2014 - Figueiredo Et Al. - Determination of The Stress Field in A Mountainous Granite Rock Mass

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 12

International Journal of Rock Mechanics & Mining Sciences 72 (2014) 3748

Contents lists available at ScienceDirect

International Journal of
Rock Mechanics & Mining Sciences
journal homepage: www.elsevier.com/locate/ijrmms

Determination of the stress eld in a mountainous granite rock mass


B. Figueiredo a,n, F.H. Cornet b, L. Lamas a, J. Muralha a
a
b

Portuguese Laboratory for Civil Engineering (LNEC), Lisbon, Portugal


Institut de Physique du Globe de Strasbourg (IPG-S-CNRS), Strasbourg, France

art ic l e i nf o

a b s t r a c t

Article history:
Received 17 June 2013
Received in revised form
20 July 2014
Accepted 28 July 2014
Available online 19 September 2014

The design of an underground hydroelectric power scheme in northern Portugal has required the
characterisation of the local stress eld. Nineteen hydraulic tests have been conducted in two, 500 m
deep, vertical boreholes. In addition twelve overcoring tests together with twelve at jack tests have
been performed from an existing adit located some 1.7 km away from the location of the hydraulic tests.
Results have been integrated into a stress model that takes into account both topography and tectonics
effects. Most of the data are consistent with a linearly elastic, gravity loaded model, provided a very soft
geomaterial is considered. This implies that the stress eld in this granite rock mass is controlled by
gravity alone and shear stress relaxation along faults and fractures but is unaffected by present-day
tectonics.
& 2014 Elsevier Ltd. All rights reserved.

Keywords:
In situ stress measurements
Regional stress eld
Elastic rock mass
Shear stress relaxation
Natural fracture network

1. Introduction
The re-powering scheme of the Paradela hydroelectric infrastructure developed on the Cvado River in Northern Portugal
involves a new 10 km long power conduit and a powerhouse
complex located about halfway in the conduit and 500 m below
ground level. It includes a new powerhouse cavern, a valves
chamber and a large surge chamber with several adits. The local
geological formation is mostly granite (Fig. 1).
The design of this new scheme requires a sound understanding
of the regional stress eld and this has prompted an in situ stress
measurement campaign. Results have raised important questions
on the relative inuence of topography and regional tectonics as
well as on that of the rock mass rheological characteristics. Such
issues have been frequently addressed in the literature [1,2] and
we describe here after results derived from a stress determination
strategy somewhat similar to that proposed for the International
Society of Rock Mechanics (ISRM) [3].
After identifying the objective of the stress determination
campaign, we present results obtained with three different methods, namely hydraulic tests in 500 m deep boreholes as well as
overcoring and at jack tests conducted at different locations.
Then we introduce the numerical model that has been developed
for integrating the various measurements in order to identify an
optimum solution. This model helps to discriminate the respective

n
Correspondence to: Uppsala University, Villavgen 16, Uppsala, Sweden.
Tel.: 46 739735500.
E-mail address: bruno.gueiredo@geo.uu.se (B. Figueiredo).

http://dx.doi.org/10.1016/j.ijrmms.2014.07.017
1365-1609/& 2014 Elsevier Ltd. All rights reserved.

contributions of gravity and tectonics and provides means to


determine the long-term rock mass rheological behaviour that
best ts observations.

2. Stress determination campaign


2.1. Objectives and design of the campaign
The design of the underground excavations planned for hosting
the new powerhouse as well as the valves and large surge system
implies a complete 3D characterisation of the stress eld at the
location of the excavations. But the design of the 10 km long
pressure tunnel requires only a sound evaluation of the minimum
principal stress magnitude all along the tunnel as well as an
estimate of its extreme local variations.
On site, two 500 m deep vertical boreholes (PD19 and PD23)
were available in the immediate vicinity of the planned underground powerhouse. Further a horizontal adit located some 1.7 km
from the vertical boreholes, with a 2.4  2.0 m2 rectangular cross
section, was excavated some 50 years ago and provided the
opportunity for further stress measurements (Figs. 13).
It was decided to run a combination of hydraulic fracturing (HF)
tests together with hydraulic tests on pre-existing fractures (HTPF)
in both vertical boreholes in order to constrain the direction and
magnitude for the three principal stress components at the
location of the excavation. Further the objective was also to
provide data on the spatial variability of these quantities along
the vertical direction.

38

B. Figueiredo et al. / International Journal of Rock Mechanics & Mining Sciences 72 (2014) 3748

Fig. 1. Layout of the Paradela II hydroelectric repowering scheme (courtesy of Energy of Portugal-EDP).

Fig. 2. Vertical cross-section AA' along the pressure tunnel showing the relative location of the adit with respect to the 500 m deep vertical boreholes PD19 and PD23.

order to combine various measuring techniques for this stress


evaluation, as recommended by the ISRM [3].
Locations of the various boreholes are shown on Figs. 2 and 3.
The 60 m deep boreholes PD1 and PD2 are 150 m apart and have
been drilled specically for overcoring measurements (Fig. 3). The
three locations SFJ1, SFJ2 and SFJ3 of at jack tests are also shown
on Fig. 3.
2.2. Hydraulic tests

Fig. 3. Vertical cross-section along the adit axis (above) and three dimensional
scheme of the adit (below) showing the location of overcoring and at jack tests.

Validity of these results for other locations would be established next, through further testing conducted in the adit. Hence it
was decided to run overcoring and at jack tests in the adit in

Hydraulic tests involved two different techniques, hydraulic


fracturing (HF) and hydraulic testing of pre-existing fractures
(HTPF), following the procedures described by Haimson and
Cornet [4].
For HF, a portion of a borehole free of pre-existing fractures is
isolated with a straddle inatable packer. The pressure is progressively raised in the isolated interval till a hydraulic fracture
develops at the so called breakdown pressure Pb. Then the fracture
is extended till it reaches zones outside the domain of inuence of
the borehole. When the fracture stops, the hydraulic injection
system is kept shut so as to monitor the subsequent pressure
decay for some time (a few minutes). This testing period is called
the shut-in period. At the end of shut-in, the system is shortly bled
off (a few seconds) and the subsequent pressure build-up is
observed for a few minutes. This part of the test is called owback and ends the rst testing cycle. This testing cycle is reproduced at least twice and sometimes more, when results show
some drifting from one cycle to the next (Fig. 4).
The pressure at which the fracture closes is called the instantaneous shut-in pressure Ps. It is equal to the normal stress
component acting on the fracture plane, i.e. the natural minimum
principal stress component when the hydraulic fracture has
propagated far enough from the borehole.
The HTPF procedure is very similar to the HF procedure except
that the isolated borehole portion is supposed to include one
single pre-existing fracture. The pressure is raised sufciently

