Uploaded by nandahuwaida123

Permeabilitas 1

advertisement
Journal of Rock Mechanics and Geotechnical Engineering 9 (2017) 383e395
Contents lists available at ScienceDirect
Journal of Rock Mechanics and
Geotechnical Engineering
journal homepage: www.rockgeotech.org
Full Length Article
Effects of temperature and thermally-induced microstructure change
on hydraulic conductivity of Boom Clay
W.Z. Chen a, b, *, Y.S. Ma b, c, H.D. Yu b, F.F. Li b, c, X.L. Li d, X. Sillen e
a
Research Centre of Geotechnical and Structural Engineering, Shandong University, Jinan, Shandong, 250061, China
State Key Laboratory of Geomechanics and Geotechnical Engineering, Institute of Rock and Soil Mechanics, Chinese Academy of Sciences, Wuhan, 430071,
China
c
University of Chinese Academy of Sciences, Beijing, 100049, China
d
European Underground Research Infrastructure for Disposal of Nuclear Waste in Clay Environment, EIG Euridice, Mol, 2400, Belgium
e
ONDRAF/NIRAS, Brussel, 1210, Belgium
b
a r t i c l e i n f o
a b s t r a c t
Article history:
Received 15 February 2017
Received in revised form
22 March 2017
Accepted 23 March 2017
Available online 21 April 2017
Boom Clay is one of the potential host rocks for deep geological disposal of high-level radioactive nuclear
waste in Belgium. In order to investigate the mechanism of hydraulic conductivity variation under
complex thermo-mechanical coupling conditions and to better understand the thermo-hydromechanical (THM) coupling behaviour of Boom Clay, a series of permeability tests using temperaturecontrolled triaxial cell has been carried out on the Boom Clay samples taken from Belgian underground research laboratory (URL) HADES. Due to its sedimentary nature, Boom Clay presents acrossanisotropy with respect to its sub-horizontal bedding plane. Direct measurements of the vertical (Kv)
and horizontal (Kh) hydraulic conductivities show that the hydraulic conductivity at 80 C is about 2.4
times larger than that at room temperature (23 C), and the hydraulic conductivity variation with
temperature is basically reversible during heatingecooling cycle. The anisotropic property of Boom Clay
is studied by scanning electron microscope (SEM) tests, which highlight the transversely isotropic
characteristics of intact Boom Clay. It is shown that the sub-horizontal bedding feature accounts for the
horizontal permeability higher than the vertical one. The measured increment in hydraulic conductivity
with temperature is lower than the calculated one when merely considering the changes in water kinematic viscosity and density with temperature. The nuclear magnetic resonance (NMR) tests have also
been carried out to investigate the impact of microstructure variation on the THM properties of clay. The
results show that heating under unconstrained boundary condition will produce larger size of pores and
weaken the microstructure. The discrepancy between the hydraulic conductivity experimentally
measured and predicted (considering water viscosity and density changes with temperature) can be
attributed to the microstructural weakening effect on the thermal volume change behaviour of Boom
Clay. Based on the experimental results, a hydraulic conductivity evolution model is proposed and then
implemented in ABAQUS. Three-dimensional (3D) numerical simulation of the admissible thermal
loading for argillaceous storage (ATLAS) III in situ heating test has been conducted subsequently, and the
numerical results are in good agreement with field measurements.
Ó 2017 Institute of Rock and Soil Mechanics, Chinese Academy of Sciences. Production and hosting by
Elsevier B.V. This is an open access article under the CC BY-NC-ND license (http://creativecommons.org/
licenses/by-nc-nd/4.0/).
Keywords:
Boom Clay
Permeability
Thermal effect
Anisotropy
Microstructure
1. Introduction
Deep geological disposal of high-level radioactive waste (HLW)
in clay formations is one of the promising methods. Boom Clay is
* Corresponding author.
E-mail address: [email protected] (W.Z. Chen).
Peer review under responsibility of Institute of Rock and Soil Mechanics,
Chinese Academy of Sciences.
considered to be one of the potential host rocks for HLW disposal in
Belgium because of its strong adsorption capacity, low permeability
(1012 m/s), self-sealing capacity and favourable creep properties
(Neerdael and Boyazis, 1997; Bernier et al., 2004). Comprehensive
investigations of the hydro-mechanical properties of this argillaceous rock have been carried out at the underground research
laboratory (URL) HADES since 1980. Due to the heat emitted by the
HLW, the temperature of the clay barrier is expected to be increased
after the installation of waste canisters. The temperature changes
http://dx.doi.org/10.1016/j.jrmge.2017.03.006
1674-7755 Ó 2017 Institute of Rock and Soil Mechanics, Chinese Academy of Sciences. Production and hosting by Elsevier B.V. This is an open access article under the CC BYNC-ND license (http://creativecommons.org/licenses/by-nc-nd/4.0/).
384
W.Z. Chen et al. / Journal of Rock Mechanics and Geotechnical Engineering 9 (2017) 383e395
would increase the hydraulic conductivity of Boom Clay and
consequently compromise the favourable properties of clay formation as a natural barrier for migration of radionuclide. A better
understanding of this issue is thus important for the repository
performance and safety assessments.
The thermal effect on hydraulic conductivity of Boom Clay has
been widely studied. Sultan (1997) observed that the hydraulic
conductivity of Boom Clay was up to two times larger when the
temperature changed from 35 C to 60 C. Delage et al. (2000) reported that the hydraulic conductivity of Boom Clay increased from
2.5 1012 m/s to 6.2 1012 m/s with temperature increasing
from 20 C to 90 C. In order to assess the self-sealing capacity of
damaged Boom Clay, Monfared et al. (2012) and Chen et al. (2014)
investigated the thermal effect on hydraulic conductivity of Boom
Clay pre-damaged by shearing and artificial fracture. They found
that the shear band or artificial fracture does not significantly affect
the permeability, because Boom Clay presents a good self-sealing
capacity. Nevertheless, the results focussing on Boom Clay are
still limited to enable a refined interpretation of the large-scale in
situ PRACLAY heater test in Belgian URL HADES (Li et al., 2010).
