Sie sind auf Seite 1von 12

Mechanical and Petrophysical Properties of

North Sea Shales


P. HORSRUD$
E. F. SNSTEB
R. BE%
One of the problems when evaluating potential borehole stability problems in
the drilling of hydrocarbon exploration and production wells is the lack of
relevant data to describe the lithologies. This relates in particular to the
shales, since these constitute most of the overburden and are also the most
troublesome. This paper presents a set of data from a variety of deeply
cored shales from the North Sea. The shales have been extensively charac-
terised and span a wide range of both petrophysical and mechanical proper-
ties, depending on, for example, burial history and diagenesis. Special
characteristics of shales have been tested and described, including pore size
distribution, permeability, anisotropy and temperature dependence. Appropri-
ate procedures for sample handling, testing and interpretation have been
developed or modied from other related sciences. # 1998 Elsevier Science
Ltd. All rights reserved
INTRODUCTION
Shales are not primary target rocks (i.e. reservoir
rocks) in petroleum exploration. Shales still play a
very important role when exploring and drilling for
petroleum. Shales act as source rocks for hydrocarbons
and as cap rocks above the reservoir, i.e. preventing
the hydrocarbons in the reservoir rock from escaping
due to the low permeability and small pores (capillary
sealing) of the shale. In addition, most of the layers
between the surface and the reservoir are shales or
mudstones. This implies that the properties and re-
sponse of shale is important for the petroleum industry
in several respects, e.g. basin modelling, interpretation
of seismic response and potential borehole stability
problems during drilling, subsidence prediction, etc.
Since the shales are not the primary target, shale
samples (cores) from deep boreholes (>1000 m) are
very scarce. The main reason for this is the additional
cost related to a coring operation in a deep borehole.
This lack of available shale samples has been the main
reason for the lack of published data on these deep
shales and many investigators have instead used out-
crop shales. Borehole stability problems may be very
costly to handle, primarily because of the extra time
needed to complete the borehole. Therefore, some op-
erators have carried out coring operations in poten-
tially troublesome shales. These shales have
subsequently been characterised and tested in the
laboratory [1]. Most of the published information is,
however, still related to special problems like the sensi-
tivity of shales to dierent water-based uids [26].
In addition to being scarce, shales from deep bore-
holes have certain characteristic features which make
them dicult to handle correctly under laboratory
conditions. The two most important characteristics are
the low permeability [79] and the sensitivity to con-
tacting uids. The fact that the samples have been
retrieved from maybe several thousand meters depth
adds complexity to this process. The shale is unloaded
from pressure and temperature, and this may cause
damage and alterations in several ways, e.g. expansion,
creation of micro-cracks, disking and reduced satur-
ation. The fact that a shale taken from a deep bore-
hole is probably never 100% saturated under
atmospheric conditions (even if all the pore water is
contained in the sample), makes the laboratory testing
susceptible to artefacts [10, 11] unless special pro-
cedures are applied. Because of the low permeability of
these shales, testing of mechanical response can also be
very time-consuming and thus also relatively expens-
ive.
This paper presents results from a research project
on borehole stability in shales, which also included
extensive characterisation of a variety of shale cores
from boreholes in the North Sea. The characterisation
programme included measurements of a wide range of
petrophysical and mechanical properties. Since at least
some of the shales were potential candidates for stab-
ility problems, this data set contains weaker shales
Int. J. Rock Mech. Min. Sci. Vol. 35, No. 8, pp. 10091020, 1998
# 1998 Elsevier Science Ltd. All rights reserved
Printed in Great Britain
0148-9062/98 $ - see front matter PII: S0148-9062(98)00162-4
{Statoil Research Centre, Postuttak, N-7004 Trondheim, Norway.
{IKU Petroleum Research, 7034 Trondheim, Norway.
1009
than those commonly available from testing of outcrop
shales.
SHALE CHARACTERISATION STRATEGY
The term shale, as used in the petroleum industry, is
not precisely dened. It may range from weak, clay
rich gumbo to strongly cemented and shaly siltstones.
In this study, the term shale may include both shale
and mudstone. What will characterise the response of
the rocks with respect to permeability, water sensi-
tivity, etc., will be determined by the amount of clay
minerals. We have therefore used the term shale if the
rock contains a certain amount of clay minerals (using
a limit of about 30% clays), thus providing a continu-
ous matrix of clay.
Most of the characteristic properties of shales are re-
lated to the pore surface area and the electric nature
of the mineral surface (which may result from ion sub-
stitutions in the crystal lattice, not completely co-ordi-
nated cations at the crystal surface or by exchangeable
cations which are free to diuse away from the crystal
surface). Due to the sheet structure of the clay min-
erals, the surface-to-volume-ratio can become very
large. Combined with pore sizes ranging from below
1 nm to some mm with the major part typically below
50 nm, this results in a large sensitivity to the compo-
sition of contacting water.
For applications like borehole stability evaluation
and subsidence prediction it is obvious that mechanical
properties need to be measured. In order to under-
stand the shale behaviour there are some other charac-
teristics of the shale that also need to be investigated.
This relates to laboratory handling and procedures,
but also to the in situ behaviour of the shale.
Petrophysical characterisation tests are often simpler
and cheaper to perform and do not require full-size
cores. In addition these data can be used to categorise
the shale, i.e. provide a means of estimating the
expected mechanical response and properties within a
certain range. This requires that a sucient amount of
reference tests are available to compare with.
Techniques and methods have also been developed for
testing of drill cuttings [12, 13]. This can provide on-
site measurements which can be used to adjust oper-
ational parameters close to real time.
The properties required to characterise a shale will
depend also on the application. During drilling of
boreholes, the borehole is lled with a uid which is in
contact with the shale and can thus interact with the
shale through a hydraulic gradient, a chemical poten-
tial gradient and a temperature gradient. This makes it
necessary to include additional tests which can help
quantify possible shaleuid interaction eects.