B. Figueiredo et al. / International Journal of Rock Mechanics & Mining Sciences 72 (2014) 3748

Fig. 4. Typical curve for the interval pressure record versus time during a hydraulic
fracturing test. Pb is the breakdown pressure and Ps is the instantaneous shut-in
pressure.

slowly to insure a uniform pressure distribution within the


fracture close to the borehole. When the injection pressure reaches
the normal stress supported by the fracture, the fracture opens.
This quasi-static fracture opening operation alters all the components of the stress eld in the vicinity of the fracture except the
stress component normal to the fracture plane. Once the fracture
has been opened up to distances far enough from the well, the
instantaneous shut-in pressure of HTPF's yields an estimate on the
normal stress component that was supported before testing by the
fracture plane away from the borehole.
The hydraulic testing equipment used for these measurements
includes an inatable straddle packer system together with an
electrical imaging device (Cornet et al. [5]). The system is used rst
to log the borehole so as to identify the optimum location of test
intervals (homogeneous rock formation for HF, single isolated
fractures for HTPF). Then hydraulic tests are conducted and an
electrical image of the tested interval is produced just after testing.
Comparison of pre- and post-fracturing images provides the
required information for determining the geometry of the fracture
that has just been tested. The uncertainty on depths measurements is close to 0.5 m.
The stress normal to a tested fracture plane is directly related to
the shut-in pressure measured during each test cycle. Aamodt and
Kuriyagawa's method [6] was applied for identifying the maximum borehole pressure for which the fracture is completely
closed, i.e. the highest pressure for which ow obeys Darcy's law.
This pressure value is an underestimate of the normal stress
supported by the fracture at closure. Hayashi and Haimson's
method [7] was applied for determining the pressure at the end
of fracture propagation, just prior to fracture closing. This provides
an overestimate of the normal stress supported by the fracture. For
each test sequence (three or more shut-in readings), the minimum
value obtained with the Aamodt and Kuriyagawa method and the
maximum value obtained with Hayashi and Haimson's method
were used to dene the 99% condence interval for the normal
stress estimate. If the error on the normal stress component
estimate is assumed to obey a normal distribution, then the extent
of the 99% condence interval is equal to six standard deviations.
The geometry and orientation of the tested fracture planes have
been determined through a comparison between the oriented
electrical images obtained before and after hydraulic tests. The
fracture planes are recognised as sinusoids on electrical imaging
logs. Two sinusoidal curves, that completely cover the extent of
the identied fracture plane, have been drawn to identify extreme
values for the azimuth and inclination of the normal to the
fracture planes. From the scatter described by the two sinusoidal

39

curves, a 99% condence interval is dened for the two angles. In


borehole PD23, a technical difculty prevented the determination
of the tool orientation during logging. This difculty was overcome
by running a properly oriented high-resolution acoustic televiewer
(HRAT) log. A comparison with the electrical log provided the
proper orientation of all the hydraulically tested fractures for this
borehole.
Tables 1 and 2 present a summary of the hydraulic test results.
In these tables, z is the depth of tests, is the azimuth of the
normal to the fracture plane with respect to the North (positive
eastward), is the inclination angle of the normal to the fracture
plane with respect to the vertical direction, and n is the normal
stress measurement for the corresponding depth interval. The
standard deviations associated with z, , and n are z, ,
and n, respectively. In Table 2, the orientations marked with n
were only obtained with the HRAT log.
The tables show that eight out of the nineteen tests are
ambiguous because more than one fracture plane is observed in
the test interval. For test number 1 in borehole PD23, the normal

Table 1
Summary of the hydraulic test results obtained from borehole PD19.
Test

1
2
3

Type

5
6
7

HF
HF
HF
HTPF
HF
HTPF
HTPF
HF
HTPF

8
9
10
11
12

HTPF
HF
HTPF
HF
HF

Depth

Azimuth

Dip

Normal stress

z (m)

z (m)

(1)

(1)

(1)

(1)

n (MPa)

n (MPa)

471.8
455.5
450.1
450.4
442.1
442.2
436.3
414.9
393.4
393.9
394.1
394.3
379.3
335.6
293.1
279.8
164.6

0.5
0.5
0.5
0.5
0.5
0.5
0.5
0.5
0.5
0.5
0.5
0.5
0.5
0.5
0.5
0.5
0.5

108
18
126
281
133
303
108
133
270
284
277
50
88
119
14
22
320

4
7
5
6
5
5
6
4
4
3
5
5
3
3
2
4
7

87
90
90
60
90
42
32
90
66
79
52
38
57
61
44
90
86

2
2
2
1
2
2
2
2
2
2
2
2
2
2
3
2
2

10.3
9.0
7.8

0.4
0.2
0.2

9.0

0.2

8.9
7.1
7.3

0.2
0.1
0.3

5.6
6.7
7.5
5.8
2.6

0.2
0.1
0.3
0.1
0.3

Table 2
Summary of the hydraulic test results obtained from borehole PD23.
Test Type

Depth
z (m)

1
2
3
4

5
6
7

HF
490.7
HF
421.8
HTPF 420.9
HF
402.4
HTPF 402.5
HTPF 377.4n
377.8n
377.8n
378.4
377.9
377.9
377.7
HF
364.7n
HTPF 364.6
HF
356.8
HTPF 357.0n
HF
176.6

Azimuth
(1)

0.5
0.5

199
5
19
5
33
5
348
5
252
5
n
215
5n
n
106
5n
46n
5n
2
5
137
5
59
5
358 10
n
325
5n
252
5
269
4
33n
5n
243
3

0.5

0.5
0.5
0.5

Normal stress

(1) (1) (1) n (MPa) n (MPa)

z (m)

0.5

Dip

90
90
42
68
34
81n
12n
20n
35
22
24
37
81n
35
90
6n
77

2
2
2
2
3
3n
3n
3n
2
2
2
2
3n
2
2
3n
2

20.2
9.9

0.5
0.2

8.9

0.2

9.7

0.2

7.0

0.2

5.9

0.1

3.2

0.2

n
orientation obtained from the re-orientation of the electrical logs by using the
HRAT log.

40

B. Figueiredo et al. / International Journal of Rock Mechanics & Mining Sciences 72 (2014) 3748

stress magnitude is signicantly larger than all the other values.


