Download as pdf or txt
Download as pdf or txt
You are on page 1of 18

Engineering Structures 102 (2015) 369386

Contents lists available at ScienceDirect

Engineering Structures
journal homepage: www.elsevier.com/locate/engstruct

Degree of restraint concept in analysis of early-age stresses in concrete


walls
Agnieszka Knoppik-Wrbel , Barbara Klemczak
Department of Structural Engineering, Silesian University of Technology, Akademicka 5, 44100 Gliwice, Poland

a r t i c l e

i n f o

Article history:
Received 6 March 2015
Revised 24 July 2015
Accepted 18 August 2015
Available online 2 September 2015
Keywords:
Degree of restraint
Early-age concrete
Numerical modelling
Restraint stresses
Soilstructure interaction
Wall

a b s t r a c t
The degree of restraint is a useful concept for characterisation of early-age thermalshrinkage stresses
occurring in externally-restrained concrete elements such as walls. It can be used not only in manual calculations, but also in numerical analysis to determine the values and distribution of stresses in walls. The
issues that must be addressed while defining the degree of restraint of the wall include the stiffness of the
restraining body (e.g. foundation), translational and rotational restraints, influence of the construction
sequence and support conditions. These issues are discussed in the paper. For the purpose of the study
a numerical model is proposed which takes into account sequential casting and interaction between
early-age structure and founding soil. The results of the study point out the factors that need be taken
into account when modelling structural behaviour of early-age walls for proper determination of the
expected stresses.
2015 Elsevier Ltd. All rights reserved.

1. Introduction
Concrete elements are subjected to early-age volume changes
due to temperature and moisture variations which characterise
the process of concrete hardening. These volume changes induce
stresses in concrete elements. In massive concrete elements, such
as foundation slabs or blocks, the stresses are induced mainly by
significant temperature differences developing between the interior and the surface of the element (self-induced stresses). In
externally-restrained elements, such as walls, thermalshrinkage
stresses result from a coupled action of self-induced (Fig. 1b) and
restraint stresses (Fig. 1c). The restraint in these elements is
exerted by the bond between the new concrete of the element
and the older concrete of the foundation or a previous lift; in a concrete wall tensile stresses result from the restraint of a potential
contraction caused by the length changes associated with decreasing temperature of the wall. In typical walls restraint stresses play
a predominant role because volumetric strains caused by the temperature and humidity gradients are relatively small in comparison
to the linear strains caused by the contraction of the element along
the line of the restraint joint [1,2]. Nevertheless, it must be remembered that with the increasing massivity of the wall the share of
the self-induced stresses increases. Surface tensile stresses occurring in thick walls (thermal gradients) and formed by early
Corresponding author.
E-mail address: agnieszka.knoppik-wrobel@polsl.pl.com (A. Knoppik-Wrbel).
http://dx.doi.org/10.1016/j.engstruct.2015.08.025
0141-0296/ 2015 Elsevier Ltd. All rights reserved.

formwork removal (both thermal and moisture gradients) may


lead to surface cracking which can further develop into through
cracking.
The magnitude of the restraint stresses depends on a degree of
restraint induced against the early-age part of the structure. The
degree of restraint can be expressed with the restraint factor, cR
a measure which in any point of the element is defined as a ratio
between the actual stress generated in the element, r, to the hypothetical stress at total restraint, rfix , [47]:

cR

r
;
rfix

and may take values between 0 at no restraint to 1 at total restraint.


It varies throughout the element with the maximum value at the
joint between the wall and the restraining body, decreasing towards
free edges of the wall.
Fig. 2 presents distribution of stresses in the mid-span crosssection of a base-restrained wall. Stresses generated by the external restraint (rres ) have a major influence on the values and character of the total stresses (rtot ). If there were no temperature and
humidity gradients within the element, the stress distribution
would be proportional to the degree of restraint as shown in
Fig. 2a (there would be no self-induced stresses, rs-ind ). Due to
the temperature and humidity gradients at the height of the wall,
which generate the self-induced stresses, the maximum stress
appears above the joint (Fig. 2b) with the maximum value of the
stress above the joint. The temperature and humidity gradients
at the thickness of the wall also cause self-induced stresses, which

370

A. Knoppik-Wrbel, B. Klemczak / Engineering Structures 102 (2015) 369386

Fig. 1. Development of early-age thermalshrinkage stresses in time in an externally-restrained element [3].

Fig. 2. Distribution of early-age stresses in a cross-section of a wall [3].

are the reason why the values of the total thermalshrinkage stresses differ in magnitude between the interior and the surface of the
wall (Figs. 1d and 2b). If the value of tensile stress in any location of
the element exceeds the tensile strength of concrete in that

location, a crack is formed (Fig. 1e). When the stress state in cooling phase looks like in Fig. 2b, which happens if the wall is kept in
formwork during the whole process of concrete hardening, development of cracks initiates from the interior of the wall (internal

A. Knoppik-Wrbel, B. Klemczak / Engineering Structures 102 (2015) 369386

stresses reach higher values than surface stresses). Nevertheless,


more often formwork is removed from the wall in cooling phase,
and then first cracks appear on the surface of the element. The
internal stresses can also become of considerable magnitude and
consequently through cracks may develop.
There are multiple examples of externally-restrained concrete
elements in which early-age thermalshrinkage cracking was
observed or which are susceptible to such cracking: structural
elements of bridges, especially bridge abutments [8,9], tank walls
[911], walls of nuclear containments and radiation protection
walls [1214], retaining walls and tunnel walls cast against hardened
foundation or cast in stages [7,15]. These elements are characterised with different massivity and restraint conditions. Their
thickness can be as little as 40 cm up to 2 m and beyond while
the length-to-height ratio may range from 1 to even 15 or more.
Yet, the observed cracking pattern is similar [3,16]. In the baserestrained walls the occurring cracks are vertical in the central part
of the wall and splay towards the ends of the element. A horizontal
crack may occur at the construction joint at the ends of the wall.
The greatest height of the crack is observed in the middle of the
wall and it declines towards the side edges of the wall. The study
presented in this paper aimed at characterisation of the stresses
generated in these externally-restrained elements taking into
account their restraint conditions, i.e. geometry and dimensions
of the wall and restraining foundation, construction sequence
and support conditions. The concept of the degree of restraint
expressed with the restraint factor was used as a measure.
2. Theoretical background
The early-age stresses in externally-restrained elements result
from a coupled action of internal and external restraints. The external restraint acts against axial deformation and flexural deformation. A free deformation of the concrete element can be
separated into deformation in an axial (X) direction (expansion
or contraction) and flexural deformation in a vertical (Z) direction.
The total stress consists then of three components: due to the
internal restraint, due to the external restraint against axial deformation and due to the external restraint against flexural deformation. The approach based on such assumptions, called
compensation plane method (CPM), was first proposed in Japan
[17] and can be found in Standard Specifications for Concrete
Structures [4]. Other methods for determination of stress state in
externally-restrained concrete elements can be referred which
base on the same assumptions as the CPM, developed by American
Concrete Institute published in ACI Report 207 [5], in Eurocode 2
Part 3 [18] and enhanced in CIRIA C660 [19], by Nilsson introduced
in his Licentiate thesis [6] and developed in further research [11].