The permeability determines the rate at which drill
uid ltrate can invade the shale and thus change the
pore pressure and interact chemically with the shale.
The permeability also sets the time frame for possible
delayed failure eects due to time-dependent pore
pressure evolution (consolidation eects). Potential
time dependent problems can be avoided if the bore-
hole is sealed with a casing or liner in due time.
The shaleuid interaction mechanisms which can
occur depend both on the composition of the drill
uid and the mineralogy of the shale. KCl-brines in
smectitic shale [14] and kaolinitic shale [15] are two
examples.
Quantication of eects of partial saturation and
possible laboratory artefacts thereof requires that the
pore size distribution of the shale is
determined [10, 15, 16].
This list is not complete, as other tests may be rel-
evant (cation exchange capacity, direct exposure tests,
etc.). There is an increasing amount of literature on
the topic of shaleuid interaction, but this is beyond
the scope of this paper. The anisotropy of shales may
be another important subject. This has not been a pri-
mary objective of this study.
Another reason for including the tests reported in
this paper is to understand fundamental shale beha-
viour in dierent environments (e.g. laboratory versus
downhole).
PETROPHYSICAL AND PETROGRAPHICAL
CHARACTERISATION
Table 1 gives an overview of the dierent shales
which have been tested and characterised in this study.
Table 1. Bulk mineralogy and porosity (fraction) of samples from deep boreholes in the North Sea which have been tested (Chl = chlorite,
Ka = kaolinite, Mi/Ill = mica and illite, ML = mixed layer, Sm = smectite)
Clay types
Shale Geological period Depth (m) Clay content Chl Ka Mi/III ML Sm Porosity
A Tertiary 1370 0.53 0.06 0.15 0.12 0.19 0.01 0.55
B Tertiary 1570 0.82 0.02 0.03 0.06 0.00 0.71 0.41
C Tertiary 1720 0.34 0.00 0.09 0.06 0.16 0.03 0.31
D Tertiary 1870 0.56 0.01 0.20 0.03 0.03 0.29 0.34
E Tertiary 1940 0.52 0.01 0.23 0.13 0.06 0.09 0.31
F Cretaceous 2090 0.46 0.03 0.12 0.12 0.18 0.01 0.29
G Jurassic 1870 0.32 0.04 0.12 0.06 0.09 0.01 0.30
H Jurassic 2630 0.58 0.06 0.36 0.10 0.06 0.00 0.10
I Jurassic 2550 0.47 0.02 0.14 0.22 0.09 0.00 0.17
J Triassic 2440 0.65 0.06 0.40 0.13 0.06 0.00 0.15
K Jurassic 4870 0.49 0.05 0.11 0.32 0.01 0.00 0.03
P. HORSRUD et al.: MECHANICAL AND PETROPHYSICAL PROPERTIES OF NORTH SEA SHALES 1010
An expected trend shown by Table 1 is that swelling
clay minerals (smectite) disappear with depth due to
illitisation of smectite as pressure and temperature
increases. The total clay content does not exhibit such
a depth trend.
Porosity and permeability
As shown by Table 1, there is considerable spread in
the composition and properties of the shales. For
Tertiary shales the porosity is generally high, with
55% porosity for the shallowest shale. As expected,
the porosity decreases with increasing depth and age,
down to 3% porosity for the deepest shale.
The porosity of the shales has been estimated from
determination of the free water content. The free water
content has been determined by drying 2040 g of
each sample in an oven at 1058C until constant sample
weight is reached. With this method for porosity deter-
mination, it is absolutely necessary that the sample has
been well preserved, i.e. that no pore water has been
lost after coring.
Some of the cores were also tested for helium-poros-
ity. In this case the sample is dried at 608C before
measuring the helium-porosity. When comparing with
porosity inferred from measurement of water content,
it was consistently found that the He-porosity was less,
see Table 2. The main reason for this is believed to be
that drying the sample causes compaction of the struc-
ture, especially in shales with high smectite content.
Thus the samples which have been subjected to the
608C drying sequence prior to testing will already be
in a compacted state during testing. The method based
on measuring the free water content is only dependent
on the initial amount of water and not on the amount
of compaction and the nal state of the sample. Note
also the good correlation between smectite content and
dierence between the two measured porosities.
Provided the sample is well preserved, we have there-
fore concluded that porosity estimated from the free
water content gives a better estimate than the He-por-
osity. Given that there are no changes in the bulk
volume, the porosity from water content is also com-
pletely independent of any structural changes of the
shale (e.g. cracks) during coring, retrieval, handling,
etc. The only requirement is that the water is not lost
prior to testing. Given that only free water is removed
in this process, the estimated value is still a lower
limit, since it is dicult to completely avoid any loss
of pore water before the sample is sealed. Another
aspect is that it may be dicult to draw a denite limit
between bound water and free water in shale which
has a range of pore sizes down to nanometer order of
magnitude [17].
The permeability of shales can vary over several dec-
ades and may be very low. Permeabilities below
10
21
m
2
(nDarcy) have been measured [7, 9, 18].
Intuitively one would think that the permeability of
shale is so low that it can be regarded as zero for all
practical purposes. That is, however, not the case. It
has been shown that the low permeability of shale can
be an important contribution in explaining the time-
delayed failure often experienced during drilling of
shales [19, 20]. The expansion of shales with nDarcy
permeability as a result of drilling causes an immediate
drop in the pore pressure. This increase in eective
stresses makes the formation more stable. Because of
the low shale permeability, it may take from hours to
weeks before the pore pressure again reaches its initial
value. This reduction in eective stress with time may
eventually cause failure of the shale.
Because of the low shale permeability, steady-state
type measurement techniques are not usually applied.