However, the measurement is reproducible and satises all the
prerequisite conditions of validity. A detailed examination of both
the electrical and the HRAT images shows a clear inclined fracture
below the lower packer. In comparison to other post-fracture
images, no other fracture is observed within the tested interval.
This result strongly suggests that ow occurred within the inclined
fracture below the packer. In such conditions, the pressure applied
by the packer is also applied on some parts of the inclined fracture
such that its opening remains very limited in azimuth. The stress
analysis for such fractures has been discussed in [8], who showed
that the shut-in pressure in such congurations does not yield the
far-eld normal stress value. This result has not been included in
the data base for the stress determination.
For test number 8 conducted in borehole PD19, a strongly
altered zone was observed in the electrical imaging log, approximately 20 m above the test location. This zone dips steeply and is
not very far from the tested fracture. It affects most probably the
stress eld measured with test 8.
Finally, we note that tests number 7 in borehole PD19 and test
number 4 in borehole PD23 resulted respectively in four and seven
different fractures observed within the tested interval. These two
tests do not bring any signicant constrain to the solution. They
have not been considered for the determination of the optimum
solution.
Fig. 5 shows a plot of the normal stress magnitudes as obtained
with HF and HTPF as a function of depth as well as the 99%
condence limit associated with the measurements.
Fig. 6 shows a plot of the azimuth of the normal to the fracture
planes as obtained with HF tests only, as a function of depth as
well as the 99% condence limit associated with the measurements. In these tests, the fracture planes are parallel to the
borehole axis which demonstrates that, at these locations, the
borehole is parallel to a principal stress direction and that the
minimum principal stress is oriented perpendicularly to the borehole axis. Two different groups of data are noted. The solid line
represents the azimuth of fractures for which the measured
normal stress is close to the expected minimum stress magnitude
at the same depth. The dashed line represents a perpendicular set
of azimuths. Results shown in Figs. 5 and 6 reveal that the normal
stress magnitudes measured for sub-vertical fractures that are
nearly perpendicular to each other are yet very similar. This result
suggests that differences between the maximum and minimum
principal horizontal stress magnitudes are likely fairly small. This
is particularly noticeable for tests 2, 3 and 4 in PD19. The three
tests are only 10 m apart, yet the fracture of tests 2 is nearly

Fig. 5. Variation of the normal stress n magnitudes as a function of depth.

Fig. 6. Variation with depth of the azimuth of the normal to the fracture planes
obtained during hydraulic fracturing tests.

normal to that of test 3 whilst shut-ins for tests 2, 3 and 4 are very
similar.
2.3. Overcoring tests
The overcoring method is based on the stress relief principle. It
yields the complete stress state at the corresponding location
provided linear elasticity applies. For these measurements the
Stress Tensor Tube (STT), initially developed by Rocha and Silvrio
[9], has been used for it provides all tensor components through a
single overcoring operation.
The STT strain measurement device is a hollow epoxy resin
cylinder with an outer diameter of 35 mm, an approximate length
of 20 cm, and a thickness of 2 mm (Fig. 7). The cell has ten
electrical resistance strain gauges embedded in positions normal
to the faces of a regular icosahedron, which enables sampling of
the displacement eld in all corresponding directions [10]. The cell
includes a metal capsule that houses the data acquisition unit as
well as a thermocouple. Readings of all strain gauges and of the
temperature are conducted at xed time intervals (60 s) and then
are stored in the local memory.
A test consists of the following operations: (1) drilling of a
140 mm diameter borehole to the depth of interest; (2) drilling a
concentric 37 mm diameter borehole from the bottom of the large
diameter hole in which the STT cell is inserted and glued against
the walls; and (3) resuming the drilling of the large diameter hole
to a depth compatible with a complete stress relief around the cell.
After overcoring, the rock core together with the STT is
recovered. Fig. 8 provides an example of the variation of strains
with time at the location of the ten strain gauges during an
overcoring test. The dashed line represents the variation of the
temperature with time. Strain readings are taken before and after
overcoring when the temperature is stabilised. The difference
between these values corresponds to the strains that result from
the overcoring stress relief.
Elastic constants of the rock core have been determined
through pressure tests conducted on the cores with a biaxial
chamber in which a radial hydraulic pressure p is applied. Three
loading and unloading cycles were performed. The rst cycle
reached a maximum pressure of 2 MPa. For the other two cycles,
a maximum pressure of 6 MPa was applied. The deformation at
the location of the ten strain gauges that resulted from the applied
pressure has been measured.
The model used for interpreting the STT tests results assumes that
the rock is homogeneous, linearly elastic and isotropic; that the
length to diameter ratio for the cell is high; and that the stiffness of
the hollow cylinder is signicantly less than the stiffness of the rock.

B. Figueiredo et al. / International Journal of Rock Mechanics & Mining Sciences 72 (2014) 3748

41

Fig. 7. STT cell and data acquisition unit.

Table 3
Principal stress (I, II, III) magnitudes and orientations obtained from interpreting
the overcoring tests.
Borehole Test Elevation (m) Principal stress
magnitudes (MPa)
I

Fig. 8. Strain versus time curves obtained during an overcoring test.

In this model, given the arrangement of the strain gauges, the strain
i at the location of strain-gauge i (i1,..,10) is linearly related to the
components of the in situ stresses j (j1,..,6):

i aij j ;

where aij are the components of a matrix with ten rows and six
columns that depends on the elastic parameters (elastic modulus E
and Poisson's ratio ) of the overcored core [10].
With the measured ten strains that result from a biaxial loading
(1 2 p and 3 4 5 6 0), a least squares method is
applied to solve the matrix Eq. (1) and determine the elastic
constants. By considering the set of six biaxial tests obtained in
each borehole, the following mean values (E and ) and associated
standard deviations (E and ) for the elastic constants have been
determined: borehole PD1: E 60.8 GPa, E 12.9 GPa, 0.30,
0.05; borehole PD2: E 57.7 GPa, E 9.5 GPa, 0.35,
0.11.
The standard deviations of the elastic modulus are approximately 21% and 16% of the mean, whereas the standard deviations
for Poisson's ratio are approximately 17% and 31% of the mean for
boreholes PD1 and PD2, respectively. This analysis outlines a
signicant dispersion, which may be compared to results from
uniaxial compression tests conducted on cores: E 44.6 GPa,
E 9.0 GPa, 0.25, 0.06.
For these tests, the standard deviations of the elastic modulus
and Poisson's ratio are approximately 20% and 24% of the mean,
respectively. Hence Poisson's ratio values as determined from
biaxial testing are found to be considerably larger than those
obtained with uniaxial tests. This may be attributed to the
development of microcracks normal to the axes of the rock
samples collected during overcoring. Indeed, it is well known that
under specic loading conditions such microcracking leads to
disking [11].
Once the strains resulting from overcoring have been measured
and the elastic constants have been determined, a least squares
method is applied for determining the six components of the
stress tensor in a co-ordinate system associated with the STT cell.
The magnitude and orientation of the principal stresses (I, II, III)
obtained in each test are presented in Table 3. In this table, the
orientations are described by two angles; the rst angle is the
direction of the principal stress component with respect to the