371

temperature and moisture concentration gradients. The increment


of stress due to the internal restraint can be determined from the
difference between the strain value at the point of the compensation line, ecomp , and the thermalshrinkage strain distribution
curve, e0 , (Fig. 3) by the equation:

rint Ec e0  ecomp ;

where Ec is a modulus of elasticity of the wall, GPa. This approach


allows to obtain the increment of free axial strain, e, and the incre (Fig. 3).
ment of curvature, u
The internal forces are generated in the element trying to return
the plane after deformation to the original restrained position
axial force N R and bending moment M R . The restraint stresses are
then caused by the coupled action of the axial force and bending
moment, as shown in Fig. 4, according to the equation:

rext

N R MR

z  zcen ;
Ac
Ic

where:
Ac , cross-section area of the wall, m2 ;
Ic moment of inertia of the cross-section; m4 ;
z  zcen distance from the centre of gravity of the crosssection, m.
The internal forces can be defined as follows:

NR RN Ec Ac e;

M R RM E c I c u

4a
4b

where the external restraint factors, RN and RM , are introduced to


represent the translational and rotational restraint of the element
by the restraining body. The Eq. (3) gets a form:

 z  zcen :
rext RN Ec e RM Ec u

2.1. Compensation plane method


The stress exerted due to the internal restraint is caused by
a differential strain in the cross section resulting from the

Fig. 4. Determination of stresses in a concrete element caused by the external


restraint according to compensation plane method after JSCE Standard [4].

Fig. 3. Determination of stresses in a concrete element caused by the internal restraint according to compensation plane method after JSCE Standard [4].

372

A. Knoppik-Wrbel, B. Klemczak / Engineering Structures 102 (2015) 369386

The values of the restraint factors vary within the element


according to the degree of restraint. In the base-restrained element
the greatest degree of restraint occurs at the centreline with the
maximum value at the joint between the early-age wall and the
restraining body, decreasing towards the top edge of the wall.
2.2. Restraint factor
There exist numerous proposals for determination of the degree
of restraint [46,1820]. All of them state that the degree of
restraint depends on the linear restraint exerted by the restraining
body expressed with the ratio of the length and height of the wall,
L=Hc , and the ratio of the stiffness of the wall and the restraining
body:

cR cR


L Ac E c
;
;
H AF E F

where Ac and Ec are cross-section area and modulus of elasticity of


the wall and AF and EF are analogically cross-section area and modulus of elasticity of the restraining body.
The most complete formulation of the restraint factor was proposed by Nilsson [6]. This approach, based on CPM, introduces a
restraint factor, cR , to determine the restraint stress, r, based on
the stress at total restraint, rfix :

r cR c0R ; dres ; dslip rfix ;

where:

restraining body, BF;eff , instead of the real width, BF , (see Eq.


(10)). Effective width defines a part of the restraining body crosssection which influences deformation of a structure. Its value is
determined in numerical analysis in such a way that the curvature
measured at the bottom of the wall obtained with the analytic
approach with the use of restraint factors complies with the
numerically calculated curvature. The resilience factor is analogical
to the structural shape restraint factor, K R , proposed by ACI Report
207.2 [5]. The values of d0res are given in diagrams in Fig. 5a or can
be approximated with a polynomial function [11]:

d0res

 i
n
X
z
ai
;
Hc
i1

where:
ai coefficients of a polynomial function describing resilience
factor distribution;
z=Hc relative location of the analysed point above the joint; z
is a location at the height of the wall above the joint, 0 6 z 6 Hc ,
and Hc is the height of the wall.
The slip factor depends on the free length, L, the width, Bc , and
the height, Hc , of the casting section. It can be determined experimentally or numerically. The values of dslip proposed by Nilsson [6]
are given in diagrams in Fig. 5b.
The restraint factor distribution at height z at the mid-span of a
wall was defined with the following expression:

c0R plane-section restraint factor, which depends on the geomrz


etry of the structure as well as the rotational, cry
R ; cR , and translational, ctR , boundary restraints;
dres resilience factor considering the non-linear effects;
dslip slip factor which depicts the reduction of the restraint
stresses as a result of slip failure when bond forces between
the early-age element and restraining body are broken and horizontal crack appears at free ends of the element.
The value of dres changes at the height of the wall and depends
on the boundary restraint. It is a product of the basic resilience factor and translational and rotational correction factors,


dres dres d0res ; d0trans ; d0rot . To simplify, the resilience factor is taken
as equivalent to the basic resilience factor, dres d0res , and the
correction to account for the translational and rotational boundary
influence is included by introduction of the effective width of the



cR z dslip  dres z  ctR z crzR z cryR z ;


where c

ctR z

t
R z;

Pn

rz
R z

and c

ai
i1 i1
EF HF BF;eff
Ec Hc Bc

ry
R z

are given as follows:



P
P
ai
ai
zcen  z zcen ni1 i1
 Hc ni1 i2
 2 
crzR z 2 
2
2  ;
H B
HF
Hc
zcen  H2c EEFc HF c F;eff
zcen H2F
12
Bc
12

11

cryR z


2
ycen  x0:5 BF;eff  Bc


 2
:
BF;eff Bc 2
EF HF BF;eff BF;eff
Bc 2
2

y

x

y
cen
cen
12
2
Ec H c Bc
12

10

12

Nilsson emphasised the influence of the subsoil on the values


of the restraint factor and analysed them thoroughly [6]. He

Fig. 5. Factors accounting for high-walls effect in determination of the restraint factor according to Nilsson [6].

A. Knoppik-Wrbel, B. Klemczak / Engineering Structures 102 (2015) 369386

concluded that a low-friction and non-cohesive foundation material would pose almost zero translational restrain (restraint against
contraction) and the restraint would increase with the increased
frictional and cohesive properties of the restraining body material.
Analogically, the stiffness of the foundation material would influence the possibility of bending of the concrete element and as such
would influence the rotational restraint to this element. The
restraint would increase with the increasing stiffness of the
restraining body material and the length of the element.
3. Strategy of analysis

q t; T
;
T_ divaTT gradT aTW gradc v
cb q
c_ divaWW gradc aWT gradT  K H qv t; T:

373

13a
13b

Analogical equations were proposed for soil; only partial coupling was assumed (it was suggested by Hillel [30] that the influence of moisture gradients on temperature development do not
need to be taken into account):

T_ divaTT gradT;
c_ divaWW gradc aWT grad T;

14a
14b

where:
The concept of the restraint factor as a measure of the degree of
restraint advocated by Nilsson was adopted in numerical analyses
of walls by Larson [21], Al-Gburi et al. [17,22] and Hsthagen et al.
[15]. In this hybrid approach proposed at the Lule Technical
University (LTU) the self-induced part of the stress, rfix , is determined numerically for the cross-section of the wall, and the total
stress, r, is calculated by introduction of the restraint factor, cR ,
analogically as in the analytic approach proposed by Nilsson. The
restraint factor is, however, calculated numerically with the
general-purpose software according to either of two methods: linear restraint method (LRM) [17,21] or equivalent restraint method
(EQM) [15,17,22].
It is common in the numerical analysis of early-age walls to
assume that the foundation is settled firmly on the subgrade total
rotational restraint of the foundation is then assumed. In the before
mentioned LTU approach the influence of the founding soil is taken
into account, but only in the analysis of cross-section (no influence
on the restraint) and limited to thermal analysis (no moisture
transport). The approach in which the founding soil block is taken
into consideration in spatial analysis is used rarely and most often
it is also limited to thermal analysis [23,24,20]. A few examples of
full cooperation can be referred [9,25]. For the purpose of this
study the model for thermalmoisturemechanical analysis of
early-age walls was proposed which takes into account the soil
structure interaction. An approach similar to the LTU method
was used, in which the numerically-determined values of stresses
were expressed with the use of the restraint factor. Nevertheless,
in the presented approach the whole spatial mechanical analysis
was performed at once. For each wall the stress state was determined for the real restraint conditions (stress r) and under the
assumption of total restraint (stress rfix ). To clarify the calculations
it was assumed that the temperature and moisture content were
uniform in the volume of the wall, so the obtained stresses were
pure restraint stresses. The degree of restraint was in each case calculated at the mid-span of the wall according to Eq. (1). Such calculated values of the restraint factor were compared to the values
of the restraint factor calculated with the Nilssons approach
according to Eq. (9).