Instead, techniques based on transient measurements
are used [7, 9, 18]. We used a technique where a thin
(6 mm) circular disk (38 mm diameter) is placed in a
pressure cell, where pore pressure and hydrostatic con-
ning pressure can be applied separately. After loading
to pre-determined levels of pore pressure and conning
pressure and ample time for the sample to consolidate,
the pore pressure is increased by 0.5 MPa on one side
of the sample and reduced by the same amount on the
other side (see Fig. 1). The decay in this pressure
dierence is then recorded by a dierential pressure
transducer and the permeability is calculated from this
time response.
Table 2. Comparison between porosity measured by sample drying
(f
w
) and by He-injection (f
He
). The table also includes fractional
smectite content (Sm)
Shale
Geological
period f
He
f
w
f
w
f
He
Sm
B Tertiary 0.30 0.41 0.11 0.71
D Tertiary 0.28 0.34 0.06 0.29
E Tertiary 0.26 0.31 0.05 0.09
F Cretaceous 0.25 0.29 0.04 0.01
I Jurassic 0.15 0.17 0.02 0.00
K Jurassic 0.01 0.03 0.02 0.00
Fig. 1. Illustration of pressure transient method used for permeability estimation
P. HORSRUD et al.: MECHANICAL AND PETROPHYSICAL PROPERTIES OF NORTH SEA SHALES 1011
Table 3 summarises the permeabilities measured by
this technique. These measurements are all performed
normal to the bedding, i.e. essentially in the vertical
direction, and at an eective conning pressure of
4 MPa. Table 3 also includes a column with a per-
meability estimated from the consolidation part of
triaxial tests [21]. These data are based on several tests
for each shale, performed at dierent eective conn-
ing pressures, normally in the range 2.5 to 10 MPa. In
the triaxial tests the samples were drained both radially
and axially during consolidation. The permeability is
typically in the range 10
21
m
2
(nDarcy), although
there is considerable spread in the data. Note that the
measurements for each shale are performed on dier-
ent samples, and the transient measurements utilise
thin disks. This makes aspects like representativity and
homogeneity of the material very important. A shale
sample of 7 cm length may contain layers which vary
signicantly in composition, texture, grain size, pore
size and hence also in permeability. Note also that in
some of the tests, the very deep shale (shale K) gave a
surprisingly high permeability. This shale has been
unloaded from an estimated eective stress of about
50 MPa. Damage in the form of microcracks during
the unloading process must be expected in this case,
and even if the sample had been reloaded to the
expected downhole stress, one may not be able to close
these cracks entirely. This may also explain why there
is so much spread in the measurements on thin disks
for this shale. The lowest values ought therefore to be
considered as most representative.
Since shales are expected to be anisotropic, it is not
unreasonable to expect the permeability to be anisotro-
pic, with the larger value parallel to the bedding. The
permeability estimated from consolidation should in
that case be completely dominated by the permeability
parallel to the bedding, but there is no systematic
dierence between the permeability estimated from
consolidation and from thin disk measurements. We
have performed some initial tests on thin disks drilled
parallel to the bedding, but there are several problems
related to this. When preparing samples the failure
rate is much higher and the results are also much more
dependent on the existence of cracks along the bedding
direction. Cracks will tend to develop more easily in
shales which are in an overconsolidated state. We
have, however, not been able to nd any data in the
literature which can shed further light on this aspect.
Pore size distribution
The pore size distribution of the shales was
measured using mercury injection (up to 200 MPa).
Small samples (about 3 g) are rst dried at 608C and
cumulative injected volume of mercury is recorded ver-
sus injection pressure. Assuming a representative shape
of the pore channels, cylindrical, the pore size distri-
bution can be calculated. In reality it is the restrictions
in the pore network, i.e. the pore throats which deter-
mine the pressure necessary to force the mercury to
pass. The stepwise increase in volume represents the
part of the pore space available for the mercury behind
these restrictions at the corresponding injection press-
ure. The injection pressure (P) can be related to the
pore radius (r) by
P =
2g cos y
r
(1)
where g is the interfacial tension between the wetting
and non-wetting phase and y is the contact angle
between the solid surface and the wetting phase.
Figure 2 shows the obtained pore size distribution
for one of the Tertiary shales (shale D), showing that
the 50% level of cumulative volume injected is at
about 15 nm. This implies that very signicant capil-
lary pressures can be induced and this has several con-
sequences, both with respect to laboratory behaviour
and downhole behaviour. Since a retrieved core will
not be fully saturated under atmospheric conditions
Table 3. Permeabilities estimated from the consolidation part of triaxial tests and measured normal to the sample bedding using a transient
technique
Shale Geological period Permeability from consolidation (10
21
m
2
) Permeability from thin disks (10
21
m
2
)
A Tertiary 22 11
B Tertiary 4 317
C Tertiary 10 3
D Tertiary 7 9
E Tertiary 7 30
F Cretaceous 3 12
K Jurassic 24 1200
Fig. 2. Pore size distribution of shale D (Tertiary) determined from
mercury injection
P. HORSRUD et al.: MECHANICAL AND PETROPHYSICAL PROPERTIES OF NORTH SEA SHALES 1012
without adding pore water, contacting a core sample
with water based uid under atmospheric conditions
may lead to artefacts which are completely determined
by the capillary eect [10, 15]. With an interfacial ten-
sion between air and water of 72.7510
3
N/m and a
contact angle of 08 (i.e. fully wetting conditions), the
capillary pressure is 4.85 MPa for a 15 nm pore size.
Thus an evaluation of shaleuid interaction eects
based on samples which are not fully saturated upon
exposure may be very misleading.