II

III

Principal stress
orientations (1)
I

II

III

PD1

1
2
3
4
5
6

618.2
614.5
599.5
598.9
570.1
569.3

6.4
7.0
8.7
9.3
10.6
11.3

5.6
3.4
3.4
6.9
8.2
6.5

4.7
3.2
3.3
5.9
7.0
6.0

325/66
249/86
354/86
36/64
94/54
26/73

85/12
147/1
231/2
188/23
200/11
150/10

179/20
56/4
141/4
283/11
298/34
242/14

PD2

1
2
3
4
5
6

619.5
617.9
598.6
597.9
580.5
579.2

11.9
8.0
6.2
5.1
 1.9
6.0

6.9
6.3
2.8
4.6
 3.6
4.2

5.9
4.3
2.0
4.3
 4.0
3.4

94/79
328/75
324/79
285/63
267/77
289/39

260/11
229/2
177/10
24/5
69/12
147/45

350/3
139/15
86/6
116/26
160/4
36/20

North, and the second angle is the inclination with respect to a


horizontal plane.
The table shows that the maximum principal stress (I) is subvertical and that the other two principal components (II and III)
are sub-horizontal and of similar magnitude. Hence the apparent
dispersion for the direction of sub-horizontal principal stress
components is not signicant. In borehole PD1, the sub-vertical
component measurements are signicantly greater than those
measured in borehole PD2, a feature which cannot be explained
by the depth difference between the tests. It has been often
observed that overcoring tests lead to high stress values in the
direction parallel to the borehole axis. This has been explained by
the large deformations that occur in this direction as a result of the
glue yield caused by the heat generated during the drilling
operation [12]. It may also result from an improper elastic model
because of the nonlinearity induced by the microcracking phenomenon already mentioned.
In borehole PD2, stress values obtained for test 5 at 580.5 m
above sea level are very questionable, considering the values
obtained with neighbouring tests. This test result has not been
considered for further analysis.
2.4. Small at jack tests
The small at jack testing method [13,14] is based on the stress
relief principle, and implies a partial stress relief followed by stress
compensation. In this technique, two pairs of pins are placed on
the rock surface, and the initial distance between the pins is
measured with digital transducers. Then, a 10 mm-thick slot is cut
perpendicularly to the rock surface, using a 60 cm diameter
diamond disk, till a 27 cm depth is reached. Due to the partial
stress relief, deformations in the direction normal to the slot occur
and the distance between the pins decreases. Subsequently a
circular at jack, consisting of two thin metal plates welded
together, is inserted into the slot and pressurised until the distance
between the pins is restored. During the test, the variation of the
relative displacements between the two pairs of pins caused by

42

B. Figueiredo et al. / International Journal of Rock Mechanics & Mining Sciences 72 (2014) 3748

the applied pressure is recorded. Fig. 9 shows a plot of displacement versus pressure as obtained during a typical test. Generally,
some non-elastic behaviour is observed and non-recoverable
displacements are detected after unloading.
The pressure required to restore the initial position of the pins
is called the cancellation pressure, and it is assumed to be equal
to the stress component normal to the slot plane. In this determination, only the loading phase observed during the rst cycle was
considered.
Results are presented in Table 4 where d is the distance
between the test location and the adit's entrance, is the azimuth
of the normal to the slot with respect to the North, is the
inclination of the normal to the slot plane with respect to the
vertical direction and n is the normal stress component.
For tests number 1, 5 and 9, somewhat similar sub-vertical
stress components have been measured, with values equal to
8.9 MPa, 9.9 MPa and 9.9 MPa, respectively. But comparison
between the horizontal stress components parallel to the adit
direction and measured during test numbers 3, 7 and 11, shows
that the normal stress value measured during test number 3 is
signicantly larger than that measured during the two other tests.
Actually, for test number 3, because of a technical problem, it was
not possible to run the measurement immediately after the slot
had been cut. The rock was observed to have a creeping behaviour.
The pressure versus displacement curve shows a signicant slot
closure during the time interval that separates the end of the
cut from the beginning of pressurisation. Consequently a high

Fig. 9. Typical pressure versus displacement curves obtained during small at


jack tests.

cancellation pressure was required for restoring the initial position


of the pins within the same time span as for the other tests.

3. The optimum stress model


3.1. Data integration and solution identication
When in situ stress measurement data of various kinds have
been acquired at diverse locations, a common procedure for
identifying the regional stress eld is to develop a numerical
model that ts best all observations [1519]. Then this numerical
model helps evaluate the stress tensor at the various locations of
interest.
The aim is to determine the model that minimises differences
between a number of observations and predictions from the
model. This determination requires a model denition, a denition
of the mist for describing the discrepancy between observed and
predicted values, and a normative measure of the mist for
quantifying the residuals for all observations [20].
3.2. Model's denition
We call model a set of unknown parameters that is looked for
in order to t observations. The objective is to dene a simple
enough model so that the number of constraining data is larger
than the number of unknown parameters (degrees of freedom) for
the model. We consider the volume of interest as perfectly
dened, i.e. topography and all structural elements like the shape
of the adit are considered to be well known. This volume is lled
with an equivalent continuous geomaterial the mechanical characteristics of which are part of the model denition. The corresponding set of differential equations that describe its mechanical
behaviour are solved with explicit nite differences using the
software FLAC3D [21].
The mesh constructed with the FLAC3D software (Figs. 1 and 10)
is composed of 600,000 elements. It is ner above sea level, with
cubic 25 m-sided elements. Below sea level, the elements are
50 m  50 m  100 m. Boundary conditions imposed on the vertical
boundaries are part of the model denition, as discussed here after,
but a condition of no vertical displacement is imposed on the
horizontal basal boundary.
In order to determine the inuence of volume size on the stress
determination, more specically the inuence of the proximity of
vertical boundaries to the points of interest, various geometries
have been considered when performing the rst gravitational
analysis (g 9.81 m s  2). Results show that the 5 km long and
3 km wide region shown in Fig. 1 is sufciently large for obtaining

Table 4
Results from the small at jack tests.
Location

Test

d (m)

(1)

(1)

n (MPa)

SFJ1

1
2
3
4

447
448
437
446

0
110
110
290

0
45
90
45

8.9
6.4
9.4
3.7

SFJ2

5
6
7
8

332
333
331
332.5

0
290
290
110

0
45
90
45

9.9
2.6
2.0
6.4

SFJ3

9
10
11
12

278
275
279
280

0
290
290
110

0
45
90
45

9.9
3.2
4.1
3.0
Fig. 10. Mesh of the FLAC3D model.