A number of studies (Wemaere et al., 1997; Bastiaens and
Demarche, 2003; Dehandschutter et al., 2004; Bastiaens et al.,
2007; Piriyakul and Haegeman, 2009; Chen et al., 2011; Lima,
2011) indicate that Boom Clay presents anisotropic properties.
Dehandschutter et al. (2005) observed bedding planes of fractured
Boom Clay by scanning electron microscope (SEM) observations.
Indeed, in the presence of sub-horizontal bedding planes, Boom
Clay can be considered as a transversely isotropic geomaterial (Yu
et al., 2014). The anisotropic property of permeability of Boom
Clay has been investigated by in situ experiments (Bastiaens et al.,
2006). For this, the anisotropic properties are further investigated
by more laboratory experiments in this study.
In most above-mentioned studies, the increase in hydraulic
conductivity with temperature has been considered due to the
decrease in viscosity of fluid (Habibagahi, 1977; Cho et al., 1999;
Delage et al., 2011). However, the changes of hydraulic conductivity with temperature are not only influenced by the changes of
water properties, but also by the thermal effect on soil-water
interaction at microstructural level (Towhata et al., 1993; Romero
et al., 2001; Villar and Lloret, 2004). As the temperature increases, the thermal effects would alter clay fabric (Romero et al.,
2001), produce larger voids between clay particles (Pusch and
Güven, 1990; Pons et al., 1994; Thomas et al., 1994), change the
effective flow cross-sectional area of porous channels (Ye et al.,
2013, 2014), and degenerate the absorbed water into free water
(Derjaguin et al., 1986). Consequently, there are different interpretations on the discrepancy of hydraulic conductivity between
the test results and the predictions by considering water properties
changing with temperature (Table 1). Towhata et al. (1993) analysed the influence of temperature on the permeability of MC clay
and bentonite, and concluded that the increment of measured hydraulic conductivity with temperature was higher than the
calculated one by using properties of free and pure water. They
attributed this discrepancy to the degeneration of absorbed water
into free water at elevated temperatures, which may result in easier
seepage through the clay. On the other hand, Houston and Lin
(1987) measured hydraulic conductivity values of illite, and Villar
and Lloret (2004) conducted hydraulic conductivity test of
bentonite at different temperatures. Their results showed that the
increase in hydraulic conductivity due to temperature evolution
was smaller than the prediction as per water viscosity change. They
suggested that this may account for the soil densification by thermal consolidation, and the variation of hydraulic conductivity with
temperature may depend on the type of material. Romero et al.
(2001) gave similar results for unsaturated Boom Clay and attributed this discrepancy to the clay fabric alteration and porosity
redistribution by thermo-chemical effects. Wan (2010) investigated
the thermal effects on the microstructure of the GMZ01 bentonite
using mercury intrusion porosimetry (MIP) technique. Results
showed that, under confined conditions, the pore structure of the
saturated GMZ01 bentonite changes slightly with temperature.
However, it reveals that the thermal effect on microstructure of clay
has not been fully understood to date, due to significant effect of
microstructure on thermo-hydro-mechanical (THM) properties of
clay. The nuclear magnetic resonance (NMR) technique is used to
investigate the microstructure change behaviour of clay without
disturbing the tested samples (Bird et al., 2005; Bayer et al., 2010).
This paper first presents the experimental results obtained using permeability tests considering heatingecooling cycle, as well as
the SEM and NMR tests conducted on heated Boom Clay samples.
These laboratory tests results help to understand the thermal effect
on hydraulic conductivity. The anisotropy of hydraulic properties of
Boom Clay as well as the importance of the thermally-induced
microstructure change and its effect on the THM behaviour of
Boom Clay is concerned. Next, based on the experimental results, a
model for hydraulic conductivity of Boom Clay in relation to temperature is proposed and then implemented in ABAQUS through
USDFLD subroutine. Finally, a three-dimensional (3D) numerical
simulation of the admissible thermal loading for argillaceous
storage (ATLAS) III in situ heating test is conducted, and the numerical results are compared with in situ measurements.
2. Materials and experimental investigations
2.1. Boom Clay samples
The Boom Clay samples are extracted at a depth of 223 m and at
dozens of metres deep from the sidewall of connecting gallery of
the URL HADES. Boom Clay is a dense plastic clay, with a total
porosity of around 39% and water content of 24%e30%. The dominant fraction (around 60%) contains illite, smectite, illite-smectite
mix layer and kaolinite. The non-clay mineral is mainly composed
of quartz (25%), feldspar with a little pyrite, and calcite (Yu et al.,
2012).
Table 1
Review of comparison of hydraulic conductivity values measured and predicted on the basis of water properties changing with temperature.
Source
Clay type
Temperature range ( C)
Test method
Test result
Reason of discrepancy
Towhata et al. (1993)
20e90
4e200
Measured < predicted
Thermally-induced degeneration of absorbed
water into free water
Soil densified by thermal consolidation
Villar and Lloret (2004)
Bentonite
20e80
Measured < predicted
Dependency on the type of material
Romero et al. (2001)
Unsaturated
Boom Clay
22e80
Calculated from cv
measurements
Calculated from cv
measurements
Constant head
permeability tests
Transient method
Measured > predicted
Houston and Lin (1987)
Bentonite and
MC clay
Illite
Measured < predicted
Thermally-induced clay fabric modification
and porosity redistribution
W.Z. Chen et al. / Journal of Rock Mechanics and Geotechnical Engineering 9 (2017) 383e395
2.2. Experimental investigations
2.2.1. Permeability tests
Permeability tests are performed in a temperature-controlled
triaxial testing machine (see Fig. 1), which is designed for investigating the THM characteristics of Boom Clay. The device consists of a
conventional triaxial apparatus and a temperature controller system. The confining pressure and the back pressure are applied by
two hydraulic pressure generators and measured through hydraulic
pressure transducers. The heater coil is installed around the outside
of the cell. The power supplied to the coil is automatically adjusted
by the temperature controller. Temperature is measured by the
temperature sensor submersed in the cell fluid. This system allows a
maximum temperature of 100 C and an accuracy of 0.5 C.
Smaller samples with standard diameter (38 mm) but reduced
height (10 mm) are used in order to ensure a measurable flow. The
hydraulic conductivities of Boom Clay measured through various
testing techniques exhibit similar values in the order of 1012 m/s
(Yu et al., 2013). Taking into account the anisotropy of Boom Clay,
we trim the samples manually with the axis parallel and perpendicular to the bedding plane, respectively.