A summary of the available pore size distributions
for the shales is presented in Table 4. In the notation
used here, 5% of the pore volume is made up of pore
radii below r
5
. The table shows that r
50
is in the range
820 nm for all the samples. The deepest shale (K) has
a larger proportion of larger pores. Again we attribute
this to more damage eect due to the large depth the
sample has been retrieved from (4870 m).
There is some uncertainty related to using this
method on shales, as it involves drying of the samples
which can aect the structure of the shale. Inability to
perform the measurements at correct downhole stress
conditions may also have an eect. With correct order
of magnitude numbers, the above conclusions should
still be valid.
TRIAXIAL TESTING
Handling and preparation procedures
The importance of preservation of the cored shale
samples has already been emphasised. This implies
that exposure to air must be minimised and the core
packing must be absolutely air-tight. Our experience is
that the standard procedure used in the oil industry
(wrapping in multiple layers of plastic foil, wrapping
in aluminium and nally dipping in wax) is adequate
at least for a few months storage. Shale cores should
never be frozen due to potential damage from pore
water expansion.
During preparation (drilling, grinding) it is common
to use an inert uid like a mineral oil as coolant to
avoid contact with water-based uids which may cause
unwanted physicochemical reactions. All the samples
were drilled with their axes normal to the apparent
bedding plane.
The potential of laboratory artefacts due to incom-
plete saturation makes it tempting to try to resaturate
the shale samples. However, this brings about several
other diculties. One is to determine the true compo-
sition of the pore water. Another is to actually obtain
full saturation without altering the properties of the
shale. Procedures for controlling saturation in cores
have been proposed (see for example Ref. [10]), using
a controlled humidity environment. To reach full sat-
uration a step of dehydration is required rst (to get a
continuous gas phase). This may induce changes in the
shale structure. It has also been shown that the shale
may start to change when the original water content is
somehow modied [17]. Our experience is that if the
core has been well preserved, satisfactory saturation is
obtained through hydrostatic loading of the sample.
This is evidenced both through the pore pressure re-
sponse during undrained triaxial loading and the lack
of capillary suction eects during uid exposure [14].
In the uid exposure tests, the samples were also
heated to 808C. This will also contribute to increase
saturation.
Test apparatus and test procedures
The test samples are 38 mm in diameter (1.5 inch)
and have a length of approximately 2 times the diam-
eter. The triaxial tests have been run as consolidated-
undrained (CU) tests in a servocontrolled loadframe.
This implies that the tests consist of the following
three segments:
(1) Loading to predetermined levels of conning
pressure and pore pressure.
(2) Consolidation, i.e. constant conning pressure
and drainage of the pore uid against a constant pore
pressure.
(3) Undrained axial loading under a constant axial
displacement rate until failure of the sample.
The internal instrumentation of the test sample is
shown in Fig. 3. Note that in addition to measurement
of external load, pressure and deformations, the pore
pressure on both ends of the sample is recorded and
acoustic wave trains in both the axial and radial direc-
tion are also recorded.
The time required to run a test depends very much
on the permeability of the shale. To run the entire test
under drained conditions may be extremely time-con-
suming. That is the main reason for running the triax-
ial loading part in undrained mode. Concepts from
soil mechanics testing can be applied to assist in deter-
mining when consolidation is completed and also to
determine the appropriate displacement rate in the
undrained part of the test [21]. Determination of the
strain rate in the undrained part is also based on the
consolidation response of the sample, in order to make
sure that pore pressure equilibration is ensured
throughout the sample. To reduce the test time as
much as possible, both axial and radial drainage of the
sample is included.
MECHANICAL PROPERTIES
A typical response from the triaxial loading part is
shown in Fig. 4 for a Tertiary shale (D). The pore
pressure increases until the sample fails, implying that
the sample contracts before failure. A brittle, macro-
Table 4. Pore size distributions (in nm)
Shale r
5
r
20
r
50
r
80
r
95
B 4 6 8 11 115
E 6 12 20 32 90
F 5 10 16 19 33
I 5 7 12 30 195
K 4 7 20 157 1360
P. HORSRUD et al.: MECHANICAL AND PETROPHYSICAL PROPERTIES OF NORTH SEA SHALES 1013
scopic shear failure is observed, with strain softening
before a relatively constant residual strength level is
established. This failure process causes the sample to
dilate, resulting also in a drop in the pore pressure.
Figure 5 shows the same test as in Fig. 4, but
including also two other tests run at dierent conning
pressures. These curves are plotted with dierential
stress (q = axial stress minus conning pressure) ver-
sus mean eective stress (p').
The two tests at the lowest conning pressures exhi-
bit a behaviour which in standard soil mechanics terms
is typical for an overconsolidated material (the mean
eective stress increases during loading). The test run
at the highest conning pressure is more vertical and
eventually the curve starts to change direction. This in-
dicates that the normal consolidation level has been
reached (Roscoe surface), i.e. the maximum stress level
the material has ever been exposed to [22]. Shales dif-
fer from soils in the way that they have some degree
of cementation. This aects the response and hence
also the possibility of applying these soil mechanics
principles.
The peak stress values in Fig. 5 fall essentially on a
straight line. In soil mechanics terms this is a projec-
tion of the Hvorslev surface onto the p'q plane (for a
given specic volume). A uniform clay obeying critical
state theory would follow the Hvorslev surface up to
the critical state line. Overconsolidated and cemented
rocks will eventually behave in a non-uniform manner.
This is clearly the case for the samples shown in Fig. 5.
As the peak stress value is approached, localisation
takes place and shear bands develop which eventually
form a macroscopic shear plane through the sample.
The behaviour after the peak is thus in this case more
dependent on characteristics typical of a rock (e.g.
cementation) rather than a soil.
The straight line in Fig. 5 can be translated into a
MohrCoulomb failure criterion, giving an extrapo-
lated uniaxial compressive strength of 14.3 MPa and a
failure angle of 49.18. Similar analyses can be per-
formed for the other shales which were tested. Table 5
Fig. 3. Triaxial cell instrumentation for consolidated-undrained triaxial testing of shale
Fig. 4. Triaxial loading part of shale D (Tertiary) tested at an initial
eective conning pressure of 5 MPa Fig. 5. Stress paths for three CU tests of shale D (Tertiary)
P. HORSRUD et al.: MECHANICAL AND PETROPHYSICAL PROPERTIES OF NORTH SEA SHALES 1014
presents a summary of the mechanical properties of
the tested shales. This shows that the extrapolated uni-
axial compressive strength for the shales varied
between 6 and 77.5 MPa. The failure angle varied
between 488 and 608. At low stress levels this is con-
siderably lower than, for example, a sandstone which
is normally in the range 608 to 708. Figure 6 shows the
interpreted uniaxial compressive strength plotted
against Young's modulus. As for many other rocks,
there is a good correlation between these two mechan-
ical properties (C
0
=6.55E, r
2
=0.99). Similar relations
but with other constants have also been found for
shale [23].
Pore pressure response
The pore pressure response during testing of shales
(and clays) may indicate whether the sample is fully
saturated or not. To quantify the pore pressure re-
sponse, the parameters A and B can be dened as [24]
Dp
f
= B[Ds
3
A(Ds
1
Ds
3
)] (2)
During hydrostatic loading (Ds
1
=Ds
3
) this is sim-
plied to
Dp
f
= BDs
3
(3)
and during triaxial loading (Ds
3
=0)
Dp
f
= ABDs
1
(4)
Introducing the change in mean eective stress
during triaxial loading yields
Dp
/
=

1
3AB
1

Dp
f
(5)
For the poro-elastic case AB can be expressed by
elastic constants [25]
AB =
1
3
1
1 jK
fr
(1 K
f
=K
s
)=aK
f
(6)
where j is the fractional porosity, a is the Biot coe-
cient (=1 K
fr
/K
s
), K
fr
is the bulk modulus of the
framework, K
s
is the bulk modulus of the solid grains
and K
f
is the uid modulus.
Normally the uid modulus is much less than the
solid grain modulus, so that
AB =
1
3
1
1 (jK
fr
=aK
f
)
(7)
This can now be related to the stress paths of the
p'q plots. With jK
fr
WaK
f
we get AB = 1/3 and
Dp' = 0, i.e. the curve in the undrained triaxial part is
vertical. This is commonly referred to as the ``weak
frame'' limit which is the common assumption for a
soil. For a soil in the elastic case it can be shown that
A = 1/3 and B = 1 [25].
If the weak frame assumption does not hold, then
AB < 1/3 and the curves will tilt to the right, as
shown in Fig. 5. The Tertiary shales are relatively
weak, and therefore the slope is not far from vertical.
The Jurassic shale (K) is much stronger, much more
brittle and is also more tilted to the right (Fig. 7).
Note that in some cases the curves tilt slightly to the
left (Dp' < 0). Examining Equation (6) we see that the
only solution to this is if K
f
>K
s
which is not realistic.
This type of behaviour thus indicates that the rock is
no longer elastic. This is for instance the case for a
normally consolidated material.
For a fully saturated soil B = 1 (weak frame limit).
If the soil is only partially saturated, B may be much
smaller than one. Thus Equation (3) can be used to
check if the sample is fully saturated or not. For a
more consolidated rock, B may be less than one even
for fully saturated samples. Thus this criterion is not
applicable to cemented rocks.
Fig. 6. Interpreted uniaxial compressive strength versus undrained
Young's modulus for the dierent shales
Table 5. Summary of mechanical properties of the tested shales
Shale Estimated uniaxial compressive strength (MPa) Failure angle (8) Young's modulus (GPa) Poisson's ratio
A 6.1 54.9 0.8 0.33
B 8.2 49.6 1.0 0.38
C 12.4 56.9 1.6 0.21
D 13.0 51.4 1.9 0.32
E 8.0 50.8 1.1 0.21
F 0.9 0.28
G 7.9 48.3 1.4 0.13
H 27.0 58.4 3.8 0.18
I 22.5 52.9 2.4 0.24
J 13.0 59.9 2.0 0.17
K 77.5 12.2 0.13
P. HORSRUD et al.: MECHANICAL AND PETROPHYSICAL PROPERTIES OF NORTH SEA SHALES 1015
Anisotropy
Due to the inherent anisotropic nature of clay plate-
lets (both texturally and mechanically), one would
intuitively expect shales to be anisotropic. Figure 8
shows the stress paths for two samples tested at the
same eective conning pressure. One of the samples
was drilled with the sample axis along the bedding and
the other was drilled with the sample axis normal to
the bedding. Figure 8 shows that the sample drilled
parallel to the bedding is both stier and stronger than
the one drilled normal to the bedding.
There are several factors which may aect the degree
of anisotropy. One would expect that the larger the
clay content, the larger the degree of anisotropy, but
this also depends on the mineral type. Both porosity
and depth should be of importance. As the depth
increases the clay platelets get compressed and more
and more aligned as the eective overburden stress
increases. The porosity also decreases, so one would
expect the degree of anisotropy to increase with burial
depth. A study of the acoustic response of the same
shales [26] showed that the shales could in general be
considered as transversely isotropic, with a P-wave vel-
ocity anisotropy in the range 0 to 25%. The P-wave
anisotropy was found to increase monotonically with
depth. The anisotropy also increased with clay content,
except for high clay content dominated by smectite.
Figure 9 is a secondary electron image of a sample of
shale D (Tertiary), illustrating the parallelism of clay
minerals. Figure 10 is from another part of the same
shale, showing the much more open texture of smec-
tite.