T temperature, K;
c moisture concentration by mass (relative humidity), kg=kg;
aTT coefficient of thermal diffusion, m2 =s;
aWW coefficient of moisture diffusion, m2 =s;
aTW coefficient representing the influence of moisture concentration on heat transfer, m2 K=s;
aWT coefficient representing the influence of thermal gradient
on moisture transport, m2 =s K;
cb specific heat, kJ=kg K;
q density of concrete, kg=m3 ;
K H watercement proportionality coefficient describing the
amount of water bounded by cement during hydration process,
m3 =J;
qv t; T rate of heat generated per unit volume of concrete,
W=m3 .
As initial conditions initial temperature and humidity of concrete mix and soil were taken. Boundary conditions were assumed
to be 3rd type boundary conditions of convective type. The coefficients of the heat and moisture exchange were calculated taking
into account the physical properties of materials and covering
material (if applied).
3.2.2. Thermophysical parameters
The coefficients of thermal diffusion, aTT , both for soil and concrete were calculated with the values of thermal conductivity, k,
and heat capacity, cv , constant over time. The values of k and cv
were calculated based on the composition of soil and concrete
mix, respectively. Thermal conductivity of concrete [3133] and
soil [34] was calculated as, respectively:

pi ki ;
X p2
k
ki ki :

15a
15b

Heat capacity of concrete [3133] and soil [35,30] was calculated as,
respectively:

cv

X
X

pi qi cb;i ;

16a

ki qi cb;i :

16b

3.1. Numerical model

cv

The model used in this study was developed based on the proposals of Klemczak [26] for thermalmoisture analysis and Majewski [27,28] and Klemczak [25,29] for stress and damage analysis. It
is an enhanced form of the model introduced by Klemczak and
Knoppik-Wrbel [9]. Formulation of the model is schematically
presented in Fig. 6.

In the equations pi is a mass fraction and ki is a volume fraction of the


ith component: cement, water and aggregate for concrete, and soil
particles and water for soil. The properties of concrete components
_
were taken after Kiernozycki
[32] and for soil after Farouki [35].
The coefficient of moisture diffusion, aWW , for concrete was calculated after Hancox [36] based on the moisture content by
volume:

3.2. Thermalmoisture analysis


3.2.1. Thermalmoisture fields
The temperature and moisture fields in early-age concrete were
defined with the coupled equations proposed by Klemczak [26]:

h
i
DWW W 4:6389  0:7 6W 2  1:0556  0:7 6W 0:3055
 1010 ;
17

374

A. Knoppik-Wrbel, B. Klemczak / Engineering Structures 102 (2015) 369386

Fig. 6. Model for analysis of early-age walls with the use of the numerical approach [3].

by applying the following relationship between the moisture content by volume and by mass:

aWW DWW

qw
:
q

18

The coefficients of the mutual influence of temperature on


moisture fields and vice versa for concrete were taken as constant
and equal to aTW 0:9375  104 m2 K=s after Gawin [37] and

aWT 2:0  1011 m2 =K s after Wyrwa and Szczesny [38]. The


coefficient of the watercement proportionality was taken as equal
to K H 0:3  108 m3 =J (for Portland cement).
Coefficients in moisture analysis for soil, aWW and aWT , were calculated based on the isothermal and thermal liquid diffusivities
determined after Clapp and Hornberger [39] and by applying the
relationship given in Eq. (18). Because of relatively small changes
in moisture content of soil the values of the diffusivities were

assumed as constant and determined under the assumption of full


saturation (S 1:0):

bK h;sat wa
;
n
K h;sat wa cT :

DWW

19a

DWT

19b

where:
K h;sat saturated hydraulic conductivity, m=s;
wa air-entry tension, m;
b fitting parameter;
cT relative change of surface tension with respect to temperature; constant value of cT 2:09  103 = C given by Philip and
de Vries [40] can be assumed.
The hydraulic properties of soil (K h;sat ; wa and b) were taken after
[39].

A. Knoppik-Wrbel, B. Klemczak / Engineering Structures 102 (2015) 369386

3.2.3. Hydration function


The rate of heat generated per unit volume of concrete was calculated based on the unit heat rate, qt (W/g), and the amount of
cement, C c (kg/m3):

qv t; T C c qt; T C c

@Qt; T
:
@t

20

The unit rate of hydration heat development, qt; T, was calculated according to Schindler and Folliard [41]:



 b  
EK
1 1
s
b
T ref T
R
qt; T Q tot
aH te e
;
te
te

21

with the parameters calculated for pure Portland cement (no SCMs
addition):
Q tot total heat of hydration, J=g; calculated based on the
known composition of cement according to the approach proposed by Schindler and Folliard [41] after Bogue [42]:

Q tot 500pC3 S 260pC2 S 866pC3 A 420pC4 AF


642pSO3 850pMgO ;

22

aH te degree of hydration at equivalent age, te , after Rastrup


[43]:

aH te aHu ete ;
s b

23

s; b hydration time and shape parameters:

0:804 0:758
66:78p0:154
p0:401
Blaine
pSO3 ;
C3 A
C3 S

24a

0:535 0:558
0:227
pSO3 ;
181:4p0:146
C3 A pC3 S Blaine

24b

aHu ultimate degree of hydration (aHu 6 1):

aHu

1:031w=c
:
0:194 w=c

25

EK activation energy, J=mol:


0:25
EK 22; 100p0:30
C3 A pC4 AF Blaine

0:35

26

pi weight ratio of the subsequent cement component with


respect to the total weight of cement.
3.3. Thermalshrinkage strain
The imposed thermalshrinkage strains, en , were treated as volumetric strains:

den den;x

den;y

den;z

0 0 0 ;

27

375

and calculated with the predetermined temperature, T (C), and


moisture content, W m3 =m3 , change:

den;x den;y den;z aT dT aW dW;

28

where aT is the coefficient of thermal dilation and aW is the coefficient of moisture dilation. For concrete the coefficient of thermal
dilation was taken after Neville [44] based on the aggregate used
in the mix. The moisture dilation coefficient was taken as equal to
aW 0:002. For soil, the thermal dilation coefficient was taken as

equal to aT 105 = C [20] and the moisture dilation coefficient


was taken as equal to aW 0:001.
3.4. Stress analysis
3.4.1. Early-age concrete
For the stress analysis in early-age concrete viscoelastovisco
plastic model was used. The applied constitutive formulation of
the consistent conception of the model was given by Klemczak
[25]. The following constitutive equations were defined in the viscoelastic area and viscoelastoviscoplastic area, respectively,
expressed with the stress and strain rates:

r_ Dv e e_  e_ n  e_ c ;
r_ Dv e e_  e_ n  e_ c  e_ v p :

29a
29b

where:
Dv e viscoelasticity matrix;
e strain matrix;
en imposed thermalshrinkage strain matrix;
ec matrix of strain representing the effect of creep;
ev p viscoelastic strain matrix.
Both the yield surface, f, and the boundary surface, F, are ratedependent and were expressed as functions of the hardening
parameter, j, and its rate, j_ :

f r; j; j_ 0;
Fr; j; j_ 0:

30a
30b

The creep function, Ct; s, was assumed according to Model


Code 1990 [45] after Gunot et al. [46]. It was assumed that creep
is symmetrical in compression and tension.
The modified 3-parameter WillamWarnke failure criterion,
MWW3, was adopted after Klemczak [29] to describe both the
yield surface and the failure surface (see Fig. 7). Final values of
the mechanical properties (compressive strength, f c;28 , tensile
strength, f t;28 and modulus of elasticity, Ec;28 ) were calculated

Fig. 7. Boundary surface according to the modified 3-parameter WillamWarnke (MWW3) failure criterion.