Samples resaturation has been done using in situ effective stress
as proposed by Yu et al. (2012) before permeability measurement.
The saturation duration for Boom Clay is about 20 d in total when a
satisfactory value of the coefficient of Skempton B (greater than
0.85) is obtained. Permeability tests are carried out during a heatingecooling cycle (23 C (room temperature)/40 C/60 C
/80 C/60 C/40 C/23 C) at hydrostatic pressure
(s1 ¼ s2 ¼ s3 ¼ 2.5 MPa, more or less representative of the in situ
effective stress). The samples are firstly gradually loaded (40 kPa/h)
to the in situ effective stress and then heated/cooled down (0.3 C/
h) following the predefined heatingecooling cycle. These rates are
defined during an experimental benchmark exercise on Boom Clay
and found to be slow enough to ensure temperature and pore
pressure equilibrium inside the sample. The back pressure is
385
maintained at 1 MPa at the bottom and atmospheric pressure at the
top of the sample. The high back pressure is found necessary in
order to obtain a satisfactory precision in the measurement of flow
rate and, hence, of the permeability (Delage et al., 2000). The injection fluid for the tests is the synthetic Boom Clay water (SBCW).
2.2.2. SEM tests
Two kinds of samples (parallel/perpendicular to bedding plane)
have been investigated using environmental SEM techniques in
order to study the anisotropy properties of Boom Clay. The environmental SEM techniques allow for imaging of fresh samples
under undried and non-vacuum condition that effectively prevents
the development of desiccation cracks and unequal swelling of the
clay particles. As the samples are extracted at tens of metres deep
from the sidewalls with special cares to prevent desaturation, the
samples can be considered as representative of the undisturbed
Boom Clay materials.
2.2.3. NMR tests
NMR tests allow providing the information about the water
content and pore distribution of soil and are realised in a NMR
analyser, as illustrated in Fig. 2. The NMR analyser is mainly
composed of permanent magnet, sample tube, radio frequency (RF)
system, and data collecting and analysing system. The strength of
the permanent magnet is 0.52 T and the magnet temperature is
maintained at (32 0.01) C in order to ensure the uniformity and
stability of main magnetic field. Its design, operation and service
condition were discussed in detail by Tian and Wei (2014).
The dimensions of the sample in the NMR analyser are the same
as those for the permeability tests. Sample resaturation has been
also done following the same procedure as the permeability tests.
After resaturation, the sample is covered by preservative film and
heated at 23 C, 40 C, 60 C, and 80 C, respectively. The equilibrium time for each temperature level is about 2 h. Then the sample
is moved rapidly to the sample tube for NMR tests. Bayer et al.
Fig. 1. Schematic diagram of the temperature-controlled triaxial cell.
386
W.Z. Chen et al. / Journal of Rock Mechanics and Geotechnical Engineering 9 (2017) 383e395
Fig. 2. Schematic diagram of the NMR analyser.
a
1
zr2
T2
R
(1) The vertical (Kv) and horizontal (Kh) hydraulic conductivities
at room temperature are 1.73 1012 m/s and 5.01 1012 m/s, respectively, which fall in the range of the in situ
hydraulic conductivity measured by Bastiaens et al. (2006),
5.0
23 C (Heating)
(Heatinig)
4.5
40
40 C
C (Heating)
(Heatinig)
4.0
Water injected (mL)
(2010) summarised the applications of NMR technique in soil science. The measured signal of NMR test is called the free induction
decay (FID). The FID curve is determined by using Carr-PurcellMeiboom-Gill (CPMG) sequence in the present study. The FID
curve contains rich information about the water content and wateroccupied pore-size distribution of the porous medium. The initial
FID peak value is directly proportional to the number of water
protons in the sample (Ishizaki et al., 1996). Hence the initial FID
peak measurements can determine water content of the sample.
Moreover, the shape of FID curve is associated with the transverse
relaxation time (T2) of pore water protons. The relaxation time T2
distribution curve can be obtained by inverse Laplace transform.
The peak area (dimensionless) under T2 distribution curve represents the water content of corresponding T2 range (Bird et al.,
2005). The transverse relaxation time T2 of the water protons is
proportional to the water-filled pore size, which means that water
in small pores relaxes faster than that in large pores:
60 C (Heating)
(Heatinig)
3.5
80 C (Heating)
(Heatinig)
3.0
60 C (Cooling)
(Coolin g)
2.5
40 C (Cooling)
(Coolin g)
2.0
23 C (Cooling)
(Coolin g)
1.5
1.0
0.5
(1)
0.0
0
2
4
6
8
10
12
Time (h)
where R is the pore radius, r2 is the transverse relaxivity, and a is
the shape factor (1, 2, and 3 for planar, cylindrical and spherical
pore geometry, respectively) (Godefroy et al., 2001). Based on the
above theory, the T2 distribution curve can reflect the information
about the water content and pore distribution of soil.
(a)
3.5
23 C (Heating)
(Heating)
40 C (Heating)
(Heating)
3.0
3. Test results and discussions
3.1. Permeability tests
Although the reduced sample height is used in association with
high water pressure gradient applied, duration of about 10 h is still
needed to achieve the target steady water flow. The steady flow
duration of each temperature level is set to 10 h. The steady water
flow characteristic at different temperatures is presented in Fig. 3. It
can be observed that a satisfactory linear relationship is obtained
for each level of temperature. The injected water volume within the
same duration increases with increasing temperature, indicating
higher permeability with respect to higher temperature. The hydraulic conductivities are calculated by Darcy’s law. The variations
in hydraulic conductivity with temperature are shown in Fig. 4. It
can be seen that:
Water injected (mL)
60 C (Heating)
(Heating)
2.5
80 C (Heating)
(Heating)
60 C (Cooling)
(Coolin g)
2.0
40 C (Cooling)
(Coolin g)
1.5
23 C (Cooling)
(Coolin g)
1.0
0.5
0.0
0
2
4
6
8
10
12
Time (h)
(b)
Fig. 3. Injected water flow-time relationships of steady stage at each temperature
level. (a) Flow direction parallel to bedding plane; and (b) Flow direction perpendicular
to bedding plane.
W.Z. Chen et al. / Journal of Rock Mechanics and Geotechnical Engineering 9 (2017) 383e395
387
Table 2
Variations of viscosity and density of pure water with temperature.
Temperature ( C)
Viscosity (103 Pa s)
Density (g/cm3)
23
40
60
80
0.9579
0.656
0.4688
0.3565
0.9985
0.9927
0.9845
0.9742
viscosity and density (Table 2), taking the experimentally measured
hydraulic conductivity at room temperature as a starting point.
Fig. 6 shows the comparison between the values determined by
experiment and prediction, which indicates that the increase in
measured hydraulic conductivity is lower than that by prediction.
Fig. 5. Variations of the intrinsic permeability of Boom Clay with temperature.
Fig. 4. Variations in hydraulic conductivity of Boom Clay with temperature. (a) Horizontal; and (b) Vertical.
i.e. (1.7e2.3) 1012 m/s for Kv and (4.1e5.2) 1012 m/s for
Kh.
(2) The hydraulic conductivity at 80 C is about 2.4 times larger
than that at room temperature, and a linear relationship
between temperature and hydraulic conductivity basically
can be found regardless of anisotropy property.
(3) There is a positive and basically reversible correlation between hydraulic conductivity and temperature during heatingecooling cycle.
In order to decouple the effect of water properties change with
temperature from other factors affecting the permeability of Boom
Clay during heatingecooling cycle, the intrinsic permeability values
k are calculated based on the measured hydraulic conductivity:
k ¼
K mw
rw g
(2)
where K is the hydraulic conductivity; mw and rw denote the water
viscosity and density, respectively; and g is the gravitational acceleration. The thermal variations of viscosity and density of “pure
water” are available in science handbooks (Table 2). Fig. 5 shows the
variations of the intrinsic permeability with temperature. It can be
observed that both the vertical and horizontal intrinsic permeabilities decrease by approximately 10% (for all tests) during the
heating phase, and then they increase slightly during the cooling
phase. Additionally, this phenomenon may be attributed to the
thermal volume change behaviour of Boom Clay. The plastic thermal contraction of samples during the heating phase results in a
decrease in the intrinsic permeability.
The variation of the hydraulic conductivity with temperature is
also predicted by considering only the variation of pore water
Fig. 6. Comparisons of hydraulic conductivity values determined experimentally and
predicted on the basis of water viscosity and density change. (a) Horizontal; and (b)
Vertical.
388
W.Z. Chen et al. / Journal of Rock Mechanics and Geotechnical Engineering 9 (2017) 383e395
This also implies that other factors including the water properties
change may contribute to the hydraulic conductivity increase with
temperature. This will be investigated by NMR tests and discussed
in Section 3.3.
3.2. SEM tests
Fig. 7 shows the microstructure of Boom Clay perpendicular to
the bedding plane. The clay particles contact with each other in
edge-to-face or face-to-face manners, and the clay presents the
bending flaky structures (Fig. 7a). The diameter of flaky structures is
about dozens of microns, and composed of smaller flaky particles.
The diameter of flaky particles is only a few microns (Fig. 7b). These
results were also observed by Yu et al. (2012) on intact Boom Clay
sample.
Fig. 8 shows the microstructure of Boom Clay parallel to the
bedding plane. The apparent bedding plane, which mainly results
from the sediment process, is observed in SEM images.
Dehandschutter et al. (2005) observed a similar microstructure of
fractured Boom Clay samples. Tens of microns long and about 2
microns wide long-strip-shaped pores have been observed between clay particles (Fig. 8b). These discontinuities which linearly
arrange long-strip-shaped pores form elongated pore zone between adjacent bedding plane (Fig. 8a). Significant difference can
be seen between the SEM images perpendicular and parallel to
bedding plane. Compared to the well-developed alignment of the
clay particles in Fig. 8, the clay particles in Fig. 7 are more irregular.
Moreover, the pore shapes are different from two perpendicular
and parallel directions with respect to the bedding plane.
The presence of the bedding plane explains the transversely
isotropic properties of Boom Clay, and these discontinuities elongating shape pores (Fig. 8) highlight the high horizontal permeability along the bedding plane (Fig. 5).
3.3. NMR tests
Fig. 9 shows the thermal effect on the T2 distribution of saturated Boom Clay. The unchanged peak area of four curves demonstrates that the sample keeps saturated during the whole test (Bird
et al., 2005). Moreover, the T2 distribution curves move to right in
parallel with the increasing temperature (Fig. 9), indicating an increase in the relaxation time T2. The increasing relaxation time T2
means that the water-occupied pores size in saturated Boom Clay
becomes larger with the increasing temperature. These results are
similar with the previous results by X-ray diffraction technique
(Pons et al., 1994; Thomas et al., 1994) and electron microscopic
examination (Pusch and Güven, 1990), showing that heating produces larger voids among clay particles. It is worth noting that the
above-mentioned phenomena occur when clays are heated under
unconfined condition. Consequently, the larger internal particle
Fig. 7. Microstructure of Boom Clay perpendicular to the bedding plane. (a) 500; and (b) 2000.
Fig. 8. Microstructure of Boom Clay parallel to the bedding plane. (a) 500; and (b) 2000.
W.Z. Chen et al. / Journal of Rock Mechanics and Geotechnical Engineering 9 (2017) 383e395
389
Fig. 9. Transverse relaxation time distribution curves of water in saturated Boom Clay.
spaces produced by heating would lead to microstructural weakening and decrease in stiffness.
However, for the THM coupled permeability tests, the samples
are heated under hydrostatic stress. The samples would be compacted due to the thermally-induced microstructural weakening
and decrease in stiffness. Therefore, the measured hydraulic conductivity is lower than predicted one (Fig. 6) and the decrease of the
intrinsic permeability (Fig. 5) is induced due to the variation of void
ratio with temperature. This means that water viscosity and density
would affect the hydraulic conductivity in addition to the thermal
volume change behaviour of Boom Clay.
4. Numerical simulation of ATLAS III in situ heating test
Based on the experimental results, a hydraulic conductivity
evolution model is proposed and implemented in ABAQUS by
USDFLD subroutine. The USDFLD subroutine is designed to update
the parameters at all material points of elements from the start of
the increment. A 3D numerical simulation of the ATLAS III in situ
heating test has been conducted to validate this model.