Shale A (Tertiary) is relatively shallow and has a
high porosity, but still the anisotropy is quite pro-
nounced. Both peak stress anisotropy and elastic ani-
sotropy is about 30%. The P-wave velocity anisotropy
for this shale was about 12%.
Fig. 7. Stress path for a CU test of shale K (Jurassic)
Fig. 8. Stress paths for shale A (Tertiary) samples drilled normal to
and parallel to sample bedding
Fig. 9. Secondary electron image from shale D (Tertiary), illustrating parallelism of platy clay minerals. The aky clay
P. HORSRUD et al.: MECHANICAL AND PETROPHYSICAL PROPERTIES OF NORTH SEA SHALES 1016
Anisotropy eects were, however, not the main sub-
ject of this study. The need for a complete description
of the anisotropy of shale mechanical properties will
depend on the application. Wells drilled at high devi-
ation, i.e. closer to horizontal, and also at larger
depths will thus be more susceptible to potential stab-
ility problems due to anisotropy. Complete mapping of
anisotropy eects requires a dedicated testing pro-
gramme with samples drilled at dierent angles.
Temperature eects
All the CU triaxial tests presented so far have been
run at room temperature. However, some tests have
also been run at the expected downhole temperature of
the shale. Figure 11 shows the resulting stress paths
for shale D (Tertiary).
At downhole temperature a signicant reduction
both of the stiness (025%) and strength (035%)
of the shale is observed. This has been reported for
other shales as well [19], and acoustic velocities have
also exhibited the same temperature
dependence [19, 27]. In general it is found that com-
pared to sandstone, the response of shales is more
dependent on temperature and less dependent on
pressure [19, 26, 27].
One possible eect of the temperature increase is to
improve the resaturation of the sample. This would
reduce capillary eects which can contribute signi-
cantly to the strength of rocks with small pores [10, 28].
Referring to Fig. 11 a signicant change of slope can
be observed when the temperature is increased.
Considering Equation (7), we note that this can be
achieved by a weakening of the rock frame and/or an
increase of the uid modulus. During undrained triax-
ial loading we have
Dp
/
=
1
3
Ds
1
Dp
f
(8)
We can thus calculate the change in pore pressure
(Dp
f
) from the slope of the curves in Fig. 11. Using
Equation (4) this provides an estimate for AB. The
tests at 208C yield AB = 0.24. Inserting this into
Equation (7) together with appropriate values for the
other parameters of this shale (j = 0.34,
K
fr
=1.7 GPa, a11) the uid modulus is estimated to
be 1.5 GPa. Note that we have assumed that the Biot
coecient is equal to one. Although we do not have
an exact value for K
s
, this is a reasonable assumption
for a weak shale like this. The pore water saturation
(S
r
) of the samples can now be estimated from
K
f
=
1
(S
r
=K
w
) (1 S
r
=p)
(9)
Fig. 10. Secondary electron image from shale D (Tertiary), showing the more open and less oriented texture of smectite clay
akes. Magnication: 4400. The white scale to the left is 10 mm
Fig. 11. Stress paths for shale D (Tertiary) samples tested at room
temperature and at 808C
P. HORSRUD et al.: MECHANICAL AND PETROPHYSICAL PROPERTIES OF NORTH SEA SHALES 1017
where K
w
is the gas-free water stiness (2.3 GPa) and
p is the pore pressure (3.5 MPa as average during the
elastic part of undrained triaxial loading). This yields a
saturation of 0.999. Even if the pores of the shale are
small and the capillary pressure eect thus can be
large, this reduction in saturation is much too small to
explain the observed dierence in stiness and strength
of shales with similar pore size distribution [10].
The observed reduction in acoustic velocities has
been discussed by Holt et al. [26]. They concluded that
changes in uid properties (density, compressibility)
could not explain the changes, as the velocity should
in fact increase with increasing temperature. The tem-
perature dependence is suggested to be related to coup-
ling eects between the uid and the solid phase. The
temperature increase will lead to a reduction in uid
viscosity, which again can result in a reduced velocity.
This should not aect the static mechanical properties
to the extent which is observed, unless the pore water
has a more fundamental eect on the shale properties.
It should be kept in mind that most of the pore water
in shale is not in a ``free'' state [29]. It is more or less
bound due to the small pore sizes and the electric
nature of the mineral surface. Heating results in ther-
mal expansion of water inside clay laminae, release of
water to the pore space and reduced interlaminar spa-
cing. This will result in altered properties of the struc-
tural water. It has been found [30] that the special
structure of water in small pores (5 nm) disappears at
about 708C. One possible explanation to the observed
eects is thus that bound water contributes signi-
cantly to both static and dynamic mechanical proper-
ties of shale.
IS SHALE A SOIL OR A ROCK?
In this paper we have used principles and methods
originating from both rock mechanics and soil mech-
anics applications. To discuss the relevance of this, let
us rst consider some denitions [31].
Clay is the term generally reserved for a material
which is plastic when wet and has no well-developed
parting along the bedding planes, although it may dis-
play banding.
Shale has a well-marked bedding-plane ssility, pri-
marily due to the orientation of the clay mineral par-
ticles parallel to the bedding planes. Shales do not
form a plastic mass when wet, although they may dis-
integrate when immersed in water.
Mudstone is a term used for rocks which are similar
to shales in their non-plasticity, cohesion and lower
water content, but lack the bedding plane ssility.
In this study, the term shale may include several of
the above denitions, in particular shale and mud-
stone. There are obvious similarities and dierences
between the clay (soil) and shale (rock):
. The main constituents of both are the clay min-
erals. What will characterise the response with respect
to permeability, water sensitivity, etc., will be very
much determined by the amount and type of clay min-
erals:
. Shales possess signicant cohesion which clays do
not.
. Shales are clearly anisotropic by nature.