B. Figueiredo et al. / International Journal of Rock Mechanics & Mining Sciences 72 (2014) 3748

43

reliable estimates. Indeed, for larger regions, the maximum difference between the computed three principal stress magnitudes at
the locations of the various tests gets smaller than 0.5 MPa. Within
the domain of interest elevation varies between 315 m and 1030 m
above sea level. A 2.5 km extension below sea level has been
assigned in the vertical direction so that boundary conditions at
the base of the model do not affect the stress evaluations where
topography is signicant.
Various models have been considered. For the rst model, the
elastic constants that characterise the geomaterial are assumed to
be those measured on the cores (2 parameters model) and only
vertical gravity loading is considered (no displacement normal to
the vertical boundaries). The adit is ignored and the data set
includes only HF and HTPF results. The second model is similar to
the rst model except for a horizontal tectonic stress, which is
introduced so as to decrease the minimum computed mist. The
third model is similar to the rst one but the Poisson's ratio for the
rock mass is taken as an unknown. Further the data set includes
both hydraulic tests and overcoring tests. In the fourth model, only
gravity loading is considered together with the optimum Poisson's
ratio identied with model 3 but the adit is introduced so as to
determine the t with data gathered close to the adit walls.

Evaluation of the uncertainty on the stress measurements is based


on considerations on the dispersion of the measurements rather
than on a rigorous error analysis procedure. We considered the
various pairs of overcoring tests that have been conducted (i.e.
tests conducted within 5 m from one other). For each pair of tests
we determined the mean difference between principal stress
components and then we determined the mean value for all the
pairs. This has led to a 1.5 MPa dispersion estimate, which is the
value chosen for characterising the uncertainty on the various
principal stress components. With this approach, no concern is
given to the systematic bias that may result from the linear elastic
hypothesis.
The hydraulic testing and overcoring methods are of different
natures and concern different rock volumes. This has been taken
into consideration for the denition of the global mist function
by introducing weighting parameters that take into account the
volume, or the area, involved by a given measurement for each of
the method. Further each measurement has been weighted
according to the relative value of its mist with respect to the
global value that may be used for characterising stress determinations conducted with hydraulic tests alone or overcoring tests
alone. The general mist HFOC
for model Mq may be expressed as
q

3.3. Denition and measurement of mist

q HFOC i HF i HF j OC j OC ;

For the ith hydraulic normal stress measurement, the mist HF


i
may be expressed as

i HF

i
i
n;mes  n;calc j
;
i
i
n f

where in;mes and in;calc are, respectively, the measured and


calculated normal stresses obtained at the location of the ith
i
hydraulic test, n is the uncertainty on the normal stress measurei
ment and f is the uncertainty on the normal stress estimate
associated with uncertainties on the orientation of the fracture
plane selected for the ith test [22].
For ambiguous tests with two or more observed fracture
planes, the fracture plane has been selected to be that which
yields the smallest difference between measured and calculated
normal stress. Uncertainties on the normal stress measurements
and on the fracture plane orientation determinations correspond
to the 99% condence interval for the measurements. Uncertainties on the normal stress component due to the uncertainty on the
fracture plane orientation determination have been estimated
using the FLAC3D code. Values have been found to be of the same
order of magnitude as those of uncertainties associated with the
shut-in pressure measurements.
For overcoring data, the uncertainties are associated with the
strain measurement technique, the orientation of the strain gauges
in the three dimensional space and the determination of elastic
properties of the overcored cores. The mist OC
that has been
j
used is simplied and is dened as

j OC

j jmes  jcalc j
;
j jbh

where jmes and jcalc are respectively the measured and calculated
value for the jth principal stress components as obtained at the
j
location of the kth overcoring test (j 3k  2 p; p 0,1,2); is the
j
uncertainty associated with the stress measurement and bh is the
uncertainty on the computed stress component because of uncertainties on the location of the test (mostly associated with
orientation of the borehole axis).
Because overcoring tests were conducted within a small depth
range (approximately 60 m) in vertical boreholes, the uncertainty
associated with the orientations of these boreholes was neglected.

i1

k1

where M and N are, respectively, the total number of hydraulic


tests and of overcoring tests and the weight factors for hydraulic
and overcoring data, HF
and OC
i
j , are given by

i HF

OC
j

AHF
REV

V OC
V

REV

HF
i
;
HF
min

OC
j
;
OC
min

where AHF and AREV denote respectively the measurement area and
the area involved in the representative elementary volume of the
rock mass (REV) for hydraulic testing; VOC and VREV are the
corresponding notations associated with the overcoring technique
(measurement volume and REV volume, respectively); and HF
min
and OC
min are respectively the minimum values for the mists
obtained when the stress determination is derived only from
hydraulic tests or only from overcoring tests, i.e. minimisation
for the sums of Eqs. (2) or (3), respectively.
The area involved by hydraulic testing depends on the area of
the opened fracture. It depends on many parameters such as the
injected ow rate but also uid losses through the walls of the
fracture. It was set somewhat arbitrarily equal to 1 m2, given the
injection tests characteristics. The volume involved by overcoring
measurements was set equal to the average volume of the
resulting hollow rock cylinder. The REV for the rock mass was
set equal to 1 m3 (i.e., area 1 m2). Thus, the suggested global mist
gives more weight to hydraulic data than to overcoring data.
Two types of normative methods are commonly used to evaluate
the mist between a model and a set of data: the l1-norm considers
the sum of the absolute values of the differences between observations and predictions; whilst the l2-norm considers the sum of the
squares of the mists. The l2-norm is generally chosen when all
uncertainties obey a normal distribution [23]. But in the present case
some of the uncertainties are not Gaussian, such as for example the
linear elastic hypothesis for the overcoring measurements. Hence the
l1-norm has been chosen.
Once a global minimum has been identied (minimum of
HFOC), the limits of the 90% condence level for a posteriori

44

B. Figueiredo et al. / International Journal of Rock Mechanics & Mining Sciences 72 (2014) 3748

Fig. 11. Variation of the magnitudes of the normal stresses obtained by hydraulic testing (n,mes) and with the FLAC3D model (n,calc) run with gravity loading only in
boreholes (a) PD19 and (b) PD23 as a function of depth, when the Poisson's ratio is taken equal to 0.25.

evaluations may be estimated by


1=2
1:645 =2  1
M N1=2 M N HFOC
HFOC

min ;
90%
M N  W

a component associated with tectonic (n,tect) loading:


7

where W is the number of unknown parameters of the model used


to describe the regional stress eld (Parker and Mcnutt [23]).
3.4. Solution with gravity loading alone and elastic parameters from
tests on core samples
A rst stress calculation was conducted to evaluate the inuence of topography on stresses at the location of hydraulic tests, on
the assumption that the rock mass is linearly elastic. Its elastic
characteristics have been assumed to be identical to those of the
rock cores tested in the laboratory (E 45 GPa, 0.25). The rock
mass density was set equal to 2650 kg/m3. No optimisation was
undertaken. A comparison between measured and computed
values is shown on Fig. 11.
For ambiguous tests with two observed fracture planes, computed values concern the plane for which the difference between
observed and computed value is the smallest. For most tests,
differences between measured and calculated values are found to
be much larger than uncertainties on the measurements. So, this
model has not been optimised and the possibility that tectonic
stresses affect the massif has been investigated.
3.5. Testing the existence of tectonic stresses
The objective here was to explore the possibility of improving
the t between model and observations by introducing so called
tectonic stress components, i.e. horizontal stress components, on
the vertical boundaries of the model.
Several assumptions are made: (i) the rock mass exhibits a
linear, isotropic and elastic behaviour; (ii) the total stress eld may
be decomposed into gravity and tectonic components; (iii) the
vertical component is equal to the weight of the overlying material
and only of gravitational origin; and (iv) with exception for the
zones close to ground level where topography effects are important, the tectonic stresses are considered independent of depth.
The normal stress (n,mes) measured on each tested fracture
plane is decomposed into a component due to gravity (n,grav) and

n;mes n;grav n;tect :

Unit normal (Sxx, Syy) and shear stress (Syy) components are
introduced in the model: uniform displacements are imposed on
the lateral boundaries of the model in order to yield unit
horizontal stress components for elements in contact with the
basal boundary [24]. As a result, non-uniform and balanced stress
distributions are generated at the lateral boundaries that simulate
the inuence of topography effects on the stress eld.
The normal stress magnitudes n,Sxx, n,Syy and n,Sxy at the
location of each tested fracture plane due to unit tectonic stress
components Sxx, Syy and Sxy are computed. The normal stress
magnitudes due to tectonic loading (n,tect) is estimated as a linear
combination of the response to unit tectonic stresses:

n;tect A n;Sxx B n;Syy C n;Sxy ;

where coefcients A, B and C are unknown parameters that


characterise the tectonic stress eld. Substituting Eq. (9) into
Eq. (8) yields

n;calc n;grav A n;Sxx B n;Syy C n;Sxy :

10

The right side of Eq. (10), n,calc, is introduced in the mist


denition [2] for determining the coefcients A, B and C. The
following values have been obtained:
A 4:9; B 5:1; C 0:0

11

A comparison between measured and computed normal stress


(n) when combining tectonic and gravity loadings is shown in
Fig. 12. Differences between measured and computed values for
hydraulic tests are found to be larger than uncertainties on the
measurements for approximately 60% of the tests. Further we note
that the values for A, B and C that best t the results are such that
parameter C is null, whereas parameters A and B are similar. This
corresponds to horizontal compressive normal stresses of similar
magnitudes and to zero shear stresses.
This is in agreement with hydraulic data collected at the
bottom of the boreholes which suggests that both horizontal
principal stress magnitudes at this depth are very similar to each
other. But, this does not agree with focal plane solutions of local
earthquakes that show an orientation for the regional tectonic

B. Figueiredo et al. / International Journal of Rock Mechanics & Mining Sciences 72 (2014) 3748

45

Fig. 12. Variation with depth of the normal stress magnitudes as measured by hydraulic testing (n,mes) and computed with the FLAC3D model (n,calc) for the combined effect
of gravity and tectonics in boreholes (a) PD19 and (b) PD23.

stress approximately N1351E [25,26] and we conclude that tectonic stresses are likely not effective at the depth of the tests.
3.6. Identication of the rock mass rheological characteristics
The analysis conducted in this section assumes a linearly elastic
behaviour for the rock mass. Because changing the elastic modulus
does not induce changes in the stress eld for a homogeneous rock
mass, only an increase of the Poisson's ratio value is tested. The
FLAC3D model has been used in the inversion of the Poisson's ratio
value to minimise the mists dened by Eqs. (2)(4). The results
presented in Table 5 show that a Poisson's ratio value of 0.47
provides the best t between results obtained from both hydraulic
and overcoring data and those computed with the model. Eq. (7)
was used (W1) to calculate a 90% condence interval for the
minimum of the global mist. The interval obtained in this way
corresponds to a Poisson's ratio value that ranges between 0.45
and 0.49.
The proles of the measured and calculated normal stresses
(n) due to gravity loading with a 0.47 Poisson's ratio are shown in
Fig. 13. The difference between measured and computed normal
stress magnitudes is less than three standard deviations for
approximately 75% of the tests which is considered acceptable
given the many simplifying assumptions implied by this model.
Comparison of measured and computed principal stress values
at the location of overcoring tests in the PD1 and PD2 boreholes
is shown in Fig. 14. Lines represent values computed with the
FLAC3D model, and dots represent overcoring results. The position
above sea level of the adit axis is also represented. At depth of the
tests, the calculated maximum principal stress is parallel to
the boreholes axis and the other two components are subhorizontal and similar magnitude. Further the FLAC3D model
results also show that the stresses magnitude at location of
boreholes PD1 and PD2 are similar. This agrees with the results
of the stress measurement results. In addition, approximately 80%
of the measured and computed principal stresses are in satisfactory agreement (the difference between both values is less than
1.5 MPa).
Hence, hydraulic and overcoring data may be explained by a
linearly elastic rock mass under gravitational load provided the
Poisson's ratio for the equivalent material is considerably larger
than that measured during short-term uniaxial compression tests.

Table 5
Variation of the mist value with the Poisson's ratio.

HF

OC

HFOC

0.25
0.35
0.45
0.46
0.47
0.48
0.49

56.7
37.1
13.2
11.8
10.9
12.2
12.6

43.9
38.5
35.3
35.1
34.9
34.3
34.6

295.0
126.3
16.4
13.2
11.5
14.2
14.9

3.7. Analysis of the test results obtained close to the adit


In this model exploration, we investigate how the linearly
elastic model dened here above ts the at jack stress measurements and overcoring tests run the closest to the adit walls.
Firstly, a comparison of measured and computed normal stress
values was made at the location of the small at jack tests. Since
the stresses provided by the small at jack technique do not
correspond to the far-eld stress components because they are
inuenced by the existing adit, a three-dimensional numerical
model of nite differences was developed using the code FLAC3D
(Fig. 15) for the interpretation of the small at jack test results.
This model is a 30 m  30 m  5 m solid and includes the rectangular cross section of the adit. Note that the elastic modulus has no
effect on the solution. A Poisson's ratio value of 0.25 obtained from
uniaxial compression tests conducted on intact cores was considered. Variations of the Poisson's ratio between 0.25 and 0.47
resulted in a maximum variation of 0.5 MPa for the calculated
normal stresses. In addition to this uncertainty, an uncertainty of
1 MPa on the normal stress measurements and an uncertainty of
0.5 MPa on the calculated normal stresses due to the uncertainty
on the geometry of the adit and on the at jack orientations were
assumed. This results in a 2.0 MPa for the maximum uncertainty
on the normal stress values. The model is used to calculate the
normal stresses at the location of the small at jack tests by
considering the far-eld stress tensor components obtained with
the large-scale model shown in Fig. 10, providing a Poisson's ratio
of 0.47 is taken for the rock mass. A comparison between the
normal stresses obtained this way and the stresses actually
measured with the small at jacks is shown in Fig. 16. The gure

46

B. Figueiredo et al. / International Journal of Rock Mechanics & Mining Sciences 72 (2014) 3748

Fig. 13. Variation with depth of the normal stress magnitudes as measured by hydraulic testing (n,mes) and computed with the FLAC3D model with 0.47, considering
gravity effects only, in boreholes (a) PD19 and (b) PD23.

