking force of the
below the tolerable limit. Following the dimensioning formulae in [27] effective area. The steel stresses, σ s , in the last secondary crack are
the required reinforcement at the top surface for limiting the crack width hereby to be determined with regard to the remaining concrete stresses
to 0.2 mm for hydration-induced cracking is ~ 22 cm2/m. between secondary cracks. It reads:
For the scientific purpose of the comparative study of this paper, ( )
however, one must assess the risk of macrocracking. Following the σ s − 0.3nfct,eff /ρeff As ≤ fct,eff hc,eff b→σs ≤ fct,eff /ρeff (1 + 0.3n) (35)
general procedure of the macrocrack index of [28] but concerning the
with:
nature of the used normal strength concrete, the present tensile strength
ρeff , the reinforcement ratio of the effective concrete area
at the time of cracking at the top surface can be determined with [42] as: ( )
As
ρeff = 2.5d 1b
,
As , the reinforcement area, m2,

9
A. Smolana et al. Construction and Building Materials 301 (2021) 124135

d1 , the surface distance of centroid of reinforcement, m, without consideration of physical processes in the internal structure of
b, the running meter, m. concrete and transitions of the gaseous, liquid, and solid phases [38].
The according crack width in the primary crack can be determined The description of the early age behaviour of a continuous medium
with: needs to apply an appropriate constitutive relationship of thermal,
moisture, and stress field.
ds ( )
wP = 0.18 σ s − 0.39fct,eff /ρeff (36) In this study, the numerical study of the considered slab was per­
Es ρeff
formed using DIANA IE 10.2 software. The FE analysis consists of two
steps, which are at first to simulate the temperature history due to the
where ds is the bar diameter, and Es is the Young’s modulus of steel.
cement hydration and subsequently the development of induced
At the same time, force equilibrium must be achieved between steel
stresses. The governing equations and an overall explanation of this
force in the primary crack where the contribution of concrete is zero and
simulation framework are provided below.
concrete force in the uncracked part between two primary cracks, where
Following a frequently used simplification and assuming the negli­
the contribution of steel is negligible. It reads:
gible influence of mechanical fields on heat and mass transport, the
hPl starting point is the determination of thermal and humidity fields. In this
σ s As = σIIt b (37)
2 regard, a complete model should describe the heat and moisture transfer
( )
The length of the whole crack system lslip depends initially on the in mass concrete as a coupled process with the heat production rate and
the reduction of the moisture content due to the hydration process.
transfer length of the steel force in the primary crack. As soon as a new
Although the fluxes of heat and moisture depend on each other [43], in
secondary crack forms, this length increases for the transfer length of the
most models this effect is neglected. The further, frequently applied
secondary crack. On basis of the steel strain distribution over the crack
simplification is the omission of the moisture fields. Such approaches
system given in [42], the total length of the crack system can be esti­
lead to zero autogenous and drying shrinkage, which seems justified in
mated with:
mass concrete at early ages [38]. As a result, the volume changes in early
( )
ds σs + 0.7nfct,eff /ρeff age concrete are predicted only on the base of thermal fields, determined
lslip = (38)
2τbm from:
1
where τbm is the mean bond stress; simplified with τbm = 1.8fct,eff . Ṫ = div(αTT gradT) + qv (39)

With this, all input parameters are connected by solving Eq. (36), Eq.
(37) and Eq. (38) in Eq. (33) and the solution can be found iteratively. with:
The iteration should start with n = 0 and σs = 500 N/mm2. From this T, the temperature, ◦ C,
point, σs can be decreased until compatibility in Eq. (33) is found. In the Ṫ = ∂∂Tt , the time derivative of temperature,
next step, the second condition acc. to Eq. (35) is to be checked for the αTT , the coefficient of thermal diffusion, m2 /s,
obtained steel stress. If not fulfilled, n is to be increased by one and σs to c, the specific heat, kJ/(kg ◦ C),
be decreased beginning from 500 N/mm2, again until compatibility is
ρ, the density of concrete, kg/m3 ,
found. This procedure is to be repeated until both conditions according
qv , the rate of heat generated by cement hydration per unit volume of
to Eq. (33) and Eq. (35) are satisfied.
concrete, W/m3 .
The result for the considered slab is summarized in Table 14.
The following boundary condition, considering the heat transfer
from the outer surfaces of the element applies to the above equation:
6. Numerical approach ( )
qeq = hcr Tsurf − Tenv (40)
6.1. General description
with:
Numerical methods allow for the full recognition of the thermal- qeq , the heat flux from the boundary, W/m2 ,
( )
moisture-mechanical fields and possible cracking risk during the entire hcr , the convection-radiation heat transfer coefficient, W/ m2 ◦ C ,
time of concrete curing. Over the past years, many approaches for Tsurf , the temperature of the boundary surface of the element, ◦ C,
simulation of the early age concrete behaviour were developed as out­ Tenv , the environmental temperature, ◦ C.
lined in the Introduction. For this purpose, the Finite Element Method The heat generation rate is calculated according to the formula:
(FEM) is typically applied. The particular advantage of the FE models is
(41)
− Ea
qv = f (αH )AT e RT
the possibility to obtain a more accurate reflection of actual conditions.
Most of the numerical models are based on the macroscopic description with:
f(αH ), the normalized heat generation rate, W/m3 ,
( )
Table 14 AT , the rate constant, J/ m3 s ,
Evaluation of the stresses and crack widths based on German guideline [27]. R, the ideal gas constant, equal to 8.314 J/(mol ◦ C),
Calculated Source/assumption Value Ea , the apparent activation energy, J/mol.
value The particular values in Eq. (41) as the rate constant AT = 1.75 × 109
( )
ΔTeff , ◦ C Eq. (29) 12.4 J/ m3 ⋅s and the apparent activation energy Ea = 38500 J/mol were
σt , MPa Eq. (30) 2.13 derived from adiabatic data available in [29].
fct,eff , MPa Eq. (31) 1.72 The increments of volumetric thermal deformations are introduced
/
σt
fct,eff
Eq. (32) 1.24 > 1 (cracking is to be to the mechanical model as imposed strains, calculated as:
expected)
wrest , mm Eq. (34) 0.42 with lcr = 6.8 m ΔεT = αT ΔT (42)
n assumptions within iterative 2
solution where αT is the coefficient of thermal deformability.
σs , MPa assumptions within iterative 136.5 Determination of the stress state and possible damage of early age
solution
mass structure involves the appropriate model of the mechanical
lslip , m Eq. (38) 1.09
behaviour of concrete. Apart from thermal imposed strains variable in
wP , mm Eq. (36) 0.12
time and space, the evolving concrete mechanical properties related to