The results presented in this study show that young
shales at relatively shallow depth have characteristics
which in many cases resemble a clay (e.g. low cohe-
sion, low stiness, high contents of swelling clay min-
erals, low degree of anisotropy). With increasing
depth, eects of compaction and diagenesis cause the
shale to deviate more and more from typical properties
and behaviour of a clay. With the rather broad de-
nition we have used for shale, shale is best described
as a rock which may be approximated to a soil if cer-
tain limiting conditions are met.
We have also shown that principles and methods
adopted from soil mechanics can be applied both to
test procedures and interpretation of test results as
there are fundamental similarities (mineralogy, per-
meability, water sensitivity). However, due to the
dierences, not all principles can be directly adopted.
Considering the interpretation of pore pressure re-
sponse as an example, we used principles which are
common in soil mechanics. Adopting the soil limit
directly would give erroneous results since the assump-
tion that jK
fr
WK
f
would already be implicit in the ex-
pressions. By including the basic expressions we can
still use the same methods and also explain dierences
in response which would otherwise be misinterpreted.
The eect of cohesion (stiness) compared to partial
saturation on the stress path is one example of this.
CONCLUSIONS
Several North Sea shales from a wide depth range
have been characterised petrophysically and mechani-
cally. The purpose of the characterisation has primar-
ily been for borehole stability evaluation, but the data
can also be used to provide a better understanding of
shale behaviour in dierent environments.
Some principles and methods adopted from soil
mechanics can be applied both to test procedures and
interpretation of test results as there are fundamental
similarities (mineralogy, permeability, water sensitivity)
between shale and clay. There are, however, also fun-
damental dierences (cementation, anisotropy). To a
varying degree this will introduce errors if the soil
limit is adopted directly.
The mechanical response of shales is sensitive to the
state of the test sample (e.g. degree of saturation, core
damage eects) and to the methods and principles
applied to test the shale. This implies that shale
samples, prior to testing, must be preserved in a man-
ner whereby loss of uids or contact with non-native
uids are avoided. For shale cores which have been
lifted from deep wells, resaturation by uid invasion
should not be attempted. Resaturation is best achieved
by reloading and reheating the sample.
P. HORSRUD et al.: MECHANICAL AND PETROPHYSICAL PROPERTIES OF NORTH SEA SHALES 1018
Shales covering several thousand meters of depth
vary signicantly in response and properties. There are
some trends which are of interest and which can be
very useful. Both strength and stiness increase with
depth. Young, Tertiary shales at depths less than
2000 m appear to have an estimated uniaxial compres-
sive strength in the range 515 MPa. Similar trends
can be found for the elastic properties, porosity and
mineralogy.
A signicant eect of temperature on the mechanical
properties of Tertiary shale has been observed.
Heating of the shale reduces the strength, stiness and
acoustic velocities signicantly. This study suggests
that this is a real eect and not an artefact due to, for
example, partial saturation during laboratory testing.
Although a complete explanation to this phenomenon
is not presented, the contribution of bound water to
the mechanical properties of shale is oered as one
contributing mechanism. An obvious implication of
such an eect is that laboratory tests should be per-
formed at downhole temperature, or tests at room
temperature somehow need to be corrected to down-
hole temperatures.
In addition to mechanical properties, some key pet-
rophysical properties have been presented and dis-
cussed. These are essential in evaluation of potential
borehole stability problems. This is primarily related
to how the shale may interact with the drilling uid
and the time-scale of this process. Some of these prop-
erties may be useful also as a general characterisation
tool, as at least some of them may be measured on
smaller samples, i.e. drill cuttings. This can reduce the
need for costly coring operations. Recommendations
have been presented both for core handling procedures
and test methods (e.g. porosity). Some test methods
are still subject to considerable uncertainty and there-
fore require further improvements (e.g. the per-
meability parallel to the lamination).
AcknowledgementsThe authors wish to thank the companies which
have participated in IKU's Shale Stability project (Amoco Norway
Oil Co., Den norske stats oljeselskap a.s, Elf Petroleum Norge AS,
Norsk Agip A/S, Norsk Hydro ASA, Norske Conoco AS, Saga
Petroleum ASA) for their nancial support and also for supplying
the cores to this study. We would also like to thank the many people
in IKU and SINTEF who have performed the measurements and
testing presented in this paper.
Accepted for publication 15 April 1998
REFERENCES
1. Steiger, R. P. and Leung, P. K., Quantitative determination of
the mechanical properties of shales. Proc. 63rd Ann. Conf. and
Exh., Paper SPE 18024. Houston, Texas, 1988.
2. Remvik, F. and Skalle, P., Shaleuid interaction under simu-
lated downhole conditions, and its eect on borehole stability.
Int. J. Rock Mech. Min. Sci. Geomech. Abstr., 1993, 30, 1115
1118.
3. Steiger, R. P., Advanced triaxial swelling tests on preserved
shale. Int. J. Rock Mech. Min. Sci. Geomech. Abstr., 1993, 30,
681686.
4. Hale, A. H., Mody, F. K. and Salisbury, D. P., Experimental in-
vestigation of the inuence of chemical potential on wellbore
stability. Proc. IADC/SPE Drill. Conf., Paper IADC/SPE 23885.
New Orleans, LA, 1821 Feb., 1992.
5. van Oort, E., Hale, A. H., Mody, F. K. and Roy, S., Critical
parameters in modelling the chemical aspects of borehole stab-
ility in shales and in designing improved water-based shale dril-
ling uids. Proc. SPE 69th Ann. Tech. Conf. and Exh., Paper
SPE 28309. New Orleans, LA, 2528 Sept., 1994.
6. van Oort, E., Ripley, D., Ward, I., Chapman, J. W., Williamson,
R. and Aston, M., Silicate-based drilling uids: Competent, cost
eective and benign solutions to wellbore stability problems.