Fig. 14. Variation with elevation above sea level of the magnitude of the principal stresses (I, II, III) obtained by overcoring testing (OC) and with the FLAC3D model (FM)
with 0.47 considering gravity effects only.

Fig. 15. Three-dimensional model used for interpretation of the small at


jack tests.

Fig. 16. Comparison of the magnitudes of the normal stresses obtained by small at
jack technique (n,mes) and with the FLAC3D model (n,calc) with 0.47 considering
gravity effects only.

B. Figueiredo et al. / International Journal of Rock Mechanics & Mining Sciences 72 (2014) 3748

47

Table 6
Comparison of the magnitudes of the principal stresses (I, II, III) obtained by overcoring testing (OC) and with the FLAC3D model (FM) with 0.47 considering gravity
effects only.
Borehole

PD1

PD2

Test

Elevation (m)

619.5

617.9

618.2

614.5

Method

OC
FM
OC
FM
OC
FM
OC
FM

Principal stress magnitudes (MPa)

Principal stress orientations (1)

II

III

II

III

6.4
6.0
7.0
6.1
11.9
5.2
8.0
5.2

5.6
4.5
3.4
4.6
6.9
3.8
6.3
3.8

4.7
3.8
3.2
3.9
5.9
3.0
4.3
3.1

325/66
225/71
249/86
225/70
94/79
193/58
328/75
193/60

85/12
127/3
147/1
127/3
260/11
293/6
229/2
293/6

179/20
36/19
56/4
36/20
350/3
27/31
139/15
26/29

shows that for approximately 75% of the small at jack tests, the
discrepancy between measured and calculated normal stresses is
smaller than the uncertainty on the normal stresses. In test
number 3, the discrepancy between the measured and calculated
normal stress value is understandable because of the already
mentioned technical problem that occurred during the test.
Secondly, the stress eld without the adit was calculated with the
model shown in Fig. 10, at the location of the shallow overcoring tests
done in boreholes PD1 and PD2. To analyse the inuence of the adit,
these stress components were used as boundary conditions in the
model shown in Fig. 15, and the stresses at the location of the shallow
overcoring tests were calculated.
Comparison between measured and calculated principal stresses
magnitude and orientation is presented in Table 6. In this table, the
orientations are described by two angles; the rst angle is the
direction of the principal stress component with respect to the
North, and the second angle is the inclination with respect to a
horizontal plane.
The table shows that the FLAC3D model results obtained at depth
of the shallow overcoring tests done in borehole PD2, are not in
satisfactory agreement with the stress measurements provided by
overcoring (the difference between measured and calculated principal
stress values is higher than 1.5 MPa). For these two tests, large subvertical stress components were measured which is also visible from
the results of at jack tests number 5 and 9. Results from FLAC3D
model and overcoring tests show that at depth of the tests done in
boreholes PD1 and PD2, one of the principal stresses is sub-vertical
and the other two components are sub-horizontal and of similar
magnitude. Due to this the dispersion observed on the direction of
sub-horizontal principal stresses is not signicant. This analysis concludes that, with an elastic solution, the stress perturbation induced by
the adit at the location of the shallow overcoring tests is negligible,
and it is not possible to explain the stresses measured by the two
shallow overcoring tests in borehole PD2 and at jack tests number
5 and 9. The source of this local stress heterogeneity is likely linked to
a local fracture zone but this possibility has not been explored further.

4. Discussion
Results from the FLAC3D model show that most of the data are
consistent with a linearly elastic equivalent geomaterial, with properties that correspond to a much softer material than suggested by
laboratory tests on cores. At this scale, the rock mass includes an
important natural fracture network (see Fig. 1, and images from
electrical and HRAT logs) and the material lling these various
fractures and faults has been subjected to creep over time so that
very small shear components are left along these planes of weakness.
Creep effects may be assessed by using a visco-elasto-plastic
model. However, our objective here is not to simulate the time
transients of the deformation but rather to extrapolate results of local

stress measurements to locations of interest for the design of openings


required by the future hydroelectric infrastructures. Our analysis
shows that choosing an equivalent linearly elastic material with very
soft elastic properties fulls this need. Further, changing the elastic
modulus does not induce any change in the stress eld for a
homogeneous rock mass, so that only an increase of the apparent
Poisson's ratio value has revealed necessary. This equivalent linearly
elastic model has been found to t approximately 75% of the in situ
stress measurements. It may be used for the design of the underground opening required by the hydroelectric repowering project.
Interestingly, this linearly elastic behaviour for the granite rock
mass is markedly different from the short-term one considered for
both interpreting overcoring tests and simulating results from the at
jack tests, which depend on the stress concentration associated with
the adit. The adit is 50 years old so our results suggest that, for this
time scale, the elastic parameters as derived from tests on core yield
satisfactory results except for the local stress measurements conducted
for tests 1 and 2 in borehole PD2 and nearby at jack tests number
5 and 9.
This suggests two possibilities for modelling this stress eld. We
may consider an elasto-viscous model with very high viscosity so that
it behaves elastically at the scale of 50 years but behaves like a uid at
the scale of million years. Or we consider a linearly elastic model with
a network of fracture sets with sufciently diverse orientations and
little long-term shear strength so that the long-term stress eld is subhydrostatic away from topographic limits. The existence of local stress
concentrations, either positive or negative, as observed for very few of
the hydraulic tests as well as for two overcoring and two at jack tests
are in qualitative agreement with models of shear motions [27]. This
suggests that the second proposition is more appropriate than the rst
one. And this has strong consequences for the design of the pressure
tunnel. Indeed, the safe design of the unlined hydraulic pressure
tunnel requires an absence of signicant water leakage as would be
observed if some hydraulic jacking occurs along the tunnel. This
requires that the planned water pressure be smaller than the smallest
minimum principal stress value encountered along the tunnel. As a
consequence, the design of the hydraulic pressure tunnel should be
made in terms of the minimum expected value for the minimum
principal stress magnitude instead of the average value, implicit in the
solution presented in Section 3.4.