10
A. Smolana et al. Construction and Building Materials 301 (2021) 124135

its aging should be considered. Moreover, creep effects cannot be fracture energy of concrete. More details concerning the concrete
omitted, both due to the nature of thermal loads and the properties of an behaviour after crack formation can be found in [52].
early age concrete, which shows more significant viscous properties The crack width is calculated by integrating crack-inducing strains
than a mature concrete [44,45]. These circumstances make the visco­ over the crack bandwidth. Although DIANA software offers three
elastic material model a primary used for the analysis of early age mass methods to determine the crack bandwidth, also a direct input of the
concrete. Therefore, in the mechanical part, a viscoelastic material crack bandwidth [52], the default value related to the volume of the
model with temperature-dependent Young’s modulus has been applied. element, as proposed by Rots [53], has been used in the presented study.
Basic creep of concrete is considered based on the Double Power Law The multi-directional fixed crack model applied in FE analysis re­
[46,47]: quires data necessary to determine the cracking occurrence. The
assumed values are shown in Table 15 and Fig. 6. It should be noted that
1 ϕ
J(t, t’ ) = + 1 (t’ )− m (t − t’ )n (43) the development of the tensile strength presented in Fig. 6 is depicted for
E(t’ ) E(t’ )
curing temperature equal to 20 ◦ C. Thus, during the numerical analysis,
with: these values were modified respectively to the calculated temperature
J(t, t’ ), the compliance function at time t for a load applied at instant distribution in the slab for each time step, using the equivalent time teq .
t’, Finally, for reinforcement, an elastic-perfectly plastic model is
E(t ’ ), asymptotic Young modulus of concrete at each loading age t’, applied. Furthermore, the linear elastic isotropic model was applied for
MPa, the subbase (lean concrete) and the soil. The technological and material
ϕ1 , m, n, material parameters. data needed for the FE analysis, not mentioned in this section, are pre­
The influence of hydration temperature on the evolving Young’s viously included in section 3 devoted to basic data for the comparative
modulus is considered in the formula [48,49]: study.
( )β 1 ( )β 2
6.2. Results of the 3D simulation
τ1 τ2
( ) − teq − teq
E(t’ ) = E teq = α1 e + α2 e (44)

with: Contrary to the analytical methods, the finite element analysis en­
α1 , α2 , τ1 , τ2 , β1 , β2 , parameters that can be determined from mea­ ables the more precise estimation of temperature, stress, and crack de­
surements, velopments in the investigated slab. Primarily, the temperature and
teq , the equivalent age considering the influence of elevated tem­ induced σ xx stress for 3 selected points in the slab, located in the inter­
perature of concrete curing [50]. section of vertical planes of symmetry of the structure (Fig. 7) are
In Eq. (43) the coefficients were based on [29,38] and adapted for discussed.
the sluice’s concrete by adjustment of the particular curves corre­ Both hardening temperature and induced σ xx stress development has
sponding to the data provided in [51]. The following values are applied: been analysed for 28-days of concrete curing. The results are presented
ϕ1 = 0.012, n = 0.263, m = 0.016. In Eq. (44) parameters determined in Fig. 8 for two performed FE analyses, considering both the reflection
from measurements are as follows: α1 = 15, τ1 = 2, β1 = 1.5, α2 = 20, τ2 = of continuous slab casting and the simplified approach assuming the
4, β2 = 1.5. casting in one phase (‘1ph’).
Furthermore, this model was enhanced with a failure criterion to Based on Fig. 8, it may be observed that the method of modelling the
control the concrete damage, demonstrated with the formation of casting process affects the hardening temperature in the center and at
smeared cracks. The simulation of crack development in the slab was the bottom of the slab. The obtained maximum temperature in these two
performed with a multi-directional fixed crack model widely described points is higher by about 3 ◦ C in the simplified model (‘1ph’), which is
in [52]. The model is based on the smeared crack approach, which perspicuous because in this case the successive cooling of subsequent
considers cracking as a distributed effect with directionality. Thus, layers of the slab is not taken into account. At the same time, the method
cracked material is simulated as a continuous medium with anisotropic of modelling the casting process has no more than a marginal effect on
characteristics. The fundamental feature of the decomposed crack model the temperature development at the top surface of the slab. Regarding
is the decomposition of the total strain ε into an elastic strain εe and a the stress development in the slab, noticeable differences were obtained
crack strain εcr : only at the bottom surface of the slab. It should be mentioned here that
the main concern of early-age slabs is the assessment of the risk of
ε = εe + εcr (42) possible cracking. The potential cracks can arise at the top surface of the
The abovementioned decomposition of the strain allows for slab primarily in the heating phase. In this context, the almost negligible
combining the decomposed crack model with, for instance, a plastic impact of the accurate mapping of the slab casting was observed. The
behaviour of the concrete in a transparent. Furthermore, the sub- difference in stress values is visible only in the first two days of concrete
decomposition of the crack strain gives the possibility of modelling curing and its maximum value is 0.32 MPa. The higher stress equal to
many cracks that occur simultaneously. The basic feature of this multi- 1.25 MPa is obtained for the “one phase” model comparing to 0.93 MPa
in the model with continuous casting (Fig. 8b). For completeness, the
directional fixed crack concept is that stress σ i and strain εicr exist in the
maps of stress distribution in the two chosen steps of the analysis are
coordinate system, aligned with each crack i.
presented in Fig. 9. Examining these maps, the inversion of the sign of
The initiation of cracks is governed by a tension cut-off criterion and
the thermal stresses can be noticed.
a threshold angle between two consecutive cracks. The successive
The development of stresses in three points, usually considered
initiation of the cracks is applied by the following two criteria which
essential for crack assessment (Fig. 8b), does not suggest the cracking, as
must be satisfied simultaneously:
a sharp drop associated with the material softening is not visible.
- the principal tensile stress exceeds the maximum stress condition,
- the angle between the existing crack and the principal tensile stress Table 15
Input data related to cracks modelling [31,54].
exceeds the value of a threshold angle αtd .
Property Value
After cracking, the concrete stress is described as a function of the Tensile strength (28 days), MPa 1.83
strain normal to the direction of the crack. The softening behaviour is Tension softening Linear based on fracture energy
expressed by tension stiffening modelling approach based on the Fracture energy, N/m 137.35

11
A. Smolana et al. Construction and Building Materials 301 (2021) 124135

7. Discussion of the results from analytical and numerical


approaches

The discussion of the results is focused on key points in the early age
cracking risks, such as hardening temperature, thermal stress, and crack
criterion. In a mass foundation slab with the predominant internal re­
straints, the difference of the self-heating temperature at the slab height
is crucial for the cracking risk. Therefore, the calculated temperature
distribution was discussed first. Aimed at the validation of the discussed
analytical and numerical methods, the results of calculations were also
compared to data obtained from on-site measurements of sensors.
Based on Fig. 13, a quite good convergence between the predicted
temperature development at the top surface of the slab and the corre­
sponding measurement of the sensor is obtained. As expected, the best
Fig. 6. Development of tensile strength [31]. compliance with the measurement results was obtained for the numer­
ical model, where the exact variation in the ambient temperature was
However, the careful analysis of the FEA results indicates the occurrence considered. The temperature development during concrete curing,
of cracks in the cross-section located at 3.45 m in Y direction from the calculated according to the iterative procedure provided in CIRIA C766,
symmetry plane (Fig. 10). The cracking area (indigo color) marked in also gave good agreement with on-site measurements, despite the
Fig. 10 includes the elements of the top and lateral surfaces of the slab. simplified average values of the ambient temperature. As expected, the
This first cracking occurs in the heating phase, after 4.23 days from the best compliance with the measurement results for the center of the slab
beginning of concrete curing, and its width reaches 0.01 mm. The sub­ was obtained for the numerical model, where the process of continuous
sequent results, presented in Figs. 11-12, show the development of casting of the slab was mapped (denoted as ‘continuous’). Furthermore,
cracks in the slab. The results for the next calculation steps show the the development of temperature measured by the sensor located in the
gradual increase of its width to 0.05 mm, reached after 7.35 days of center of the slab is remarkably close to the peak temperature predicted
concrete curing (Fig. 11). It is worth noting, the cracks in the interior of with the CIRIA C766 method. Despite a noticeable shift in temperature
the slab, with a width of 0.05 mm, appeared in the cooling phase as well. plots, the characters of the temperature development are very close to
The maximum crack with a width of 0.36 mm is located near the lateral each other. According to Table 16, the difference between the discussed
surface, in the center area of the slab cross-section (Fig. 12). values is 3.5 ◦ C. The reason for this difference is presumably related to
the fact that the analytical method does not reflect the continuous

Fig. 7. Location of the reference points at the slab thickness.

Fig. 8. The development of hardening temperature (a) and induced stress σxx (b) at the top, in the center, and at the bottom of the slab.

12
A. Smolana et al. Construction and Building Materials 301 (2021) 124135

Fig. 11. Maps of stresses σxx (SXX) and crack-widths in the cross-section A-A in
the distance of 3.45 m from the symmetry plane corresponding to 7.35th day of
analysis (one phase FE model - ‘1ph’).
Fig. 9. Maps of stresses σxx (SXX) obtained for the heating phase - corre­
sponding to 4.23th day of analysis (a) and the cooling phase - corresponding to
25.52th day of analysis (b), view on the planes of symmetry (one phase FE
model - ‘1ph’).

Fig. 10. Maps of stresses σ xx (SXX) and crack widths in the cross-section A-A in Fig. 12. Maps of stresses σxx (SXX) and crack-widths in the cross-section A-A in
the distance of 3.45 m from the symmetry plane corresponding to 4.23th day of the distance of 3.45 m from the symmetry plane corresponding to 25.52th day of
analysis (one phase FE model - ‘1ph’). analysis (one phase FE model - ‘1ph’).

process of casting. At the same time, the discrepancy between the The registration of stresses in the center of the discussed foundation
analytical and one-phase (‘1ph’) numerical model is only 1.4 ◦ C and the slab was performed with the stressmeter, aiming at the measurement of
plot shift is lower. Similar remarks can be formulated for the bottom part concrete stresses without any post-processing [55]. Fig. 14 presents the
of the slab. Additionally, a very similar value of the maximum temper­ comparison of the stress development determined using a numerical
ature is visible, comparing JCI method and other results (Table 16), model with measurements of the corresponding sensor. Based on the
excluding the German/Austrian guideline which provides no physical shown plots, a quite good convergence may be observed by the 8th day
temperature but only stress-effective value. of concrete curing, both for ‘one phase’ and ‘continuous’ FE model. After

13
A. Smolana et al. Construction and Building Materials 301 (2021) 124135

Fig. 13. Comparison of temperature development based on the analytical, numerical results with on-site measurements.

Table 16
A final comparison of analytical and numerical methods with real measurements.
Value Method

CIRIA C766 ACI JCI German/Austrian Numerical model Real values


**
Maximum temperature in the center, C ◦
47.7 47.7 45.0 – 43.2* (46.4) 44.2
Maximum differential (center - top), ◦ C 28.7 28.7 – 12.4*** 24.3* (24.8)** 21.5
Maximum differential (center - bottom), ◦ C 13.4 13.4 – – 13.3* (8.13)** 12.6
Maximum stress at the top surface, MPa – 4.17 – 2.13 1.8 –
Maximum strain at the top surface, με 94 – – – 126 137
Crack width at top surface, mm 0.054 0.065 0.067 0.12 0.05 not greater than 0.25
*
model with continuous casting.
**
model ‘one phase’ (‘1ph’) casting.
***
no physical temperature but stress-effective value.

Finally, the cracking risk estimated based on the discussed methods


was compared. The results visible in Table 16 show cracks with a small
width, which is generally coherent to reality as no incompliant cracks
greater than 0.25 mm are reported on the surfaces of the executed slab.
The great consistency of the analytically (CIRIA, JCI, ACI) and numer­
ically calculated crack widths is astonishing considering the significant
discrepancies in the calculated stresses and strains (Table 16). Only the
German/Austrian gives the higher crack width. Nevertheless, although
in the presented example a good agreement of the analytically deter­
mined crack width with the numerical calculation results has been ob­
tained, the guideline methods can be further verified for slabs with other
dimensions and technological data.

8. Conclusions
Fig. 14. Comparison of stress development based on the numerical results and
on-site measurements. The assessment of early age cracking risk is not only an interesting
topic of scientific research but also an important issue for engineering
practice. Preliminary theoretical identification of the cracking possibil­
that time, the presented stress development diagrams do not coincide. ity allows taking measures to modify the casting technology or the
Furthermore, a significant drop of on-site measured stresses is observed concrete mix before starting the construction process. This undoubtedly
on the 15th day of the measurement, which is usually recognized as reduces the cost of possible cracks’ repairs or additional measures dur­
cracking. The lower registered stresses between the 8th and 15th days ing construction, such as top surface insulation or installing the cooling
can be explained by cracking occurring in an area close to the stress­ pipes.
meter’s location and affecting the stress profile in the center. It seems the In this regard, the simple analytical methods provided by four
results of the numerical study confirm that, especially the cracking area guidelines, and the numerical approach were assessed in terms of their
in the center of the slab visible in Fig. 12. Nevertheless, the good stress reliability. The results of the analytical and numerical computation were
coincidence in the first days of concrete curing can be regarded as a also compared to the measurement results made on the construction site.
reliable prediction of all stresses using FEA. As the evolved hydration temperature is the main reason for cracking in

14
A. Smolana et al. Construction and Building Materials 301 (2021) 124135

slabs with considerable thickness, only thermal loads are considered in - finally, considering the evaluated widths of cracks at the slab top
the study. The following comments result from the conducted compar­ surface, remarkably high compliance of the results from CIRIA C766,
ative study for the mass foundation slab: ACI and JCI with the numerical results was obtained. This agree­
ment, however, should not be seen as a strong validation for two
- considering the thermal loads, CIRIA C766 provides an iterative reasons. Firstly, the determined crack widths are very small so that
method for predicting the temperature development at the slab the contribution of concrete in the cracked section is still appearing -
thickness. The ACI guidelines suggest using Schmidt’s method which this is reflected in the numerical approach but not included in the
concept is similar to that proposed in CIRIA 766. The maximum analytical solutions. And secondly, especially in CIRIA C766 [25] the
temperature obtained from the iterative analytical procedure is crack spacing is derived from experiments with specimens being in
satisfying, the deviation from the measured temperature was 3.5 ◦ C the stabilized crack stage – which is definitely not present in this
and from the numerical model with input data close to the analytical stage.
approach (one phase modelling, ‘1ph’) was 1.4 ◦ C. Simultaneously, - Nevertheless, the guidelines procedures enable the calculation of the
the difference in temperature between the center and the top of the crack width only at the top surface, in the heating phase. In this
slab, which is crucial for cracking risk assessment, was 7.2 ◦ C higher context, the solution of [27] and [28], as well as the FE analysis,
comparing to on-site measurements and 3.9 ◦ C relating to the one- delivers considerably greater opportunities. The initiation of crack
phase numerical model. These differences can be justified because and its later development in FE models can be observed at each point
in the analytical model the variations of ambient temperature are of the slab and throughout the entire concrete hardening period.
considered in a simplified manner. Discussing the numerical model, These capabilities made it possible to detect cracks not only at the
better compliance with the actual measurements provides the model top surface but also in the center of the slab thickness, which
with the accurate mapping of the casting process (denoted as occurred in the cooling phase. Besides, the solution proposed in [30]
‘continuous’). In the JCI procedure, only maximum temperature rise and [31] enable also the calculation of cracking risk and crack widths
is calculated based on a set of coefficients given in the guideline. arising from the bottom up to through cracks at later stages.
Despite such a simple method, the obtained maximum temperature is
very close to on-site measurement. In German/Austrian guidelines Nowadays, the general tendency is to use FE modelling for most civil
the calculated temperature is only used for the determination of the engineering problems. The presented study shows that comprehensive
effective stress and has no physical sense. calculations including temperature and stress development as well as
- considering the analytical evaluation of the cracking risk, the main the full recognition of developing cracks in the concrete curing process
difference is related to the crack criterion. CIRIA C766 [25] proposes can be done using the FE method. Even some guidelines explicitly
cracking risk prediction through an evaluation of tensile strains 3 recommend the use of numerical methods [6,26], the users should be
days after slab casting, assuming the occurrence of a maximum aware of the demand for thermo-mechanics knowledge, necessary for
temperature difference between the center and the top of the slab at the proper creation of the FE element model. Additionally, preparing of
that time. In ACI [24], the cracking risk is identified based on the data needed for numerical analysis, similarly to the duration of the
stress criterion and the tensile stresses are evaluated after 7 days of calculations, is usually time-consuming. In this context, the simple
concrete curing. Therefore, the condition given in CIRIA C766 seems analytical methods provided by the discussed CIRIA C766, ACI, JCI, and
to be more restrictive because the limit value of the acceptable strain German/Austrian guidelines can be still useful for structural engineers.
not causing cracks is taken for 3-day concrete. At the same time, both The results presented in this comparative study confirm that guidelines’
CIRIA and ACI do not consider the faster development of mechanical methods can be considered in the evaluation of cracking risk in the mass
properties accelerated by hydration temperature. This problem does foundation slabs, at least as its preliminary recognition.
not arise in the case of FE analysis, because both the gradual increase
in thermal loads and the progressive development of mechanical CRediT authorship contribution statement
properties are taken into account. Therefore, the obtained results
allow tracing the entire cracking process including the crack for­ Aneta Smolana: Software, Investigation, Validation, Data curation,
mation and propagation. The solution proposed in German/Austrian Writing - original draft, Writing - review & editing. Barbara Klemczak:
differentiates from the other solutions in a way that only the risk of Conceptualization, Methodology, Writing - review & editing, Supervi­
macroscopic cracking due to bending, tension, or any combination of sion. Miguel Azenha: Conceptualization, Methodology, Writing - orig­
it is assessed. inal draft, Supervision, Resources. Dirk Schlicke: Conceptualization,
- it should be mentioned that many design models determine the crack Methodology, Writing - original draft, Supervision, Resources.
widths from the strain difference between concrete and steel over the
crack spacing in a stabilized crack stage. Cracking due to imposed Declaration of Competing Interest
deformations, however, usually causes only single cracks whereby
the distance between them is not relevant for the resulting crack The authors declare that they have no known competing financial
width, see e.g. [56]. In this regard, German and Austrian approaches interests or personal relationships that could have appeared to influence
[27,28,57,58] make a clear distinction between single crack stage the work reported in this paper.
and stabilized crack stage. Another good example are the JCI
guidelines, which also do not rely on the concept of crack spacing but Acknowledgment
determine the maximum crack width based on the thermal cracking
index and the reinforcement ratio. The guidelines justify the choice Funding provided by: Silesian University of Technology (project 03/
of the thermal cracking index as a major parameter for crack width 060/RGJ20/0081); the Portuguese Foundation for Science and Tech­
by the opinion that thermal crack widths depend mainly on thermal nology (FCT) to the Research Project IntegraCrete PTDC/ECM-EST/
stress due to heat of hydration and tensile strength of concrete. 1056/2014 (POCI-01-0145-FEDER-016841), the Research Unit ISISE
Overall, the assessment of thermal cracks the authors’ perspective an (POCI-01-0145-FEDER-007633). The support of COST Action TU1404 in
oversimplification. Finally, in FE models the crack width calculation the networking actions (particularly the Short-Term Scientific Mission of
is using the typical equations based on the crack spacing, such as in the 1st author) necessary for this work is also acknowledged.
CIRIA C766 and ACI, are from based on the cracking strain and the
crack bandwidth.

15
A. Smolana et al. Construction and Building Materials 301 (2021) 124135

References [29] N.V. Tue, D. Schlicke, J. Bödefeld, Beanspruchungen in dicken Bodenplatten


infolge des Abfließens der Hydratationswärme, Bautechnik 84 (10) (2007)
702–710.
[1] ACI Committee, ACI 207.1R-05 - Guide to Mass Concrete, American Concrete
[30] A. Żmij, Numerical study on early age thermal-shrinkage stresses in massive
Institute, Farmington Hills, 2005.
foundation slabs, PhD Thesis, Silesian University of Technology, 2020.
[2] F. Barre, P. Bisch, D. Chauvel, J. Cortade, J. Coste, J. Dubois, S. Erlicher, E. Gallitre,
[31] Materialprüfanstalt (MPA) für das Bauwesen, MPA Braunschweig. Prüfberiicht Nr
P. Labbe, J. Mazars, C. Rospars, A. Sellier, J.M. Torrenti, F. Toutlemonde,
1268/5535, 2005.
Hydration Effects of Concrete at an Early Age and the Scale Effect, in: Control of
[32] CEN, EN 1992-1-1 European Standard Eurocode 2 - Design of concrete structures -
Cracking in Reinforced Concrete Structures: Research Project CEOS.Fr, 2016: pp.
Part 1-1: General rules and rules for buildings, 2008.
27–45. https://doi.org/10.1002/9781119347088.ch2.
[33] CEN, EN 1992-3 European Standard Eurocode 2 - Design of concrete structures -
[3] D.J. Shen, X.Z. Liu, Q.Y. Li, L. Sun, W.T., Early-age behavior and cracking
Part 3: Liquid retaining and containment structures, 2006.
resistance of high-strength concrete reinforced with Dramix 3D steel fiber, Constr.
[34] R. Weil, The Nature and Properties of Soils. 15th edition: Appendix C: Properties of
Build. Mater. 196 2019 307–316. https://doi.org/10.1016/j.
soils, (2016) 1396–1400.
conbuildmat.2018.10.125.
[35] F. Kanavaris, A. Jedrzejewska, I.P. Sfikas, D. Schlicke, S. Kuperman, V. Šmilauer, T.
[4] I. Maruyama, P. Lura, Properties of early-age concrete relevant to cracking in
Honório, E.M.R. Fairbairn, G. Valentim, E.F. Faria, M. Azenha., Enhanced
massive concrete, Cem. Concr. Res. 123 (2019) 105770, https://doi.org/10.1016/
massivity index based on evidence from case studies: Towards a robust pre-design
j.cemconres.2019.05.015.
assessment of early-age thermal cracking risk and practical recommendations,
[5] M. Azenha, C. Sousa, R. Faria, A. Neves, Thermo–hygro–mechanical modelling of
Constr. Build. Mater. https://doi.org/10.1016/j.conbuildmat.2020.121570.
self-induced stresses during the service life of RC structures, Eng. Struct. 33 (12)
[36] K. Flaga, B. Klemczak, Konstrukcyjne i technologiczne aspekty naprężeń termiczno-
(2011) 3442–3453, https://doi.org/10.1016/j.engstruct.2011.07.008.
skurczowych w masywnych i średniomasywnych konstrukcjach betonowych
[6] ACI Committee, ACI 231 - Report on early-age cracking: causes, measurement, and
(Structural and technological aspects of thermal-shrinkage stresses in mass and
mitigation., American Concrete Institute, Farmington Hills, Mich., 2010.
medium-mass concrete structures), Cracow University of Technology, Monograph,
[7] M. Briffaut, F. Benboudjema, J.-M. Torrenti, G. Nahas, Effects of early-age thermal
2016.
behaviour on damage risks in massive concrete structures, Eur. J. Environ. Civ.
[37] D. Schlicke, Untersuchung zu Temperatur- und Steifigkeitsentwicklung im jungen
Eng. 16 (5) (2012) 589–605, https://doi.org/10.1080/19648189.2012.668016.
Beton am Beispiel der Schleuse Sülfeld, Master Thesis, University of Leipzig, 2006.
[8] Z. Bofang, Thermal stresses and temperature control of mass concrete, Elsevier
[38] E.M.R. Fairbairn, M. Azenha, Thermal cracking of massive concrete structures.
(2014), https://doi.org/10.1016/C2012-0-06038-3.
State of the Art Report of the RILEM Technical Committee 254-CMS, Springer
[9] B. Klemczak, Analytical method for predicting early age thermal effects in thick
Berlin Heidelberg, Cham, 2018.
foundation slabs, Materials 12 (22) (2019) 3689, https://doi.org/10.3390/
[39] B. Klemczak, A. Żmij, Reliability of standard methods for evaluating the early-age
ma12223689.
cracking risk of thermal-shrinkage origin in concrete walls, Constr. Build. Mater.
[10] B. Klemczak, A. Żmij. External restraint factors in early-age massive foundation
226 (2019) 651–661, https://doi.org/10.1016/j.conbuildmat.2019.07.167.
slabs, ACI Struct. J. 117 2020 45–54. https://doi.org/10.14359/51720192.
[40] ACI Committee, ACI 209R-92 - Prediction of Creep, Shrinkage, and Temperature
[11] Y. Sargam, M. Faytarouni, K. Riding, K. Wang, C. Jahren, J. Shen, Predicting
Effects in Concrete Structures, American Concrete Institute, Farmington Hills,
thermal performance of a mass concrete foundation – A field monitoring case
1992.
study, Case Stud. Constr. Mater. 11 (2019) e00289, https://doi.org/10.1016/j.
[41] D. Schlicke, L. Matiašková, Advanced computational methods versus analytical and
cscm.2019.e00289.
empirical solutions for determining restraint stresses in bottom-restrained walls,
[12] J. Gajda, M. Vangeem, Controlling temperatures in mass concrete, Concr. Int. 24
J. Adv. Concr. Technol. 335 (2019), https://doi.org/10.3151/jact.17.6.335.
(2002) 59–62.
[42] D. Schlicke, K. Hofer, V.T. Nguyen, Restraint-induced Crack Formation and Crack
[13] A. Golda, Concrete resistance to the environment in massive structures, on an
Widths in Thick Walls. in Proceedings of the International Conference on
example of construction of blocks no. 5 and 6 of Opole Power Plant, PhD Thesis (in
Sustainable Materials, Systems and Structures (SMSS2019): Challenges in Design
Polish), Silesian University of Technology, 2017.
and Management of Structures (Band 128, S. 138 - 144). RILEM Publications S.a.r.
[14] D.J. Shen, C. Liu, C. Li, X. Zhao, G. Jiang, Influence of Barchip fiber length on early-
l.
age behavior and cracking resistance of concrete internally cured with super
[43] B. Klemczak, Prediction of coupled heat and moisture transfer in early-age massive
absorbent polymers, Constr. Build. Mater. 214 (2019) 219–231, https://doi.org/
concrete structures, Numerical Heat Transfer, Part A: Appl. 60 (3) (2011) 212–233,
10.1016/j.conbuildmat.2019.03.209.
https://doi.org/10.1080/10407782.2011.594416.
[15] D.J. Shen, C. Liu, Z. Feng, S. Zhu, C. Liang, Influence of ground granulated blast
[44] A.M. Neville, Properties of concrete, 5th ed, Pearson, Harlow, England ; New York,
furnace slag on the early-age anti-cracking property of internally cured concrete,
2011.
Constr. Build. Mater. 223 (2019) 233–243, https://doi.org/10.1016/j.
[45] Z. Zhao, K. Wang, D.A. Lange, H. Zhou, W. Wang, D. Zhu, Creep and thermal
conbuildmat.2019.06.149.
cracking of ultra-high volume fly ash mass concrete at early age, Cem. Concr.
[16] R. Faria, M. Azenha, J.A. Figueiras, Modelling of concrete at early ages: application
Compos. 99 (2019) 191–202, https://doi.org/10.1016/j.
to an externally restrained slab, Cem. Concr. Compos. 28 (6) (2006) 572–585,
cemconcomp.2019.02.018.
https://doi.org/10.1016/j.cemconcomp.2006.02.012.
[46] Z.P. Bažant, E. Osman, Double power law for basic creep of concrete, Mat. Constr. 9
[17] G. De Schutter, Finite element simulation of thermal cracking in massive hardening
(1) (1976) 3–11, https://doi.org/10.1007/BF02478522.
concrete elements using degree of hydration based material laws, Comput. Struct.
[47] Z.P. Bažant, Material Models for Structural Creep Analysis, in: Mathematical
80 (27-30) (2002) 2035–2042, https://doi.org/10.1016/S0045-7949(02)00270-5.
Modeling of Creep and Shrinkage of Concrete, John Wiley & Sons, New York, 1988:
[18] M. Azenha, R. Faria, D. Ferreira, Identification of early-age concrete temperatures
pp. 99–215.
and strains: monitoring and numerical simulation, Cem. Concr. Compos. 31 (6)
[48] J. Carette, Towards early age characterisation of eco-Concrete containing blast-
(2009) 369–378, https://doi.org/10.1016/j.cemconcomp.2009.03.004.
furnace slag and limestone filler, PhD Thesis, Université Libre de Bruxelles, 2015.
[19] S. Wu, D. Huang, F.-B. Lin, H. Zhao, P. Wang, Estimation of cracking risk of
[49] J. Granja, Continuous characterization of stiffness of cement-based materials:
concrete at early age based on thermal stress analysis, J Therm Anal Calorim. 105
experimental analysis and micro-mechanics modelling, phdthesis, Universidade do
(1) (2011) 171–186, https://doi.org/10.1007/s10973-011-1512-y.
Minho, 2016.
[20] B. Klemczak, A. Knoppik-Wróbel, Reinforced concrete tank walls and bridge
[50] M. Azenha, Numerical simulation of the structural behaviour of concrete since its
abutments: Early-age behaviour, analytic approaches and numerical models, Eng.
early ages, PhD Thesis, University of Porto, 2009.
Struct. 84 (2015) 233–251, https://doi.org/10.1016/j.engstruct.2014.11.031.
[51] B. Klemczak, M. Batog, Z. Giergiczny, A. Żmij, Complex effect of concrete
[21] Md. Safiuddin, A. Kaish, C.-O. Woon, S. Raman, Early-age cracking in concrete:
composition on the thermo-mechanical behaviour of mass concrete, Materials 11
causes, consequences, remedial measures, and recommendations, Appl. Sci. 8 (10)
(2018) 1–18, https://doi.org/10.3390/ma11112207.
(2018) 1730, https://doi.org/10.3390/app8101730.
[52] J. Gomes, R. Carvalho, C. Sousa, J. Granja, R. Faria, D. Schlicke, M. Azenha, 3D
[22] D. Schlicke, N.V. Tue, Minimum reinforcement for crack width control in
numerical simulation of the cracking behaviour of a RC one-way slab under the
restrained concrete members considering the deformation compatibility, Struct.
combined effect of thermal, shrinkage and external loads, Eng. Struct. 212 (2020)
Concr. 16 (2) (2015) 221–232, https://doi.org/10.1002/suco.v16.210.1002/
110493, https://doi.org/10.1016/j.engstruct.2020.110493.
suco.201400058.
[53] J.G. Rots, Computational Modeling of Concrete Fracture, PhD thesis, Delft
[23] M. Zych, Modification of the simplified method of crack control included in EN
University of Technology, 1988.
1992–3, Struct. Concr. 17 (2016) 553–563, https://doi.org/10.1002/
[54] CEB Comitte’ Euro - International du Beton, CEB - FIB Model Code 2010. Bulletin
suco.201500077.
D’Information. Final draft., 2011.
[24] ACI Committee, ACI 207.2R-07 - Report on thermal and volume change effects on
[55] Gekon, Instruction Manual - Model 4370 Concrete Stressmeter, 2015. https
cracking of mass concrete, American Concrete Institute, Farmington Hills, 2007.
://www.geokon.com/content/manuals/4370_Toyuku_Elmes_Type_Concrete_Stress
[25] P.B. Bamforth, Construction Industry Research and Information Association,
meter.pdf (accessed October 5, 2019).
Control of cracking caused by restrained deformation in concrete, CIRIA C766
[56] D. Schlicke, E.M. Dorfmann, E. Fehling, N.V. Tue, Calculation of maximum crack
(2018).
width for practical design of reinforced concrete, Civ. Eng. Des. 3 (3) (2021) 45–61,
[26] JCI, Guidelines for Control of Cracking of Mass Concrete 2016, Japan Concrete
https://doi.org/10.1002/cend.v3.310.1002/cend.202100004.
Institute, Tokyo, 2017.
[57] DIN EN 1992-1-1/NA:2013-04: National Annex to EC2 Design of concrete
[27] DIN EN 1992-1-1/NA:2013-04: National Annex to EC2 Design of concrete
structures - Part 1 1: General rules and rules for buildings. Beuth Verlag GmbH,
structures - Part 1 1: General rules and rules for buildings. German Institute for
Berlin 2013.
Standardisation, 2013.
[58] OENORM B 1992-1-1:2018-01-01: National Annex to EC2 Design of concrete
[28] OENORM B 1992-1-1:2018-01-01: National Annex to EC2 Design of concrete
structures - Part 1 1: General rules and rules for buildings. Beuth Verlag GmbH,
structures - Part 1 1: General rules and rules for buildings.
Berlin 2013.

16

View publication stats

You might also like