Proc. 1996 IADC/SPE Drilling Conf., Paper IADC/SPE 35059.
New Orleans, LA, 1215 March, 1996.
7. Katsube, T. J., Mudford, B. S. and Best, M. E., Petrophysical
characteristics of shales from the Scotian shelf. Geophysics, 1991,
56, 16811689.
8. Best, M. E. and Katsube, T. J., Shale permeability and its signi-
cance in hydrocarbon exploration. The Leading Edge, March,
1995, pp. 165170.
9. Chenevert, M. E. and Sharma, A. K., Permeability and eective
pore pressure of shales. Proc. SPE/IADC Drill. Conf., SPE/
IADC 21918. Amsterdam, The Netherlands, 1114 March, 1991.
10. Schmitt, L., Forsans, T. and Santarelli, F. J., Shale testing and
capillary phenomena. Int. J. Rock Mech. Min. Sci. Geomech.
Abstr., 1994, 31, 411427.
11. Forsans, T. M. and Schmitt, L., Capillary forces: The neglected
factor in shale stability studies? Proc. Eurock `94. A. A.
Balkema, Rotterdam, 1994, pp. 7184.
12. Santarelli, F. J., Marsala, A., Brignoli, M., Rossi, E. and Bona,
N., Formation evaluation from logging on cuttings. Proc. 1996
European Petr. Conf., SPE 36851. Milan, Italy, 2224 Oct., 1996.
13. Nes, O.-M., Horsrud, P., Snsteb, E. F., Holt, R. M., Ese, A.
M., kland, D. and Kjrholt, H., Rig-site and laboratory use of
CWT acoustic velocity measurements on cuttings. Proc. 1996
SPE Eur. Petr. Conf., SPE 36854. Milan, Italy, 2224 Oct.,
1996.
14. Snsteb, E. F. and Horsrud, P., Eects of brines on mechanical
properties of shales under dierent test conditions. Proc. Eurock
`96. A. A. Balkema, Rotterdam, 1996, pp. 9198.
15. Santarelli, F. J. and Carminati, S. Do shales swell? A critical
review of available evidence. Proc. 1995 SPE/IADC Drilling
Conf., Paper SPE/IADC 29421. Amsterdam, 28 Feb.2 March,
1995.
16. Carminati, S., Brignoli, M., Di Marco, A. and Santarelli, F.
J., The activity concept applied to shales: Consequences for oil,
tunnelling and civil engineering operations. Int. J. Rock Mech.
Min. Sci.(Paper No. 038), 1997, 34, 34.
17. Santos, H., Diek, A., Roegiers, J.-C. and Fontoura, S., Can
shale swelling be (easily) controlled? Proc. Eurock `96. A. A.
Balkema, Rotterdam, 1996, pp. 99106.
18. van Oort, E., A novel technique for the investigation of drilling
uid induced borehole instability in shales. Proc. Eurock `94. A.
A. Balkema, Rotterdam, 1994, pp. 293308.
19. Horsrud, P., Holt, R. M., Snsteb, E. F., Svan, G. and
Bostrm, B., Time dependent borehole stability: Laboratory stu-
dies and numerical simulation of dierent mechanisms in shale.
Proc. Eurock `94. A. A. Balkema, Rotterdam, 1994, pp. 259266.
20. Charlez, Ph. and Heugas, O., Evaluation of optimal mud weight
in soft shale levels. Proc. 32nd U.S. Rock Mech. Symp., ed. J.-C.
Roegiers. A. A. Balkema, Rotterdam, 1991, pp. 10051014.
21. Head, K. H., Manual of Soil Laboratory Testing, Vol. 13. ELE
International, 1984.
22. Atkinson, J. H. and Bransby, P. L., The Mechanics of Soils. An
Introduction to Critical State Soil Mechanics. McGraw-Hill,
London, 1978.
23. Lashkaripour, G. R. and Dusseault, M. B., A statistical study
on shale properties: Relationships among principal shale proper-
ties. In Probablistic Methods in Geotechnical Eng., eds. Li and
Lo. A. A. Balkema, Rotterdam, 1993, pp. 195200.
24. Skempton, A. W., The pore pressure coecients A and B.
Geotechnique, 1954, 4, 143147.
25. Fjaer, E., Holt, R. M., Horsrud, P., Raaen, A. M. and Risnes,
R., Petroleum related rock mechanics. Developments in
Petroleum Science, Vol. 33. Elsevier, 1992.
26. Holt, R. M., Snsteb, E. F. and Horsrud, P., Acoustic vel-
ocities of North Sea shales. Proc. EAGE 58th Conf. and Techn.
Exh. Amsterdam, 37 June, 1996.
27. Johnston, D. H., Physical properties of shale at temperature and
pressure. Geophysics, 1987, 52, 13911401.
P. HORSRUD et al.: MECHANICAL AND PETROPHYSICAL PROPERTIES OF NORTH SEA SHALES 1019
28. Papamichos, E., Brignoli, M. and Santarelli, F. J., An exper-
imental and theoretical study of a partially saturated collapsible
rock. Mech. Cohesive-Frictional Mater., 1997, 2, 251278.
29. Baldi, G., Hueckel, T. and Pellegrini, R., Thermal volume
changes of the mineral-water system in low-porosity clay soils.
Can. Geotech. J., 1988, 25, 807825.
30. Derjaguin, B. V., Karasev, V. V. and Khromova, E. N., Thermal
expansion of water in ne pores. J. Colloid Interface Sci., 1986,
106, 586587.
31. Whitten, D. G. A. and Brooks, J. R. V., Dictionary of Geology.
Penguin Books, 1972.
P. HORSRUD et al.: MECHANICAL AND PETROPHYSICAL PROPERTIES OF NORTH SEA SHALES 1020

Das könnte Ihnen auch gefallen