Vous êtes sur la page 1sur 9

Materials and Structures/Matriaux et Constructions, Vol.

35, December 2002, pp 614-622

Determination of isothermal unsaturated capillary flow


in high performance cement mortars by NMR imaging
K. Hazrati1,2, L. Pel3, J. Marchand 1, K. Kopinga3 and M. Pigeon1
(1) CRIB-Department of Civil Engineering, Laval University, Quebec City, Quebec, Canada, G1K 7P4
(2) Grace Construction Products, 62, Whittemore Avenue, Cambridge, MA 02140, USA
(3) Department of Physics, Eindhoven University of Technology, P.O. Box 513, 5600 MB Eindhoven, the Netherlands

A B S T R A C T

R S U M

The time-dependent liquid water distribution in


cement mortar mixtures during water absorption was
determined using a proton nuclear magnetic resonance
imaging (NMRI) technique. The variation of the material
water diffusion coefficient with the water content was
established on the basis of these results. Eight different
mortar mixtures were prepared and tested. Test variables
included water/binder ratio (0.60, 0.40 and 0.25), type of
cement (an ASTM Type 1 cement and a commercial
white cement) and the use of silica fume. In addition, the
effect of drying treatments on the water transport properties was also investigated. The absorption of water in dry
mortars is found to be well described by a non-linear diffusion equation based on the extended Darcy's law of continuum mechanics. Results also confirm that the macroscopic water diffusion coefficient (D()) of mortar is
strongly dependent on the material water content. The
relationship between the two parameters can be modeled
by two exponential equations. Test results finally show
that a reduction of the mortar water/binder ratio and the
use of silica fume tend to signif icantly decrease the
absorption of water. This phenomenon is well described
by the water penetration coefficient max.

Lvolution de la distribution de la teneur en eau dans


divers mlanges de mortier de ciment durant un essai
dabsorption capillaire a t tudie laide dune technique dimagerie par rsonance magntique nuclaire du
proton. Les rsultats ainsi obtenus ont par la suite permis
dvaluer la variation du coefficient de transport de lhumidit en fonction de la teneur en eau de ces diffrents matriaux. Huit diffrents mlanges de mortier ont t tests.
Les variables tudies concernaient le rapport eau/liant
(0.60, 0.40 et 0.25), le type de ciment (un ciment portland ordinaire de type I selon lASTM et un ciment
blanc), lajout de fume de silice et le conditionnement des
prouvettes avant lessai (type de schage). Les rsultats
dmontrent que le transport de leau lors dun essai
dabsorption capillaire peut tre dcrit par une quation de
diffusion non-linaire. La variation du coefficient de transport de lhumidit peut tre modlise laide dune combinaison de deux quations exponentielles. Les donnes
exprimentales indiquent galement que la rduction du
rapport eau/liant et lajout de fume de silice contribuent
ralentir le transport de leau par absorption. Ces phnomnes sont correctement dcrits par le coefficient de pntration de leau max.

1. INTRODUCTION
Over the past decades, it has been clearly established
that cement-based materials are moisture-sensitive
solids. For instance, it is now well known that the
mechanical properties of concrete are strongly inf luenced by its moisture state [1]. Moisture also has a predominant influence of the volume stability of cementbased materials. Creep and drying shrinkage are both
moisture-dependent phenomena [2].

Numerous investigations have also indicated that


moisture can have a marked influence on the durability
of concrete structures. Numerous deterioration phenomena, such as frost damage, ion ingress, reinforcingsteel corrosion and alkali-silica reactions, are intensified
by high water contents [3-5]. Furthermore, the performance of protective measures, such as the application of
paints, adhesives or sealants, is clearly dependent on the
water content of the concrete surface. Knowledge of the
water content and its distribution in the pore system is

Editorial Note
Laval University (Canada) is a RILEM Titular Member.
Prof. J. Marchand was awarded the 2000 Robert L'Hermite Medal. He is Editor in Chief for Concrete Science and Engineering and Associate Editor for
Materials and Strutures. He participates in RILEM TC 186-ISA 'Internal Sulfate attack'.
1359-5997/02 RILEM

614

Hazrati, Pel, Marchand, Kopinga, Pigeon

thus of paramount importance for the development of


predictive long-term durability models.
In order to provide a quantitative basis for understanding the mechanisms of water transfer in unsaturated
porous materials, it is essential to have reliable experimental methods for measuring local water content.
Given the importance of the problem, a great deal of
effort has been spent, over the past decade, to develop
new experimental techniques or to refine existing ones.
For instance, numerous authors have tried to
improve the accuracy of direct relative-humidity assessment methods [6, 7]. These techniques consist in placing
capacitance-sensitive probes in small holes previously
cast or drilled in the concrete element. The main advantage of these techniques lies in their relative simplicity.
Their reliability is, however, quite limited. The conversion of the actual relative humidity (measured by the
probes) into water content can only be made on the basis
of an equilibrium sorption isotherm. The latter is sensitive to the presence of salts in the concrete pore solution
[5]. Furthermore, direct relative humidity measurements
can only be performed at discrete intervals and their
validity is restricted to low water contents.
Other authors have relied on the traditional gravimetric method to assess the distribution of evaporable
(or free) water in concrete [6, 8]. The technique involves
the splitting and the weighting of concrete slices in wet
and oven-dry states. The procedure is straightforward
but special care should be taken during the weight measurements to avoid undesired evaporation of water during the manipulations [6]. The method being destructive, measurements have to be made on a series of
samples. Furthermore, the method is plagued by a very
coarse spatial resolution.
More recently, alternative non-destructive methods
have been tested on cement-based materials. For instance,
some authors have tried to apply dielectric methods to the
determination of liquid water content in concrete [9]. The
fact that these methods are extremely sensitive to any variation of the ionic concentration of the concrete pore solution limits their application in practice (at least for moisture content determinations) [9].
Other techniques such as neutron radiography
(which is based on the interaction of neutrons with the
material nuclei) and gamma densimetry (which is based
on the absorption of radiation by the electrons of the
material) have successfully been tested for concrete.
These two attenuation methods are relatively sensitive to
water but, in both cases, the determination of the water
content of the material is indirect [10-12].
Unlike the last two non-destructive methods, nuclear
magnetic resonance (NMR) imaging has the advantage
of being selective of a particular nucleus, e.g. the hydrogen nucleus for the study of water. The NMR signal is
directly proportional to the amount of hydrogen nuclei
in the free liquid water present in the material [13-16].
However, serious complications occur if the material
under investigation contains large amounts of paramagnetic ions, e.g. Fe or Mn. This is the case of ordinary
Portland cement, which contains approximately 3% of

Fe2O3 (in mass). First, the presence of paramagnetic ions


causes a drastic decrease in the so-called transverse relaxation. Secondly, local magnetic field gradients will be
induced, which broaden the resonance line and thereby
limit the spatial resolution. Therefore no longer standard
NMR equipment can be used to study these materials.
However, recently Pel [13] reported the development
of a specially designed NMR imaging scanner capable of
quantitative measurements of water profiles in building
masonry materials, such as fired-clay bricks, containing
relatively large amounts of paramagnetic impurities
(about 6% Fe ions). He obtained one-dimensional water
content distributions with a submillimeter spatial resolution.
Over the past decade, NMR imaging has been predominantly used for the study of water transfer in highly
porous inorganic building materials such as brick
masonry [13] or building stones. The application of the
technique to water-content measurements in hydrated
cement systems has, so far, been limited to high
water/cement ratio mortars [16] and aerated concrete
elements [17]. Very little work has been done on lowwater/cement ratio mixtures.
The objective of this project was to investigate the
mechanisms of the isothermal water transport in unsaturated high performance mortars. The one-dimensional
time-dependent liquid water distribution and the water
content dependent diffusion coefficient were experimentally determined by means of H1 nuclear magnetic
resonance imaging (NMRI). Both ordinary and highperformance mortars were tested. The influence of various parameters such as type of cement, water/binder
ratio, use of silica fume and drying pretreatment were
investigated.

2. MATERIALS AND SAMPLE PREPARATION


2.1 Materials and mixture characteristics
Eight different mortar mixtures were prepared. Test
variables included water/binder ratio (0.60, 0.40 and
0.25), type of cement (an ASTM Type 1 cement and a
commercial white cement) and the use of silica fume.
The use of silica fume (6% of the total mass of binder)
was limited to the 0.25-water/binder ratio mixtures. The
specimens containing silica fume are referred to in the
text with the letters SF. The Fe2O3 contents of the normal Portland cement and the white cement were 3%
and 0.3% respectively. In the text, the mixtures made
with ASTM Type I Portland cement are labeled by the
letter I and those prepared with the white cement are
identified by the letter W. Information on the chemical
and mineralogical compositions of the various binders is
given in Table 1.
All mortars were made with demineralized water and
the sand volume fraction was kept at 50%. The sand was
a standardized crushed siliceous sand (Ottawa silica) with
a density of 2.60. Information on the chemical and physical properties of the sand is given in Tables 2 and 3. For

615

Materials and Structures/Matriaux et Constructions, Vol. 35, December 2002


Table 3 - Grain size distribution of the sand

Table 1 - Chemical and mineralogical compositions of the


cements on a mass basis
Constituents

ASTM I
(%)

White
(%)

Silica Fume
(%)

SiO2

20.72

24.29

94.86

Al2O3

4.15

1.71

0.27

TiO2

0.20

0.07

0.01

P2O5

0.23

0.13

Fe2O3

3.02

0.32

0.80

CaO

62.15

68.60

0.33

SrO

0.26

0.13

MgO

2.53

0.54

0.06

0.03

Na2O

0.19

0.14

0.11

K2O

0.90

0.03

0.36

SO2

3.43

2.11

Loss on ignition

2.36
4616

% Passing

20

100.0

30

100

40

72

50

30

60

17

100

Pan

Table 4 - Mixture characteristics

0.26

Mn2O3

Blaine (cm2/g)

Sieve (ASTM E-11)

1.13

Mixture

W/C

Cement Silica fume Water


Sand Superplast.
(kg/m3) (kg/m3) (kg/m3) (kg/m3)
(l/m3)

I60*

0.60

545

327

1325

I40*

0.40

697

279

1325

I25*

0.25

881

220

1325

61

I25SF*

0.25

828

53

220

1325

61

*or W60, W40, W25 and W25SF for mixtures made with the white cement.

4020

Mineralogical composition (Bogue)

Table 5 - Compressive strength test results (28 days)

C3S

54

77

C2S

19

12

Mixture

Compressive strength* (MPa)

C3A

I60

41.5

I40

71.0

I25

92.6

I25SF

118.0

W60

33.7

W40

81.6

C4AF

Table 2 - Chemical composition and physical properties


of the sand
Constituents

% Weight

W25

114.3

SiO2

99.709

W25SF

146.6

Fe2O3

0.028

Al2O3

0.122

CaO

0.008

TiO2

0.012

MgO

0.005

K2 O

0.015

Na2O

0.003

Absorption (%)

0.2

Density

2.6

all the 0.25 mixtures, a melanine-based superplasticizer


(Melment L10) was used at a dosage of 2.5% of dry
material per unit mass of cement.
All mixtures were prepared in a high-speed mixer
placed under vacuum (10 mbar) to prevent, as much as
possible, the formation of air voids during mixing.
Rectangular bars 200 x 35 x 35 mm were then cast in
plastic molds, cured for 24 hours under damp cloths,

demolded and then water cured in a Ca(OH)2-saturated


solution (kept at room temperature) for a minimum
period of two years. The characteristics of all mixtures
are given in Table 4.
The average compressive strength of each mixture, as
determined from 3 cubes after 28 days of curing, is given
in Table 5. All measurements were done according to
the specifications of ASTM C 109.
The degree of hydration of each mixture was determined after one year of immersion in the Ca(OH)2-saturated water. Non-evaporable water content or the
amount of chemically bound water was assessed by measuring the loss of mass between 105C and 1000C.
Corrections were made for the loss on ignition of the
sand and of the unhydrated cement. The amount of
chemically bound water at full hydration ( = 100%) was
assumed to be 25% of the cement mass for all mixes
except for those containing silica fume where the value
was fixed at 24%. At full hydration, silica fume pastes
tend to bind less water than OPC pastes [18].

616

Hazrati, Pel, Marchand, Kopinga, Pigeon

period of 4 months. The propan-2-ol was renewed regularly during the storage time. Then, the propan-2-ol
saturated specimens were oven-dried for 11 days at 40C
and then placed in a second oven kept at 105C for
another 3 days. In the text, the oven-dried and the
propan-2-ol dried specimens are labeled with the letters
O and P respectively.
Table 7 presents the results of the dry bulk density of
each mixture calculated from the dry mass and the overall physical dimensions. The results in this table correspond to the average dry bulk density of 10 prisms (35 x
35 x 200 mm).

Table 6 - Degree of hydration and cement gel content


Mixture

Wn
(g/g of
cement)

Degree of
hydration
(%)

Wgel
(g/g
sample)

I60

0.21

85

0.29

I40

0.18

72

0.29

I25

0.13

50

0.24

I25SF

0.13

58

0.26

W60

0.22

90

0.31

W40

0.19

76

0.31

W25

0.12

49

0.23

W25SF

0.12

53

0.24

3. EXPERIMENTAL PROCEDURE

The total cement gel content by mass of a given mixture can be estimated according to Equation (1):
W gel =

(1 + Wno ) f c
W
1 + Wno f c

(1)

where W no , is the constant equal to 0.25 for cement mixtures and 0.24 for silica fume mixtures, , the degree of
hydration, fc, the cement content and W the total mass of
the sample.
The non-evaporable water content (Wn ), the calculated degree of hydration values () and the cement gel
contents (Wgel) calculated according to Equation (1) are
given in Table 6. As can be seen, the results show, at
constant time, an increase in the degree of hydration as
the water to cement ratio is increased. The increase in
the water/cement ratio results in larger space for hydration products to grow.

2.2 Drying pretreatments


After the curing period, 20 x 40 mm cylinders were
drilled from the central part of the 200 x 35 x 35 mm
bars with a diamond corer. The mortar cylinders were
then dried using two different procedures. A first series
of specimens was oven dried at 105C for 30 days. The
second series under went a liquid replacement by
immersing the specimens in a generous amount of
reagent grade anhydrous propan-2-ol for a minimum
Table 7 - Dry bulk density of the mixtures
Mixture

Dry bulk density (kg/m3)

W60

1953

W40

2187

W25

2308

W25SF

2273

I60

2029

I40

2106

I25

2285

I25SF

2244

In a NMR experiment, the magnetic moments of


the hydrogen nuclei are manipulated by suitably chosen
alternating radio frequency fields, resulting in a so called
spin-echo signal. The amplitude of this spin-echo signal
is proportional to the amount of hydrogen nuclei excited
by the radio frequency field. The resonance condition
for the nuclei is given by f = Bo. Here f is the frequency
of the alternating field, is the gyromagnetic ratio of the
nuclei ( = 42.58 MHz/T for 1H), and Bo is the magnitude of the externally applied static magnetic f ield.
Because of the resonance condition, the method can be
made sensitive to only hydrogen and therefore to water.
When a known magnetic gradient is applied, the resonance condition will depend on the spatial position of
the nuclei. By varying the frequency f, the moisture distribution can be measured without moving the sample.
An extensive introduction and the basic concepts of
NMR techniques can be found in the literature [19, 20].
For the experiments described in this study, a NMR
scanner built at the Eindhoven University of Technology
was used. This instrument was especially designed for
quantitative measurements of moisture in porous materials with short transverse relaxation (T2) times (unlike
standard Magnetic Resonance Imaging (MRI), which is
generally used in a qualitative way). An extensive
description of the NMR apparatus used can be found in
references [13] and [21].
This NMR machine operates at a main field of 0.8
Tesla corresponding to a frequency of 34 MHz. This
field is provided by a conventional water-cooled iron
cored electromagnet of which the poles are 50 mm apart.
The NMR setup for the water absorption experiments is
presented in Fig. 1. A coil, which forms part of a tuned
LC circuit, is placed around the sample for creating and
receiving the radio frequency fields during the NMR
experiments. A specially designed Faraday shield is
placed between the coil and the sample to suppress the
effects of the changes of the dielectric permittivity by
variations of the moisture content, thereby making the
NMR experiments quantitative.
In the set-up an initial dry mortar cylinder of 20 mm
diameter and a length of 40 mm is put in contact with
water. The reduced height of the sample allows to
neglect gravity effects. The water level is maintained

617

Materials and Structures/Matriaux et Constructions, Vol. 35, December 2002

Fig. 1 - Experimental set-up of the NMR equipment for determining water distributions in a mortar sample during water absorption.

constant during absorption by communicating vessels,


thereby allowing the sample to freely absorb water.
During absorption, the water level is maintained at 1
mm above the bottom surface of the sample.
A well-defined magnetic field gradient of 0.3 T/m is
applied by the gradient coils, offering a one-dimensional
resolution of 1.0 mm. A complete moisture profile can
be measured by performing spin-echo experiments at
various fixed frequencies. In this study, straightforward
Hahn spin-echo sequences (90x--180x) were used.
Measuring an entire moisture profile over 25 mm with
an accuracy of 1% and an equally spaced grid of 0.15
mm takes typically about 2 min. This time is strongly
dependent on the spin-lattice time (T1) of the material
under investigation. The spin-lattice time determines
the repetition time of the spin-echo experiment. For the
mortar samples used in this study, the spin-lattice time is
of the order of 3 ms. During the measurements, a time
stamp is added to each point of the experimental water
profiles. The maximum length of a NMR water absorption tests was 14 hours. All NMR experiments were at
room temperature (23C).
In order to convert the NMR signal amplitude into
absolute water content values, a calibration had to be
performed using a gravimetric method. Typical calibration results are given in Fig. 2. As can been seen from
this figure, a perfect linear behavior is found between the
cumulative water signal obtained by NMR (by the integration of the NMR water profiles) and the mass of
water absorbed by the sample obtained gravimetrically.
The scatter of the data is comparable to the observed signal to noise ratio ( 2%). The slope of the lines corresponds to the calibration coefficient, , of the material.
The absolute water content is then calculated from the
following equation:
=

NMR
w A

Fig. 2 - Typical calibration curves of the NMR signal for two


types of mortar representing the integrated water profiles for
samples with diameter of 20 mm plotted against the corresponding mass of water, determined gravimetrically.

In this report, the values of the water content are


expressed in m3 of water per m3of dry material. This can
be converted into m3 of water per kg of dry material
using the dry bulk density results given in Table 7. The
values of the calibration coefficient are given in Table 8
for all the materials studied in this project. The samples
made with ASTM Type I Portland cement (I) have
lower calibration coeff icient values and thus lower
NMR signals than the ones made with white cement
(W). This phenomenon is explained by a higher content
of Fe ions in the ASTM Type I Portland cement.

4. THEORETICAL CONSIDERATIONS
The liquid water transport in an unsaturated porous
material can be described by the extended Darcy's first
law [22]:
r
U = K( )( )
(3)
where U is the flux of water in the liquid phase, K is the
hydraulic conductivity and is the hydraulic capillary
pressure originating from the curved menisci. Both K
and are a function of the volumetric water content .
A more fundamental basis for this law can be given on
the basis of volume averaging techniques. In this equation, the effects of gravity on the capillary rise of water
are neglected. Combining Equation (3) with the conser-

(2)

where is the absolute moisture content (m3/m3), NMR


the intensity of the NMR signal, the calibration coefficient (g - 1), w the water density (considered to be
always 0.001 g/mm3) and A the cross section of the sample (mm2).

618

Table 8 - Calibration constants of the NMR signal


Mixture

Calibration (g-1)
Oven-dried (O)

Propan-2-ol treated (P)

W60

17.662

13.415

W40

14.275

13.676

W25

14.883

12.500

W25SF

14.999

12.700

I60

14.354

I40

13.525

I25

12.480

I25SF

11.842

Hazrati, Pel, Marchand, Kopinga, Pigeon

vation of mass, the liquid water transport in an unsaturated material can be described by:

(4)
= ( K( )( ))
t
where t is the time.
In general, the potential is difficult to quantify in a
specimen. Instead, by a change of variables, the equation
can be transformed into an equation where the macroscopic moisture content behaves in a diffusive manner:

( ( ) ( ))

= D
t

(5)

The material coefficient D is the isothermal liquid


diffusivity:
d
D =K
d

()

()

(6)

5. RESULTS AND DISCUSSION

The diffusion coefficient D completely characterizes


the response of a material as a function of the water content.
During water absorption the following boundary
conditions can be applied: = 0 at x = 0 for t = 0 and
= cap at x = 0 for t 0. In this study, cap is the capillary
water content corresponding to the maximum water
content at the inf low face of the tested sample under
atmospheric conditions and 0 is the initial uniform
water content of the sample. In our experiments, the
mortars were initially completely dry and 0 was thus
considered to be equal to zero.
The partial differential Equation (3) is nonlinear and
has no general analytical solution. By contrast, for onedimensional transport, the Boltzmann transformation
= xt-1/2 may be applied to Equation (6) reducing it to an
ordinary differential equation [22-26]:

d d
d
=
D

2 d d
d

()

(7)

With the following boundary conditions for the


absorption: = cap at = 0 and = 0 at . Hence
during water absorption all profiles are related by a simple t1/2 scaling. In this study, the function is determined experimentally using the water profiles data (x,
t), obtained by the NMR technique. The water diffusivity coefficient can now be determined from the experimental data by the integration of Equation (7):
1 d
D =
2 d

()

Fig. 3 - Typical water content profiles measured during the capillary absorption of water by the W40O sample. The times are
given as an indication of the elapse time. In the inset a schematic
diagram of the experiment is given.

()d

Typical water concentration profiles measured at different times are given in Fig. 3 for the W40O sample. As
shown in this figure, rather steep wetting fronts were
developed when water was absorbed into the initially dry
mortars. Because of the influence of the free water bath
on the NMR signal in the lower part of the mortar sample, no accurate results could be measured for x 3mm.
The corresponding Boltzmann transformed moisture
profiles are shown in Fig. 4. As can be seen, for each
specimen, the profiles generally collapse into a single
master curve of versus . Indicating that the water
transport can be described by a non-linear diffusion
equation and that the water diffusion coefficient is not
dependent on the position [29].
For all materials investigated, a sharp wetting front is
observed. Hence the so-called water penetration coefficient max (also identified by the letter B in some publications) [30, 31], i.e. the position of the wetting front
divided by t1/2, is a well-defined parameter for these
materials.
In Table 9, the water penetration coefficient, max,
corresponding to the leading edge of the fronts in Fig. 4

(8)

In the literature, the liquid diffusivity function D()


for soils and a number of construction materials is commonly described by the equation D() = D0exp() where
D0 and are parameters determined, for each material,
from the experimental data [13, 24, 25, 27]. The knowledge of D() is quite limited for cement-based materials
and the validity of the approximate equation has yet to be
confirmed for different mixture designs.

Fig. 4 - Boltzmann transformation of the measured water profiles


(master curves) for different water/binder ratio specimens made
with white cement (W) and dried according to the two treatments (O and P).

619

Materials and Structures/Matriaux et Constructions, Vol. 35, December 2002

Table 9 and in Fig. 4.


In Table 3, the capillary water contents corresponding to the maximum water content obtained
during the actual NMR test for each mixture are
Mixture Experimental cap Experimental max Error cap Error max
given. As expected, cap decreases with a reduction
(m3/m3)
(mm/s1/2)
(%)
(%)
of the water/cement ratio. The addition of silica
W60O
0.210
0.400
0.9
3.7
fume increases slightly the cap values for the 0.25water/binder ratio mixtures. The drying treatment
W40O
0.155
0.330
0.9
3.8
does not influence significantly the value of cap.
W25O
0.100
0.129
0.6
3.4
Overall, test results clearly emphasize the strong
W25SFO
0.117
0.095
1.0
4.6
inf luence of moisture history on the liquid water
diffusivity of concrete. This strong influence of the
W60P
0.21
0.085
1.0
8.7
exposure conditions should always be kept in mind
W40P
0.150
0.090
1.2
5.2
while assessing the moisture transport properties of
the material. Furthermore, the fact that the funcW25P
0.104
0.050
1.4
9.5
tions and are apparently not singleW25SFP
0.118
0.038
1.3
11.6
valued should also be considered during the analysis
of any capillary absorption test data.
I60O
0.195
0.300
0.3
4.1
Using Equation (8), the liquid water diffusivity
I40O
0.156
0.265
0.3
4.6
can be calculated numerically from the transformed
I25O
0.106
0.139
0.3
3.5
experimental profiles. The coefficients obtained by
I25SFO
0.144
0.132
0.2
3.5
this method are given in Fig. 5 for samples W40O
and W25O. The scatter of the data at low water
are given for the materials investigated in this study. For
contents originates from the calculation of the numerical
the same drying treatment, max decreases with a reducderivative in the evaluation of Equation (8). As can be
tion of the water/cement ratio and the addition of silica
seen, the water diffusivity coefficient is strongly depenfume.
dent on the material water content () This is in good
The reduction of the water/cement ratio results in a
agreement with the numerical simulations of Martys
refinement of the pore structure by the elimination of
[28] who studied the mechanisms of moisture transport
part of the large capillary pores and more importantly,
in various porous systems.
the reduction in the pore space connectivity, which plays
The experimental D() data (Fig. 5) is, for both speca major role in the capillary absorption of water.
imens, well described by two exponential equations
The drying treatment used prior to the absorption
D() = (D1exp(1)) + (D2exp(2)) each valid for a diftests has also a marked influence in the water penetration
ferent water content range. The coefficients D and
coefficient. The max of the propan-2-ol replaced W60P
were fitted using computer simulations. The computer
samples is 5 times lower than that of the oven-dried
simulations were performed using a standard NAG
W60O samples. These results are in good agreement
Fortran library routine (D03PGF) [32]. In Fig. 6, typical
with the microstructural results of previous investigaresults of the simulations of the water transport (full
tions, where direct oven drying was found to coarsen the
lines) are given for two different materials (W40O and
pore structure of hydrated cement systems [32-34]. The
W25O). As can be seen, for both water/cement ratio
use of propan-2-ol replacement technique has been
mixtures, the simulations obtained by the two exponenshown to be less damaging to the microstructure of the
tial functions yield a better approximation of the experimaterial.
If the reduced values of max measured for the solvent-treated samples is also probably inf luenced by the
fact that there is always a certain amount of propan-2-ol
that remains strongly adsorbed to the material pore surface even after several days of oven drying. The presence
of residual propan-2-ol can inf luence the capillary
absorption process in two manners: by blocking the
entries of some pores, and by changing the local
hydraulic capillary pressure () (see Equations (3) to
(6)). Both phenomena tend to decrease the water penetration of the propan-2-ol treated samples. The relative
importance of these phenomena requires further study.
Test results indicate that the harmful effect of direct
oven dr ying is attenuated by a reduction of the
Fig. 5 - The diffusion coefficient D() as function of the moisture
water/cement ratio. This is clearly reflected in the results
content determined for two water/cement ratio mixtures. Points
of max corresponding to the 0.25 water/binder ratio
correspond to experimental data and the solid lines to approximation
of the results by D() = (D1exp(1)) + (D2exp(2)).
mixtures dried with the two methods (O and P) in
Table 9 - Values of the capillary moisture content, cap
and the water penetration factor, and their corresponding
experimental error limits

620

Hazrati, Pel, Marchand, Kopinga, Pigeon

mental results than those obtained by using only a single


exponential function as proposed in the literature
(dashed lines in Fig. 6). The four parameters of the diffusivity coefficient of all the specimens described in this
report are given in Table 10. Table 11 gives the range of
water content where each exponential equation is valid.

6. CONCLUSION
The time-dependent moisture distribution in the
connected pore network of unsaturated mortars is a cruFig. 6 - Typical experimental data of Boltzmann transformation
cial issue for all durability aspects. It has been shown that
for two water/cement ratio mixtures (points) and typical
accurate quantitative water absorption data can be
Boltzmann transformation of simulated moisture profiles.
obtained in a non-destructive manner by the use
of the nuclear magnetic resonance imaging techTable 10 - Parameters of the two exponential isothermal
nique (NMRI). The use of a specially designed
hydraulic diffusion coefficient
NMR scanner even allowed the study of mortars
D() = (D1exp(1)) + (D2exp(2)) as measured by NMR
with a relatively high iron content (3% Fe ions)
Mixture D()=(D1exp(1))+(D2exp(2))
with a spatial resolution of 1mm.
1
D2
2
D1
The ingress of water by capillary absorption in
the unsaturated mortars investigated in this study
W60O
1.0E-09
15.4 1.0E-12 66.6
can be well described by a non-linear diffusion
W40O
5.0E-10
25.3 1.0E-14 115.1
equation based on Darcy's law. The water transW25O
1.6E-10
25.3 5.0E-18 239.0
port can be numerically predicted when the water
diffusivity coefficient D(), a fundamental propW25SFO
3.2E-11
36.8 1.3E-15 149.7
erty of the material, is determined experimentally.
W60P
4.0E-11
20.7 1.0E-26 206.1
This study provides new D() data necessary
for the mathematical modeling of water ingress in
W40P
1.6E-10
17.3 1.0E-21 208.4
unsaturated mortars with different mixture
W25P
1.6E-11
43.7 1.0E-23 337.0
designs. The water absorption diffusivity coeffiW25SFP
1.3E-11
32.2 1.0E-22 274.0
cients D() is strongly dependent on the water
content and can approximated by the sum of two
I60O
3.2E-10
27.6 1.0E-29 272.0
exponential functions. The commonly used single
I40O
2.5E-10
32.2 5.0E-25 268.9
exponential form of D() used in the modeling of
I25O
1.0E-11
73.7 8.0E-25 377.6
soils and highly porous masonry building materials
does not apply for the mortars investigated in this
I25SFO
3.2E-11
34.5 1.0E-23 262.0
study.
The movement of the wetting front, well described
Table 11 Range of validity of each exponential equation
by the water penetration coefficient max, ref lects the
characteristics of the continuous pore network of the
Mixture
D1exp(1)
D2exp(2)
Water content
Water content
dif ferent mater ials studied. The decrease of the
(m3/m3)
(m3/m3)
water/cement ratio of the mixture and the use of silica
Range 1
Range 2
fume results in a significant decrease in the water peneW60O
0.135
0.135
tration coefficient. The marked influence of the drying
treatment on the capillary absorption characteristics of
W40O
0.126
0.126
mortars has also been underlined. Propan-2-ol dried
W25O
0.081
0.081
samples (P) present significant lower water penetration
W25SFO
0.089
0.089
coefficient than their oven dried (O) counterparts.
W60P

0.193

0.193

W40P

0.135

0.135

W25P

0.096

0.096

W25SFP

0.108

0.108

I60O

0.184

0.184

I40O

0.143

0.143

I25O

0.099

0.099

I25SFO

0.126

0.126

ACKNOWLEDGEMENTS
The authors would like to acknowledge the financial
support of Concrete Canada and the Dutch Technology
Foundation (STW). The authors also wish to thank
H.J.P. Brocken for his help in conducting part of the
NMR tests and Mr. P. Henocq for his assistance in
preparing the final version of this paper.

621

Materials and Structures/Matriaux et Constructions, Vol. 35, December 2002

REFERENCES

wall and analysis of internal condensation, Int. Seminar Heat and


Mass Transfer, Dubrovnik, Vol. 1 (1978) 45-58.
[18] Justness, H., Sellevold, E.J. and Lundevall, G., High strength
concrete binders- Part A: Reactivity and composition of cement
pastes with and without condensed silica fume, ACI SP-132,
(1992) 873-890.
[19] Kopinga, K. and Pel, L., One-dimensional scanning of moisture
in porous materials with NMR, Rev. Sci. Instrum. 65 (1994)
3673-3681.
[20] Farrar, T. and Becker, E.D., Pulse and Fourrier Transform
NMR. Introduction to theory and methods, (Academic Press,
New York, 1971).
[21] Abragam, A., Principles of nuclear magnetism, (Oxford
University Press, New York., USA, 1985).
[22] Gummerson, R.J., Hall, C. and Hoff, W.D., Capillary water
transport in masonry structures; building construction application of Darcy's law, Construction Papers 1 (1980) 17-27.
[23] Bear, J. and Bachmat, Y., Introduction to modeling of transport
phenomena in porous media, Vol. 4, (Kluwer, Dordrecht, The
Netherlands, 1990).
[24] Philip, J.R. and Knight, J.H., On solving the unsaturated flow
equation: 3. New quasi-analytical technique, Soil Science 93
(1974) 405-412.
[25] Parlange, J.-Y., Theory of water-movement in soils: 1. One
dimensional absorption, Soil Science 111 (1971) 134-137.
[26] Philip, J.R., Numerical solution of equations of the diffusion
type with diffusivity concentration-dependent, Trans. Faraday
Soc. 51 (1955) 885-892.
[27] Gardner, W. and Mayhugh, M., Solutions and test of the diffusion equation for the movement of water in soil, Soil Science Soc.
Amer. Proc. 22 (1958) 197-201.
[28] Martys, N.S., Diffusion in partially saturated porous materials,
Mater. Struct. 32 (1999) 555-562.
[29] Hall, C., Barrier performance of concrete: A review of fluid
transport theory, Mater. Struct. 27 (169) (1994) 291-306.
[30] Kohonen, R., A method to analyze the transient hygrothermal
behaviour of building materials and components, Dissertation,
Publication 21, Technical Research Centre of Finland, Espoo, 1984.
[31] CIB, Quantities symbols and units for the description of heat
and moisture transfer in buildings; conversion factors, Report
No. 37, Rotterdam, 1977.
[32] The NAG Numerical Fortran Library, Mark 15, NAG Ltd,
Wilkinson House, Jordan Hill Road, Oxford, U.K., 1991.
[33] Bager, D.H. and Sellevold, E.J., Ice formation in hardened cement
paste - Part II: Drying and resaturation of room temperature cured
paste, Cement and Concrete Research 16 (6) (1986) 835-844.
[34] Feldman, R.F., Diffusion measurements in cement paste by
water replacement using propan-2-ol, Cement and Concrete
Research 17 (4) (1987) 602-612.
[35] Parrott, L.J., Effect of drying history upon the exchange of pore
water with methanol and upon subsequent sorption behaviour in
hydrated alite paste, Cement and Concrete Research 11 (1981) 651-658.

[1] Galloway, J.W., Harding, H.M. and Raithby, K.D., Effects of


moisture changes on flexural and fatigue strength of concrete,
Transport and Road Research Laboratories, Report 864, U.K.
(1979) 26 p.
[2] Hanson, J.A., Effect of curing and drying environments on splitting tensile strength of concrete, ACI Journal 65 (1968) 535-543.
[3] MacInnis, C. and Beaudoin, J.J., Effect of degree of saturation on
the frost resistance of mortars mixes, ACI Journal 65 (3) (1968)
203-209.
[4] Nilsson, L.O., Hygroscopic moisture in concrete -drying, measurements and related material properties, Report TVBM 1003,
Lund Institute of Technology, (1980) 162 p.
[5] Tuutti, K., Corrosion of steel in concrete, Swedish Cement and
Concrete Research Institute, Stockholm, Sweden, (1982) 160 p.
[6] Hashiba, H., Tanaka, K. and Koike, M., Moisture distribution in
concrete before and after application of the finish, Building
Research and Practice 5 (1990) 303-308.
[7] Parrott, L.J., Factors influencing relative humidity in concrete,
Magazine of Concrete Research 43 (154) (1991) 45-52.
[8] Bruce, R.R. and Klute, A., The measurement of soil moisture
diffusivity, Soil Science Soc. Am. Proc. 20 (1956) 458-462.
[9] Rose, M., Non intrusive moisture analyser for concrete,
National Science Foundation, Report No. NSF/ISI-88156,
Washington, DC, (1989) 85 p.
[10] Pel, L., Ketelaars, A.A.J. and Adan, O.C.G., Determination of
moisture diffusivity in porous media using scanning neutron
radiography, Int. Journal of Heat and Mass Transfer 36 (5) (1993)
1261-1267.
[11] Justnes, H., Bryhn-Ingegrigtsen, K. and Rosvold, G.O.,
Neutron radiography: an excellent method of measuring water
penetration and moisture distribution in cementitious materials,
Advances in Cement Research 6 (21) (1994).
[12] Qunard, D. and Salle, H., A gamma-ray spectrometer for
measurement of the water diffusivity of cementitious materials,
Materials Research Society Symposium Proceedings, Vol 137,
(1989) 165-169.
[13] Pel, L., Moisture transport in porous building materials, Ph.D.
Thesis, Eindhoven University of Technology, The Netherlands,
(1995) 125 p.
[14] Gummerson, R.J., Hall, C. and Hoff, W.D., Unsaturated water
flow within porous materials observed by NMR imaging, Nature
281 (1979) 56-57.
[15] Carpenter, T.A., Davies, E.S., Hall, C., Hall, L.D., Hoff, W.D.
and Wilson, M.A., Capillary water migration in rock: process
and material properties examined by NMR imaging, Mater.
Struct. 26 (1993) 286-292.
[16] Hall, C., Water sorptivity of mortars and concretes: a review,
Magazine of Concrete Research 41 (147) (1989) 51-61.
[17] Matsumoto, M., Simultaneous heat and mass transfer in porous

622

Vous aimerez peut-être aussi