Download as pdf or txt
Download as pdf or txt
You are on page 1of 34

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

net/publication/225917466

Experimental and numerical studies of thermo-hydrous transfers in concrete


exposed to high temperature

Article  in  Heat and Mass Transfer · December 2007


DOI: 10.1007/s00231-006-0212-9

CITATIONS READS

51 206

5 authors, including:

Mulumba Kanema Marcus V. G. de Morais


Université du Havre University of Brasília
6 PUBLICATIONS   265 CITATIONS    107 PUBLICATIONS   354 CITATIONS   

SEE PROFILE SEE PROFILE

Albert Noumowé Jean-Louis Gallias


Université de Cergy-Pontoise Université de Cergy-Pontoise
97 PUBLICATIONS   2,933 CITATIONS    71 PUBLICATIONS   641 CITATIONS   

SEE PROFILE SEE PROFILE

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

SHM Techniques to Damage Identification on Bridges using Wavelet View project

Structural Control with TMDs View project

All content following this page was uploaded by Jean-Louis Gallias on 30 January 2017.

The user has requested enhancement of the downloaded file.


INTERNATIONAL JOURNAL OF HEAT AND MASS TRANSFER, ELSEVIER

EXPERIMENTAL AND NUMERICAL STUDIES OF THERMO-


HYDROUS TRANSFERS IN CONCRETE EXPOSED TO HIGH
TEMPERATURE

M. Kanema, M.V.G. de Morais, A. Noumowe, J.L. Gallias, R. Cabrillac

Université de Cergy-Pontoise – L2MGC


5 mail Gay-Lussac, Neuville sur Oise, 95031 Cergy-Pontoise, France
Tel. : (00-33) 1 34 25 69 16
Fax. : (00-33) 1 34 25 69 41
e-mail : albert.noumowe@iupgc.u-cergy.fr

ABSTRACTS: The compression strength of concrete can be as high as 80 MPa at


28 days. High strength concrete can be obtained by decreasing porosity and
lowering permeability. Concrete, especially high strength concrete (HSC),
performs poorly when subjected to fire. This is attributed to high thermal stresses
and water vapour pressure. High thermal gradient induces high thermo-
mechanical stresses in the concrete system. Low permeability prevents water from
escaping and induces high water vapour pressure causing cracking and spalling.
The aim of this study is both experimentally and numerically study the coupled
heat and mass transfers in concrete exposed to elevated temperature. Five
concrete mixtures with various cement contents and water cement ratios of a
constant aggregate content were studied before and after heating–cooling cycles.
The concrete cylindrical specimens were subjected to several tests: compression
and splitting tensile tests, measurement of modulus of elasticity, heating – cooling
cycles, thermal field and mass loss during the heating – cooling cycles, and
permeability tests. Comparisons between the numerical and experimental results
on the thermo-hydrous behaviour were reported. Parametric analyses were
carried out in order to underline main parameters involved in concrete behaviour
at high temperature. The numerical and experimental results included thermal
gradient, water vapour pressure, relative humidity, concrete mass losses due to
dehydration, and water content for concrete elements heated from 20 to 600 °C.
The results show the degrees of damage due to the concrete chemical
transformations at high temperature.

1
INTERNATIONAL JOURNAL OF HEAT AND MASS TRANSFER, ELSEVIER

KEYWORDS: Concrete, spalling, thermal gradient, vapor pressure, coupled heat


and mass transfers.

Introduction

Concrete is one of the most used materials in buildings and civil engineering
constructions. The wide use of concrete is attributed to its high strength,
availability and other important mechanical properties. A concrete should
possesses low porosity and permeability in order to achieve high mechanical
properties. . Unfortunately, concrete, in certain circumstances, performs poorly
when subjected to high temperature. This was seen in the recent tunnel fires,
which have provoked cracking and spalling in the concrete structures. These
phenomena are mainly due to high thermal stresses and water vapour pressure.
High thermal gradient induces high thermo-mechanical stresses in the concrete
element. Low permeability prevents water from escaping and induces high water
vapour pressure in the concrete element.
A better knowledge of the behaviour of this material, in particular of high
performance concrete, at high temperature has been acquired thanks to various
works and research projects carried out for a few years at the national and
international level - BHP 2000 (Collectif Presses de l'ENPC, 2005), HITECO
(Brite, 1999), etc . In spite of its good performances at ambient temperature,
concrete can present risk of spalling or bursting at high temperature. The stake is
to understand the thermo-hydro-mechanical phenomena which occur when
concrete is subjected to high temperature. The analysis of the effects of high
temperature on concrete microstructure is a necessary step to understand the
complex macroscopic phenomena.
The research tasks dealing with this problem can be divided into two
groups: one group of experimental works aiming at the determination of the
thermal, hydrous and mechanical properties of concretes (ordinary concrete and
high performance concrete) in function of temperature (Noumowé, 1995; Galle et
al., 2000; Galle and Sercombe, 2000; Noumowe et al., 2003) and a group of
theoretical works aiming at the modelling of the coupled mechanisms (thermal,
hydrous and mechanical) which occur when concrete is subjected to high
temperature (Bazant and Thongutai, 1978; Bazant and Kaplan, 1996; Mainguy et

2
INTERNATIONAL JOURNAL OF HEAT AND MASS TRANSFER, ELSEVIER

al., 1999; Feraille-Frenet, 2000; Nechnech, 2000; Mounajed and Zakhem, 2001;
Gawin et al., 2001; Sercombe et al., 2001). The present work makes a junction
between the both groups.
Several authors have made the study of heat and mass transfer in concrete
when subjected to high temperature. Bazant and Thongutai (1978) have presented
the first numerical model for the prediction of moisture migration in normal and
high performance concretes based to water thermo behaviours. Some
experimental results are included for comparison with model predictions. Other
models (Gawin et al., 2001; and Bazant and Kaplan, 1996) are proposed to
simulated behaviours of a heated concrete elements. These methods propose
complex models that treat thermo-hydrous-chemical-mechanical coupling with
three fluids (liquid water, vapour, dry air) into a solid skeleton. These models
seem to adapt complex problems with different geometries and different
materials. Less complex models are most adapted to understand the physical
behaviours of a problem like a concrete cylindrical specimen exposed to elevated
temperatures.
The aim of this work is to carry out a numerical study on the coupled heat
and mass transfers in a concrete element exposed to elevated temperature in order
to explain the behaviour observed during the experimental study. This simple
numerical model, proposed at first by Sercombe et al. (2003), simulate concrete
element behaviour exposed to temperature under to 650°C. This model simulates
chemical degradation of cement paste by porosity increase (Feraille-Frenet, 2000).
Our experimental procedure determines five formulations of concrete
characterized by their cement content. Each formulation was heated in an electric
furnace for four heating–cooling cycles where eighteen cylindrical specimens
(∅16x32cm, ∅15x30cm and ∅11x22cm) were tested. Comparisons were drawn
on numerical and experimental results on the thermo-hydrous behaviour of a
concrete element. Parametric analyses were carried out in order to underline main
parameters involved in concrete behaviour at high temperature. The numerical and
experimental results included thermal gradient, water vapour pressure, relative
humidity, concrete mass losses due to dehydration, and water content for a
concrete element heated from 20 to 600°C. The results show high thermal
gradients and high vapour pressure in the concrete element in addition to the
damage due to the concrete chemical transformations at high temperature.

