Anda di halaman 1dari 13

Geomechanics for Energy and the Environment ( ) –

Contents lists available at ScienceDirect

Geomechanics for Energy and the Environment


journal homepage: www.elsevier.com/locate/gete

Consolidated-undrained triaxial testing of Opalinus Clay: Results and


method validation
Silvio B. Giger a, *, Russell T. Ewy b , Valentina Favero c , Rudy Stankovic d , Lukas M. Keller e
a
National Cooperative for the Disposal of Radioactive Waste, Wettingen, Switzerland
b
Chevron Energy Technology Co., Richmond, CA, USA
c
Swiss Federal Institute of Technology, EPFL, Lausanne, Switzerland
d
Rock Soil Testing and Development Company, Park City, UT, USA
e
ZHAW, University of Applied Sciences, Winterthur, Switzerland

highlights

• An alternative procedure to ‘‘conventional’’ testing of clay shales is presented.


• The robustness and advantages of the procedure is demonstrated.
• The impact of inappropriate strain rate on test results is highlighted.
• The procedure decreases test complexity and rig time (10 to 20 days) significantly.

article info a b s t r a c t
Article history: Specific equipment and procedures developed for geomechanical testing of hydrocarbon caprocks were
Received 14 July 2017 adopted to conduct truly undrained triaxial tests with Opalinus Clay. The amount of pore pressure
Received in revised form 16 January 2018 development during consolidation, and the resulting effective stress, is managed by equilibrating the
Accepted 24 January 2018
samples in vacuum desiccators of different relative humidities (vapor equilibration technique) prior to
Available online xxxx
assembling into the test apparatus. A drained consolidation test was first conducted to determine the
appropriate strain rate for consolidated-undrained (CU) triaxial testing. Opalinus Clay samples were then
consolidated in the triaxial rig to mean effective stresses in the range from 3 to 52 MPa and eventually
sheared. Within the explored stress range elastic and pore pressure coupling parameters were found
to be stress dependent. The different stress paths to peak indicate a transition from overconsolidated
to rather normally consolidated state, yet failure was in all tests dilatant, i.e. associated with a drop in
pore pressure and strain-softening (more so at low effective stress). Accurate pore pressure monitoring
enabled the discrimination of different deformation stages during deviatoric loading. In terms of Mohr–
Coulomb strength parameters, transition from peak to post-peak strength is manifested by a reduction in
the effective cohesion whereas the effective friction angle remains nearly constant. The robustness of the
CU testing methodology is demonstrated by (i) diagnostic analyses, (ii) inconsistency of CU tests with two
CU tests deliberately loaded faster to explore the effect of strain rate, and (iii) consistency of CU tests with
two consolidated-drained tests. Finally, test results of two caprock shales are also shown for comparison.
The caprocks are of similar basic properties as the Opalinus Clay and stem from a large data base of tests
conducted using the same methodology.
© 2018 Elsevier Ltd. All rights reserved.

1. Introduction testing of Opalinus Clay, the designated host rock for high-level
radioactive waste in Switzerland. The planned repository depth
Triaxial testing of low-permeability argillaceous rocks (shales, range is approximately 400–900 m below ground. Opalinus Clay
claystones) is challenging as elements of traditional soil mechan- exhibits a very low hydraulic conductivity of the order of 10−14 –
ics testing must be adapted in the stress realm more typical of 10−12 m/s and a porosity of approximately 10–20% Refs. 1,2.
rock mechanics. Such procedures are required for geomechanical Specific to Opalinus Clay, procedures for careful triaxial testing
were proposed in the past.3,4 More recently, Wild et al.5 and Favero
et al.2 outlined the challenges of sound triaxial testing of Opalinus
* Corresponding author.
E-mail address: silvio.giger@nagra.ch (S.B. Giger). Clay. In their proposed procedure, Wild et al.5 focused in particular

https://doi.org/10.1016/j.gete.2018.01.003
2352-3808/© 2018 Elsevier Ltd. All rights reserved.

Please cite this article in press as: Giger S.B., et al., Consolidated-undrained triaxial testing of Opalinus Clay: Results and method validation, Geomechanics for Energy and
the Environment (2018), https://doi.org/10.1016/j.gete.2018.01.003.
2 S.B. Giger et al. / Geomechanics for Energy and the Environment ( ) –

Fig. 1. Overview of core integrity and heterogeneity: (a) Visualization of the core outer surface. (b) Visualization of the 3D compositional layering. (c) Virtual cross-section
through the core. Blue rectangle indicates interval selected for sub-coring. White arrows denote cracks.

on appropriate sample saturation prior to the deviatoric loading 2. Experimental methods


phase. Favero et al.2 performed drained triaxial testing with Opali-
nus Clay, with particular care of maintaining constant sample vol- 2.1. Sample handling
ume during the saturation phase. This aspect is important to avoid
The core material used for this study stems from a borehole
possible sample disturbance (and thus modifications of stiffness
drilled in the village of Lausen (Switzerland) and was sourced from
and strength) from contacting the sample with a synthetic pore
a shallow depth of approximately 33 m below ground. Opalinus
fluid at low effective stress (‘‘swelling’’). Clay was cored using a water-based drilling fluid with a polymer
The hydrocarbon industry faces similar challenges for geome- additive. Within approximately fifteen minutes after recovery core
chanical assessment of caprock shales and growingly also for tight sections of approximately 100 cm in length and 10 cm in diameter
shales from unconventional reservoirs. Special handling of the were wrapped and sealed in aluminum foil and resin-impregnated
material,6 and equipment and procedures for testing were there- into a plastic barrel to keep it preserved.
fore developed (e.g. Refs. 7,8). Prior to shipping to the testing laboratory and 11 days after
Here we report on consolidated-undrained (CU) triaxial test extraction from the ground, the preserved cores were checked for
results of Opalinus Clay by adopting the same established pro- integrity and internal variability using medical X-ray tomography
(XCT). Imaging was performed with a Somatom Definition AS 64
cedures, which have been continuously used and improved as
and tube voltage was 140 kV. All scans were performed using the
Chevron’s and RSTD’s shale testing standard for nearly 20 years.9
automatic dose modulation software (CARE Dose 4D). Collimation
The procedure precludes contacting the sample with a synthetic was 64×0.6 mm. All image reconstructions were performed using
pore fluid. Instead, the samples are pre-conditioned by exposure the soft tissue kernel (I31f). The investigated samples had a diam-
to prescribed relative humidity in desiccators outside the defor- eter of about 10 cm and the length of the entire cores was scanned.
mation vessel, assuring that full saturation develops from appli- This volume in combination with the used XCT device yielded im-
cation of confining stress. The allowable strain rate is constrained age data with a voxel size of 0.25 × 0.25 × 0.4 mm. The gray-value
from a consolidation test (the only test for which a synthetic of certain XCT image voxel visualizes a ‘‘mean’’ material density,
pore fluid is used) prior to the actual deformation test program. which is related to the volume fractions of different minerals and
Eight CU tests were performed, two of which were conducted pores filling the volume of that particular voxel. The local density
is controlled by the respective volume fractions of clay content,
at strain rates of up to two orders of magnitudes faster than
porosity and non-clay minerals. From the 3D XCT images a virtual
considered appropriate. This was done deliberately to explore the
cross-section through the center of the core was generated (Fig. 1).
effect of inappropriate loading rates on test results, notably on This cross-section allowed the detection of cracks with a crack
the apparent pore pressure evolution. Two consolidated-drained spacing wider than voxel-size. These cracks were carefully avoided
(CD) tests were also conducted for comparison with the CU during the selection of the subsample that was subsequently in-
results. vestigated. The outer surface as well as the compositional layering

Please cite this article in press as: Giger S.B., et al., Consolidated-undrained triaxial testing of Opalinus Clay: Results and method validation, Geomechanics for Energy and
the Environment (2018), https://doi.org/10.1016/j.gete.2018.01.003.
S.B. Giger et al. / Geomechanics for Energy and the Environment ( ) – 3

To evaluate its geomechanical properties, Opalinus Clay is mostly


sourced from the Mont Terri underground laboratory in West-
ern Switzerland (approximately 250 m below ground)11 , or from
deep boreholes in the repository candidate sites in North-eastern
Switzerland.10,12
The material used in this study stems from a more shallow
depth. The borehole was drilled to study the potential unloading
effects on barrier integrity of Opalinus Clay. The core section shown
in Fig. 1 was sourced from 32.4 to 33.4 m. Extensive analyses
including structural mapping, in situ hydraulic testing and basic
properties (mineralogy and permeability) suggests that Opalinus
Clay at Lausen site is mechanically affected by unloading and
weathering processes to a depth of approximately 28 m.13,14 The
bulk wet density (ρb ) measured on the sub-cores is 2.47 g/cm3 on
average (Table 1). This value is in good agreement with the values
measured on Opalinus Clay from greater depths (e.g. Refs. 1,11)
and lends further support that the samples were not significantly
mechanically affected by unloading, despite the shallow present
depth.
Bulk mineralogy (QXRD) was constrained on one spare sam-
ple from the blue section in Fig. 1. The three main constituents
are clay minerals (accounting for 59 wt%), quartz (24 wt%) and
carbonates (13.6 wt%, mainly calcite with some dolomite/ankerite
and siderite). Feldspars (2 wt%) and pyrite (<0.2 wt%) were much
less abundant. The bulk mineralogy is consistent with the average
mineralogy constrained for Opalinus Clay from some 125 samples
across Northern Switzerland1 .
After deformation in the rig the samples were gently pushed
out of the jacket into a pre-weighed container to obtain post-test
and oven-dried weights. Samples did maintain integrity during
this step and all material was extruded from the jacket into the
container. Water content (Table 1) was constrained by oven drying
(105–110 o C for 1–2 weeks) the samples after triaxial testing and
compare the oven-dried weight with the initial weight (immedi-
ately pre-testing). Water-loss porosity was calculated following
two approaches. First it was assumed that all samples were fully
saturated with a fluid of density equal to 0.998 g/cm3 . This leads to
an average porosity of 11.6% and a grain density (ρs ) of 2.663 g/cm3
(Table 1). If samples were not fully saturated then both the esti-
Fig. 2. Steps of sample preparation: (a) Core section after cutting. Note annular mated porosity and the derived grain density values would be too
space between core (and aluminum foil) and the black plastic barrel filled with low. Therefore, porosity was also calculated using the average ρs =
resin. (b) Sub-cored plugs submerged in decane.
2.720 g/cm3 from Nagra’s Opalinus Clay data base (constrained
by He-pycnometry on 87 samples,1 ). This approach results in an
average porosity of 13.5% and an apparent saturation of typically
of the core was visualized in 3D using the volume-rendering tool less than 90% (Table 1). Grain density was also constrained by He-
implemented in the Avizo software (Fig. 1). On the basis of a visual pycnometry in the one sample used for bulk mineralogy, yielding
inspection of the virtual cross-section through the core a crack-free a value of 2.681 g/cm3 . This suggests that the first approach with
and homogeneous subsample was selected for sub-coring of test the lower porosity values in Table 1 is more appropriate. This
specimens (blue in Fig. 1). interpretation is corroborated by an independent observation and
This 9 cm long subsample was extracted from the 98 cm long will be discussed in Section 2.4.
core by cutting directly through the plastic core holder (Fig. 2a). The potential difference in porosity is highlighted here because
Samples of 19 mm diameter were then sub-cored in a collar ar- small changes of porosity appear to be particularly relevant in this
rangement parallel to the core axis and perpendicular to bedding. class of material as will be shown later. In any case, from the XCT
Decane was used as a lubricant for cutting and sub-coring. The
scans (Fig. 1) and also thin section analyses it is apparent that
edges of the core were avoided, due to possible alteration by the
mineralogy (and hence grain density) is relatively constant in all
coring fluid. After cutting, the samples were stored in sealed jars of
the plugs taken from the extracted section of the core (cf. Fig. 1).
decane to prevent moisture loss (Fig. 2b). During the end-grinding
Another important aspect of the tested material is the low
process, samples were kept wet with decane at all times and were
never contacted with brine or water. All prepared samples were content of biogenic components which could serve as source of
approximately 38 mm in length, except for one sample (13 mm) calcite cement during lithification. In a micro-facies analysis based
used for a consolidation test (cf. Section 2.3). on thin section in five of the deformed specimens the biogenic
components (e.g. echinoderms) are estimated to account for only
2.2. Sample characterization and basic properties 5–7 vol.% and calcite cement for only 2–4 vol.%, most of which
even appears to be localized in lenses of quartz and biogenic
Opalinus Clay is an overconsolidated claystone with an esti- components (communication by H. Bläsi, April 2017). Hence on the
mated past burial depth of up to 1700 m in Northern Switzerland.10 scale of thin sections, there appears to be virtually no cement.

