You are on page 1of 71

IN-SITU STRESS

MEASUREMENT

ITA-WORKING GROUP 2
9/19/22
TABLE OF CONTENTS

1. INTRODUCTION.....................................................................................................................................1
2. SCOPE....................................................................................................................................................1
3. THE CONCEPT OF STRESS AND STRESS TENSOR.........................................................................2
4. STRESS CLASSIFICATION FOR A ROCK MASS.................................................................................3
5. RELEVANCE, ACCURACY AND UNCERTAINTY..................................................................................4
5.1. RELEVANCE.............................................................................................................................................4
5.2. ACCURACY AND UNCERTAINTY...........................................................................................................6
5.3. INFLUENCE OF GEOLOGICAL FEATURES ON THE STATE OF STRESS...........................................7
6. METHODS FOR MEASUREMENT OF IN-SITU STRESS......................................................................9
6.1. CLASSIFICATION.....................................................................................................................................9
6.2. HYDRAULIC METHODS.........................................................................................................................11
6.2.1. Hydraulic Fracturing (HF)..........................................................................................................................11
6.2.2. Hydraulic Tests on Pre-existing Fractures (HTPF)...................................................................................19
6.3. RELIEF METHODS.................................................................................................................................22
6.3.1. Borehole relief methods............................................................................................................................22
6.3.2. Overcoring methods – Overview..............................................................................................................23
6.3.3. Surface relief methods..............................................................................................................................33
6.3.4. Relief of large rock volumes. Back analysis.............................................................................................35
6.4. JACKING METHODS..............................................................................................................................36
6.4.1. Procedure.................................................................................................................................................36
6.4.2. Assumptions, limitations and advantages of the flat jack.........................................................................37
6.4.3. Test interpretation – Flat jack...................................................................................................................38
6.4.4. Stress computations in the flat jack method.............................................................................................38
6.5. CORE-BASED METHODS......................................................................................................................39
6.5.1. Strain recovery methods...........................................................................................................................39
6.5.2. Anelastic strain recovery (ASR)................................................................................................................40
6.5.3. Differential strain curve analysis (DSCA)..................................................................................................40
6.5.4. Acoustic Emissions (AE) Method..............................................................................................................41
7. APPLICATIONS.....................................................................................................................................42
7.1. MINING....................................................................................................................................................42
7.1.1. Mine design and long-term planning.........................................................................................................43
7.1.2. Increase production..................................................................................................................................43
7.2. TUNNELS................................................................................................................................................43
7.2.1. Numerical analysis....................................................................................................................................43
7.2.2. Selection of the optimal shape..................................................................................................................43
7.2.3. Optimization of excavation sequences.....................................................................................................44
7.2.4. Support and lining design.........................................................................................................................44
7.2.5. Prediction of rockbursts............................................................................................................................45
7.3. CAVERNS...............................................................................................................................................45
8. CASE STUDIES....................................................................................................................................45
8.1. CASE HISTORY – PORCE III HYDROPOWER PLANT – HYDRAULIC FRACTURING AND
OVERCORING TESTS..................................................................................................................................45
8.1.1. General Characteristics............................................................................................................................45
8.1.2. Hydraulic Fracturing..................................................................................................................................46
8.1.3. Results......................................................................................................................................................48
8.1.4. Conclusions..............................................................................................................................................50
8.2. CASE HISTORY – IN SITU STRESS MEASUREMENTS AT THE UNDERGROUND RESEARCH
LABORATORY (URL), MANITOBA, CANADA..............................................................................................53
8.2.1. General Characteristics and geological framework..................................................................................53
8.2.2. In situ stress measurement.......................................................................................................................55
8.2.3. Results......................................................................................................................................................56
8.2.4. Conclusions..............................................................................................................................................58
8.3. CASE HISTORY IN SITU STRESS MEASUREMENTS FOR UNLINED PRESSURE TUNNEL DESIGN
IN QUEBEC, CANADA...................................................................................................................................58
8.3.1. General Characteristics and geological framework..................................................................................58
8.3.2. In situ stress measurement campaign......................................................................................................59
8.3.3. Results......................................................................................................................................................60
8.3.4. Conclusions..............................................................................................................................................63
9. REFERENCES......................................................................................................................................64
1. INTRODUCTION

During the past decades, important advances and breakthrough research have been made in the field of rock
mechanics, as a response to the growing demand for electric energy, new transportation networks and raw
materials that are transformed trough industrial processes into modern day consumer goods.

In the field of geotechnical engineering, such advances refer in particular to the development of techniques for
measuring the natural state and properties of the rock mass. Of particular interest is the measurement of the
in-situ state of stress of the rock mass, given its influence in the overall behavior of underground excavations.
Acknowledging the relevance stresses have on the rock mass in the excavation of different types of
underground structures (e.g. tunnels, mines, radioactive waste deposits and caverns), Work Group No. 2, of
the International Tunneling Association, has prepared a document in which the more relevant aspects
regarding the conceptual principles, techniques and methods developed for evaluating the in-situ state of
stress have been gathered and summarized.

To such effect, this document has been structured into nine (9) chapters, where the first two chapters refer to
an introduction and a general scope on the theory and practice of stress measurements, including a case
history. