Sunteți pe pagina 1din 20

Effects of the early age thermal behaviour on long term

damage risks in massive concrete structures

Matthieu Briffaut*,** — Farid Benboudjema* — Jean-Michel


Torrenti ***— Georges Nahas*,**

* Laboratoire de Mécanique et Technologie, ENS de Cachan


CNRS/Université Pierre et Marie Curie
61 avenue du Président Wilson, F-94235 Cachan cedex
briffaut@lmt.ens-cachan.fr
** Institut de Radioprotection et de Sureté Nucléaire
Fontenay-aux-Roses, France
*** Université Paris Est, Laboratoire Central des Ponts et Chaussées
Paris, France

ABSTRACT. At early-age, temperature in massive concrete structures may overcome 70°C


because hydration is an exothermic chemical reaction. This important temperature evolution
can lead to long term damage risks. Firstly, cracking due to self restrained strains can occur
(which increase the transport properties and so the kinetics of degradation) and secondly,
delayed ettringite formation can appear. Therefore, if autogenous and thermal strains are
restrained, compressive stresses and then tensile stresses rise, which can cause crossing
cracks but this paper will not deal with this phenomena. In the first part, sensibility to delayed
ettringite formation and early age cracks by self restrained with regards to the environmental
conditions are studied by a numerical approach. This part shows that the external
temperature has a significant impact on the maximal temperature reached but that the
temperature difference between the core and the surface is mainly impacted by the wind
velocity. Then, a parametric study on the effect of the variability of thermal properties at
early age has been achieved and shows that this variability needs to be taken into account.
Finally, visco-elastic mechanical calculations show the impact of thermal properties
variability on the stresses generated by self restraint.
KEYWORDS: Early age, thermal properties, concrete, massive structures.

Revue. Volume X – n° x/année, pages 1 à X


2 Revue. Volume X – n° x/année

1. Introduction

At early-age (during the construction), massive concrete structures (bridge piers,


dams, nuclear containments for example) are submitted to important thermal
evolutions (maximal temperature may overcome 70°C) because hydration is an
exothermic chemical reaction. This important temperature evolution can lead to long
term damage risks. Indeed, cracking due to self restrained strains can occur because
there is strains gradient between the core and the surface (due to thermal gradient).
For structures like tanks or nuclear containments, this cracking may increase
significantly the concrete permeability (reduce the tightness) and diffusivity and
thus the kinetics of concrete deterioration (steel corrosion, chloride ions penetration,
leaching…).
Moreover, reaching important temperatures in the core of massive structure can
lead at long term to delayed ettringite formation (DEF), possible source of
deterioration.
A good knowledge of thermo-chemical phenomena and mechanisms which
occur at early age in the concrete is thus necessary to predict long term behaviour
and durability of massive structures.
Modeling the temperature field in massive structures will allow us to study the
effect of environmental conditions on the risks of DEF and cracking by thermal
gradient. Besides, a parametric study will highlight the effect of the variation
thermal properties. Finally, consequences of this variability on the stresses generated
by self restraint will be studied by numerical calculations (using a visco-elastic
model).
Effects of autogenous and drying shrinkage are not considered in this study.

2. Thermo-chemical behaviour of a massive element

2.1. Modeling presentation

The temperature evolutions are obtained by solving the heat equation. This
equation includes the heat release due to the hydration reaction (through a heat
source term):
CT  kT   L [1]
where L is the total heat release [J.m ], k is the thermal conductivity [W.m .K-1]
-3 -1

and C is the thermal volumetric capacity [J.m-3.K-1].


To take into account the thermo-activation of the hydration reaction, an
Arrhenius type law is used with the notation proposed by (Ulm and al., 1998):
 Ea 
  A exp  
~
 [2]
 RT 
Massive structures long term damage risks: early age thermal behaviour effect 3

where  is the hydration degree, A  is the chemical affinity, T is the


~
temperature[K], Ea is the activation energy [J.mol-1] and R is the perfect gas
constant.
Thermal boundary conditions are of convective type. Heat flux can be written as
following:
  hTs  Text  [3]
where h is the coefficient of global exchange convection which takes into account
convective and radiation exchanges [W.m-2.K-1], Ts is the temperature at the surface
of the element [K] and Text is the external temperature [K].
2.2 Evolution of the convective coefficient

In the literature, (for example (Boussa, 2000), (Azenha and al., 2009), (Jeon and
al., 2008), (Lee and al., 2009)), a value of about 10W.m-2.K-1 for the coefficient of
exchange by convection is generally used. Besides, to simplify the model, this
parameter includes the exchange by natural or forced convection but also the
radiation one ((Buffo-Lacarrière, 2007) for example). The calculation of the
convective coefficient can be done through the number of Nusselt (Nu):
hl
Nu  c [4]
kf

where lc is the characteristic length [m] of the convective flux and kf is the thermal
conductivity of the fluid [W.m-1.K-1].
Although, the Nusselt number can also be calculated from the following
equations:

Nu  0,13Pr  Gr 
1/ 3
For natural convection: [5]

For forced convection: Nu  0,664  Re1/ 2  Pr1/ 3 [6]


Theses equations involve the numbers of Grashof (Gr), Prandt (Pr) and
Reynolds (Re):

g    T  lc,Gr
Gr  [7]
2
where g is the gravitational acceleration [m.s-2],  is the thermal dilatation
coefficient of the fluid [K-1], lc,Gr is the characteristic length of the natural
convection [m], is the fluid kinematic viscosity [m².s-1].

Pr  [8]

4 Revue. Volume X – n° x/année

where  is the thermal diffusivity [m².s-1]

V l
c, Re
Re  [9]

where lc,Re is the characteristic length of the forced convective flux [m], V is the fluid
velocity [m.s-1].
Because the temperature variation remains little compared to the radiation heat
transfer, radiation heat transfer can be linerarized (Hernot and al., 1984) in order to
be integrated to the global convective coefficient:

hray  4 Tmoy


3
[10]

where hray is the equivalent convective coefficient which corresponds to the


radiation [W.m-2K-1], T is the temperature [K],  is the surface emissivity [-],  is the
Stefan-Boltzmann constant [W.m-2.K-4].
The final expression of the global convective coefficient can therefore be written as
following:
1/ 3
 g    T  l   
 c, Gr   kf  3 [11]
  4 Tmoy
For natural convection: h  0,13  
      lc, Gr
   

1/ 2 1 / 3 
 V  lc, Re     kf 
  3 [12]
h  0,664      4 Tmoy
For forced convection:
      l
   c, Re 

For example, for a massive wall (thickness: 1.2m; height: 2m; length: 20m) and
with the assumption that the wind direction is parallel to the wall, figure 1 shows the
evolution of the global convective coefficient. One can see on this figure that for a
wind velocity equal to 0, usual values are included between our maximal and
minimal value (the evolution begins with 5.3W.m-2.K-1 and reaches 13.3 W.m-2.K-1
for a high difference between the surface and the external temperature). The
framework could be integrated in the convective coefficient (by the use of an
equivalent conduction coefficient for instance).
Massive structures long term damage risks: early age thermal behaviour effect 5

30,00

25,00
Text
20,00
Twall
h 15,00
(W/m²K)
10,00
5,00
25,00-30,00 Wind
20,00-25,00 0,00
0

10

20

15,00-20,00
30

40 0 Wind
50

60

70
10,00-15,00

80
velocity
5,00-10,00 (Km/h)
0,00-5,00 Text-Twall (°C)

Figure 1. Global convective coefficient evolution


Then, a comparison between two numerical simulations (performed with Cast3m
finite elements code, (Castem)) with a constant convective coefficient or with a
variable convective coefficient (natural convection) can be performed (the mesh is
displayed in figure 2). The results of these simulations are displayed on figure 3.
These simulations show that the temperature evolutions (of the wall studied) are
slightly impacted by the effect of the convective coefficient variability (which stands
for the temperature measurement accuracy). Thus, it seems accurate to use a
constant value for the convective coefficient if the wind velocity is low.

X
Z

Figure 2. Finite element mesh


6 Revue. Volume X – n° x/année Levée 2
120 Dimensions en cm
90
#7
10
10
65 #6
10 10
Lift 11
Levée 95
60
#A
55 #4 #3 #5
95
Temperature (°C)

50
#2
10
45 10
#1
Raft fondation
Fondation
40
T1 h variable
35 T2 h variable
T3 h variable
30
T5 h variable
25 T6 h variable
T1 h constant
20 T2 h constant
0 20 40 60 80 100 120 T3 h constant
time (h) T5 h constant
T6 h constant

Figure 3. Temperature evolution in a massive wall: effect of a variable global


convective coefficient (for natural convection)

2.3 Effect of the environnemental conditions (wind velocity and external


temperature) on the DEF risks

Investigating the maximal temperature reached at early age can be a good way to
study the massive structure sensibility to the risks of delayed ettringite formation
(DEF). Indeed, even though DEF is a long term deterioration (Taylor, 1997),
(Lawrence, 1998) the dissolution of ettringite composed phases is a necessary
conditions for the appearance of DEF. It may appear if at early age the maximal
temperature exceeds 65-70°C. Of course, this condition is not sufficient and DEF
needs other factors to occur (Collepardi, 2003) (LCPC recommendation, 2007).
Nevertheless, in this paper, a study of maximal temperature reached in a massive
wall (thickness: 1.2m; height: 2m; length: 20m), cast on a massive slab, is proposed
(this case corresponds to the first lift of a nuclear containment construction).
In this study, only the external conditions (wind velocity and external
temperature) are source of variability but for an application to a real case, materials
properties variability must be taken into account (total heat release, thermal concrete
conduction,…). Besides, the initial concrete temperature has a significant effect on
the maximal temperature but to avoid this third parameter, we used a relation
between the external temperature (Text) and the initial concrete one (Tini), proposed
by (Torrenti and al., 2009). The initial concrete temperature is linked to the external
temperature because aggregates are generally kept in stock outside but it often is
higher than external temperature because cement is generally at a higher
temperature. Figure 4 shows the comparison between external temperature
measurement and initial concrete temperature measurement on two building sites.
Massive structures long term damage risks: early age thermal behaviour effect 7

Initial concrete temperature


30
y = 0,6x + 10,9
25 2
R = 0,5
(°C) 20
15
10
5
0
-5 0 5 10 15 20
External temperature (°C)

Figure 4. Temperature evolution in a massive wall: effect of a global variable


convective coefficient (for natural convection)

The results of this study are presented on figure 5. One can see on this figure a slight
effect of wind velocity on maximal temperature. This can be explained by the fact
that, as the thermal concrete conductivity is rather low, the thermal exchange
conditions at the core are close to adiabatic conditions. On the contrary, the
evolution of the maximal temperature due to the external temperature is quasi-linear.
This can be explain by the fact that the initial temperature is a linear function of the
external temperature.

60-62
60

55-60
55

Tmax (°C)
50-55
50

45 45-50

40 40-45

35
35-40
40

20
30

15
10
20

Wind velocity (Km/h) 5


10

0
0

External temperature (°C)

Figure 5. Maximal temperature (Tmax) evolution reached in a massive wall


(thickness 1.2m) : Effect of wind velocity and external temperature
8 Revue. Volume X – n° x/année

2.4 Effect of wind velocity and external temperature on the risks of cracking due
to thermal gradient

The early age cracking can be a consequence of a high temperature gradient


between the core and the surface of an element. In this case, at the beginning of the
hydration reaction, the core tends to expand more than the surface. Therefore tensile
stresses are generated, which can lead to cracking. This non-crossing cracking will
have important effect on transfer phenomena in the concrete cover area. Then, a
reinforcement premature corrosion may appear. During the temperature decrease,
the stresses state is inversed and tensile stresses appear on the core. Jeon and al.
(Jeon and al., 2008) try to specify a criteria based on the temperature difference
between the core and the surface instead of the temperature gradient. Indeed, the
stresses profile, generated by the thermal strains which respect the Euler-Bernoulli
assumption, is almost parabolic. Therefore the gradient is not constant. With a
model which takes into account an age-dependent elastic modulus during hydration,
Jeon and al. finally defined a cracking criterion: Icr = 25/ T (Icr> 1.5 to prevent
crack, 1.5>Icr>1.2 to limit cracks and 1.2>Icr>0.7 to limit harmful cracks). The study
results are shown on figure 6. On contrary to the risks of DEF, the wind velocity has
a high effect whereas the external temperature has a slight one. This indicates that
with regards to these two phenomena, there are no optimal casting conditions. It is
recalled that autogenous and drying shrinkage are not considered.

29
28-29
28
27-28
27 26-27

26 25-26
Tsurface - Text
(°C) 24-25
25
23-24
24
22-23
23 21-22

22

21 20
40

30

10
20

10

0 External temperature (°C)


0

Wind velocity (Km/h)

Figure 6. Temperature difference evolution between the core and the surface (Tsurface
–Text) of a massive wall (thickness 1.2m): effect of the wind velocity and of the
external temperature
Massive structures long term damage risks: early age thermal behaviour effect 9

2.5 Effect of the evolution of the concrete thermal properties evolutions at early
age

In the previous simulation the concrete thermal properties (the thermal


conductivity and the thermal volumetric capacity) have been considered constant
(Waller, 2000), (Loukili and al., 2000) but at early age this is not completely exact
(Mounanga and al., 2004). Effectively, these parameters depend on the concrete
temperature but also on the hydration degree. To quantify the error on the
temperature field evolution induced by the use of constant values, a parametrical
study has been achieved.
2.5.1 Effect of temperature on thermal conductivity and thermal capacity:
constitutive equations

Morabito (Morabito, 2001) performed measurements in hardened concrete


samples under temperature variations. These tests enlighten that thermal
conductivity decreases with the increase of concrete temperature and that this
decrease depends more on the aggregate type than on the cement type. On the
contrary, these tests show that the thermal capacity increases with the concrete
temperature and that the variations are more significant in concrete with limestone
aggregate than with gravel (the evolution also seems to be independent from the
cement type). The author proposed the following relationship to link the thermal
conductivity and the thermal capacity to temperature:
X
T
X
X

0   T T
0
 [13]
0
where XT is the thermal property (capacity or conductivity) at temperature T,  is the
slope of the relationship determined from the experimental data (for limestone
aggregate =-0.0015°C-1 for the thermal conductivity and =0.0016°C-1 for the
thermal capacity), X0 is the thermal property at the reference temperature and
T0=20°C.
2.5.2 Effect of hydration degree on thermal conductivity and thermal capacity:
constitutive equations

During cement hydration, thermal conductivity and thermal capacity endure


significant changes due to the evolution of the products which compose concrete.
According to Faria and al. (Faria and al. 2006) the concrete thermal conductivity
in the first hours after mixing is about 20–30% greater than in hardened concrete.
Ruiz and al. (Ruiz and al. 2001) proposed a variation according to:

  
k    k 1,33  0,33 [14]
  
 
10 Revue. Volume X – n° x/année

where k∞ stands for the value of thermal conductivity for hardened concrete [W.m-
1 -1
.K ] (ξ= ξ∞).
In the IPACS project (Improved Production of Advanced Concrete Structures),
Lura and al. (Lura and al. 2001), proposed a thermal capacity decrease linked to the
hydration degree:

cbindwwcement Cwater
C    C0  [15]
 concrete

where C0 is the thermal capacity calculated from the concrete mix,  is the hydration
degree [-], cbindw is a coefficient which takes into account how chemically and
physically bond water influences the thermal capacity (equal to 0.2), Wcement is the
cement content of the mix [Kg.m-3], Cwater is the thermal capacity of the water,
concrete is the concrete density .
2.5.3 Effect of temperature and hydration degree on thermal conductivity and
thermal capacity: simulation results
Simulations have been performed on the same geometry than the simulations
studying the effect of a variable convective coefficient (figure 3).
Figure 7 displays the results obtained with different thermal conductivity
evolution. The evolution represented with white dots (square, circle and triangle) are
the results of the simulations with a constant thermal conductivity (equal to 2.5
W.m-1.°C-1). The upper and the lower values, represented by the vertical bar,
correspond to simulations with a constant thermal conductivity (equal 2 and 3 W.m-
1
.°C-1 respectively). The black dots (square, circle and triangle) correspond to the
evolution which takes into account effect of temperature and hydration degree
(eq.13 and 14). For this simulation, the final thermal conductivity used is equal to
2.15 in order to obtain a mean value of 2.5 at ∞=0.42 (half of the hydration
reaction). One can see that temperature at the core of the concrete element (T3) is
slightly decreased by the use of a variable thermal conductivity.
Massive structures long term damage risks: early age thermal behaviour effect 11

65
T1_k=2.5
60 T1_k=f(T,ξ)

55 T2_k=2.5

T2_k=f(T,ξ)
50
T3_k=2.5
Temperature (°C)

45 T3_k=f(T,ξ)

40

35

30

25

20
0 20 40 60 80 100 120

Time (hours)

Figure 7. Temperature evolution of a massive wall (thickness 1.2): effect of thermal


conductivity

The study of the thermal capacity variation is more complex because this one is
also used in the determination of the total heat release and the chemical affinity. The
concrete thermal capacity is used in the analysis of semi-adiabatic test which is the
base of the determination of the adiabatic temperature evolution (Ulm et al. 98).

In this test, the heat release q is calculated from the concrete temperature
evolution Tconcrete and the external temperature Text :
qt sa   Cconcrete Tad t sa   Tad t sa  0
t sa [16]
   a  b t  t dt
 Ctot Tconcrete t sa   Tconcrete t sa  0 
0

where Cconcrete is the concrete thermal capacity [J/°C], Ctot is the total thermal
capacity (concrete + calorimeter), a [W] & b [W.°C-1] are the heat losses coefficient
of the calorimeter, tsa [s] is the real time of the semi-adiabatic test,  = Tconcrete - Text,
and Tad is the theoretical adiabatic temperature.
Table 1. Value of adiabatic temperature obtained for different thermal capacity
value

Thermal capacity (J.Kg-1.°K-1) 900 1050 1200 C = f (T); 1050 at 20°C

Tad (°C) 79.43 70.72 64.75 67.3


12 Revue. Volume X – n° x/année

Table 1 gives the value of adiabatic temperature and figure 8 represents the
affinity evolution obtained for different thermal capacity values. In this calculation,
the effect of the hydration degree variation on the thermal capacity is not taken into
account because its evolution is calculated from the results of the semi-adiabatic
test.
A(ξ) (C=900 J/(Kg.K))
1600
A(ξ) (C=1050 J/(Kg.K))
1400
A(ξ) (C=1200 J/(Kg.K))
Chemical affinity (1/s)

1200 A(ξ) (C=f(T))

1000 Polynomial (A(ξ) (C=900 J/(Kg.K)))

Polynomial (A(ξ) (C=1050 J/(Kg.K)))


800
Polynomial (A(ξ) (C=f(T)))
600 Polynomial (A(ξ) (C=1200 J/(Kg.K)))

400

200

0
0,00 0,10 0,20 0,30 0,40 0,50 0,60 0,70 0,80 0,90

Hydration degree (-)

Figure 8. Evolution of the chemical affinity

Figure 9 displays the effect of thermal capacity. The evolutions represented with
white dots (square, circle and triangle) are the results of simulations with a constant
thermal capacity (equal to 1050J.Kg-1.K-1). The upper and the lower value,
represented by a vertical bar, correspond to simulations with constant thermal
capacity (values of 900 and 1200 J.Kg-1.K-1 respectively). The black dots (square,
circle and triangle) correspond to the evolution taking into account the effects of
temperature and hydration degree (eq. 13 and 15). For this simulation, the initial
thermal capacity used is equal to 1111J.Kg-1.K-1 in order to obtain a value of
1050J.Kg-1.K-1 at half of the hydration reaction (∞=0.84). A decrease of about 8%
on the temperature at the core of the concrete element (T3) by the use of a variable
thermal capacity is obtained. Besides, one can see that the effect of a variation of the
thermal capacity (from 900 to 1200J.Kg-1.K-1) has a significant non-linear influence
on the maximal reached temperature.
Massive structures long term damage risks: early age thermal behaviour effect 13

70
T1_C=1050 J.Kg-1.K-1
65 T1_C=f(T,ξ)
T2_C=1050J.Kg-1.K-1
60
T2_C=f(T,ξ)
55 T3_C=1050J.Kg-1.K-1
Temperature (°C)

T3_C=f(T,ξ)
50

45

40

35

30

25

20
0 20 40 60 80 100 120

Time (hours)

Figure 9. Temperature evolution in a massive wall (thickness 1.2): effect of the


thermal capacity

In a real case, both thermal capacity and conductivity previously presented


evolve at early age. Therefore, a simulation, taking into account the evolution of
thermal conductivity and thermal capacity with respect to temperature and hydration
degree, was performed (see figure 10). It shows a decrease of about 10.4% on the
reached maximal temperature and a decrease of about 14.4% of the difference of
temperature between the core and the surface (with regards to the simulation with
constant thermal properties). The slightly more important impact on the temperature
difference between the core and the surface is explained by the fact that the surface
temperature is slightly affected by the evolution of the thermal properties whereas
the reached maximal temperature decreases. That is why, the study of the DEF risks
or cracking due to global thermal shrinkage can be performed with constant thermal
properties because this type of study is mainly focused on the reached maximal
temperature. Nevertheless, the study of cracking by thermal gradient with constant
parameter slightly overestimates the cracking risk.
14 Revue. Volume X – n° x/année

65 Temperature at surface C,k = constant


Temperature at core C,k = constant
60
Temperature at surface C,k = f( T,ξ)
55 Temperature at core C,k = f( T,ξ)

50
Temerature (°C)

45

40

35

30

25

20
0 20 40 60 80 100 120
Time (h)

Figure 10. Temperature evolution of a massive wall (thickness 1.2) at the surface or
at the core: effect of the variation of thermal properties with respect to hydration
degree and temperature

3. Impact of thermal properties variability on stresses generated by self


restraint
3.1 Mechanical visco-elastic model presentation

3.1.1 Strains decomposition


The relationship between apparent stresses σ , elastic stiffness tensor E, elastic
strains ε e , basic creep strains ε bc , transient thermal creep strains ε ttc , autogenous
strains ε au , thermal strains ε th , total strains ε , reads:

σ  E ε el  E ε  ε bc  ε ttc  ε au  ε th  [17]

The Young modulus E increases due to hydration as follows (De Schutter 1999)
(Stefan 2010):

  0
E    E    [18]
with    0 

Where  is the hydration degree,  0 the mechanical percolation threshold,   is


the final hydration degree (which depends on the concrete mix (Waller 2000)), E is
the final Young modulus (i.e. when    )  
is the positive part operator, β is a
constant material parameters.
3.1.1 Autogenous and thermal shrinkage
Massive structures long term damage risks: early age thermal behaviour effect 15

Autogenous shrinkage is directly related to the evolution of hydration in the


material. Experimental results show that autogenous shrinkage evolution is almost
linear with respect to hydration degree as soon as a percolation threshold has been
overcame ((Laplante 1993) ; (Mounanga 2006)). Therefore, autogenous shrinkage
ε au can be modelled by the following equation (Ulm and al. 1998):

ε ijau     0   ij [19]

where  is a constant material parameter.


An expansion can also be measured at the beginning of hydration, depending on the
w/c ratio and type of cement. This is not taken into account here.
Thermal strain ε th is linked to temperature variation, and to the coefficient of
thermal expansion  (considered constant, (Loukili and al. 2000)):

ε ijth   (T  T0 ) ij [20]

where  is the thermal dilatation coefficient and T0 is the reference temperature.

3.1.1 Creep strain


In our model, basic creep is described with 3 Kelvin-Voigt units combined in serial.
For each Kelvin-Voigt unit, the differential equation can be written as
(Benboudjema and al. 2008) :

 kbci    i 
 bci bci   bci  1bc  i [21]
 kbc   
i
kbc  

where  bc
i
is the characteristic time (constant), kbc
i
  is the stiffness of the spring
(increasing with the hydration degree),  is the apparent stress.
The stiffness parameter for each unit is calculated with the following equation (De
Schutter 1999):

kbci    kbci _ 
0.473
 0.62 [22]
2.081  1.608

where kbci _  is the final stiffness.


The characteristic time is assumed to be constant for each Kelvin-Voigt units, so the
dashpot viscosity depends on the hydration degree:

bci    kbci   bci [23]

The effect of temperature on creep must be taken into account. Experimental results
highlight that temperature increases creep at early-age by a factor of 2 or 3
(Arthanari 1967). On the one hand, temperature evolution increases the rate of
hydration degree, and therefore reduces creep strains at higher temperature. Indeed,
16 Revue. Volume X – n° x/année

if temperature increases, the hydration degree (eq.2) and, therefore, each spring
stiffness increase (eq. 22). The result is a decrease of the creep strain rate with
respect to temperature. This is in contradiction with experimental evidences. On the
other hand, temperature increases creep strains due to the two following factors
(Hauggard and al. 1999):
 At constant temperature, creep strains rate increases. This is due to the
decrease of water viscosity with temperature;
 Transient temperature history also increases creep strains. The obtained
deformation is called transient thermal creep or load induced thermal
strains. Such kind of strain is observed at very high temperature (above
100°C) but exists at lower ones (Bažant and al 1997). Bažant and al
(Bažant and al 1997) suggest that this strain corresponds to drying creep.
The effect of constant temperature on creep strain is taken into account by
modifying the spring stiffness and the dashpot viscosity of Kelvin-Voigt units as
follows:
Eac  1 1 
  
kbci  , T   kbci  , T0 e R T T0
[24]
Eac  1 1 
  
bci  , T   bci  , T0 e R T T0
[25]

where Eac is the creep activation energy and T 0 = 293 K. In this way, the
characteristic time is kept independent of the temperature.
Hauggard and al.(Hauggard and al. 1999) have modeled transient thermal creep by
using the model developed by Bažant and al. (Bažant and al 1997). Here, a simpler
model is used which is similar to the one used to predict thermal transient creep of
concrete at high temperatures (Thelandersson, 1987). Thermal transient creep strain
 ttc reads:

ttc   T ~ [26]

where  [Pa-1.K-1] is a material parameter.


However, if several heating/cooling cycles are considered, eq. [26] predicts a
continuous increase of strains, which may not be consistent. Therefore, this model
should not be used in that case.
The creep models are then extended to multiaxial state of stresses by mean of a
creep Poisson ratio taken equal to the elastic one.

3.1 Mechanical results of simulation

Mechanical simulations have been performed on geometry presented on figure 2. To


highlight the effect of self restraint, the structural restraint due to the slab has been
Massive structures long term damage risks: early age thermal behaviour effect 17

removed and the wall is free to expand or to shrink. Nevertheless, the temperature
fields have been calculated on the complete geometry.
Four simulations for four cases were performed:
1. The thermal capacity and the thermal conduction are constant values
2. The thermal capacity is function of hydration degree and temperature but
the thermal conduction is a constant value
3. The thermal capacity is a constant value but the thermal conduction is
function of hydration degree and temperature
4. The thermal capacity and the thermal conduction are function of hydration
degree and temperature
The other thermal parameters and all the mechanical parameters are the same for
each simulation.
For example, figure 11 displays the stress fields for the case N°3 when temperature
at the core reaches the peak.

Figure 11. Example of stress fields (MPa) generated by self restraint for a wall of
1.2m width.

Table 2. Stress variability generated by thermal properties variability


due to self-restraint

Case 1 2 3

I (% of the case 4) 29 2,3 21

XX (% of the case 4) 26,4 3,2 22,1

YY(% of the case 4) 29 1 19

(% of the case 4) 30,9 4,9 21,4


18 Revue. Volume X – n° x/année

Table 2 summarizes the stress variability obtained for each direction and for each
case with respect to the case which is closer to reality (the thermal capacity and the
thermal conductivity depend on the hydration degree and the temperature).
The results show that, not taking into account the evolution of thermal properties
can lead to an error on the maximal stress calculated of about 30%. Besides, two
conclusions can be drawn. Firstly, although in this configuration the range of
generated stress by self restraint is low, thermal properties variability cannot be
neglected in an accurate study of cracking due to self restraint. Secondly, even
though it is conservative, constant values for the thermal properties induce a
significant overestimation of the stresses.

4. Conclusions
The thermal early age behaviour influence on the concrete damage mechanisms
can be declined under two main axes:
 The sensibility to DEF risks
 The sensibility to concrete skin cracking due to thermal gradient

In this paper, the effects of environmental conditions during the casting and the
hydration reaction were studied. The results show that, if the wind velocity is low,
the use of a constant convective coefficient gives temperature field with sufficient
accuracy. Besides, it has been shown that the wind velocity effect on the DEF risks
is weak but non-negligible on the cracking by thermal gradient risks. As the effects
of external temperature are the opposite, there are no optimal casting conditions.
A parametrical study highlights the effect of the thermal properties (concrete
thermal conductivity and thermal heat capacity) variation (in function of the
temperature and the hydration degree) on the maximal temperature reached and on
the maximal temperature difference between the core and the surface
(overestimation of 10.4% on the maximal temperature and of 14.4% on the
temperature difference). Moreover, with the use of constant thermal parameter, the
value of the latter could have a significant and non-linear influence on the reached
maximal temperature.
Finally, a mechanical study with a visco-elastic model show that the
overestimation of the temperature difference between the core and the surface with
the use of constant parameter leads to an overestimation of about 30% on the
maximal generated stress. As the overestimation is significant, thermal properties
evolution cannot be neglected for the study of massive structure self-restraint.

5. Bibliographie
(Azenha and al., 2009) Azenha M., Faria R., Ferreira D. (2009), Identification of
early-age concrete temperatures and strains: Monitoring and numerical
simulation. Vol. 31. Cement & Concrete Composites, p. 369-378.
Massive structures long term damage risks: early age thermal behaviour effect 19

(Boussa, 2000) Boussa O. (2000), Structures en béton soumises à des sollicitations


thermomécaniques sévères : évolution des dommages et des perméabilités.
Memoire de thèse LMT Cachan, in French.
(Buffo-Lacarrière, 2007) Buffo-Lacarrière L. (2007), Prévision et évaluation de la
fissuration précoce des ouvrages en béton. Memoire de thèse, Université de
Toulouse, LMDC, in French.
(Castem) Commissariat à l'Energie Atomique CEA - DEN/DM2S/SEMT, Cast3m
finite element code, available at http://www-cast3m.cea.fr/
(Collepardi, 2003) Collepardi M. (2003), A state-of-the-art review on delayed
ettringite attack on concrete. Vol. 25. Cement & Concrete Composites.
(Faria and al., 2006) Faria R., Azenha M., Figueiras J.A. (2006), Modelling of
concrete at early ages: Application to an externally restrained slab, Cement
and concrete composites, 28 p. 572–585
(Hernot, 1984) Hernot, Porcher (1984). Thermique appliquée aux bâtiments.
Editions parisiennes, ISBN2862430153.
(Jeon and al., 2008) Jeon S.J., Choi M.S., Kim Y.J. (2008). Advanced Assessment of
Cracking due to Heat of Hydration and Internal Restraint. Vol. 105. ACI
materials journal.
(Lawrence, 1998) Lawrence C.D.(1998). Physiochemical and mechanical properties
of Portland cements. Lea's chemistry of cement and concrete. 4th ed.
(Lee and al., 2009) Lee Y., Choi M.S., Yi S.T., Kim J.K. (2009). Experimental study
on the convective heat transfer coefficient of early-age concrete. Vol. 31.
Cement & Concrete Composites.
(Loukili and al., 2000) Loukili A., Chopin D., Khelidj A., Le Touzo J.Y. (2000) A
new approach to determine autogenous shrinkage of mortar at an early age
considering temperature history. Cement and Concrete Research. Vol. 30.
(Lura and al., 2001) Lura P., Van Breugel K. (2001), Thermal Properties of
Concrete_Sensitivity Studies, Projet IPACS.
(Morabito, 2001) Morabito P. (2001), Thermal Properties of Concrete. Variations
with the temperature and during the hydration phase, Projet IPACS.
(Ruiz and al., 2001) Ruiz J, Schindler A, Rasmussen R, Kim P, Chang G. (2001),
Concrete temperature modeling and strength prediction using maturity concepts
in the FHWA HIPERPAV software, 7th international conference on concrete
pavements, Orlando, USA.
(Taylor, 1997) Taylor H.F.W.(1997). Cement chemistry. 2nd ed. London:Telford
Publishing, 1997.
(Torrenti and al., 2009) Torrenti J.M., Buffo-Lacarrière L. (2009), On the variability
of temperature fields in massive concrete structures at early age,
(Ulm and al., 1998) Ulm F.J., Coussy O. (1998), Couplings in early-age concrete:
from material modelling to structural design. Edited by International Journal of
Solids and Structures. Vol. 35.
20 Revue. Volume X – n° x/année

(Waller, 2000) Waller V. (2000), Relations entre composition des bétons,


exothermie en cours de prise et résistance en compression. Edited by série «
Ouvrages d’Art » ( in French) collection Etudes et recherches des Laboratoires
des Ponts et Chaussées.
(De Schutter, 1999) De Schutter G. (1999), Degree of hydration based Kelvin model
for the basic creep of early age concrete, Materials and Structures, pp. 260-265,
Vol. 32.
(Stefan, 2010) Stefan L., Benboudjema F., Torrenti J.M., Bissonnette B. (2010),
Prediction of elastic properties of cement pastes at early ages, Computational
Materials Science, pp. 775-784, Vol. 47.
(Hauggaard, 1999) Hauggaard A.B., Damkilde L.,Hansen P.F. (1999), Transitional
thermal creep of early age concrete, Journal of Engineering Mechanics, pp. 468-
475, Vol. 125.
(Bažant, 1997) Bažant Z.P., Hauggaard A.B., Baweja S., Ulm F.J. (1997),
Microprestress-solidification theory for concrete creep. I: Aging and drying
effects, Journal of Engineering Mechanics, pp. 1188-1194, Vol. 123, 1997.
(Arthanari, 1967) Arthanari S., Yu C.W. (1967), Creep of concrete under uniaxial
and biaxial stresses at elevated temperatures, Mag. of Concrete Research, pp.
149–156, Vol. 19, 1967.
(Thelandersson, 1987) Thelandersson S. (1987), Modelling of combined thermal and
mechanical action in concrete, Journal of Engineering Mechanics, pp. 893-903,
Vol. 113, 1987.
(Benboudjema, 2008) Benboudjema F., Torrenti J.M. (2008), Early-age behaviour
of concrete nuclear containments, Nuclear Engineering and Design, pp. 2495-
2506, Vol. 238.
(Laplante, 1993) Laplante P. (1993), Propriétés mécaniques des bétons durcissants :
Analyse comparée des bétons classiques et à très hautes performances, PhD
Thesis ENPC, in French.
(Mounanga, 2006) Mounanga P., Baroghel-Bouny V., Loukili A., Khelidj A. (2006),
Autogenous deformations of cement pastes: Part I. Temperature effects at early
age and micro–macro correlations, Cement and Concrete Research, , pp. 110–
122, Vol. 36.

S-ar putea să vă placă și