3
INTERNATIONAL JOURNAL OF HEAT AND MASS TRANSFER, ELSEVIER

Concrete Spalling Phenomenon

In situation of fire, constructions (buildings, tunnels, etc) are exposed to high


temperature that provokes important damages. For concrete, spalling or bursting
generally occurs at temperature ranging between 250°C and 400°C. This
instability can take place in two forms: by bursting, detachments of pieces of
concrete one after the other or by explosive spalling of the structural elements.
Figure 1 shows concrete damage observed on a 16x32 cm concrete
specimen heated at 1°C/min up to 300°C (Noumowé et al., 1994). The specimen
was reconstructed from the spalled pieces collected in the furnace after the test
completed.

Figure 1 – Spalling: Pieces of one spalled concrete specimen heated up to 300°C, see Noumowé et
al. (1994).

Two main phenomena seem to be at the origin of the concrete spalling or bursting:
o The heating of the concrete causes a high thermal gradient, which involves
important deformations. The deformations, when they are prevented,
induce high thermal stresses in the element.
o When the temperature increases, free or/and unbound water vaporizes. A
part of this vaporized water is evacuated towards the heated surface of the
concrete. The other part migrates towards the interior of the concrete
where the temperature is still low and condenses. Thus; it forms a quasi -
saturated zone with low water permeability. It is near this zone that vapour
pressure reaches its maximum. The peak can reach high values and induce
high stresses in the concrete element.
These phenomena seem to occur in a simultaneous way and the magnitude
depends on several parameters:
4
INTERNATIONAL JOURNAL OF HEAT AND MASS TRANSFER, ELSEVIER

o The heating rate; with heating rate, the risk of concrete bursting increases.
o The concrete surface exposed to the heat source.
o Initial water content of the concrete; tests showed that concrete element
cracks less in a dry state than in a wet state.
o The mix composition and the aggregate type of the concrete; the risk of
bursting is greater for high performance concrete than for ordinary
concrete. Low permeability of high performance concrete makes more
difficult the vapour escaping.
o Geometry of the concrete element (dimensions, shape).
o The mechanical loading; concrete elements subjected to a large loads are
more likely to burst.
o The position of the steel reinforcements in a reinforced concrete element.

Concrete Mixes

Materials

Cement CEM I 52.5 was used. Three granular phases were also used:
• a semi-crushed siliceous gravel of class 10/20 mm.
• a rolled siliceous fine gravel of class 5/10 mm.
• a rolled siliceous sand of class 0/5 mm.
• a high water reducing superplasticizer CIMFLUID 2002 containing modified
polycarboxylate was employed (with different contents) to ensure a
satisfactory workability of the concrete mixtures. Its density at 20°C is 1100
kg/m3. Its pH is 7 ± 1. The dried material is 35.0 ± 1,7 %. The advised content
of CIMFLUID 2002 is in the range from 0.2 to 2kg for 100kg of cement.

Formulations

We determine five formulations of concrete characterized by their cement content


325, 350, 400, 450 and 500 kg/m³, see Table 1. The B325 concrete formulation
was determined. In order to obtain the other compositions, the cement content was
increased and the aggregate content kept constant. Superplasticizer was adjusted
so as to obtain a good workability of the fresh concrete (between 8 and 20 cm at
the slump test).

5
INTERNATIONAL JOURNAL OF HEAT AND MASS TRANSFER, ELSEVIER

During the experimental study of the behaviour of these concretes at high


temperature, several B400 concrete specimens spalled explosively. All the
spalling took place during the heating at about 300°C. So, the experimental results
of B400 concrete was chosen to compare to numerical results and to analyze the
spalling phenomenon.

Table 1 – Concrete compositions studied (Kanema et al., 2005).

Components (kg/m³) B325 B350 B400 B450 B500


Cement 325 350 400 450 500
Gravel 960 960 960 960 960
Fine gravel 89 89 89 89 89
Sand 740 740 740 740 740
Water 202 194 177 160 143
Superplasticizer 0,00 0,35 1,04 1,73 2,43
Water/Cement 0,62 0,55 0,44 0,36 0,29
Admixture/Cement 0,00% 0,10% 0,26% 0,39% 0,49%
Theoretical Vol. Mass (kg/m³) 2316 2333 2367 2401 2435

Specimen preparation

Cylindrical specimens (16x32 cm, 15x30 cm and 11x22 cm) were made. Two
Type K thermocouples were embedded in fresh concrete 16x32 cm specimens.
These thermocouples make it possible to measure the temperature in the centre
and semi-ray of the specimens and assume the thermal gradients. After casting,
the specimens were wrapped in a plastic sheet and cured at ambient temperature
during three days, before being demoulded and stored in water. The mechanical
and thermal tests were carried out when the specimens were 28 days old.

Heating and Mechanical Tests

Heating–cooling cycles

The specimens were heated in an electric furnace. Four heating–cooling cycles


from 20°C to 150°C, 300°C, 450°C and to 600°C were applied. The first phase of
the cycle is heating rise ramp. The temperature increases at 1°C/min from 20°C to
the maximum target temperature. The second phase is a constant temperature
6
INTERNATIONAL JOURNAL OF HEAT AND MASS TRANSFER, ELSEVIER

dwell at the maximum target temperature for one hour. The last phase is a cooling
ramp. The temperature decreases at approximately 1°C/min until the room
temperature. The chosen heating and cooling rates refer to the recommendations
of the technical committee RILEM TC-129 (1995). The four heating – cooling
cycles are presented on Figure 2.
600
T (°C)
150°C
500
300°C
400
400°C

300 450°C

200 600°C

100

0
0 5 10 15 t (h) 20

Figure 2 – Schematic representation of the heating – cooling cycles (surface temperature in


function of time).

Concrete specimens disposition into test-furnace

For each heating – cooling cycle, various sizes of cylindrical specimens


(∅11x22cm, ∅16x32cm, ∅15x5cm) were tested. The placement of the specimens
in the furnace was carried out so as to get a quite homogeneous temperature
around the specimens (see Figure 3).

7
INTERNATIONAL JOURNAL OF HEAT AND MASS TRANSFER, ELSEVIER

concrete
specimen
∅11×22

concrete
specimen
∅11×22
(suspended)
concrete
specimen
∅15×5
(cut)
concrete
specimen
∅16×32

Figure 3 – Schematic view of test oven

The specimens were divided into different groups as shown in Table 2.

Table 2 – Tested specimens


Number Dimensions Test
6 ∅15×30cm Permeability
3 ∅16×32cm Thermal gradient
3 ∅16×32cm Modulus of elasticity
4 ∅11×22cm Compressive strength
4 ∅11×22cm Tensile strength
1 ∅11×22cm Weight loss
2 ∅11×22cm Thermal gradient

Additional thermocouples were placed on the surface of the specimens. All


the thermocouples (those placed on the surface and those inserted in the
specimens) were connected to an automatic acquisition unit. A regulator-
programmer controls the furnace temperature and carries out the desired cycle has
been used.

8
INTERNATIONAL JOURNAL OF HEAT AND MASS TRANSFER, ELSEVIER

In addition one ∅11×22cm cylindrical specimen was suspended to a load


cell in order to measure the weight of the concrete specimen during the heating –
cooling cycle. The use of this device makes it possible to measure, plot and store
the specimen mass variation during the heating – cooling cycle.

Mechanicals tests

Compression and splitting tensile tests

The French standard NF P 18-406 and NF P 18-408 for compressive and tensile
splitting tests respectively were used. The ∅11×22cm cylindrical specimens were
used to characterize the strength of the concrete before and after the heating –
cooling cycles. Four specimens were tested for each cycle.
Measurement of the modulus of elasticity was carried out on three
∅16×32cm specimens for each heating – cooling cycle. The secant modulus of
the stress-strain curve was measured by applying 70 % of the specimen yield load.

Permeability test

CEMBUREAU permeameter with constant pressure of Nitrogen was used to


determine the concrete apparent permeability. The apparent permeability of
concrete was measured. Intrinsic permeability Kv can be obtained by considering
the relation shown in equation (1). The equation takes into account the slip of the
gas molecules on the concrete pores and cracks walls.

 b *
K v = K 1 +  (1)
 p 

Kv is the intrinsic permeability (m²), b* is slope at the curve origin, and p is


average pressure (Pa).
In order to measure the concrete permeability ∅15×30cm cylindrical
specimens in which were cut out 15x5 cm cylindrical samples were made. The
samples were then tested in the permeameter before or after the heating – cooling
cycle. Non-heated samples undergo a simple drying at 80°C until constant mass
before testing.

9
INTERNATIONAL JOURNAL OF HEAT AND MASS TRANSFER, ELSEVIER

Concrete mass loss during the heating – cooling cycle

The knowledge of the specimen mass loss during the heating – cooling cycle
makes it possible to determine the part of eliminated water according to the
applied heating – cooling cycle.

Temperature differences and thermal gradient

The two thermocouples embedded in the concrete specimens made it possible to


measure the temperature at the centre and at the middle-ray during cycling. A
third thermocouple was placed on the surface of the specimen; thus it made it
possible to measure the temperature field within the specimen during the heating –
cooling cycle and to calculate the thermal gradient.

Experimental Results

Mechanical strength

Figure 4 shows the residual compressive strength as a function of the maximum


temperature of the heating – cooling cycles. It can be noticed that all the curves of
residual compressive strength included two parts. The first part, from 20°C until
300°C or 350°C, various evolutions of strength were observed. B325 concrete
exhibited no decrease in the compressive strength. B350 and B400 concretes
exhibited an insignificant decrease in the compressive strength when subjected to
temperature from 20°C to 150°C and then an equivalent increase between 150°C
and 300°C. B450 and B500 concretes presented a quasi continuous increase
between 20°C at 300°C or 350°C. On the other hand for temperatures higher than
300°C or 350°C, compressive strength decreased significantly. When the
concretes were heated at 600°C, the residual compressive strength was less than
10 MPa.
The behaviour of the residual tensile strength was similar to that of the
compressive strength as shown in Figures 4 and 5. However, there was no
increase observed for the tensile strength curves, as shown in Figure 5.
Furthermore, the concretes with low cement content (B325 and B350) seemed to
preserve their initial tensile strength up to 150°C.

10
INTERNATIONAL JOURNAL OF HEAT AND MASS TRANSFER, ELSEVIER

100
σ ck (MPa)
90 B300
80 B350
70 B400
60 B450
50 B500
40
30
20
10
0
0 100 200 300 400 500 T (°C) 600

Figure 4 – Residual compressive strength as a function of temperature.

7
σ tk (MPa) B300
6
B350
5 B400
B450
4
B500
3

0
0 100 200 300 400 500 T (°C) 600

Figure 5 – Residual tensile strength as a function of temperature.

It appeared that when the concretes were heated up to 300°C or 350°C, the
strength properties were slightly affected by the heating whereas from 300°C or
350°C to 600°C the strength properties were decreased significantly and became
almost negligible at 600°C.

Modulus of elasticity

The residual modulus of elasticity of the tested concrete as a function of


temperature was presented in Figure 6. It is observed that contrary to the strength
properties, the modulus of elasticity was regularly deteriorated by heating.

11
INTERNATIONAL JOURNAL OF HEAT AND MASS TRANSFER, ELSEVIER

50
E (GPa)
B300
40
B350

30 B400

20

10

0
0 100 200 300 400 500 T (°C) 600

Figure 6 – Residual modulus of elasticity as a function of temperature.

Permeability

Figure 7 shows the permeability as a function of temperature. It was observed


that, regardless of the concrete mix, the permeability decreased between 20°C and
150°C. However, it increased in an exponential way for higher temperatures. For
the heating at 600 °C, the permeability was quite high (>10-13m²); it exceeded the
possibility of the permeameter. The concrete specimens presented a dense
network of open cracks.

0 100 200 300 400 T (°C) 500


1E-13
Kv (m²)
1E-14

B300
1E-15
B340
B400
1E-16 B450
B500

1E-17

Figure 7 – Intrinsic permeability as a function of temperature.

For a heating at 300°C, the concrete permeability was similar to that


measured at 80°C. The development of permeability as a function of temperature
was similar to that of compressive strength and to a lesser degree to that of tensile

12
INTERNATIONAL JOURNAL OF HEAT AND MASS TRANSFER, ELSEVIER

strength. It can be seen that an improvement in the strength properties was often
observed at 150°C. A favourable modification of the microstructural properties of
the tested concrete cementing matrix could be assumed. However, an
investigation on the concrete matrix microstructure is necessary to confirm such
an assertion.

Concrete mass loss

Figures 8 and 9 present the evolution of mass loss during the heating - cooling
cycles. It was observed that before 150°C, the variation of mass is very weak. The
mass loss in this temperature range corresponds mainly to the escape of the
concrete interstitial water. Between 150 and 300°C, one could observe a strong
mass loss for all the tested concretes. It corresponds to the decomposition of
cement hydrates and represents about 65 to 80 % of the total water in the concrete.
Each concrete lost between 5 and 6 % of its initial mass. Knowing that the total
water in the concretes is between 6 and 9 % of the concrete mass, we could notice
that the main part of the water contained in each concrete escaped during the
heating between 150 and 300°C. The other part of water (between 1 and 3 % of
the concrete mass) was evacuated at temperatures beyond 300°C.

9%
PM (%)
8%
7%
%-B300-600°C
6%
%-B400-600°C
5%
4%
3%
2%
1%
0%
0 2 4 6 8 10 12 t (h) 14

Figure 8 – Concrete mass loss as function of time (heating–cooling cycle at 600°C).

13
INTERNATIONAL JOURNAL OF HEAT AND MASS TRANSFER, ELSEVIER

8%
PM (%)
7%

6%

5%

4%

3% %-B450-300°C
%-B450-450°C
2%
%-B450-600°C
1%

0%
0 100 200 300 400 500 T (°C) 600

Figure 9 – Concrete mass loss as function of the specimen surface temperature


(for B450 concrete subjected to three heating – cooling cycles)

By comparing the mechanical strength, the permeability and the mass loss,
it can be noticed that a strength decrease after 300°C takes place after a mass loss.
This confirms that the heating between 150°C and 300°C induced the
decomposition of the cement hydrates without a significant effect on the concrete
strength. The decomposition of cement hydrates beyond 300°C resulted in an
increase in permeability and a decrease in mechanical strength.

Temperature differences and thermal gradients

Figures 10 and 11 show the temperature differences between the centre and the
surface of the specimens during the heating – cooling cycle at a maximum
temperature of 300°C.

14
INTERNATIONAL JOURNAL OF HEAT AND MASS TRANSFER, ELSEVIER

120
B325-300°C
100
B350-300°C
80 B400-300°C
B450-300°C
60
B500-300°C
40

20

-20

-40
0 1 2 3 4 5 6 7 8 9 t (h) 10

Figure 10 – Temperature difference between the surface and the centre of the specimen as a
function of time.

120
∆ T (°C) B325-300°C
100
B350-300°C
80 B400-300°C
B450-300°C
60
B500-300°C
40

20

-20
Tcenter (°C)
-40
0 50 100 150 200 250 300 350

Figure 11 – Temperature difference between the surface and the centre of the specimen as a
function of the temperature at the centre (b).

As shown in Figure 11, for the five studied concrete mixtures the temperature
difference between the surface and the centre of the specimens increased and
reached a peak at about 200°C. Also as shown in Figure 11, the temperature
difference decreased after 200°C. The peak value was located near 105°C and
corresponded to a global thermal gradient in the specimen of approximately
13°C/cm. High thermal gradient may cause high thermal stresses in the concrete
specimen. Numerical calculations are being carried out to confirm this
assumption.

15
INTERNATIONAL JOURNAL OF HEAT AND MASS TRANSFER, ELSEVIER

Discussion

The effect of heating on the concrete specimens was analyzed. The experimental
results showed two temperature ranges:
• During the first temperature range, from the ambient temperature to 300°C,
the mechanical strength and permeability did not vary significantly. However,
between 150 and 300°C, there was a small decrease in the modulus of
elasticity and a large decrease in mass loss. The concrete mixtures showed
small changes in the permeability and mechanical properties. The properties
of the heated concrete at 150°C remained very close to those of the non-heated
concrete. For concrete mixtures with high cement content there was even an
improvement in the compressive strength. Between 150 and 300°C, a large
amount of water (interstitial water and hydrates bound water) had escaped
from the specimens; however the mechanical properties of the concrete
remained unchanged. Even with water escaping the concrete specimens, the
mechanical properties have not changed. This might be attributed to the
reorganization of the cementing matrix. Further microscopical studies need to
be carried out.
• During the second temperature range, from 300 to 600°C, there was a large
decrease in the mechanical properties, a small mass loss, and high thermal
gradients. This may lead to elevated thermal stresses and deterioration in the
concretes.

Modelling of Coupled Heat and Mass Transfers

With the aim of looking further into the knowledge of concrete behaviour at high
temperature, Sercombe J. et al. (2001) developed a model of coupled heat and
mass transfers. The model is based on the theory of the mechanics of unsaturated
porous media and simplifying assumptions in the case of concrete. The model is a
numerical formulation including a space and temporal discretization for nonlinear
calculations. It was applied in the finite elements code Cast3M developed at the
CEA (French Atomic Energy Agency) (Sercombe Galle and Ranc, 2001). The
calculation procedure of the model was used in the present study. Simulation of
the heating of 16x32 cm cylindrical specimens was carried out. Numerical results
were then compared with experimental results.

16
INTERNATIONAL JOURNAL OF HEAT AND MASS TRANSFER, ELSEVIER

Description of the model

This model is based on a porous medium macroscopic scale description,


containing three continuous mediums (Coussy, 1995; Mainguy Coussy and
Eymard, 1999): solid skeleton, liquid phase water and gas phase. With the
intention to establish a simplified model, the deformability of solid skeleton is not
taken into account. The gas phase is supposed to be only made up of vapour, the
dry air being neglected. The role of this component in the development of the pore
water pressures at high temperature is still relatively unknown (Kalifa
Mennenteau and Quenard, 2000; Tenchev et al., 2001).

Assumptions

The following assumptions are made:


o The porous medium includes three phases:
o Solid phase (index s),
o Liquid phase (index l) containing only pure water,
o Gas phase (index g) containing only steam (index v).
o The porous medium is homogeneous and isotropic,
o The deformation of the solid skeleton is not taken into account,
o The hysteresis phenomenon of the sorption curve is not taken into account.

Constitutive equations

At high temperature, the fluid mass balance must take into account both the
quantity exchanged by vaporization or condensation and the quantity of water
(liquid) released by dehydration. The mass conservation equations of water and
vapour take the following form (Sercombe J. et al. (2001)):

∂m l
= −∇w l − µ l→ v + d& (4)
∂t

∂m v
= −∇w v + µ l → v (5)
∂t

where ml and mv correspond to the mass quantities of liquid water and vapour per
unit volume dΩ. These quantities are expressed according to saturation of water Sl
and of total porosity φ by:

17
INTERNATIONAL JOURNAL OF HEAT AND MASS TRANSFER, ELSEVIER

m l = φρ l S l and, m v = φρ v (1 − S l ) (6)

In Equation (6), ρl(T) and ρv(T, pv) are respectively the water and vapour
densities, which depend on temperature T and vapour pressure pv. In Equations
(4) and (5), µl→vdt represents liquid water mass transformed into vapour per unit
volume dΩ. d& is the quantity of dependent water transformed into liquid water
per unit volume dΩ and time dt.
The water and vapour transfers are supposed to be mainly controlled by
the pore-water pressure gradients (Bazant and Thongutai, 1978; Bazant and
Kaplan, 1996), which in agreement with Darcy equation:

Ki
w i = −ρ i k ri (S l )∆p i (7)
ηl

where i = l,v (liquid or vapour), ηi, kri, pi and Ki are respectively dynamic
viscosity, relative permeability, pressure and the intrinsic permeability associated
with fluid i.
The balance between water and vapour is theoretically well adapted to
transfer in partially saturated mediums. In the capillary pores, the thermodynamic
balance between water and vapour is modified by the surface forces at the
interface between liquid phases and capillaries. With the assumption that vapour
is a perfect gas, a general relation between water pressure and vapour pressure can
be obtained by the integration of the chemical potential of water (Mainguy Coussy
and Eymard, 1999; Feraille-Frenet, 2000). By choosing a state of reference where
water and vapour are in balance with saturated vapour pressure, liquid balance –
vapour can be described by the Clapeyron’s equation (Feraille-Frenet (2000):

ρ l RT  p v 
p l = p vs (T) + ln  (8)
Mv  p vs (T) 

where Mv is molar mass of vapour and R is a perfect gas constant. The difference
of liquid water pressure pl and vapour pressure pv caused by surface forces is
defined as a capillary pressure pc (p c = p v − p l ) .
For the combination of the entropy conservation equations for a porous
medium including three phases (liquid, solid, vapour), the thermodynamic balance
can be given by (Coussy O., 1995):

18
INTERNATIONAL JOURNAL OF HEAT AND MASS TRANSFER, ELSEVIER

 dS L 
T − S ml + m& l→v  = − ∇ ⋅ q + r (9)
 dt T 

where heat flow flux q is given by Fourier analysis ( q = −λ (T ) ⋅ ∇ T ) with


material thermal conductivity λ(T).
By relating the entropy conservation equations for a porous medium with
the conservation equations of water vapour and liquid water, the following
equation of heat can be obtained (Sercombe et al., 2001):

∂T
c(S l ,d ) + (w l Cpl + w v C pv )∇T − v ∇p v = ∇q − L l→ v µ l→ v − L s→l d
w .
(10)
∂t ρv

Where:
c(Sl, d) = mds Cds + φρv Sl Cpl + φρv (1-Sl ) Cpv + (do - d) Cll is the concrete
volumetric heat capacity which depends on the saturation and on the
quantity of water released by dehydration d. do is the quantity of the initial
water. The constants Cds , Cpl , Cpv and Cll are respectively the heat
capacities of the dry solid phase (aggregates and anhydrous cement), the
liquid water, the vapour and the bound water.
(w C l pl + w v C pv )∇T is the heat transported by fluids convection, generally

negligible for less permeable materials like concrete (Bazant and Kaplan,
1996).
wv
(c) ∇p v is the heat dissipation associated with the compressibility of
ρv
the vapour phase which is also neglected.
(d) L l→ v = T (S v − S l ) is the latent heat of vaporization (thermodynamic
balance into non isothermal condition).
(e) L s→l = T (S l − S ll ) is the latent heat of dehydration (heat consumed by the
concrete for one kg of unbound water).
Thus, the first equation of the field (heat conservation) is written as:

∂T
c(S l ,d ) = ∇q − L l →v µ l→v − L s→l d& (11)
∂t

By combining the mass conservation equations for liquid and water vapour
(4–5), the Darcy’s flow equation (7) and Clapeyron’s equation (8) (Sercombe et
al., 2001), the following equation is obtained:

19
INTERNATIONAL JOURNAL OF HEAT AND MASS TRANSFER, ELSEVIER

[ρ l (T ) − ρ v (T)] φ(d ) ∂S l = ∇[D(S l , d, T ) ∆S l ] + d& (12)


∂t

where:
ρ ρ  ∂p
D(S l , d, T ) =  v k rg (S l ) + l k rl (S l ) K (d ) v is the hydrous conductivity.
 ηv ηl  ∂S l

Thus, the problem consists of finding the variables T and Sl for a porous
medium, knowing the geometry, the initial state and the external thermo-hydrous
medium. The unknown parameters are temperature fields T and liquid saturation
Sl (in function of pv).

Materials data

Data from the thermodynamic properties of water and vapour at high


temperature, Eurocode 2 and various studies (Sercombe et al.(2001) were used in
this model. This data is presented in Table 3 and Figure 12.
Initial saturation was obtained from measurements of relative humidity.
The quantities of initial bound water and interstitial water are calculated based on
the saturation level and the quantity of mixing water.

waterbound = 0,9 ⋅ M cement ⋅ 0,21 and water = waterMixing − waterbound (13)

where Mcement (kg) and waterMixing(kg) correspond respectively to the cement mass
and the mixing water mass. The quantity of released bound water is obtained
experimentally (Galle and Sercombe, 2000). The choice of the other data of the
procedure is presented elsewhere Ranc et al (2003).

Table 3 – Materials data for the calculation of coupled moisture and heat transfers.

20
INTERNATIONAL JOURNAL OF HEAT AND MASS TRANSFER, ELSEVIER

Data Value
Initial saturation Sl0 (%) = Vl0/Vpores
(0≤Sl0≤100%), where Vl0 and Vpores are initial 97
volumes of liquid and total pores respectively

Saturation Sl (%) S l = Sl (t o ) ⋅ p vap (To ) p vap (T )

RT p vap (T )
Liquide pressure pl (Pa) p l = p vsat (T ) + ρ l (T ) ln
M vap p vsat (T )

Vapour pressure pvap (Pa) (Desorption curve) p vap = Sl ⋅ p vsat (T )

Capillary pressure pc (Pa) p c = p vap − p l

Quantity of initial bound water d0 (kg/m3) 53

Water released by dehydration d (kg/m3) d = (T − 60) ⋅ d o 540

Total porosity (%) 10

Permeability related to liquid water k l = Sl (1 − (1 − S1l / m ) m ) 2

m
Permeability related to vapour k v = 1 − Sl (1 − S l ) 2 m

Intrinsic permeability to liquid water (m2) 5.10-21

Intrinsic permeability to gas (m2) 1.10-17e0.126d

Desorption curve S l = p vap p vsat (T )

Heat capacity of dry concrete (kJ/°C/kg) 0.92

Heat capacity of interstitial water (kJ/°C/kg) 3.76

Heat capacity of bound water (kJ/°C/kg) 3.76

  θ   θ 
λ inf = 1,36 − 0,136 − 0,0057   
Thermal conductivity of concrete (W/m/°C)   100  100 
for 20°C ≤ θ ≤ 1200°C   θ   θ 
λ sup = 2 − 0,2451+ 0,0107   
  100  100 

Latent heat of dehydration (kJ/kg) 2500

21
INTERNATIONAL JOURNAL OF HEAT AND MASS TRANSFER, ELSEVIER

1,20 3,5
µ eau x 10-3 µ vapeur x 10
-5

1,00 (kg/m/s) (kg/m/s)


3,0

0,80
2,5
0,60
2,0
0,40

0,20 1,5

0,00 1,0
0 100 200 300 400 500 T (°C) 600 0 100 200 300 400 500 T (°C) 600

(a) Water viscosity (Pa s) (b) Water vapour viscosity (Pa s)

1100 1000
ρ (kg/m )
3
pv apeur (MPa)
1000 100

900 10

800 1

700 0,1

600 0,01
T (°C)
500 0,001
0 50 100 150 200 250 300 350 400 10 100 T (°C) 1000

(c) Water density ρ (kg/m3) (d) Saturation vapour pressure pvsat (MPa)

1000 3000
3
ρ (kg/m ) L (kJ/kg)
2500
100

2000
10
1500
1
1000

0,1
500

0,01 0
10 100 T (°C) 1000 0 100 200 300 400 500 T (°C) 600

(e) Vapour density ρ (kg/m3) (f) Latent heat L (kJ/kg)

Figure 12 – Physical parameters in function of temperature

Numerical Results

Simulation were carried out on a 16×32 cm specimen heated on all the faces. The
heating – cooling cycles go from the ambient temperature (20°C) to a maximum
temperature of 300 or 600°C at 1°C/min (see Figure 2). Due to symmetry, a
quarter of the specimen was modelled in 2D and in an axisymmetric mode. Thus,
the geometry is a rectangle made up of quadratic elements (8 nodes) which can be
seen in Figure 13.
The initial condition is assumed that the specimen is in equilibrium at
ambient temperature. The surface temperature of the specimen increases at

22
INTERNATIONAL JOURNAL OF HEAT AND MASS TRANSFER, ELSEVIER

1°C/min (see Figure 2). The desorption curve was used in which Sl = h (Bazant
and Kaplan, 1996). At the day of testing, the concrete moisture content was 39%.
Consequently, liquid saturation Sl of 39% was imposed on the surface, and 99%
inside the concrete specimen.
y

T = f(t) and Sl=39%

T = f(t) and Sl=39%


q = 0 and ∂nSl=0
32 x

16 q = 0 and ∂nSl=0

Figure 13 – Modelled specimen, the boundary conditions (∂nSl = ∂Sl/∂n).

5300 2,00
Cp (J/kg/°C) u = 0% λ (W/m/°C)
4800 λ sup (W/m/K)
lsup (W/m/°C)
u = 1,5% 1,75
4300
u = 3,0% λ inf(W/m/K)
linf (W/m/°C)
3800 1,50
u = 6,0%
3300
u = 9,0% 1,25
2800
2300 1,00
1800
0,75
1300
800 0,50
0 200 400 T (°C) 600 0 200 400 600 800 1000 T (°C) 1200

(a) (b)

1,05
ρ(θ)/ρ(20°C)

1,00

0,95

0,90

0,85

T (°C)
0,80
0 200 400 600 800 1000 1200
(c)

Figure 14 – (a) Concrete specific heat Cp , (b) concrete thermal conductivity λ ,(c) concrete ratio of
density ρ(θ)/ρ(20°C) as a function of temperature according to the EUROCODE 2.

23
INTERNATIONAL JOURNAL OF HEAT AND MASS TRANSFER, ELSEVIER

The concrete thermal properties used are those recommended by the


Eurocode 2 (see Figure 14). The numerical results of temperature field for the
600°C cycle and temperature profiles at the specimen central ray are presented on
Figure 15.

(a) (b)

Figure 15 - (a) Isotherms in the specimen heated at 600 °C on the surface, (b) temperature profiles
in the specimen in function of the surface temperature Tsurface.

In Figure 16, the numerical results were compared with the experimental
results obtained by Kanema et al. (2005) for the 300°C and 600°C heating-cooling
cycles.

300 600
T (°C) T (°C) Tsurface
250 500
B400-600°C23c

200 400

150 300

100 200
Tsurface
50 B400-300°C21c 100

0 0
0 2 4 6 8 t (h) 10 0 5 10 15 t (h) 20

(a) T = 300°C (b) T = 600°C

Figure 16 – Experimental and numerical heating-cooling cycles at (a) 300°C and (b) 600°C.

The elevated temperature at the surface of the specimen suggests that an


external boundary condition of vapour pressure is imposed. The vapour pressure
is fixed at an initial value, i.e. 50% of the vapour pressure of saturation at 20°C.
24
INTERNATIONAL JOURNAL OF HEAT AND MASS TRANSFER, ELSEVIER

Taking into account the significant development of the vapour pressure at


saturation, this condition is equivalent to imposing a relative humidity less than
5% when the surface temperature exceeds 60°C. The initial state is characterized
by a temperature equal to 20°C and a water content equal to 3.9 % (by mass) and
corresponds to a capillary pressure equal to 5.93×10-7 Pa.
The numerical results, obtained for the maximum temperature of 300°C,
are shown in Figure 17. Figure 17 (a to d) shows the vapour pressure, liquid
pressure, capillary pressure and saturation rate at the central ray of the specimen
as a function of temperature. Figure 17a shows a peak on the vapour pressure
curves. This peak is due to a moisture clog (Bazant and Thongutai, 1978) and is
attributed to the phase change of liquid water to water vapour at high temperature.
The peak moves from the surface to the inner part of the specimen during heating
and provokes a migration of part of the water towards the centre of the specimen.
High water vapour pressure contributes to the specimen spalling.

(a) Pvapeur (b) Pliquide

25
INTERNATIONAL JOURNAL OF HEAT AND MASS TRANSFER, ELSEVIER

(c) Pcapillaire (d) Sl

Figure 17 – (a) Vapour pressure, (b) liquid pressure, (c) capillary pressure and (d) saturation rate at
the central ray of the specimen for a heating up to 300°C.

Figure 17a shows the maximum vapour pressure obtained for a surface
temperature between 250 and 300°C. The numerical results are in line with those
experimental at 300°C. Furthermore, there is a sharp decrease in the saturation
rate within the specimen at temperature between 150 and 300°C as shown in
Figure 17d. At the temperature 300°C, the specimen is almost dry with a
saturation rate Sl of near 0 %.
The migration of liquid water towards the inner part of the specimen due
to moisture clog delays the increase of temperature within the specimen. This
phenomenon has the consequence of increasing the thermal gradient in the
concrete element, then the increase of thermal stresses. At the temperature 250°C,
the liquid water mass wedged in the centre of the specimen changes phase (from
liquid to vapour) (Figures 17a, 17b and 17d). The main variations of liquid
pressure (Figure 17b) and capillary pressure (Figure 17c) within the concrete
element took place at temperature between 200 and 300°C. This numerical result
compared well with the experimental observed spalling phenomenon. After the
temperature of 300°C, the concrete dehydration is almost completed. Vapour
pressure and saturation rate decrease to zero.
Figure 18 (a and b) shows the numerical results of the temperature
difference, ∆T, as a function of the temperature at the centre of the specimen (a)
and as a function of the temperature at the surface of the specimen (b), for a

26
INTERNATIONAL JOURNAL OF HEAT AND MASS TRANSFER, ELSEVIER

heating up to 600°C. The shape of the curves shows the effect of the water content
on the heat transfer. The peak value increases with the concrete water content.
This is due to the fact that water increases the concrete thermal heat capacity
resulting in high thermal gradient during heating.
140 140
∆ t (°C) ∆ t (°C)
120 120

100 100

80 80

60 60

40 T - u = 0,0% Thyd - Cd sup 40 T - u = 0,0% THyd - Cd sup


T - u = 3,0% Thyd - Cd inf T - u = 3,0% THyd - Cd inf
20 20
T - u = 6,0% T - u = 9,0% T - u = 6,0% T - u = 9,0%
0 0
0 100 200 300 400 Tcentre
500(°C) 600 0 100 200 300 400 Tsurface
500 (°C) 600

(a) (b)

Figure 18 – Temperature difference (∆T) versus the temperature at the centre and at the surface of
the specimen.