Please cite this article in press as: Giger S.B., et al., Consolidated-undrained triaxial testing of Opalinus Clay: Results and method validation, Geomechanics for Energy and
the Environment (2018), https://doi.org/10.1016/j.gete.2018.01.003.
4 S.B. Giger et al. / Geomechanics for Energy and the Environment ( ) –

Table 1
Overview of basic properties of tested samples.
Test-ID Bulk wet density (ρb ) Water content Assuming full saturation Assuming ρs = 2.72 g/cm3
Porosity Grain density (ρs ) Porosity Saturation
(g/cm3 ) (wt.%) (vol.%) (g/cm3 ) (vol.%) (vol.%)
LSN1-1-1 2.458 4.58 10.8 2.634 13.6 79.2
LSN1-1-2 2.464 4.34 10.3 2.632 13.2 77.8
LSN1-1-3 2.505 4.78 11.5 2.700 12.1 94.5
LSN1-1-4 2.447 5.34 12.4 2.653 14.6 85.0
LSN1-1-5 2.476 4.87 11.5 2.669 13.2 87.1
LSN1-1-6 2.459 5.61 13.1 2.679 14.4 90.7
LSN1-1-7 2.465 4.90 11.5 2.656 13.6 84.6
LSN1-1-8 2.477 4.93 11.7 2.672 13.2 88.1
LSN1-1-9 2.479 4.75 11.3 2.666 13.0 86.4
LSN1-1-10 2.462 5.23 12.3 2.667 14.0 87.5
Average 2.47 4.93 11.6 2.66 13.5 86
StDev 0.02 0.4 0.8 0.02 0.7 5

2.3. Determination of appropriate strain rate

The allowable loading rate in triaxial testing is controlled by the


capacity of the claystone to equilibrate pore fluid pressure within
the sample. Head15 provides equations that can be used to calculate
allowable axial strain rate for drained and undrained testing. These
are reproduced in Ref. 16, and tables of allowable strain rate for a
range of conditions can be found in Ref. 9.
Therefore a special test was performed to measure the consoli-
dation coefficient, cv (perpendicular to bedding). A sample 19 mm
diameter by 13 mm length was subjected to 10 MPa isotropic
confining stress and 3 MPa pore pressure using 8.29 wt% NaCl
brine. This test setup allows for applied fluid pressure at one
end, while sample pore pressure is measured at a fully-undrained
boundary at the other end, as described and illustrated in Ref. 16.
Once pore pressure was stable, it was increased to 6 MPa at the ap-
plied pressure end. After reaching a stable pressure at the no-flow
end it was decreased back to 3 MPa at the applied pressure end.
Fig. 3. Pressure diffusivity (consolidation coefficient) test, with fitted values.
This is essentially a one-dimensional pressure diffusion test. The
consolidation coefficient is approximated as the value of pressure
diffusivity that explains the observed pressure change vs. time.
The results of this test are shown in Fig. 3. The pressure increase 2.4. Sample preparation
behavior and pressure decrease behavior are fit by slightly differ-
ent values of cv . The average value is 0.0034 mm2 /s. By using the
equations in Ref. 15 the allowable axial strain rate for undrained After temporary storage in decane (Fig. 2b) each sample was
triaxial compression is 6 × 10−8 s−1 without side drains and equilibrated to a specific relative humidity between 92 and 98%
1.3 × 10−6 s−1 with side drains, for cylindrical plugs of radius (Table 2) in vacuum desiccators. These values were achieved
of 19 mm (and 38 mm length) and with strain at failure of 1%. with a closed air space above the following saturated salt solu-
For drained triaxial compression tests, the allowable strain rate is tions: 92%—sodium tartrate, 96%—potassium phosphate monoba-
3.5 × 10−9 s−1 for drainage from one end only and 1.5 × 10−7 sic, 98%—copper sulfate. This allows the water content of the
specimens to be changed through vapor adsorption/desorption,
s−1 with side drains. The side drains consist of thin, 3 mm wide
hence without direct contact with water. The purpose of this step
strips of fine-mesh metallic screen. They are on two opposite sides
is twofold: (1) it assures a very high level of fluid saturation in the
and extend from near the top all the way to the bottom. Another
sample, and (2) the subtle differences in relative humidity are used
screen provides the drainage along the bottom of the sample to
to control the amount of pore pressure (and hence effective stress)
connect the side drains to the pore fluid hole in the bottom end
developing during consolidation. This is required as the set-up does
platen (Fig. 4).
not use a pore fluid line which would allow dynamic control of fluid
Since the side screens do not provide complete circumferential pressure during testing (see below).
coverage, compromise strain rates of 2.0 × 10−7 and 2.0 × 10−8 s−1 Samples were left in the desiccators until they had reached
were used for the undrained and the drained tests, respectively. weight equilibrium, which would usually take 2–4 weeks. The
The allowable strain rates are actually 50% (or more) greater than relative humidity of the material from the fresh core was also
those quoted above, as the samples reached peak strain at ∼1.5% measured by using a small jar with a rubber stopper and an
strain rather than the assumed 1%. inserted humidity probe (cf. Ref. 6). The scrap material used for
To explore the effect of loading rate on the sample response, ‘‘native’’ relative humidity measurement initially gave a value of
two undrained tests were also conducted at faster strain rates, 94%, which by the next day had reduced to 91%. The specimens
i.e. 2.0 × 10−6 and 2.0 × 10−5 s−1 (cf. Section 3.2). For failure at 1.5% for compression testing are therefore allowed to equilibrate at
strain and with complete radial drainage, the allowable strain rate relative humidity values that are in the range of the ‘‘native’’
happens to be 2.0 × 10−6 s−1 . This was therefore a test of whether sample. It is also noted that none of the samples gained weight in
or not the two thin side screens provided adequate radial drainage. the desiccators, even when exposed to 98% relative humidity, and

Please cite this article in press as: Giger S.B., et al., Consolidated-undrained triaxial testing of Opalinus Clay: Results and method validation, Geomechanics for Energy and
the Environment (2018), https://doi.org/10.1016/j.gete.2018.01.003.
the Environment (2018), https://doi.org/10.1016/j.gete.2018.01.003.
Please cite this article in press as: Giger S.B., et al., Consolidated-undrained triaxial testing of Opalinus Clay: Results and method validation, Geomechanics for Energy and

S.B. Giger et al. / Geomechanics for Energy and the Environment (


Table 2
Overview of test conditions and test results.
Sample-ID RH [%] Strain rate Initial mean Skempton-B value [–] Poisson’s ratio Young’s Modulus Deviator stress Effective mean stress
effective stress p′ in at failure at failure
[/sec] [MPa] After 1 h equilibration After 8 h equilibration vu (CU) or vd (CD) [–] Eu (CU) or Ed (CD) [MPa] qf [MPa] p′f [MPa]
Standard undrained tests (CU)
LSN1-1-4 98 2 × 10−7 3.3 n.a. 0.30 2100 22.4 9.6
LSN1-1-9 98 2 × 10−7 5.3 0.76 0.90 0.48 3100 19.4 8.4
LSN1-1-10 98 2 × 10−7 16.5 0.79 0.87 0.49 5100 28.0 18.2
LSN1-1-7 96 2 × 10−7 18.1 0.60 0.80 0.41 5100 28.4 19.4
LSN1-1-8 96 2 × 10−7 43.5 0.62 (0.50–0.70) 0.36 7400 43.4 39.3
LSN1-1-5 92 2 × 10−7 51.5 0.52 0.71 0.38 9300 50.3 48.6
‘‘Fast’’ undrained tests (CU)
LSN1-1-3 96 2 × 10−5 22.9 n.a. 0.16 5000 45.8 24.4
LSN1-1-1 96 2 × 10−6 24.1 (0.52) (0.6–0.84) 0.49 9700 37.4 26.3
Drained tests (CD)
LSN1-1-6 98 2 × 10−8 6.4 n.a. 0.28 2000 21.6 13.3
LSN1-1-2 96 2 × 10−8 18.4 n.a. 0.29 5000 46.5 33.4

)
Note: Cylindrical samples were cored with the axis perpendicular to bedding.


5
6 S.B. Giger et al. / Geomechanics for Energy and the Environment ( ) –

pump was set to a value just slightly below the sample pore pres-
sure value, and the valve was then opened. After equilibrium was
ensured, deviatoric loading commenced. For these tests, the pore
pressure transducer located in the bottom end platen measures
only the applied pore pressure if the valve is open.
The combination of preconditioning the samples in a desiccator
and using small samples with side drains enabled relatively short
test duration between typically 10–20 days for CU and up to 30
days for CD tests, depending on the targeted stress range.

3. Results

Fig. 4. Schematic diagram of sample assembly. In the following, test results are discussed by referring to the
effective mean and the deviator stresses defined as:

p′ = (σ1 ′ + 2 · σ3 ′ )/3 (2)


therefore any microscopic damage of the samples due to swelling q = σ1 ′ − σ3 ′ (3)
is extremely unlikely.
where σ1 ′ and σ3 ′ are the vertical and lateral effective stress
2.5. Testing procedure magnitudes during triaxial compression tests with axial (vertical)
loading. Failure is defined here where the maximum value of q
The apparatus used for the CU test is described in detail in Ref. 8. is achieved during compression, and accordingly the respective
It does not contain a pore fluid line to ensure the external pore values of p′ and q are indexed as p′f and qf .
fluid volume is a small fraction of the sample pore volume. The Test conditions and test results for the various experiments are
key innovation of the apparatus is that pore fluid pressure can still summarized in Table 2. The tests are grouped by type or boundary
be measured continuously right near the sample face (Fig. 4). In conditions (‘‘undrained’’, ‘‘fast undrained‘‘—i.e. with higher strain
this way truly undrained tests can be conducted at the expense rate, and ‘‘drained’’) and the test order in the various groups is
of the capability to actively adjust pore fluid pressure during the according to the increasing initial isotropic effective mean stress
consolidation stage. This is managed by adjusting the samples to at the end of consolidation (p′in ). The Skempton-B parameter was
slightly different water contents as described above. The ‘‘consol- measured at the end of consolidation prior to deviatoric loading.
idation’’ stage requires that confining stresses are raised at least The elastic parameters (Poisson’s ratio and Young’s modulus)
until a positive pore fluid pressure is recorded. So in contrast to correspond to either undrained (vu , Eu ) or drained (vd , Ed ) values
conventional triaxial testing of soils or weak porous rocks, the depending on the type of test, and the values were constrained
consolidation in the procedure used here is undrained. on first loading at approximately 10–25% of peak strength. The
Before applying deviatoric loading, the Skempton-B value17 was alternative results from unload–reload cycles are presented in
evaluated in undrained tests. It is defined as Section 4.1.
∆u
B= (1)
∆σc 3.1. Results from consolidated-undrained (CU) testing
where ∆u and ∆σc are the incremental changes in pore fluid pres-
sure (u) and confining stress (σc ), respectively. For the Skempton-B In this section only CU tests performed at a strain rate of
tests σc was increased three times by 1 MPa, holding for equilibra- 2 × 10−7 s−1 are documented. Skempton-B values after 1 h equi-
tion typically for 1 day per step (Fig. 5 and Table 2). Based on the co- libration are consistently lower than after 8 h of equilibration
efficient of consolidation (Fig. 3), pore fluid pressure is expected to (Table 2). Whereas the values after 1 h can be considered as lower
stabilize approximately eight hours after drained isotropic loading; bound (i.e. possibly insufficient pressure equilibration), the values
however, due to the undrained boundary conditions, required pore after 8 h may be slightly elevated due to temperature variations
fluid pressure equilibration time may be one order of magnitude (cf. Fig. 5b). From the robust Skempton-B values a trend can be de-
faster.15 In addition, shorter time intervals also limit the potential tected towards lower B values with greater initial (cf. at the end of
impact of diurnal temperature variations (Fig. 5b). Based on these consolidation) mean effective stress (p′in ). This trend is observed for
considerations, the B-value was evaluated after two time steps measurements after 1 and 8 h alike and the generally high absolute
(Table 2): i) one hour after a confining step increase (to limit effect values (especially when constrained after 8 h) are good indicators
of temperature variations), and ii) eight hours after a confining step for full sample saturation. In this context it is emphasized that in
increase (to allow for maximum pore pressure equilibration). Both comparison with soft clays, the grain framework in Opalinus Clay
values are reported in Table 2. It is noted that temperature was is stiffer and Skempton-B values are expected to be <1 (e.g. Ref. 9).
not monitored inside the vessels, but temperature in the laboratory For example Favero et al.2 derived values between 1.0 and 0.8 for
was controlled to within approximately 1◦ C. effective mean stresses ranging from 2 MPa down to 12 MPa.
Deviatoric loading was then applied (perpendicular to bedding) The Young’s modulus correlates positively with p′in , whereas the
using a constant loading rate as determined from the consolidation Poisson’s ratio only shows a weak tendency towards lower values
test (see above). Loading was also reversed at a deviator stress of with increasing p′in . The general trend of all these predominantly
typically 10–15 MPa (25–50% of peak deviator stress) to perform (poro-)elastic parameters constrained immediately before or early
an unload–reload cycle for alternative evaluation of the Young’s during deviatoric loading indicates that the material becomes
Modulus (cf. Section 4.1). stiffer with greater effective stress.
For drained tests a different bottom end platen was used, con- To track deformation in undrained testing, the pore pressure
taining a pore line and a valve located just outside the pressure coupling parameter AB is introduced17 :
vessel. Initial consolidation was still performed undrained (valve ∆u
closed). Prior to deviatoric loading, the external pore pressure AB = (4)
∆q

Please cite this article in press as: Giger S.B., et al., Consolidated-undrained triaxial testing of Opalinus Clay: Results and method validation, Geomechanics for Energy and
the Environment (2018), https://doi.org/10.1016/j.gete.2018.01.003.
S.B. Giger et al. / Geomechanics for Energy and the Environment ( ) – 7

Fig. 5. (a) Stress/pressure vs. time sequence for one sample (LSN 1-1-10). (b) Detail for Skempton-B determination.