5. Conclusion
Several in situ stress measurements were conducted at the
Paradela II site for the design of an underground reinforcement
power scheme that includes a large powerhouse cavern and a
hydraulic pressure tunnel. The measurements include nineteen
hydraulic tests in two 500 m deep vertical boreholes, twelve
overcoring tests in two 60 m deep vertical boreholes drilled from

48

B. Figueiredo et al. / International Journal of Rock Mechanics & Mining Sciences 72 (2014) 3748

an existing adit and twelve small at jack tests in the walls of


the adit.
Analysis of hydraulic and overcoring data demonstrates that
one principal stress component is sub-vertical within most of the
volume that was tested. The other two components are subhorizontal and of similar magnitude. Some local zones of heterogeneity were encountered, which have been attributed to preexisting inclined fractures, as observed on the electrical imaging
logs. Sub-equality between the two horizontal principal stress
components and the local heterogeneities explain the dispersion on orientations observed for horizontal principal stresses
direction.
75% of the measurements are consistent with a linearly elastic
rock mass under gravity loading if the Poisson's ratio of the
equivalent geomaterial is taken equal to 0.47, a value quite larger
than the value measured on cores and used for interpreting both
overcoring and at jack tests.
It has been concluded that the observed large-scale stress eld
results from the shear stress relaxation over a large number of preexisting fractures and faults with very variable orientations rather
than from the shear stress relaxation to be expected with a very
long-term viscous behaviour for this rock mass. Indeed only slip
along weakness planes is consistent with observed large-scale
stress relaxations and local stress concentrations that are either
positive or negative.
Acknowledgements
This paper is a part of the PhD thesis by Bruno Figueiredo at the
Strasbourg University. The work was funded by the Portuguese
Laboratory for Civil Engineering (LNEC) and the Foundation for
Science and Technology (FCT) PhD grant SFRH/BD/68322/2010.
Authorisation by EDP-Energies of Portugal to publish the stress
measurement results obtained at the Paradela II site are
acknowledged.
References
[1] Amadei B, Stephansson O. Rock stress and its measurements. London: Chapman & Hall; 1997.
[2] Stephansson O, Zang A. ISRM suggested methods for rock stress estimation
Part 5: establishing a model for the in situ stress at a given site. Rock Mech
Rock Eng 2012;45(6):95569.
[3] Hudson JA, Cornet FH, Christiansson R. ISRM suggested methods for rock
stress estimationPart 1: strategy for rock stress estimation. Int J Rock Mech
Min Sci 2003;40:9918.
[4] Haimson BC, Cornet FH. ISRM suggested methods for rock stress estimation
Part 3: Hydraulic fracturing (HF) and/or hydraulic testing of pre-existing
fractures (HTPF). Int J Rock Mech Min Sci 2003;40:101120.

[5] Cornet FH, Doan ML, Fontbonne F. Electrical imaging and hydraulic testing for
a complete stress determination. Int J Rock Mech Min Sci 2003;40:122542.
[6] Aamodt RL, Kuriyagawa M. Measurement of instantaneous shut-in pressure in
crystalline rock. In: Zoback MD, Haimson B, editors. Hydraulic fracturing stress
measurements. Washington, DC: National Academy Press; 1983. p. 13942.
[7] Hayashi K, Haimson BC. Characteristics of shut-in curves in hydraulic fracturing stress measurements and the determination of the in situ minimum
compressive stress. J Geophys Res 1991;96(B11):1831121.
[8] Cornet FH, Li L, Hulin JP, Ippolito I, Kurowski PF. The hydromechanical
behaviour of a single fracture: an in situ experimental case study. Int J Rock
Mech Min Sci 2003;40:125770.
[9] Rocha M, Silvrio A. A new method for the complete determination of the
state of stress in rock masses. Gotecnique 1969;19(1):11632.
[10] Pinto JL, Cunha AP. Rock stresses determinations with the STT and SFJ
techniques. In rock stress and rock stress measurements. Lulea: Centek
1986:25360.
[11] Li F, Schmitt DR. Drilling induced core fractures and in situ stress. J Geophys
Res 1998;103:522539.
[12] Ask D. New developments of the integrated stress determination method and
application to the spo hard rock laboratory, Sweden. (PhD thesis). Stockholm: Royal Institute of Technology; 2004.
[13] Rocha M, Lopes B, Silva N. A new technique for applying the method of at
jack in the determination of stresses inside rock masses. In: Proceedings of the
1st international congress rock mechanics. Lisbon; 1966, 2: p. 5765.
[14] Habib P, Marchand R. Mesures des pressions de terrains par l'essai de verin
plat. Ann Instit Tech Bt Trav Publ, Sols Found 1952;58:96671.
[15] Sousa L, Martins C, Lamas L. Development of the techniques of measurement
and the interpretation of the state of stress in rock masses. Application of the
Castelo do Bode tunnel. In: Proceeding of the 5th international congress IAEG,
Buenos Aires; 1986.
[16] Tonon F, Amadei B, Pan E, Frangopol DM. Bayesian estimation of rock mass
boundary conditions with applications to the AECL underground research
laboratory. Int J Rock Mech Min Sci 2001;38:9951027.
[17] Hart R. Enchancing rock stress understanding through numerical analysis. Int J
Rock Mech Min Sci 2003;40:108997.
[18] Matsuki K, Nakama S, Sato T. Estimation of regional stress by FEM for a
heterogeneous rock mass with a large fault. Int J Rock Mech Min Sci
2009;46:3150.
[19] Li G, Mizuta Y, Ishida T, Li H, Nakama S, Sato T. Stress eld determination from
local stress measurements by numerical modelling. Int J Rock Mech Min Sci
2009;46:13847.
[20] Gephart W, Forsyth W. An improved method for determining the regional
stress tensor using earthquake focal mechanism data: application to the San
Fernando Earthquake Sequence. J Geophys Res 1984;89:930520.
[21] Itasca. FLAC3D. Version 4.0. User's Manual. Minneapolis: Itasca Consulting
Group; 2009.
[22] Cornet FH, Yin JM. Analysis of induced seismicity for stress determination and
pore pressure mapping. Pure Appl Geophys 1995;145:677700.
[23] Parker RL, Mcnutt MK. Statistics for one-norm mist measure. J Geophys Res
1980;85:442930.
[24] Figueiredo B, Cornet FH, Lamas L, Muralha J. Effect of topography on
distribution of in situ stresses due to gravity and tectonic loadings at Paradela
site. In: Proceedings of the EUROCK 2012 rock engineering and technology for
sustainable underground construction. Stockholm; 2012.
[25] Heidbach O, Tingay M, Barth A, Reinecker J, Kurfe D, Mller B. The release of
the World Stress Map (available online at www.world-stress-map.org);
2008.
[26] Bezzeghoud M, Borges JF. Focal mechanisms of earthquakes in Portugal (in
Portuguese). Fs Tierra 2003:2294515 2003:22945.
[27] Scotti O, Cornet FH. In situ stress elds and focal mechanism solutions in
central France 1994. Geophys Res Lett 1994;21:23458.

You might also like