120 160

100
∆ T (°C) Cd sup 140 ∆ T (°C) Cd sup
Cd inf 120 Cd inf
80
B400-300°C21c 100 B400-600°C23c
60 80
40 60

20 40
20
0
0
-20 -20
-40 -40
0 1 2 3 4 5 6 7 t 8(h) 9 0 2 4 6 8 10 12 14 t (h) 16

(a) T = 300°C (b) T = 600°C

Figure 19 – Temperature difference (∆T) for the heating cycles up to 300°C (a) and 600°C (b)

Figure 19 shows the comparison between the numerical results and the
experimental results for the 300°C and 600°C heating cycles. In both cases, the
lower thermal conductivity (λinf) of Eurocode 2 compared satisfactory with the
experimental results. Therefore, as shown in Figure 19, the results are clearly in
accordance with the choice of the Eurocode 2 lower conductivity (λinf)
recommended in the French Appendix.

27
INTERNATIONAL JOURNAL OF HEAT AND MASS TRANSFER, ELSEVIER

9% 9%
PM (%) PM (%)
8% 8%
7% 7%
6% 6%
5% 5%
4% 4%
PM - Cd sup
3% 3% PM - Cd sup
PM - Cd inf
2% 2% PM - Cd inf

1% %-b400-300°C
1% %-b400-600°C
0% 0%
0 100 200 300 400 Tsurface
500 (°C) 600 0 100 200 300 400 Tsurface
500 (°C) 600

(a) T = 300°C (b) T = 600°C

Figure 20 – Concrete mass loss (%) as a function of the temperature at the surface (Tsurface °C) of
the specimen for the heating cycles up to (a) 300°C and up to (b) 600°C

Figure 20 shows the concrete mass loss as a function of the temperature at


the surface of the specimen for heating up to 300°C and 600°C. The % mass loss
PM is the ratio of the difference between the specimen initial mass and the
specimen current mass divided by the specimen initial mass (PM = 100 (mi –
mc)/mi). The difference between the specimen initial mass and the specimen
current mass is the absolute mass loss and is supposed to be equal to water mass
escaping during heating. The main mass loss occurs at the specimen surface
temperature from 150 to 300°C, which corresponds to CSH, hydrated calcium
silicate, decomposition (CaO-SiO2-H2O C2S and C3S + H2O). It is known that
the endothermic decomposition of the CaSO42H2O gypsum and the dehydration
of CSH respectively take place towards 130°-170°C and towards 180°-300°C
respectively. At the temperature stage between 450 and 600°C, there is portlandite
decomposition (Ca(OH)2 CaO + H2O).
Figure 21 shows the mass loss and temperature difference data. The figure
shows that the experimental and numerical results compared favourably. The
thermo-hydrous model describes the coupled heat and mass transfers in the
concrete during heating. It is also observed that the slope change of the mass loss
curve between 4 and 5 hours of heating corresponds to the peak of the temperature
difference curve. For the heating cycle up to 600°C, the peak of the experimental
temperature difference is higher than that obtained numerically, as shown in
Figure 21(b). This is attributed to heat absorption due to the beginning of concrete
decarbonation (CaCO3 CaO + CO2).

28
INTERNATIONAL JOURNAL OF HEAT AND MASS TRANSFER, ELSEVIER

160 9% 160 9%
∆ T (°C) PM (%) ∆ T (°C)
140 8% 140 PM (%)
8%
120 7% 120 7%
100 100
6% 6%
80 B400-300°C21c 80
5% 5%
60 60
Diff. T - Cd inf 4% 4%
40 40
3% 3%
20 %-b400-300°C 20 B400-600°C74i Diff. T - Cd inf
0 2% 0 2%
PM - Cd inf 1% %-b400-600°C PM - Cd inf 1%
-20 -20
-40 0% -40 0%
0 2 4 6 8 10 t12
(h) 14 0 2 4 6 8 10 t12
(h) 14

(a) T = 300°C (b) T = 600°C

Figure 21 – Temperature difference between the surface and the centre of the specimen (∆T) and
concrete mass loss (%) as a function of time for heating up to (a) 300°C and (b) 600°C.

Figure 22 shows the temperature difference between the centre and the
surface of the specimen as a function of the surface temperature. The figure also
shows that the experimental mass loss data compared well with that of numerical.
The temperature difference data also compared well with that of numerical up to
the temperature of 350°C. At a temperature range from 350°C to 500°C, the
experimental temperature difference curve was higher than that of the numerical.
This might be attributed to concrete decarbonation as mentioned earlier.
Experimental observation indicated that spalling occurred during heating when
the specimen surface temperature was about 300°C. The numerical results show
that the vapour pressure, the concrete mass loss and the thermal gradient within
the specimen are maximum at a temperature of about 300°C. The good correlation
between the experimental and numerical (up to 350°C) allows the determination
of the temperature range at which concrete spalling takes place.
160 10%
∆ T (°C) PM (%)
140 9%

120 8%

100 7%

80 6%

60 5%

40 4%

20 3%
B400-600°C74i Diff. T - Cd inf
0 2%

-20 PM - Cd sup % -b400-600°C 1%

-40 0%
0 100 200 300 400 500 T (°C) 600

29
INTERNATIONAL JOURNAL OF HEAT AND MASS TRANSFER, ELSEVIER

Figure 22 – Evolution of temperature difference between the surface and the centre of the
specimen (∆T) and concrete mass loss (%) in function of the temperature at the centre (Tcentre °C).

The influence of the mechanical damage is not considered in the model.


Experimental studies show an increase in concrete permeability and a decrease in
the concrete mechanical properties at a temperature range between 300 and 600°C
(Kanema et al., 2005). These phenomena have consequences on the heat and mass
transfers within the concrete element. Thus, it would be beneficial to include the
mechanical effects in the model.

Conclusions

During the first temperature range from room temperature to 300°C, the following
were observed: a) mechanical strength and permeability did not vary significantly,
b) a weak decrease in the modulus of elasticity and c) a high mass loss between
150 and 300°C. The properties of the heated concrete at 150°C remained very
close to that of the non-heated concrete. For concrete mixtures with high cement
content there was an improvement in the compressive strength. Between 150 and
300°C, a large part of water escaped (interstitial water and hydrates bound water)
but the mechanical properties of the specimens remained unchanged. The water
transfer may be accompanied by a reorganization of the cementing matrix at the
microstructural level making it possible to preserve the mechanical strength.
For the second temperature range from 300 to 600°C, there was a) a major
decrease in the mechanical properties b) a small mass loss, c) a large thermal
gradient leading to elevated thermal stresses and deterioration in the concrete
matrix.
This research validates the thermo-hydrous model and provides a better
understanding of the behaviour of concrete subjected to high temperature. The
thermo-hydrous calculations show a good agreement between the numerical and
experimental results. For a heating cycle from 20 to 300°C, the thermo-hydrous
behaviour such as temperature gradient, mass loss can be simulated accurately by
knowing the thermal and hydrous properties of the concrete. Between 300 and
600°C, the thermo-hydrous model remains relevant even if it does not consider
the mechanical damage of the concrete system. The numerical results showed that

30
INTERNATIONAL JOURNAL OF HEAT AND MASS TRANSFER, ELSEVIER

the lower thermal conductivity curve of Eurocode 2 (recommended in the French


Appendix) is valid for thermal characterization of the studied concrete.
Results showed that spalling occurred at a temperature about 300°C for the
studied specimen. Experimental and numerical results showed that vapour
pressure, concrete mass loss and the thermal gradient within the specimen (then
the thermal stresses) are maximum at a temperature about 300°C.