where ∆u and ∆q denote incremental changes in pore fluid pres- All tests have in common that curvature is to the right after
sure and deviator stress, respectively. The parameter AB is not maximum AB is reached, suggesting that failure is associated with
unique for a material, but depends on several factors such as the dilation in all tests. It is also interesting to note that the sample of
amount of stress applied, the initial stress state, the stress history test LSN 1-1-7, consolidated at p′in = 18.1 MPa, exhibits virtually
and stress-path applied.18 no volume change and maximum u and q coincide. Hence this may
Based on the above defined variables, the path to failure during be considered as an indication that a value of approximately 18
undrained compressive loading is subdivided into three distinct MPa relates to maximum pre-consolidation stress of the samples,
stages: fitting also the overall behavior observed in the stress paths as
described above. In this sense samples equilibrated at p′in ≤ 15–
– Stage 1: Maximum value of AB, i.e. the rate at which the 18 MPa may be considered overconsolidated, samples equilibrated
pore pressure increase with deviatoric loading reaches a at greater stresses as normally consolidated with respect to maxi-
maximum. It is an indicator of the onset of damage and mum pre-consolidation stress in situ.
crack development, as pore pressure increase is thereafter The stress–strain behavior to peak is in all tests clearly non-
counteracted by volume increase. linear (Fig. 8). All tests do exhibit strain-softening associated with
– Stage 2: Maximum value of pore fluid pressure u. failure. This finding is consistent with the dilative behavior associ-
– Stage 3: Maximum deviator stress qf . ated with peak in the stress paths above. Strain-softening is much
more pronounced for tests deformed at lower effective confining
stresses. Axial strain to failure is fairly constant around 1.5–2%. At
The different stages are illustrated in a collection of plots in
the highest effective mean stresses (p′in > 40 MPa), deviator stress
Fig. 6 for one representative test example (LSN 1-1-10). Besides
peaks in a much more ductile manner (prolonged peak) before
the variables defined above, the plots also examine the volumetric
eventually dropping at approximately 3% (for p′in = 52.5 MPa,
behavior (axial εa , radial εr and volumetric strain εv , respectively).
Fig. 8).
Note that AB shown in Fig. 6f depicts the curvature of the line
shown in Fig. 6e. The maximum AB value also broadly corresponds
to the minimum p′ achieved during testing (Fig. 6b), and to the 3.2. Results from ‘‘fast’’ consolidated-undrained (CU) testing
stress level at which volumetric deformation starts to deviate from
a linear increase with q. The last point is not very well resolved in Based on the sample dimensions and the consolidation co-
Fig. 6d, but is clearer in other tests, especially at high stresses. efficient the maximum strain rate for robust undrained testing
The maximum pore fluid pressure is reached well before the (with complete radial drainage) is calculated as 2 × 10−6 s−1 , and
maximum q, and it coincides with an acceleration of both the 2.0 × 10−7 s−1 was selected for the CU tests documented above,
negative volumetric deformation (dilation, Fig. 6d) and hardening since complete radial drainage was not certain. Two undrained
(Fig. 6a) with increasing loading. tests were conducted at deliberately faster loading rate to explore
A summary of stress paths of all CU tests deformed at a strain the effect on test results. One test (LSN 1-1-1) was conducted with
rate of 2 × 10−7 s−1 is given in Fig. 7 (cf. Fig. 6b). Also indicated one order of magnitude higher strain rate (2.0 × 10−6 s−1 ), and
in the figure are the three stages of deformation as defined above one test (LSN 1-1-3) even with two order of magnitude higher
(max. AB, max. u and max. q). All peak strengths (qf ) appear to strain rate (2.0 × 10−5 s−1 ). Based on the theoretical calculations
line up nicely on a straight line (see Section 4.3). Stress paths in Section 2.3, test LSN 1-1-1 would be considered as critical
in all tests show a predominantly negative slope until reaching (very close to calculated threshold) and test LSN 1-1-3 clearly as
maximum AB values. However tests with samples consolidated at inappropriate.
p′in approximately equal or lower than 15 MPa are dominated by a As expected from theory, the stress path of LSN 1 − 1 − 1 is
curvature to the right, whereas the tests with samples consolidated indeed similar as those from the other CU, except for the maximum
at greater stresses are dominated by a curvature to the left. Using AB value which appears to be somewhat elevated (Fig. 9). In
the same bulk cut-off of p′in ≈ 15 MPa, the bulk of deviatoric load- contrast, LSN 1 − 1 − 3 shows a clearly different behavior to all
ing to failure in the lower stress range is dominated by the interval other tests. The most noticeable difference is the apparently much
of maximum AB to qf , and the maximum pore fluid pressure is greater strength compared to the trend of all other CU tests. Fur-
achieved well before qf (cf. Fig. 6). In contrast, loading of samples thermore, all three criteria of defined test stages actually coincide,
consolidated at greater p′in values is dominated by the phase up to that is maximum AB and maximum u values are reached at peak
maximum AB, and maximum fluid pressure and qf nearly coincide. strength (Fig. 10). Also, curvature prior to reaching peak is curved

Please cite this article in press as: Giger S.B., et al., Consolidated-undrained triaxial testing of Opalinus Clay: Results and method validation, Geomechanics for Energy and
the Environment (2018), https://doi.org/10.1016/j.gete.2018.01.003.
8 S.B. Giger et al. / Geomechanics for Energy and the Environment ( ) –

Fig. 6. Example of diagnostic test analysis (LSN 1-1-10) with key stages during deformation. See text for definition of variables. Note: dashed line in Gray depicts development
during unload–reload cycle.

to the left, indicating contraction, yet pore fluid pressure is still clearly greater strength than what could be expected from the CU
increasing at this stage (Fig. 10). tests.
This apparent discrepancy of strength between CU and CD
3.3. Results from consolidated-drained (CD) testing tests can largely be reconciled with consideration of initial water
content or sample porosity (Table 1). Consistent with the relative
The drained tests (CD) were conducted at a strain rate of deviation from average strength trend in CU tests, samples used
2.0 × 10−8 s−1 , one order of magnitude lower than for the standard in LSN 1 − 1 − 6 and LSN 1 − 1 − 2 exhibit highest and lowest
CU tests. The stress paths are linear since no pore pressure change water contents (or porosities) of all samples when the more likely
is generated (assumed, as pore fluid pressure was measured only assumption of full initial saturation is considered. In addition,
at the drained end of the sample), and the slope of the stress path sample volumes can change during drained testing. Although it is
is simply q/p′ = 3/1 (Fig. 11). Peak strength of test LSN 1-1-6 estimated that changes are only of the order of half a percent in
consolidated at 6.4 MPa is slightly lower than the trend from the CU porosity, the effects of initial porosity anomalies can be amplified.
tests. In contrast, test LSN 1-1-2 consolidated at 18.4 MPa exhibits It is indeed found that in LSN 1-1-6, the sample dilates prior to peak

Please cite this article in press as: Giger S.B., et al., Consolidated-undrained triaxial testing of Opalinus Clay: Results and method validation, Geomechanics for Energy and
the Environment (2018), https://doi.org/10.1016/j.gete.2018.01.003.
S.B. Giger et al. / Geomechanics for Energy and the Environment ( ) – 9

drained tests are considered as reasonably consistent with the


undrained test results.
The elastic moduli for the drained tests have been analyzed both
on first loading paths and on unloading–reloading branches. The
Young’s moduli yield values of 2000 MPa and 5200 MPa for tests
LSN 1-1-6 and LSN 1-1-2, respectively, and the Poisson’s ratios 0.28
and 0.29 for tests LSN 1-1-6 and LSN 1-1-2, respectively. The results
are in good agreement with the elastic properties of CD tests on
Opalinus Clay reported in Ref. 2.

4. Discussion

4.1. Young’s modulus

In Section 3 the Young’s moduli were constrained from primary


(first) loading for both drained and undrained tests. Unload-reload
Fig. 7. Undrained stress paths for all CU tests conducted at a strain rate of 2 × 10−7 cycles were also performed and are used here for discussion of
s−1 . the appropriate methodology for derivation of Young’s moduli. To
this end the points at the start of unloading as well as of the start
of reloading were connected to derive the unload–reload (UL/RL)
E-modulus. It is noticeable from Fig. 12a, that the curve during
an unload–reload loop evolves from very steep to flat. Therefore,
the evolution of the E-modulus was investigated by plotting it as
a function of the strain increment. As can be seen from Fig. 12b,
the values of the E-moduli computed based on increasing strain
increments during an unloading path, line up nicely, and a value
for the E-modulus can be derived by extrapolating to a zero-
strain increment. This approach follows a method proposed by
Fjær et al.,19 and although empirical in nature, it is considered as a
traceable approach to derive the E-modulus.
The results of the three approaches are illustrated in Fig. 13.
The derived E-moduli from unload–reload cycles are comparable
to the values from first loading at lower stresses (p′in ≤ 20 MPa)
but deviate more strongly at greater stresses. In the lower stress
regime, values from the ‘‘zero-strain analysis’’ are approximately
a factor of two greater than the values from the other approaches,
and the discrepancy increases with increasing p′in .
When normalized by the undrained shear strength (cu = qf /2)
Fig. 8. Stress–strain curve for all CU tests conducted at a strain rate of 2 × 10−7 s−1 .
Deviator stress is normalized to p′in . it is found that the value of Eu :cu is fairly constant around 350:1,
varying by less than 10% in five out of six CU tests. The exception
is found only in the test consolidated at the lowest initial mean
stress (LSN1-1-4, p′in = 3.3 MPa), yielding a Eu :cu ratio of only
185:1. In contrast, for the unload–reload cycles and the zero-strain
approach, the Eu :cu ratios increase by approximately 30–40% in the
explored stress range.

4.2. Effect of strain rate on CU test results

In Section 3.2 it was demonstrated that test results are clearly


affected for strain rates faster than allowed by theory. Here a few
additional diagnostic analyses are provided to highlight that the
apparently anomalous strength results can be related to a lack
of pore fluid pressure equilibration. Fig. 14 shows the pore fluid
pressure dissipation after peak. It can be seen that at a strain rate of
2.0 × 10−7 s−1 pore fluid pressure mimics in detail the decrease in
deviator stress, whereas this is not the case for larger strain rates.
Most importantly, stabilization of pore fluid pressure along with
stable values of ∆q at large strain is not achieved in tests of higher
Fig. 9. Undrained stress paths for ‘‘fast’’ CU tests. Stress paths of ‘‘regular’’ tests are strain rates.
shown in Green.
The anomalous behavior of the test with the highest strain rate
(2.0 × 10−5 s−1 , LSN 1-1-3) can also be illustrated by plotting the
amount of fluid pressure increase to qf and the derived Poisson’s
ratio as a function of the initial consolidation stress (Fig. 15). The
(increase in porosity), whereas in LSN 1-1-2, volumetric change is
excess fluid pressure is clearly much greater than expected from
contraction up to failure. Taking these aspects into account, the
the other tests, and hence the apparent mean effective stress at

Please cite this article in press as: Giger S.B., et al., Consolidated-undrained triaxial testing of Opalinus Clay: Results and method validation, Geomechanics for Energy and
the Environment (2018), https://doi.org/10.1016/j.gete.2018.01.003.
10 S.B. Giger et al. / Geomechanics for Energy and the Environment ( ) –

Fig. 10. Diagnostic test analysis of ‘‘fast’’ test (LSN 1-1-3) loaded at 2 × 10−5 s−1 with key stages during deformation. See text for definition of variables.

failure is too low. Probably the greatest discrepancy is seen in the the two drainage strips on opposite sides of the sample almost
Poisson’s ratio with a value three times lower than the trend from provided full radial drainage.
other tests. Note that assuming isotropic material properties (for
simplification) and (i) a Biot coefficient alpha equal to 0.9, (ii) a 4.3. Derived strength properties
Skempton-B value of 0.9 (Table 2) and (iii) a drained Poisson’s ratio
of 0.28 (Table 2) the undrained Poisson’s ratio would yield a value
In Section 3.3 it was shown that CU and CD tests result in
of vu = 0.45 (e.g. Ref. 20), in good agreement with the reported
broadly consistent strength values taking into account the slightly
values of tests conducted at a lower strain rate.
different properties of the initial samples and the volume changes
Although test LSN 1 − 1 − 1 would be considered a ‘‘robust’’
during testing. The impact of small differences in porosity for this
test based on strain rate only for complete radial drainage, the
class of material was also reported by Menaceur et al.21 A Mohr–
effects on the derived values are small and noticeable only on
Coulomb based strength fitting is therefore performed only for the
some diagnostic analyses (cf. maximum AB in Fig. 9). This suggests
CU tests (excluding the ‘‘fast’’ CU tests), i.e. all tests with samples

Please cite this article in press as: Giger S.B., et al., Consolidated-undrained triaxial testing of Opalinus Clay: Results and method validation, Geomechanics for Energy and
the Environment (2018), https://doi.org/10.1016/j.gete.2018.01.003.
S.B. Giger et al. / Geomechanics for Energy and the Environment ( ) – 11

Fig. 13. Comparison of Young’s moduli derived from three different approaches.
Fig. 11. Drained stress paths in comparison with undrained stress paths (in Green). UL/RL denotes average value derived from unload–reload cycles.

Table 3
Mohr–Coulomb strength parameters for peak and post-peak conditions. approx. 20o (using the Mohr–Coulomb relationship of θ = ϕ /2
Mohr–Coulomb strength parameters +45o ). Furthermore, Fig. 16 also highlights that post-peak strength
Effective friction angle φ’ Effective cohesion c’ and maximum AB values (Fig. 7) in the CU tests are in relatively
(◦ ) (MPa) good agreement.
Peak 19 6.8
Post-peak 18 3.2
4.4. Comparison with caprock shales

of a water content (4.93 ± 0.4%) or porosity (11.6 ± 0.8%, using


From a dataset of over 90 caprock shales tested using these
the assumption of full saturation) within the range of one standard
same protocols, covering a porosity range of ∼10% to ∼30%,
deviation (Fig. 16). The drained tests are also plotted in Fig. 16 for
two examples are selected here for comparison. Both shales are
comparison, but were not considered for the regression analysis.
roughly similar to the tested Opalinus core; Shale X has bulk
The regression for peak values yields an effective friction angle of
density 2.49 g/cc, porosity (oven-drying) 14%–15%, is 64% clay (all
19o and an effective cohesion of 6.8 MPa (Table 3). This translates
illite/smectite) and has a bulk cation exchange capacity (CEC) of
to a computed unconfined compressive strength of approximately
∼25 meq/100 g. Shale Y has bulk density 2.43 g/cc, porosity ∼16%,
19 MPa.
is 77% clay (half kaolinite and half illite/smectite) and has a bulk
Regression of post-peak strength values suggests that the same
CEC of ∼18 meq/100 g. Additional test results on Shale Y, including
slope as for the peak values is also suited and hence the transition
strength anisotropy, are presented as ‘Shale E’ in Ref. 23.
from peak to post-peak mainly affects the intercept value, but
Undrained stress paths for both shales are shown in Fig. 17. Sev-
not the slope (Table 3). This is qualitatively and quantitatively
eral similarities with the Opalinus Clay (Fig. 7) are apparent. Firstly,
(friction angle of approximately 18–20o ) consistent with drained
the interpreted cohesion and apparent UCS of these shales are
test results by Favero et al.2 , and also with recent studies conducted
roughly similar to the tested Opalinus Clay (cf. Fig. 16). Secondly,
with the Callovo–Oxfordian claystone from the Bure Underground
the friction angle for these two shales is ∼18.6◦ , also similar to the
Laboratory.21,22
Opalinus Clay (this may be mostly coincidence; measured friction
It is noted also that failure planes and bedding form typically
angles for clay-rich shale are usually found to be anywhere in the
an angle of θ = 55o (±5 o ), consistent with a friction angle of
range 10◦ –25◦ in the claystone dataset). Thirdly, at high stress

Fig. 12. Derivation of Young’s Modulus (LSN 1-1-10): (a) Close-up depicting the difference of primary loading and unload–reload cycle. (b) Projection to ‘‘zero-strain’’ by
considering incremental changes during unload–reload cycle.

Please cite this article in press as: Giger S.B., et al., Consolidated-undrained triaxial testing of Opalinus Clay: Results and method validation, Geomechanics for Energy and
the Environment (2018), https://doi.org/10.1016/j.gete.2018.01.003.
12 S.B. Giger et al. / Geomechanics for Energy and the Environment ( ) –

Fig. 16. Peak and post-peak strength values (excluding the fast CU tests). Note
consistency of slope of regression for both peak and post-peak values, and broad
match of maximum AB values with post-peak values.

Fig. 17. Stress paths from two example caprock shales with similar basic properties
as the tested Opalinus Clay.

the undrained stress paths tend to the left, exhibiting normally-


Fig. 14. Rates of pore fluid pressure dissipation (∆u) and decrease in strength (∆q)
after peak for different strain rates (tests LSN1-1-7, LSN1-1-1 and LSN1-1-3 from
consolidated behavior. Fourthly, a strength drop following peak
top to bottom). stress occurs for all stress conditions, just as for the Opalinus Clay.
In addition, Shale X shows an overconsolidated stress path
direction (to the right) at low stress, similar to Opalinus Clay,
although Shale Y does not. A small difference compared to Opalinus

Fig. 15. Effect of strain rate on CU test results: (a) Increase of pore fluid pressure at peak (∆uf ). (b) Poisson’s ratio. Note clear deviation especially of test conducted at strain
rate 2 × 10−5 s−1 (LSN 1-1-3) from general trends.

Please cite this article in press as: Giger S.B., et al., Consolidated-undrained triaxial testing of Opalinus Clay: Results and method validation, Geomechanics for Energy and
the Environment (2018), https://doi.org/10.1016/j.gete.2018.01.003.
S.B. Giger et al. / Geomechanics for Energy and the Environment ( ) – 13