376

A. Knoppik-Wrbel, B. Klemczak / Engineering Structures 102 (2015) 369386

based on the concrete class according to Model Code 2010 [47] as


mean values (denoted with index m):

f cm;28 MPa f ck;28 8



2=3
f ctm;28 MPa 0:3 f cm;28  8
;

1=3
f cm;28
Ecim;28 GPa 21:5aE
;
10

31a
31b
31c

where the coefficient aE depends on the type of aggregate. The biaxial compressive strength was calculated as [48]:

f t
f t:
f cc t 1:2  cm
1000 cm

32

Yield surface evolves according to the applied hardening law


until it reaches failure surface. When failure surface is reached,
the material exhibits softening behaviour. Deviatoric and volumetric softening was applied. Hardening and softening laws were
adopted after Majewski [28].
3.4.2. Subsoil
For soil the elastoplastic material model with a modified
DruckerPrager failure criterion was used. The constitutive equations ware given by Majewski [27]. In the elastic phase the constitutive equation has a form:

dr De de;

36

Development of the surfaces in time with progressing maturity


of concrete was defined by the time-development of the material
properties, Pt; T; ageing of concrete was defined as a function of
the equivalent age of concrete:

Pt; T bc n P 28 ;

33

where D is elasticity matrix. Beyond the elastic phase the constitutive equation was defined:

dr Dep dee dep ;

where the magnitude of the plastic strain was determined by the


law of plastic flow and:

Dep De  Dp :

where:
Pt; T; P 28 material property (f c ; f t or Ec ) in time t and at the
actual maturity level, and at the age of 28 days, respectively;
bc time development function [47]:

h pi

s 1

bc t; T e

28
te

34

n exponent dependent on the property to describe and concrete composition;


s coefficient dependent on the type of cement.

37

38

The plasticity matrix, Dp , was given by the equation:


p

D D

@f
@r

T "
 T  #1
@f
@f
@f
e @f
T @f
D
r

De
:
@r
@j @r
@r
@r



39

The yield surface, f, has a form:

 0;
f f 1 rm ; j f 2 r

40

where:

j hardening parameter;
Possibility of cracking and its influence on development of
stresses was taken into account. A smeared cracking image was
used in the model. A possibility that a crack occurs in an analysed
point was defined with a damage intensity factor given by the
equation:

0 6 sl

soct
6 1;
f
soct

35

f
where soct is the actual stress state in the analysed point and soct
represents stress states on the failure surface. The damage intensity
factor equal to 1 is equivalent to the stress reaching the failure surface and damage of the element. The character of this damage
depends on the location where the failure surface was reached
[28]. In the analysed cases failure surface was always reached in
the range of hydrostatic tensile stresses which was equivalent to
formation of the splitting crack in the plane perpendicular to the
direction of the maximum principal stress.

rm mean stress;
r stress intensity.
The values of bulk modulus, K, and shear modulus, G, required
for definition of the failure surface were assumed in a manner similar to Duncan and Chang model [49] after Majewski [27]. The bulk
modulus, K, and the initial value of shear modulus, G1 , were given
by the following expressions:


K K o pa

rm

0:5


G1 Go pa

pa

rm

41a

0:5

pa

41b

where K o and Go are material constants, rm is a mean stress and pa


 , the value
is the atmospheric pressure. For a given stress intensity, r
of the shear modulus, G, was calculated as:

Fig. 8. Scheme of the software architecture [3].

377

A. Knoppik-Wrbel, B. Klemczak / Engineering Structures 102 (2015) 369386


2
r
G G1 1  rf ;

42

rf


where r f is a material parameter and rf is a limit value of stress r
for a given mean stress, rm .
The values of the cohesion and internal friction angle were
taken as average values according to the literature data.
3.4.3. Contact elements
A layer of contact finite elements was introduced in the mesh
between the concrete element and soil to allow detachment of
the rotated wall from the soil. Contact elements were assigned
with the material properties of the modified soil. This material
has no tensile strength and limited ability to transfer shear stresses. Such concept was proposed by Majewski [27].
3.5. Implementation
Implementation of the model has a modular architecture
(Fig. 8). The mesh is defined with the use of the MAFEM3D module
created by Wandzik (see [50]). The same mesh is used in thermal
moisture and stress analysis. The main calculations are performed
with the use of two computational modules. The first module,
TEMWIL, is used for thermalmoisture analysis. This module
implements the thermalmoisture part of the numerical model.
The second module, MAFEM, is used for stress and damage analysis. The results are presented with open-source software PARAVIEW.

equal to 4:80  105 (this shrinkage can be correlated with autogenous shrinkage).

4. Analysis and results


The factors that have major effects on the early-age stresses in
walls are dimensions and support conditions of the walls
[16,25,51]. The presented study was focused on the character
and magnitude of stresses in the view of these factors. The analysis
was performed on 12 walls which dimensions are listed in Table 1.
It has been reported that the stresses differ with respect to L=Hc
ratio of the wall; in this study it was investigated if the effect of the
restraint is the same in the walls with equal L=Hc but different
dimensions L and Hc . Dimensions of the walls were chosen in such
a way that wide range of L=Hc was covered (from 1.4 to 10) but
with different walls representing the same L=Hc ratios. AF =Ac ratio
Table 1
Geometrical data for the analysed walls.
No.

01
02
03
04
05
06
07
08
09
10
11
12

in each case was equal to 1.0, given that HF Bc and BF Hc . The


maximum temperature, autogenous shrinkage and temperature
difference in walls with simple geometry is the same for equal
thickness of the wall [52], so all the walls were assigned with the
same thickness.
The ambient temperature was set at T a 20  C and the initial
temperature of concrete at T i T a 5  C 25  C (assumption of
5  C temperature increase due to mixing is recommended by JCI
[20]). Relative humidity of air was set at 55%. The wall was
assumed to be executed 2 weeks after the foundation. Both elements were kept in formwork during the whole process of curing.
The same material was assumed in all the analysed cases, and both
for the wall and its foundation. The composition of the assumed
concrete was as follows: cement CEM I 42.5N 365 kg=m3 , water
160 l=m3 (w=c = 0.44), aggregate sand 641 kg=m3 and granite
1289 kg=m3 . The design class of concrete was C30/37. Composition
of cement is shown in Table 2. Properties of concrete are listed in
Tables 3 and 4.
The meshing of the wall and time discretisation of the analysis
were made according to the recommendations of JCI Guideline
[20]. Fig. 9 presents a scheme of the assumed model of an exemplary wall (No. 09) with its finite element mesh. Because the material model is not regularised, individual finite elements of the mesh
were always of the same dimensions. The analysis was performed
for 28 days in total (14 days before execution of the wall and
14 days after its casting).
The maximum temperature reached in the core of the wall was
equal to 47:0  C. The shrinkage strain at the end of the analysis was

Wall

4.1. Influence of construction sequence


Fig. 10a presents development of the temperature in the core of
the wall with respect to the temperature development in the supporting foundation at different depth in the mid-span longitudinal
section of the wall. The temperature rise in the wall led to reheating of the foundation, which at the moment when the wall
was executed had already cooled down to the temperature of the
surrounding air (T f T a ). Such a temperature increase in the foundation resulted in the strain of the foundation fibres. The strain
change due to re-cooling of the foundation at given depth d was
equal to:
Table 3
Thermo-physical parameters of concrete used in the parametric study.