4.1. A model for predicting hydraulic conductivity of Boom Clay
based on experimental studies
The experimental results show that not only the water viscosity
and density would affect the hydraulic conductivity of Boom Clay,
but also affect the thermal volume change behaviour. As a porous
medium, the variation of the void ratio of Boom Clay is controlled
by the deformation of skeleton:
De ¼ D
Vp
Vs
(3)
¼
DVp
De
Vs De
e e0
¼
¼
¼
Vs ð1 þ e0 Þ
1 þ e0
Vs ð1 þ e0 Þ
1 þ e0
1 n0
1 þ 3v
(5)
where n is the porosity, and n0 is the initial porosity.
Based on the Kozeny-Carman equation, the expression of the
hydraulic conductivity under constant temperature is
K ¼ C
g
n3
mw rw ð1 nÞ2 S2 G2s
(4)
where e0 is the initial void ratio, and e is the void ratio. Based on
Eq. (4), the porosity can be obtained:
(6)
where C is a constant and influenced by the tortuosity and shape of
the flow channels; S and Gs denote the mass specific surface area
and the specific weight of solids, respectively. Substituting Eq. (5)
into Eq. (6) to consider the thermal effect on the hydraulic conductivity and assuming that the C, g, S and Gs in Eq. (6) are constant,
the hydraulic conductivity of Boom Clay varies indirectly with
temperature and the volumetric strain can be obtained:
K ¼ KðT0 Þ
rw ðT0 Þmw ðT0 Þ ð3 v þ n0 Þ3
rw ðTÞmw ðTÞ n30 ð1 þ 3 v Þ
(7)
where T0 is the initial temperature; K(T0) is the hydraulic conductivity at initial temperature; mw ðT0 Þ and rw ðT0 Þ are the water dynamic viscosity in Pa s and the water density in g/cm3 at initial
temperature, respectively.
The dynamic viscosity mw ðTÞ and density rw ðTÞ of water
decrease with elevated temperature T. The following expressions
are adopted to describe the variations of dynamic viscosity and
density with temperature (Gong, 2015):
mw ¼ 2:9 1011 T 4 9:17 109 T 3 þ 106 T 2 5:56
1011 T þ 1:7862 103
where De is the increment of void ratio, Vp is the pore volume, and
Vs is the solid volume. Assuming that the clay particles are
incompressible, the volume strain 3 v can be obtained as
3v
n ¼ 1
rw ¼ 5:5 106 T 2 þ 2:28 105 T þ 0:99997
(8)
(9)
4.2. The ATLAS III in situ heating test
The ATLAS in situ heating test series have been performed since
1992 by SCK$CEN, which aimed at assessing the THM effects on the
390
W.Z. Chen et al. / Journal of Rock Mechanics and Geotechnical Engineering 9 (2017) 383e395
Boom Clay for hosting heat-emitting HLW. The ATLAS III in situ
heating test had been performed from April 2007 to April 2008 to
obtain a more accurate picture of the pore water pressure and
broaden the THM characterisation of the Boom Clay at different
temperature levels (Chen et al., 2011).
The ATLAS III test consists of a horizontal heater borehole
(AT89E, 19 m long, 160 mm in internal diameter; heater section
runs from 11 m to 19 m deep), three horizontal (AT85E, AT93E and
AT98E) and a downward inclined (AT97E) observation boreholes.
The temperature sensors and piezometer filters were installed in
Fig. 10. Schematic diagrams of instrumentation of the ATLAS III in situ heating test. (a) Spatial location of the 6 boreholes; (b) Horizontal view; and (c) Vertical view.
W.Z. Chen et al. / Journal of Rock Mechanics and Geotechnical Engineering 9 (2017) 383e395
391
the AT85E (15.3 m long, 60 mm in internal diameter), AT93E
(15.3 m long, 60 mm in internal diameter) and AT98E (21 m long,
100 mm in internal diameter) to measure the temperature and pore
pressure, respectively. However, the AT97E (21 m long, 40 mm in
internal diameter) only allows the temperature measurement.
Fig. 10 shows the test field of ATLAS III. The specifications and
location of sensors in the boreholes are shown in Table 3. The
heater was switched on from April 2007 to April 2008 with a
stepwise power increase (see Fig. 11), followed by an instantaneous
shutdown.
4.3. 3D finite element modelling for ATLAS III in situ heating test
The geometry and mesh of the 3D modelling are illustrated in
Fig. 12. The computational domain covers 60 m in the direction
parallel to the axis of the heater borehole, 40 m in the direction
perpendicular to the axis of the heater borehole (both horizontal
and vertical planes). It consists of 355,350 elements and 371,904
nodes. The actual central steel tube (19 m long, 95 mm in external
radius and 15 mm in thickness) is included in the geometry. The
impermeable characteristic of the steel is considered by using 294
thermo-mechanical coupled elements (C3D8T), and the heater is
attached to its innermost 8 m-long part. The THM coupled elements (C3D8PT) are used to simulate the THM behaviours of Boom
Clay.
The thermo-elastoplastic damage constitutive model based on
the Drucker-Prager Cap model proposed by Gong (2015) is verified
to capture the mechanical behaviours of Boom Clay at various
temperatures. It is therefore used in the hydraulic conductivity
model for numerical simulations proposed in this paper. The basic
THM coupling theories for deformable saturated porous medium is
based on the momentum conservation equation, mass conservation equation, energy conservation equation, constitutive equations
and the Terzaghi’s effective stress principle (Gong, 2015). ABAQUS
solves these equations in a fully coupled manner.
The initial conditions are described as following: sZZ ¼ rgz is
the initial vertical stress in Z-direction, sXX ¼ lX sZZ is the initial
horizontal stress in X-direction, and sYY ¼ lY sZZ is the initial horizontal stress in Y-direction. The lateral total pressure coefficient
lX ¼ lY ¼ 0:85. The pore pressure has been initialised according
to uw ¼ gw z, where gw is the unit weight of water. The initial
temperature is 16.6 C.
The boundary conditions are described as follows. On the top
surface of the model, vertical stress equal to the weight of the
Table 3
Summary of the instrumented sensors for the ATLAS III in situ test.
Borehole
Sensor no.
Distance to borehole
entrance (m)
Sensor type
AT85E
PP-AT85E
TC-AT85E
PP-AT93E
TC-AT93E
TC-AT97E1-12
TC-AT98E1
TC-AT98E2
TC-AT98E3
TC-AT98E4
TC-AT98E5
TC-AT98E6
TC-AT98E7
TC-AT98E8
TC-AT98E9
TC-AT98E10
PP-AT98E1
PP-AT98E2
PP-AT98E3
14.64
15.04
14.64
15.04
21-10
20
19
17
16
15
14
13
11
10
9
19
15
11
Piezometer
Temperature
Piezometer
Temperature
Temperature
Temperature
Temperature
Temperature
Temperature
Temperature
Temperature
Temperature
Temperature
Temperature
Temperature
Piezometer
Piezometer
Piezometer
AT93E
AT97E
AT98E
sensor
sensor
sensor
sensor
sensor
sensor
sensor
sensor
sensor
sensor
sensor
sensor
sensor
Fig. 11. Heating procedure of ATLAS III in situ heating test.
overlying strata is applied. On the right surface of the model,
symmetrical boundary condition is imposed. The normal deformation is fixed on other surfaces. All boundaries are impervious
with the exception of symmetrical boundary condition. The
impermeable characteristic of the steel tube is considered by using
impermeable element. All boundaries except the steel tube surfaces
where the heater is located are defined to be adiabatic.
The THM parameters of the Boom Clay are presented in Table 4
(Gong, 2015). When considering transversely isotropy, higher
values of Young’s modulus, Poisson’s ratio, thermal conductivity
and hydraulic conductivity in the horizontal plane (Eh0, nhh, Kh(T0),
lh) than those in the vertical plane (Ev0, nvh, Kv(T0), lv) are used.
4.4. Comparison between numerical results and in situ
measurements
4.4.1. Temperature evolution
The temperature sensors in boreholes AT85E and AT93E are
placed at a depth of 15 m, and those in boreholes AT97E and AT98E
are placed at the depths of 12 m and 10 m, respectively (Table 3).
The comparisons between measured and simulated temperatures
are shown in Figs. 13e15. As shown in Fig. 13, the maximum temperature change in borehole AT93E (24 C) is larger than that in
borehole AT85E (22 C) due to the shorter distance of AT93E to the
heater tube. The decrease in temperature is only visible 2 d after
switching off the heater. The maximum temperature change of
sensors in borehole AT98E is only 13 C (Fig. 14). Moreover, temperatures in borehole AT98E show a delay of the temperature
decrease after switching off the heater. The furthest temperature
sensor TC-AT98E10 starts to drop nearly 2 months after switching
off the heater. Temperatures in the inclined borehole AT97E (Fig. 15)
show a similar delay phenomenon. The maximum temperature
change of sensors is determined by its distance to the midpoint of
heater tube. Thermal convection can be neglected in Boom Clay, as
a kind of low-permeability dense plastic clay (1012 m/s). Thermal
conduction is the determinant way of heat transfer (Gens Sole et al.,
2007). As shown in Figs. 13e15, the numerical results are in good
agreement with in situ experimental results.
4.4.2. Thermally-induced pore pressure variation
The piezometers in boreholes AT85E and AT93E are used to
measure the pore pressure parallel to bedding plane at a depth of
14.6 m. There are three piezometers in borehole AT98E to measure
the pore pressures parallel to bedding plane at the depths of 11 m
(PP-AT98E3), 15 m (PP-AT98E2) and 19 m (PP-AT98E1), respectively.
392
W.Z. Chen et al. / Journal of Rock Mechanics and Geotechnical Engineering 9 (2017) 383e395
Fig. 12. 3D finite element mesh in the modelling of ATLAS III in situ heating test.
Table 4
THM coupled parameters for Boom Clay (Gong, 2015).
Young’s modulus (MPa)
Eh0
Ev0
1400
700
Thermal conductivity (W/(m K))
lh
lv
2.33
1.77
Cohesion, c0 (MPa)
0.3
Friction angle, 40 ( )
15
Hydraulic conductivity (m/s)
Poisson’s ratio
Kh(T0)
Kv(T0)
nhh
nvh
6 1012
3 1012
0.25
0.125
Dry density (kg/m3)
Porosity, n0 (%) Specific heat
capacity (J/(kg K))
1640
39
740
Thermal expansion Density of pore water, Viscosity of pore
rw ðT0 Þ (kg/m3)
water, mw ðT0 Þ (Pa s)
coefficient (K1)
Specific heat capacity of
pore water (J/(kg K))
Thermal conductivity of Thermal expansion coefficient of
pore water (W/(m K))
pore water (K1)
1 105
4200
0.57
999
1.101 103
These three piezometers are in the symmetric plane perpendicular
to the heated section of borehole AT89E and pass through its startpoint, midpoint and endpoint. The mechanism underlying the hydraulic behaviour is a competition between the generation of pore
pressure by the differential thermal expansion of liquid and solid
Fig. 13. Comparisons of measured and simulated temperature changes in boreholes
AT85E and AT93E.
3.3 104
and the dissipation of pore pressure, the rate of which is mainly
controlled by the value of permeability (Gens Sole et al., 2007). The
comparisons between measured and simulated pore pressure
changes in sensors PP-AT85E and PP-AT93E are shown in Fig. 16. At
the first two heating phases (0e400 W and 400e900 W), the generation of pore pressure by the differential thermal expansion of
liquid and solid dominates the hydraulic behaviour, and the pore
pressure increases with elevated temperature. However, during the
third heating phase (900e1400 W), the domination of the generation of pore pressure fades away with elapsed time and the dissipation of pore pressure begins to dominate. The pore pressure starts
to decrease about 2 months after the heating power is increased to
1400 W. It is interesting to note that at the beginning of each heating
phase, a temporary decrease of the pore pressure occurs. Furthermore, a temporary increase of the pore pressure also occurs when
the heater power is shutdown. Chen et al. (2011) suggested that this
phenomenon is resultant from mechanical anisotropy properties of
Boom Clay. After that, the pore pressure changes in sensors PP-AT85
and PP-AT93 drop to a maximum value of around 0.31 MPa, i.e.
0.31 MPa lower than initial pore pressure before heating, and then
recover slowly to their initial states before heating. The pore pressure variation of three piezometers in borehole AT98E is similar to
the sensors PP-AT85 and PP-AT93 (see Fig. 17). Because of the longer
W.Z. Chen et al. / Journal of Rock Mechanics and Geotechnical Engineering 9 (2017) 383e395
393
Fig. 14. Comparisons of measured and simulated temperature changes in borehole AT98E. (a) TC-AT98E1TC-AT98E5; (b) TC-AT98E6TC-AT98E10.
Fig. 15. Comparisons of measured and simulated temperature changes in borehole AT97E. (a) TC-AT97E1TC-AT97E6; (b) TC-AT97E7TC-AT97E12.
distance to the heating borehole, the magnitude of pore pressure
change is smaller than those in boreholes AT85E and AT93E. The
instantaneous but temporary pore pressure decrease (increase) after each power increase step (cooling) is again observed both from
field measurement and numerical results. The pore pressure change
throughout the duration of the in situ test can be reproduced,
especially the temporary decrease after increasing power, as well as
the temporary increase after cooling.
Fig. 16. Comparisons of measured and simulated pore pressure changes in boreholes
AT85E and AT93E.
Fig. 17. Comparisons of measured and simulated pore pressure changes in borehole
AT98E.
394
W.Z. Chen et al. / Journal of Rock Mechanics and Geotechnical Engineering 9 (2017) 383e395
5. Conclusions
The hydraulic conductivity of Boom Clay during a heatinge
cooling cycle has been successfully represented by taking into account the anisotropy property. The hydraulic conductivity
measured by constant head permeability tests is consistent with
that from field measurements. The hydraulic conductivity at 80 C
is about 2.4 times larger than that at room temperature, and a linear
relationship between temperature and hydraulic conductivity can
be found regardless of anisotropy property. The SEM tests highlight
the transversely isotropic characteristics of intact Boom Clay, and
the observed bedding features explain that the horizontal permeability is higher than the vertical one. Data analyses show that the
measured increment in hydraulic conductivity with temperature is
slightly lower than the predicted one on the basis of thermal
change in water viscosity and density, indicating that the variation
of porosity also has effect on the hydraulic conductivity during
heating.
The thermal effect on clay microstructure is investigated by
NMR technique to better understand the microstructure effect on
THM properties of clay. Larger internal particles spaces produced by
heating lead to microstructural weakening of Boom Clay, and the
samples would be compacted during THM coupled permeability
tests because of this phenomenon. Furthermore, based on the
experimental results, a hydraulic conductivity evolution model is
proposed and developed in ABAQUS by USDFLD subroutine. A 3D
numerical simulation of the ATLAS III in situ heating test is conducted to validate this model.
Conflict of interest
We wish to confirm that there are no known conflicts of interest
associated with this publication and there has been no significant
financial support for this work that could have influenced its
outcome.
Acknowledgments
The authors gratefully acknowledge the financial support of the
National Science Foundation for Distinguished Young Scholars
(Grant No. 51225902), National Natural Science Foundation of
China (Grant No. 51479190) and EURIDICE (European Underground
Research Infrastructure for Disposal of Nuclear Waste in Clay
Environment, Mol, Belgium) for the work presented in this paper.
The authors would like to thank Ms. Huihui Tian for the NMR tests
technical support. The authors are also grateful to the editors and
the reviewers for their valuable comments, which have significantly improved this paper.
References
Bastiaens W, Bernier F, Li XL. An overview of long-term HM measurements around
HADES URF. In: Van Cotthem A, Charlier R, Thimus J-F, Tshibangu J-P, editors.
EUROCK 2006-multiphysics coupling and long term behavior in rock mechanics. London, UK: Taylor and Francis Group; 2006.
Bastiaens W, Bernier F, Li XL. SELFRAC: experiments and conclusions on fracturing,
self-healing and self-sealing processes in clays. Physics and Chemistry of the
Earth, Parts A/B/C 2007;32(8e14):600e15.
Bastiaens W, Demarche M. The extension of the URF HADES: realization and observations. In: Proceedings of Waste Management 2003 Symposium. WM
Symposia, Inc.; 2003.
Bayer JV, Jaeger F, Schaumann GE. Proton nuclear magnetic resonance (NMR)
relaxometry in soil science applications. The Open Magnetic Resonance Journal
2010;3:15e26.
Bernier F, Demarche M, Bel J. The Belgian demonstration programme related to the
disposal of high level and long lived radioactive waste: achievements and
future works. In: Proceedings of Waste Management 2004 Symposium. WM
Symposia, Inc.; 2004.
Bird NRA, Preston AR, Randall EW, Whalley WR, Whitmore AP. Measurement of the
size distribution of water-filled pores at different matric potentials by stray field
nuclear magnetic resonance. European Journal of Soil Science 2005;56(1):135e
43.
Chen GJ, Maes T, Vandervoort F, Sillen X, Van Marcke P, Honty M, Vanderniepen P.
Thermal impact on damaged Boom Clay and opalinus clay: permeameter and
isostatic tests with mCT scanning. Rock Mechanics and Rock Engineering
2014;47(1):87e99.
Chen GJ, Sillen X, Verstricht J, Li XL. ATLAS III in situ heating test in Boom Clay: field
data, observation and interpretation. Computers and Geotechnics 2011;38(5):
683e96.
Cho WJ, Lee JO, Chun KS. The temperature effects on hydraulic conductivity of
compacted bentonite. Applied Clay Science 1999;14(1):47e58.
Dehandschutter B, Vandycke S, Sintubin M, Vandenberghe N, Gaviglio P, Sizun JP,
Wouters L. Microfabric of fractured Boom Clay at depth: a case study of brittleductile transitional clay behaviour. Applied Clay Science 2004;26(1e4):389e
401.
Dehandschutter B, Vandycke S, Sintubin M, Vandenberghe N, Wouters L. Brittle
fractures and ductile shear bands in argillaceous sediments: inferences from
Oligocen Boom Clay (Belgium). Journal of Structural Geology 2005;27(6):1095e
112.
Delage P, Sultan N, Cui YJ. On the thermal consolidation of Boom Clay. Canadian
Geotechnical Journal 2000;37(2):343e54.
Delage P, Sultan N, Cui YJ, Ling LX. Permeability changes in Boom Clay with temperature. 2011. arXiv preprint arXiv: 1112.6396.
Derjaguin BV, Kasasev VV, Khromova EN. Thermal expansion of water in fine pores.
Journal of Colloid and Interface Science 1986;109(2):586e7.
Gens Solé A, Vaunat J, Garitte B, Wileveau Y. In situ behaviour of a stiff layered clay
subject to thermal loading: observations and interpretation. Géotechnique
2007;57(2):207e28.
Godefroy S, Korb JP, Fleury M, Bryant RG. Surface nuclear magnetic relaxation and
dynamics of water and oil in macroporous media. Physical Review E
2001;64(2):021605.
Gong Z. Long-term thermo-hydro-mechanical coupled behaviour of Belgium Boom
Clay. PhD Thesis. Wuhan, China: Institute of Rock and Soil Mechanics, Chinese
Academy of Sciences; 2015 (in Chinese).
Habibagahi K. Temperature effect and the concept of effective void ratio. Indian
Geotechnical Journal 1977;7(1):14e34.
Houston SL, Lin HD. Thermal consolidation model of pelagic clays. Marine Geotechnology 1987;7:79e98.
Ishizaki T, Maruyama M, Furukawa Y, Dash JG. Premelting of ice in porous silica
glass. Journal of Crystal Growth 1996;163(4):455e60.
Lima A. Thermo-hydro-mechanical behaviour of two deep Belgian clay formations:
Boom and Ypresian Clays. PhD Thesis. Barcelona, Spain: Universitat Politècnica
de Catalunya; 2011 (in Spanish).
Li XL, Bastiaens W, Van Marcke P, Verstricht J, Chen GJ, Weetjens E, Sillen X. Design
and development of large-scale in situ PRACLAY heater test and horizontal
high-level radioactive waste disposal gallery seal test in Belgian HADES. Journal
of Rock Mechanics and Geotechnical Engineering 2010;2(2):103e10.
Monfared M, Sulem J, Delage P, Mohajerani M. On the THM behaviour of a sheared
Boom Clay sample: application to the behaviour and sealing properties of the
EDZ. Engineering Geology 2012;124:47e58.
Neerdael B, Boyazis JP. The Belgium underground research facility: status on the
demonstration issues for radioactive waste disposal in clay. Nuclear Engineering and Design 1997;176(1):89e96.
Piriyakul K, Haegeman W. Stiffness anisotropy of Boom Clay. In: Proceedings of the
17th International Conference on Soil Mechanics and Geotechnical Engineering:
The Academia and Practice of Geotechnical Engineering. IOS Press; 2009.
p. 167e70.
Pons CH, Franzetti N, Tchoubar D. Analyse de la structure multiéchelle des dispersions de montmorillonite Na. C.R. Rech, 92 N 800094, PIRSEM et ARTEP.
1994 (in French).
Pusch R, Güven N. Electron microscopic examination of hydrothermally treated
bentonite clay. Engineering Geology 1990;28(3):303e14.
Romero E, Gens A, Lloret A. Temperature effects on the hydraulic behaviour of an
unsaturated clay. Geotechnical & Geological Engineering 2001;19(3e4):311e32.
Sultan N. Etude du comportement thermo-mécanique de l’argile de Boom: Expériences et modélisation. PhD Thesis. Ècole Nationale des Ponts et ChausseesSpecialite; 1997 (in French).
Tian HH, Wei CF. A NMR-based testing and analysis of adsorbed water content.
Scientia Sinica Technologica 2014;44(3):295e305 (in Chinese).
Thomas F, Jin HM, Cases JM. Définition du comportement des argiles au cours des
forages. C.R. Rech, 92 N 80003, PIRSEM et ARTEP. 1994 (in French).
Towhata I, Kuntiwattanakul P, Seko I, Ohishi K. Volume change of clays induced by
heating as observed in consolidation tests. Soils and Foundations 1993;33(4):
170e83.
Villar MV, Lloret A. Influence of temperature on the hydro-mechanical behaviour of
a compacted bentonite. Applied Clay Science 2004;26(1):337e50.
Wan M. Study of soil-water characteristics and hydraulic conductivity of highly
compacted GMZ bentonite under temperature control. PhD Thesis. Shanghai,
China: Tongji University; 2010 (in Chinese).
Wemaere I, Marivoet J, Labat S, Beaufays R, Maes T. Core manipulations and
determination of hydraulic conductivities in the laboratory. Geological disposal
of high-level and long lived radioactive waste. Mol, Belgium: Waste & Disposal
Department, SCK∙CEN; 1997.
W.Z. Chen et al. / Journal of Rock Mechanics and Geotechnical Engineering 9 (2017) 383e395
Ye WM, Borrell NC, Zhu JY, Chen B, Chen YG. Advances on the investigation of the
hydraulic behaviour of compacted GMZ bentonite. Engineering Geology
2014;169:41e9.
Ye WM, Wan M, Chen B, Chen YG, Cui YJ, Wang J. Temperature effects on the
swelling pressure and saturated hydraulic conductivity of the compacted
GMZ01 bentonite. Environmental Earth Sciences 2013;68(1):281e8.
Yu HD, Chen WZ, Jia SP, Cao JJ, Li XL. Experimental study on the hydro-mechanical
behavior of Boom Clay. International Journal of Rock Mechanics and Mining
Sciences 2012;53:159e65.
Yu HD, Chen WZ, Li XL, Sillen X. A transversely isotropic damage model for Boom
Clay. Rock Mechanics and Rock Engineering 2014;47(1):207e19.
Yu L, Rogiers B, Gedeon M, Marivoet J, Craen MD, Mallants D. A critical review of
laboratory and in situ hydraulic conductivity measurements for the Boom Clay
in Belgium. Applied Clay Science 2013;75:1e12.
395
Dr. Weizhong Chen is the distinguished professor of
School of Civil Engineering, Shandong University, Jinan,
China. His research interests cover rock mechanics and
underground engineering. He is the author of 4 books and
has published more than 200 scientific papers.
Download