Clay is that the high-stress stress paths for both of these shales samples. Andreas Bauer introduced the authors to the ‘‘zero-strain’’
continue to the left at peak and post-peak, whereas they turn to concept to derive values for Young’s moduli. S. Giger appreciates
the right for the Opalinus. The slight differences in behavior among the advice and compelling discussions on improving mechani-
these two caprock shales, and compared to the Opalinus, could be cal testing of Opalinus Clay with Paul Marschall, Derek Martin,
due to differences in mineralogy or to fabric, including possible Lyesse Laloui and Alessio Ferrari. Bruno Kunz and Claudia Frei are
effects of anisotropic deformation behavior, and burial history. But acknowledged for assistance with graphical illustrations. Thank
all three shales demonstrate both clay-like behavior and rock-like you to two anonymous reviewers for their helpful suggestions to
behavior. In general, the more a shale has been compacted to low improve an earlier version of the manuscript.
porosity, the more its behavior deviates from that of high-porosity
clay soils. Some further examples, including stress–strain curves as References
well as stress paths, are given in Refs. 8and 9.
1. Nagra: SGT Etappe 2: Vorschlag weiter zu untersuchender geologischer Stan-
dortgebiete mit zugehörigen Standortarealen für die Oberflächenanlage. Geol-
5. Conclusions ogische Grundlagen. Nagra Tech. Ber. NTB 14-02. 2014.
2. Favero V, Ferrari A, Laloui L. Anisotropic behaviour of Opalinus Clay through
An alternative procedure to ‘‘conventional’’ triaxial testing of consolidated and drained triaxial testing in saturated conditions. Rock Mech Rock
low-permeability argillaceous rocks was presented with examples Eng. 2018. http://dx.doi.org/10.1007/s00603-017-1398-5.
3. Bellwald P. A contribution to the design of tunnels in argillaceous rock. Doctor of
of Opalinus Clay core. It consists of the following main steps:
Science in Civil Engineering at the Massachusetts Institute of Technology; 1990.
4. Aristorenas GV. Time-dependent behavior of tunnels excavated in shale. Doctor of
– A consolidation test to derive the consolidation coefficient Philosophy at the Massachusetts Institute of Technology; 1992.
and compute appropriate strain rate for compression tests. 5. Wild KM, Barla M, Turinetti G, Amann F. A multi-stage triaxial testing procedure
– Pre-conditioning the samples at specified relative humidity for low permeable geomaterials applied to Opalinus Clay. J Rock Mech Geotech
in a desiccator (control of water content). Eng. 2017;9:519–530.
6. Ewy RT. Shale/claystone response to air and liquid exposure, and implications
– ‘‘Undrained’’ consolidation at isotropic stresses in the defor-
for handling, sampling and testing. Internat J Rock Mech Min Sci. 2015;80:388–
mation rig. The effective mean stress develops as a function 401.
of sample water content. 7. Steiger RP, Leung PK. Consolidated undrained triaxial test procedure for shales.
– Undrained axial compressive loading to failure. In: Proc. 32nd U.S. Rock Mech. Symp. Balkema. 1991; 637–646.
8. Ewy RT, Stankovich RJ, Bovberg CA. Mechanical behavior of some clays and
Compared to ‘‘conventional’’ testing protocols, the procedure shales from 200 m to 3800 m depth. In: 39th US Rock Mech Symp/12th Panamer-
ican Conf Soil Mech & Geotech Eng. Cambridge, USA: MIT; 2003.
has the following advantages: 9. Ewy RT. Practical approaches for addressing shale testing challenges associated
with permeability, capillarity and brine interactions. Geomech Energy Environ.
– Decrease in test complexity: (1) No synthetic pore fluid 2018. http://dx.doi.org/10.1016/j.gete.2018.01.001.
added, avoiding effects of osmotic suction. (2) Waiving of 10. Mazurek M, Gautschi A, Marschall P, Vigneron G, Lebon P, Delay J. Transferability
the pore fluid line facilitates interpretation of poroleastic of geoscientific information from various sources (study sites, underground rock
parameters. laboratories, natural analogues) to support safety cases for radioactive waste
repositories in argillaceous formations.- Clay in natural and engineered barriers
– Decrease in rig time: (1) Pre-conditioning of the samples is for radioactive waste confinement (Lille 2007). Phys Chem Earth. 2008;33(suppl
done outside the deformation vessel (in a desiccator). (2) 1):S95–S105.
As consolidation in the rig is ‘‘undrained’’, no equilibration 11. Bossart P, Thury, M., eds. Mont Terri Rock laboratory –Project, Programme
phase is required to evacuate fluid out of the sample. This is 1996 to 2007 and results. Reports of the Swiss geological survey No. 3 –Swiss
geological survey, Wabern. 2008.
typically the most time-consuming step in ‘‘conventional’’
12. Giger SB, Marschall P, Lanyon GW, Derek Martin, C. Transferring the geome-
testing with low-permeability argillaceous rocks. (3) Small chanical behaviour of Opalinus Clay observed in lab tests and the Mont Terri
sample size (38 mm long and 19 mm diameter) in conjunc- URL to assess engineering suitability at a potential repository site. In: 49th US
tion with side screens enables for shear phase of several days Rock Mechanics / Geomechanics Symposium 2015, Vol. 3. 2015; 2384–2391.
even for low-permeability shales. Total test time reduces 13. Crisci E, Ferrari A, Giger S, Laloui L. One dimensional consolidation and perme-
ability of Opalinus Clay from shallow depth. In: Advances in Laboratory Testing
from several weeks/months (e.g. 2 and 5) to approximately and Modelling of Soils and Shales (ATMSS). 2017:338–344. Springer Series in
10–20 days. Geomechanics and Geoengineering [Proceedings ATMSS Workshop, Villars (CH), 18.-
20.01.2017].
The robustness of the testing procedures is documented in this 14. Vogt T, Hekel U, Ebert A, Becker JK, Traber D, Giger S, Brod M. Hydrogeologische
study by Untersuchungen im oberflächennahen Opalinuston (Bohrloch Lausen, Schweiz).
Grundwasser. 2017. http://dx.doi.org/10.1007/s00767-017-0363-2.
– Diagnostic analyses (evolution of pore pressure and 15. Head KH. Manual of Soil Laboratory Testing, Volume 3: Effective Stress Tests.
Chichester, UK: John Wiley & Sons; 1998.
strength). 16. Ewy RT, Daniels EJ, Stankovich RJ. Behavior of a reactive shale from 12000 ft
– Consistency of test results (line-up of failure points to nearly depth. ARMA-01-0077.- DC Rocks 2001. In: The 38th U.S. Symposium on Rock
perfect line) for samples from a core section with low min- Mechanics (USRMS), 7–10 July, Washington, D.C. 9. 2001.
eralogical variability. 17. Skempton A. The pore-pressure coefficients A and B. Geotechnique. 1954;4(4):
143–147.
– Broad consistency of test results between undrained and
18. Lambe TW, Whitman RV. Soil Mechanics. SI Verison. Series in Soil Engineering. John
drained tests. Wiley & Sons; 1979.
– The deviation of test results when conducted deliberately 19. Fjær E, Stroisz AM, Holt RM. Elastic dispersion derived from a combination of
‘‘too fast’’. static and dynamic measurements. Rock Mech Rock Eng. 2013;46:611–618.
– Good agreement of results with other recent, careful testing 20. Rice JR, Cleary MP. Some basic stress diffusion solutions for fluid-saturated
elastic porous media with compressible constituents. Rev Geophys Space Phys.
of Opalinus Clay in similar stress range, when considering 1976;14(2):227–241.
the differences in tested samples (water content/porosity). 21. Menaceur H, Delage P, Tang A-M, Conil N. The thermo-mechanical behaviour of
the Callovo-Oxfordian claystone. Internat J Rock Mech Min Sci. 2015;78:290–303.
Acknowledgments 22. Hu DW, Zhang F, Shao JF. Experimental study of poromechanical behavior of
saturated claystone under triaxial compression. Acta Geotech. 2014;9:207–214.
23. Ewy RT, Bovberg CA, Stankovic RJ. Strength anisotropy of mudstones and shales.
Martin Mazurek (University of Berne) is acknowledged for pro- In: 44th U.S. Rock Mech Symp/5th U.S.-Canada Rock Mechanics Symp, Salt Lake City,
viding data on mineralogy and grain density, and Hansruedi Bläsi USA, 27–30 June, 2010, paper ARMA 10-114. 2010.
for his analysis of the microfacies in the tested Opalinus Clay

Please cite this article in press as: Giger S.B., et al., Consolidated-undrained triaxial testing of Opalinus Clay: Results and method validation, Geomechanics for Energy and
the Environment (2018), https://doi.org/10.1016/j.gete.2018.01.003.

Anda mungkin juga menyukai