Foundation

L; m

Hc ; m

Bc ; m

Ac ; m2

HF ; m

BF ; m

AF ; m2

L=Hc

Parameter

Unit

Value

15
15
15
10
10
10
7
7
7
5
5
5

1.50
2.14
3.00
1.42
2.00
3.33
1.40
2.33
3.50
1.67
2.50
3.57

0.7
0.7
0.7
0.7
0.7
0.7
0.7
0.7
0.7
0.7
0.7
0.7

1.05
1.50
2.10
0.99
1.40
2.33
0.98
1.63
2.45
1.17
1.75
2.50

0.70
0.70
0.70
0.70
0.70
0.70
0.70
0.70
0.70
0.70
0.70
0.70

1.50
2.14
3.00
1.42
2.00
3.33
1.40
2.33
3.50
1.67
2.50
3.57

1.05
1.50
2.10
0.99
1.40
2.33
0.98
1.63
2.45
1.17
1.75
2.50

10
7
5
7
5
3
5
3
2
3
2
1.4

Thermal conductivity, k
Specific heat, cb
Density, q

W=m K
kJ=kg K
kg=m3

2.5
0.95
2455

Coefficient of heat exchange, ap

W=m2 K

3.6 1:8 cm plywood


0.8 soil

Coefficient of moisture exchange, bp

m=s

0:18  108  1:8 cm


plywood
0:01  108 soil

Thermal dilation coefficient, aT

1= C

Moisture dilation coefficient, aW

10  106
0.002

Table 2
Mineral composition of cement used in the parametric study.
Component

C3 S

C2 S

C3 A

C4 AF

SO3

CaO

MgO

Blaine; m2 =kg

Amount (%)

64.0

15.0

10.0

8.0

3.3

0.8

0.6

367

378

A. Knoppik-Wrbel, B. Klemczak / Engineering Structures 102 (2015) 369386

Table 4
Mechanical parameters of concrete used in the parametric study.
Parameter

Unit

Value

Compressive strength, f cm
Tensile strength, f ctm
Modulus of elasticity, Eci
Coefficient s
Coefficient n for tensile strength
Coefficient n for modulus of elasticity

MPa
MPa
GPa

38
2.9
33.0
0.25
0.67
0.50

eT;d T max;d  T f aT :

43

At the same time cooling of the wall resulted in the strain


change in the wall:

eT;wall T max;wall  T f aT :

44

temperature difference during cooling. Analogical observation


can be made with respect to the shrinkage strain the restrained
shrinkage strain that generates stresses in the wall is the difference
between shrinkage strain in the foundation and in the wall, as
shown schematically in Fig. 11. This shows that sequential casting
must be taken into account in the analysis, otherwise the value of
the stress would be overestimated.
Another issue is the relative stiffness of the restraining body.
Stiffness change is caused by development of the modulus of elasticity. Assuming that the wall and the foundation are made of the
same material and that the modulus of elasticity development in
time can be described with the following function [47]:

p
n
28
Ec t e es 1 t
Ec;28 ;

47

The strain change that generates the restraint stresses is the difference between the strain in the walls fibres and foundations
fibres:

r cR eT;wall  eT;d Ec :

45

The ratio between the thermal strain during cooling of the wall
and the foundation at the depth of the foundation, shown in
Fig. 10b, was calculated as:

eT;wall  eT;d
:
eT;wall

46

A mean value of c 0:7 can be assumed for this particular case.


It means that the linear strain that would produce the restraint
stress is about 70% of the strain in the wall itself caused by the

Fig. 11. Shrinkage strain difference between the wall and the restraining body.

Fig. 9. Model and FE mesh of an exemplary wall used in the parametric study.

Fig. 10. Relative free thermal strain in an exemplary wall [3].

379

A. Knoppik-Wrbel, B. Klemczak / Engineering Structures 102 (2015) 369386

It can be observed that at the moment when cooling begins


(usually 13 days) the stiffness of the wall can be as little as 20%
of the foundation stiffness (for cements with low rate of strength
development) but should not be less than 50% (after 1 day) to even
75% (after 3 days) in concretes with Portland cement. In such concretes 90% of the foundation stiffness can be obtained after 7 days
of curing. The increased temperature accelerates the rate of the
modulus of elasticity development and Ec =EF is even higher in massive concrete elements. Nevertheless, at the moment when tensile
stresses start to develop the ratio between the stiffness of the wall
and the restraining foundation is non-negligible.
4.2. Influence of dimensions of the wall
Fig. 12. Relative stiffness of early-age wall in time.

where s and n are material coefficients, than the change of the relative stiffness of the wall, calculated as:

p
28

s 1

Ec
e
h pi ;
EF
28
s 1
td
e
t

48

where d is a time between execution of the foundation and the wall,


is shown in Fig. 12 for typical values of s (0.200.40) and d (7, 14
and 28 days).

In the stress analysis it was assumed that the temperature and


moisture content are uniform in the wall this allowed to elimiTable 5
Polynomial coefficients for calculation of dres used in the parametric study.
L=Hc

a0

a1

a2

a3

a4

7
5
3
2
1.4

1
1
1
1
1

0.185
0.387
0.912
1.238
2.362

0.222
0.036
0.041
0.541
0.931

0.253
0.132
0.189
1.158
1.581

0.127
0.031
0.054
0.441
1.224

Fig. 13. Degree of restraint distribution along mid-span section of the walls [3].

380

A. Knoppik-Wrbel, B. Klemczak / Engineering Structures 102 (2015) 369386

Table 6
Slip factor dslip used in the parametric study.
Wall no.
dslip

01

02

03

04

05

06
0.83

07
0.84

08
0.77

09
0.73

Fig. 14. Degree of restraint for walls with equal lengths and different L=Hc ratios [3].

10
0.74

11
0.68

12
0.60

A. Knoppik-Wrbel, B. Klemczak / Engineering Structures 102 (2015) 369386

Fig. 15. Degree of restraint for walls with equal L=Hc ratios and different lengths [3].