REFERENCES
Bazant ZP, Thongutai W (1978) Pore pressure and drying of concrete at high temperature. ASCE
Journal of Engineering Mechanics 104:1058-1080.
Bazant ZP, Kaplan MF (1996) Concrete at high temperatures - Material properties and
mathematical models. Longman House, Burnt Mill, England.
Brite (1999), Final report of the BRITE EURAM III-BRPR-CT95-0065 ªHITECOº research
programme, Understanding and industrial Application of High Performance Concrete in
High Temperature Environment, sponsored by European Community,
http://cordis.europa.eu/search/index.cfm?fuseaction=proj.simpledocument&PJ_RCN=1450878&CFID=9232918&CFTOK
EN=31031765.
CEN TC 250 (2004) Eurocode 2 - Design of concrete structures Part 1-2 General rules - Structural
fire design.
Collectif Presses de l'ENPC (2005) Synthèse des travaux du projet national BHP 2000 sur les
bétons à hautes performances (in french), Eyrolles Editions, Paris, ISBN : 2-85978-408-X.
Coussy O (1995) Mechanics of porous continua. John Wiley & Sons Ltd, Chichester, England.
Feraille-Frenet A (2000) Le rôle de l’eau dans le comportement à haute température des bétons.
Thèse de doctorat, Ecole Nationale de Ponts et Chaussées, Paris, France.
Galle C, Sercombe J, Pin M, Arcier G, Bouniol P (2000) Behaviour of high performance concrete
under high temperature (60-450°C) for surface long-term storage: thermo-hydro-mechanical
residual properties. Acts de la conference international Materials Research Symposium (MRS),
Sydney, Australia.
Galle C, Sercombe J (2000) Permeability and pore structure evolution of silica-calcareous and
hematite high-strength concretes submitted to high temperatures. Materials and Structures 34:619-
628.
Gawin D, Majorana CE, Pesavento F, Schrefler BA (2001) Modelling thermo-hydro-mechanical
behaviour of high performances concrete in high temperature environments. Acts de la conference
international Fracture Mechanics of Concrete Structures (FRANCOS’4), pp. 199-206, Cachan,
France.
ISO (1993) Guide to the expression of uncertainty in measurement, prepared by ISO Technical
Advisory Group 4 (TAG 4), Working Group 3 (WG 3), October 1993. ISO/TAG 4 has as its
sponsors the BIPM, IEC, IFCC (International Federation of Clinical Chemistry, ISO, IUPAC

31
INTERNATIONAL JOURNAL OF HEAT AND MASS TRANSFER, ELSEVIER

(International Union of Pure and Applied Chemistry, IUPAP (International Union of Pure and
Applied Physics), and OIML, the Guide is published by ISO in the name of all seven
organizations.
Kalifa P., Mennenteau F.D., Quenard D. (2000) Spalling and pore pressure in HSC at high
temperatures, Cement and Concrete Research, 30:1-13.
Kanema M, Noumowe A, Gallias J-L, Cabrillac R (2005) Influence of the mix parameters and
microstructures on the behaviour of concrete at high temperature. 18th international conference on
Structural Mechanics in Reactor Technology (SMiRT 18), SMiRT 18-H07_7, Beijing, China.
Kessel W. (1998) European and international standards for statements of uncertainty, Enginnering
Science and Education Journal, 7(5):201-207.
Kessel W. (1999) ISO/BIPM Guide: Uncertainty of measurement, Metrodata,
http://www.metrodata.de/papers/resistor_en.pdf.
Mainguy M, Coussy O, Eymard R (1999) Modélisation des transferts hydriques isothermes en
milieu poreux. Application au séchage des matériaux à base de ciment. Etudes et Recherche des
Laboratoires des Ponts et Chaussées, Série Ouvrage d’Art, n°32, Paris, France.
Mounajed G., Zakhem D (2001) Modélisation du comportement thermo-hydro-mécaniques des
bétons à hautes températures. Rapport d’étude BHP2000, Paris, France.
Nechnech W (2000) Contribution à l'étude numérique du comportement du béton et des structures
en béton armé soumises à des sollicitations thermiques et mécaniques couplées. Thèse de doctorat
de l’INSA, Lyon, France.
Noumowé NA, Clastres P, Debicki G, Bolvin M (1994) High temperature effect on high
performance concrete: strength and porosity. 3rd CANMET/ACI international conference on
Durability of Concrete, Nice, France.
Noumowe A, Ranc G, Hochet C (2003) Moisture migration and thermo-mechanical behaviour of
concrete at high temperature up to 310°C. 17th international conference on Structural Mechanics
in Reactor Technology (SMiRT 17), Prague, Czech Republic.
Noumowe A, Debicki G (2002) Effect of elevated temperature from 200 to 600°C on the
permeability of high performance concrete. 6th international symposium on Utilization of High
Strength/Performance Concrete, vol. 1, Leipzig, Germany.
Noumowe A (1995) Effet des hautes températures (20-600°C) sur le béton. Cas particulier du
béton à hautes performances. Thèse de doctorat de l’INSA, Lyon, France.
Ranc G, Sercombe J, Rodrigues S (2003) Comportement à haute température du béton de
structure. Impact de la fissuration sur les transferts hydriques. Revue française de génie civil,
7(4) :397-424.
RILEM TC 129-MHT (1995) Test methods for mechanical properties of concrete at high
temperatures, Part 1: Introduction, Part 2: Stress-strain relation, Part 3: Compressive strength for
service and accident conditions, Materials and Structures, 28(181):410-414.
Sercombe J, Galle C, Ranc G (2001) Modélisation du comportement du béton à haute température:
Transferts des fluides et de chaleur et déformations pendant les transitoires thermiques. Note
Technique SCCME n°81, CEA Saclay, France. Cited in the reference Ranc G. et al. (2003).

32
INTERNATIONAL JOURNAL OF HEAT AND MASS TRANSFER, ELSEVIER

Tenchev R.T., Li L.Y., Purkiss J.A., Khalafallah B.H. (2001) Finite element analysis of coupled
heat and mass transfer in concrete when it is in fire, Magazine of Concrete Research, 53(2):117-
125.

33

View publication stats

You might also like