381

382

A. Knoppik-Wrbel, B. Klemczak / Engineering Structures 102 (2015) 369386

nate the influence of the self-induced stresses. The temperature


and shrinkage strain in each time step were assumed as equal to
the temperature and shrinkage strain in the core of the element.
The influence of the linear restraint, expressed with the L=Hc ratio,
on the degree of restraint was analysed.
Fig. 13 presents the results of stress analysis for time
t 16:5 days (age of wall t 2:5 days). The diagram in Fig. 13a
presents distribution of the degree of restraint determined in
numerical analysis as a ratio between the stress, r, and the stress
caused by the same strain but under total end restraint,
rfix 1:5 MPa, i.e. cR r=rfix . For comparison the degree of
restraint determined with the analytic approach is shown in
Fig. 13b. The degree of restraint acc. to Nilsson [6] was calculated
taking into account the resilience and possibility of slip failure;
the coefficients of polynomial functions used for definition of the
resilience factors, dres , are listed in Table 5 while the values of slip
factors, dslip , are listed in Table 6. No rotation was taken into
account (total base restraint). The ratio between the moduli of
elasticity was Ec =EF 0:9.
Comparing the results obtained with the two calculation methods it can be observed that they comply only to some extent. The
Nilssons approach reproduced a characteristic that although in
the walls with lower L=Hc the degree of restraint is generally lower,
the influence of the restraining body is more pronounced in the
close vicinity of the joint [53]. Nevertheless, when analysing the
degree of restraint obtained with the numerical analysis it is easily
noticed that it is not equal in the walls with the same L=Hc . Therefore, the influence of length, L, and height, Hc , of the wall on the
degree of restraint was further investigated.
Firstly, the diagrams of the degree of restraint distribution for
the walls with the same length were compared to analyse the
influence of the height of the wall. This comparison is shown in
Fig. 14. The diagrams on the left show the restraint factor distribution while the diagrams on the right the normalised value of the
restraint factor. The normalised restraint factor, cR;1 , was calculated
by introduction of the modification factor, M 1 :

cR;1

cR
M1

49

In the previous study carried out by the authors [53] it was


observed that when the L=Hc was changing by varying length at
constant height the diagrams of the restraint factor exhibited the
following behaviour: the diagrams crossed at the level of 0.25 Hc
and with the increasing L=Hc the value of the restraint factor
decreased at the joint and increased at the top of the wall. Rotation around the crossing point was observed. Based on that observation the value of M 1 was determined so that the diagrams of the
restraint factor show the before mentioned characteristic. In this

case the length was constant while such a behaviour was observed
when L=Hc was changing due to the change of length. Therefore, the
relationship between the value of the modification factor, M1 , and
the heights of the walls was investigated. The walls height ratio,
fH , was calculated for each length, L, in such a way that the shortest
wall was set as a basic wall with the height Hbas and the heights of
the other walls, Hi , were taken as relative heights, i.e. the walls
height ratio for the ith wall was calculated as:

fH;i

Hbas
;
Hi

50

with Hbas < Hi ; for the basic wall fH 1. Comparing M 1 and fH


(Fig. 16a) it was observed that with the increasing height of the wall
the magnitude of the restraint decreases. This relationship becomes
more pronounced as the length of the wall increases. 2nd order
polynomial approximation was proposed to show the trend. It can
be observed that in very long walls (10 and 15 m) the characteristic
is identical and it starts to differentiate as the length of the wall
decreases. This is most probably the influence of the slip at the ends
of the wall.
Analogical analysis was performed on the walls with the same
L=Hc ratios but different dimensions. This comparison is shown
in Fig. 15. The diagrams on the left show the actual restraint factor
distribution while the diagrams on the right the normalised
value of the restraint factor. The normalised restraint factor, cR;2 ,
was calculated by introduction of the modification factor, M 2 :

cR;2

cR
M2

51

It is suggested by analytic approaches that the restraint factors for


the walls with equal L=Hc are also equal, which means their diagrams cover. Thus, for each L=Hc group the M2 was determined so
that the diagrams of the restraint factors covered. It was also
observed that there is a direct relationship between the value of
the modification factor, M 2 , and dimensions of the wall. Since in
each case L=Hc were constant, both length and height of the walls
could have been compared. Thus, for comparison the surface areas
were used (Ai Li Hi ). The walls area ratio, fA , was calculated for
each L=Hc ratio in such a way that the longest wall was set as a basic
wall with the surface area Abas and the areas of the other walls were
taken as the relative areas, i.e. the walls area ratio for ith wall was
calculated as:

fA;i

Abas
;
Ai

52

with Abas > Ai . Comparing M2 and fA (see Fig. 16b) it was observed
that with the increasing area of the wall the magnitude of the
restraint decreases. The influence of the walls area increases with

Fig. 16. Modification factors for determination of the restraint factor.

383

A. Knoppik-Wrbel, B. Klemczak / Engineering Structures 102 (2015) 369386

the increasing L=Hc ratio. Because of a small number of data points


linear approximation was proposed, which suggests directly proportional relationship, but such relationship must be confirmed.
4.3. Influence of support conditions
In the previous analyses it was assumed that the walls had no
possibility to rotate, and to lose contact with the restraining soil,
and that the base had infinite stiffness. In this part of the study the
soil block was introduced to simulate the founding soil. Dimensions
of the soil block were chosen based on the recommendations of JCI
Guideline [20]. Contact elements were introduced between the concrete structure and soil to allow for lifting of walls ends. The influence of the subsoil on the degree of restraint was analysed.
Two walls with distinctly different geometries were chosen a
short wall (No. 12, L=Hc 1:4) and a long wall (No. 02, L=Hc 7).
Two types of soil were investigated: soft soil and hard soil.
Mechanical parameters of the soils were taken after Majewski
[27]. The Model and the finite element mesh are shown in
Fig. 17 on the example of the long wall. Thermophysical and
mechanical properties of the soils are listed in Tables 7 and 8.
The same thermophysical parameters were assumed for both
types of soils to clarify the comparison of stresses. Properties of
concrete were taken from Tables 3 and 4.
Diagrams in Fig. 18 present distribution of the degree of
restraint in the walls with different support conditions, calculated
with the use of numerical and analytic model. In the numerical calculations the comparison was made for the degree of restraint

Table 7
Thermo-physical parameters of soil used in parametric study.
Parameter

Unit

Hard soil and soft soil

Thermal conductivity, k
Specific heat, cb
Density, q
Liquid diffusivity, DWW

W=m K
kJ=kg K
kg=m3
m2 =s

1.5
1.0
2600

Liquid diffusivity, DWT

m2 =s K

Coefficient of heat exchange, ap


Coefficient of moisture exchange, bp

W=m2 K
m=s

109
1.5

Thermal dilation coefficient, aT

1= C

Moisture dilation coefficient, aW

105

0:01  108
10  106
0.001

Table 8
Mechanical parameters of soil used in parametric study.
Parameter

Unit

Hard soil

Soft soil

Coefficient for bulk modulus, K o


Coefficient for shear modulus, Go
Cohesion, c
Internal friction angle, u

MPa

900
405
0.02
40

100
45
0.05
25

calculated assuming total rotational base restraint as well as taking


into account the real stiffness of the founding soil. Analogical comparison was made for analytic calculations, however, in case when
rotation of the wall was allowed the founding soil was assumed to
be infinitely stiff.

Fig. 17. Model and FE mesh of an exemplary wall with real support conditions.

384

A. Knoppik-Wrbel, B. Klemczak / Engineering Structures 102 (2015) 369386

It can be observed that both the numerical and the analytic


model present the same character of the walls behaviour: the
effect of the rotation of the wall on the restraint factor is observed
and it is pronounced in the long wall while in the short wall there

is almost no influence of rotation. The difference in results between


the numerical and analytic model is caused by the fact that the
length of the wall itself is not taken into account in determination
of the restraint factor in analytic calculations (which was discussed

Fig. 18. Degree of restraint in the walls with different support conditions [3].

Fig. 19. Deformation and

rxx stresses, MPa, in walls with different support conditions [3].

A. Knoppik-Wrbel, B. Klemczak / Engineering Structures 102 (2015) 369386

in the previous section). It can also be observed that the stiffness of


the soil influences distribution of the restraint factor: when the soil
is softer, larger rotation occurs. Also in the long wall it can be
observed that the magnitude of the restraint is smaller when the
wall is founded on the soft soil which poses smaller friction.
Deformations of both types of walls are shown in Fig. 19. In the
walls with total restraint the foundation has only possibility of
contraction; when real founding subsoil is taken into account,
the rotation of the foundation occurs. The effect of the rotation is
also visible on the maps of stresses, and is more pronounced on
the soft soil and in the long wall. As it can be deduced from the
degree of restraint diagram, by allowing the possibility of rotation
the value of the tensile stress near the joint increases. Due to that
fact in the long wall founded on the hard soil cracking reaching
approx. 60% of the walls height occurred; concentration of stresses
in the upper part of the wall results from the consequent redistribution of stresses in the cracked wall. In the wall supported on the
soft soil the magnitude of the restraint was generally reduced. As a
result, the value of the stress decreased significantly (with respect
to the wall founded on hard soil) and only a part of the wall (60%)
was subjected to tensile stresses (with respect to 100% in the wall
without the possibility of rotation). Nevertheless, the values of
these stresses were lower than in the wall on the hard soil and
did not lead to cracking.
5. Discussion and conclusions
Dangerous early-age tensile stresses are induced in concrete
walls mainly by limitation of the contraction of the wall in the
cooling phase (restraint of linear strain). Although walls in various
engineering structures differ in geometry, concrete design or technology of execution, the character of stresses in these elements is
similar. The magnitude of these stresses depends on the degree
of restraint exerted by the restraining body (e.g. foundation) to
the restrained wall. There are multiple proposals for computation
of the degree of restraint, which declare that the degree of restraint
depends on the L=Hc ratio of the wall and relative stiffness of the
restraining body. These methods are discussed in the paper.
For proper analysis of stresses in early-age concrete elements
spatial numerical analysis is recommended. Nevertheless, when
such analyses are performed for walls, modelling of the external
restraint is often not appropriately approached and the obtained
degree of restraint is incorrect. The most common shortcomings
of the models are: assumption of total rotational restraint of the
base, infinite stiffness of the soil and negligence of different maturity development of the early-age element and the restraining concrete element resulting from the construction sequence.
The presented study focused on the analysis of stresses in
early-age concrete walls in relation to the restraint conditions.
The concept of the degree of restraint was used as a measure for
characterisation of these stresses. A numerical model was proposed for simulation of the behaviour of early-age walls including
casting sequence and soilstructure interaction. This model was
used in further study for determination of stresses in walls with
different dimensions and support conditions. The results obtained
with the use of the numerical model were compared with the
results of the analytic method proposed by Nilsson [6]. The following conclusions can be drawn from the study:
1. Construction sequence must be taken into account for two
major reasons: (1) The strain that produces stress in a wall is
a differential strain, ediff , resulting from the difference in strain
in the wall (due to cooling and autogenous shrinkage), ewall , and
foundation (due to re-heating and autogenous shrinkage), efound ,
which is less than the strain in the wall itself (ewall > ediff ). (2)

385

Following layers of concrete in the element are characterised


with different maturity and stiffness. In typical walls made of
OPC the stiffness of foundation may be even 2 times higher at
the moment when tensile stresses start to develop in the wall,
and this ratio may increase when cements with lower rate of
strength development are used. Consideration of relative stiffness was also postulated by Nilsson.
2. The restraint factor can be used as a measure to characterise the
stresses induced in the walls, however, the degree of restraint
obtained with the numerical and analytic model comply only
to some extent. The degree of restraint increases with an
increasing L=Hc ratio, which agrees with Nilssons approach
and other analytic formulations. Nevertheless, in contrary to
them, for the same L=Hc ratios it is not equal but has lower values in walls with larger L and Hc dimensions. Hence, in determination of the degree of restraint not only the L=Hc ratio but also
the individual dimensions of the wall must be taken into
account.
3. Real support conditions of the wall must be simulated to properly determine the degree of restraint. This requires introduction of the soil block to simulate (1) real stiffness and friction
of the restraining body (soil) and (2) possibility of loss of contact between the rotating foundation and the soil. When possibility of rotation is taken into account, the value of the restraint
factor increases at the joint but decreases at the top edge of the
wall; when soil is softer, larger rotation occurs. The influence of
the rotation of the wall on the degree of restraint increases with
an increasing length of the wall and it is practically unnoticed in
short walls. In the long walls it is also visible that the decrease
of frictional properties of soil leads to reduction of the translational restraint. This observation is in agreement with Nilssons
observations.

Acknowledgements
The co-author of this paper, Agnieszka Knoppik-Wrbel, was a
scholar under the Project DoktoRIS, co-funded by the European
Union under the European Social Fund.
References
[1] Knoppik-Wrbel A. The influence of self-induced and restraint stresses on
crack development in reinforced concrete wall subjected to early-age thermalshrinkage effects. In: Proceedings of the 14th international conference of
postgraduate students juniorstav, Brno, Czech Republic; 2012. p. 162.
[2] Klemczak B, Knoppik-Wrbel A. Analysis of early-age thermal and shrinkage
stresses in reinforced concrete walls. ACI Struct J 2014;111(2):31322. http://
dx.doi.org/10.14359/51686523.
[3] Knoppik-Wrbel A. Analysis of early-age thermalshrinkage stresses in
reinforced concrete walls. PhD thesis. Silesian University of Technology,
Gliwice, Poland; 2015.
[4] JSCE. Guidelines for concrete no. 15: Standard specifications for concrete
structures. Design; 2011.
[5] ACI Committee 207. ACI 207.2R-07: Report on thermal and volume change
effects on cracking of mass concrete; 2007.
[6] Nilsson M. Thermal cracking of young concrete. Partial coefficients, restraint
effects and influence of casting joints. Masters thesis. Lule University of
Technology, Lule, Sweden; 2007.
[7] Xiang Y, Zhang Z, He S, Dong G. Thermalmechanical analysis of a newly cast
concrete wall of a subway structure. Tunn Undergr Space Technol
2005;20:44251. http://dx.doi.org/10.1016/j.tust.2005.02.005.
[8] Flaga K, Klemczak B, Knoppik-Wrbel A. Early-age thermalshrinkage crack
formation in bridge abutments experiences and modelling. In: Proceedings of
the fib symposium, Tel-Aviv, Israel; 2013. p. 4358.
[9] Klemczak B, Knoppik-Wrbel A. Reinforced concrete tank walls and bridge
abutments: early-age behaviour, analytic approaches and numerical models.
Eng
Struct
2015;84:23351.
http://dx.doi.org/10.1016/j.
engstruct.2014.11.031.
[10] Seruga A, Zych M. Research on thermal cracking of a rectangular RC tank wall
under construction. I: Case study, J Perform Constr Facil. doi:http://dx.doi.org/
10.1061/(ASCE)CF.1943-5509.0000704.

386

A. Knoppik-Wrbel, B. Klemczak / Engineering Structures 102 (2015) 369386

[11] Nilsson M. Restraint factors and partial coefficients for crack risk analyses of
early age concrete structures. PhD thesis. Lule University of Technology,
Lule, Sweden; 2003.
[12] Benboudjema F, Torrenti J-M. Early-age behaviour of concrete nuclear
containments. Nucl Eng Des 2008;238:2495506. http://dx.doi.org/10.1016/j.
nucengdes.2008.04.009.
[13] Prochaczek M. Technical issues of the design of shielding bunker for 235 MeV
cyclotron. In: Proceedings of the 9th central european congress on concrete
engineering, Wrocaw, Poland; 2013.
[14] Knoppik-Wrbel A. Cracking due to restraint stresses in early-age radiation
shielding wall. ArchCivil EngEnviron 2014;7(3):4961.
[15] Hsthagen A, Jonasson J-E, Emborg M, Hedlund H, Wallin K, Stelmarczyk M.
Thermal crack risk estimations of concrete tunnel segments Equivalent
Restraint Method correlated to empirical observations. Nordic Concr Res
2014;49:12743.
[16] Klemczak B, Knoppik-Wrbel A. Early-age thermal and shrinkage cracks in
concrete structures description of the problem. ArchCivil EngEnviron
2011;4(2):3548.
[17] Al-Gburi M, Jonasson J-E, Nilsson M, Hedlund H, Hsthagen A. Simplified
methods for crack risk analyses of early age concrete. Part 1: Development of
Equivalent Restraint Method. Nordic Concr Res 2012;46:1738.
[18] PN-EN 1992-3. Eurocode 2 Design of concrete structures. Part 3: Liquid
retaining and containment structures; 2008.
[19] Bamforth PB. CIRIA C660: Early-age thermal crack control in concrete. Tech
rep. CIRIA, London, UK; 2007.
[20] Japanese Concrete Institute. Guidelines for control of cracking of mass
concrete; 2008.
[21] Larson M. Thermal crack estimation in early age concrete. Models and
methods for practical application. PhD thesis. Lule University of Technology,
Lule, Sweden; 2003.
[22] Al-Gburi M, Jonasson J-E, Yousif ST, Nilsson M. Simplified methods for crack
risk analyses of early age concrete. Part 2: Restraint factors for typical case
wall-on-slab. Nordic Concr Res 2012;46:3956.
[23] Waller V, dAloa L, Cussigh F, Lecrux S. Using the maturity method in concrete
cracking control at early ages. Cement Concr Compos 2004;26:58999. http://
dx.doi.org/10.1016/S0958-9465(03)00080-5.
[24] Kwak H-G, Ha S-J. Non-structural cracking in RC walls. Part II: Quantitative
prediction model. Cement Concr Res 2006;36:76175. http://dx.doi.org/
10.1016/j.cemconres.2005.12.002.
[25] Klemczak B. Modeling thermalshrinkage stresses in early age massive
concrete structures comparative study of basic models. Arch Civil Mech
Eng 2014;14(4):72133. http://dx.doi.org/10.1016/j.acme.2014.01.002.
[26] Klemczak B. Prediction of coupled heat and moisture transfer in early-age
massive concrete structures. Numer Heat Transfer, Part A: Appl 2011;60
(3):21233. http://dx.doi.org/10.1080/10407782.2011.594416.
_
[27] Majewski S. Sprezystoplastyczny
model wsppracujacego ukadu budynek
_
podoze
poddanego wpywom grniczych deformacji terenu. Silesian
University of Technology, Gliwice, Poland; 1995.
[28] Majewski S. MWW3 elasto-plastic model for concrete. Arch Civil Eng
2004;50(1):1143.
[29] Klemczak B. Adapting of the WillamWarnke failure criteria for young
concrete. Arch Civil Eng 2007;53(2):32339.
[30] Hillel D. Fundamentals of soil physics. 1st ed. San Diego (USA): Academic
Press; 1998.

ervenka J. Multiscale hydro-thermo-mechanical model


[31] Jendele L, milauer V, C
for early-age and mature concrete structures. Adv Eng Software
2014;72:13446. http://dx.doi.org/10.1016/j.advengsoft.2013.05.002.
_
[32] Kiernozycki
W. Betonowe konstrukcje masywne. Polski Cement, Cracow,
Poland; 2003.
[33] Zreiki J, Bouchelaghema F, Chaouchea M. Early-age behaviour of concrete in
massive structures experimentation and modelling. Nucl Eng Des
2010;240:264354. http://dx.doi.org/10.1016/j.nucengdes.2010.07.010.
[34] Endrizzi S, Quinton WL, Marsh P. Modelling of spatial pattern of ground thaw
in a small basin in the arctic tundra. Cryosphere Discuss 2011;5:367400.
http://dx.doi.org/10.5194/tcd-5-367-2011.
[35] Farouki OT. Thermal properties of soils. Monograph 81-1. US Army Cold
Regions Research and Engineering Laboratory, Hanover, USA; 1981.
[36] Hancox NL. The diffusion of water in concrete. Tech rep. UKAEA, Winfrith, UK;
1966.
[37] Gawin D. Numerical solution of coupled heat and moisture transfer problems
with phase changes in porous building materials. Arch Civil Eng 1993;34
(4):393412.
[38] Wyrwa J, Szczesny J. Migracja wilgoci wywoana gradientami temperatury w
pytach betonowych poddanych obrbce termicznej. Arch Civil Eng 1986;32
(1):42132.
[39] Clapp RB, Hornberger GM. Empirical equations for some soil hydraulic
properties. Water Resour Res 1978;14(4):6014.
[40] Philip JR, de Vries DA. Moisture movement in porous materials under
temperature gradients. Trans Am Geophys Union 1957;38:22232.
[41] Schindler AK, Folliard KJ. Heat of hydration models for cementitious materials.
ACI Mater J 2005;102(1):2433. http://dx.doi.org/10.14359/14246.
[42] Bogue JH. The chemistry of Portland cement. New York (USA): Reinhold
Publishing Corporation; 1947.
[43] Rastrup E. Heat of hydration in concrete. Mag Concr Res 1954;6(17):114.
[44] Neville AM. Properties of concrete. 5th ed. San Francisco (USA): Prentice Hall;
2012.
[45] CEB-FIP fib. Model Code 1990; 1991.
[46] Gunot I, Torrenti J-M, Laplante P. Stresses in early age concrete: comparison
of different creep models. ACI Mater J 1996;93(3):2548. http://dx.doi.org/
10.14359/9809.
[47] CEB-FIP fib. Model Code 2010; 2013.
[48] CEB-FIP fib. Bulletin 70. State-of-the-art report: code-type models for concrete
behaviour. Background of MC2010. Tech rep; 2013.
[49] Duncan JM, Chang CY. Nonlinear analysis of stress and strain in soil. ASCE J Soil
Mech Found Div 1970;96(SM5):162953.
[50] Majewski S, Wandzik G. MWW3 model for concrete adjustment of failure
and yield surface for use in computational FEM system MAFEM3D. ArchCivil
EngEnviron 2009;2(2):3344.
[51] Klemczak B, Knoppik-Wrbel A. Early-age thermal and shrinkage cracks in
concrete structures influence of geometry and dimensions of a structure.
ArchCivil EngEnviron 2011;4(3):5570.
[52] Klemczak B, Knoppik-Wrbel A. Comparison of analytical methods for
estimation of early-age thermalshrinkage stresses in RC walls. Arch Civil
Eng 2013;59:97117.
[53] Klemczak B, Knoppik-Wrbel A. Comparative study of methods for analysis of
early-age thermalshrinkage stresses in externally restrained concrete
structures. In: Proceedings of the 10th international conference on new
trends in statics and dynamics of buildings, Bratislava, Slovakia; 2012. p. 414.

You might also like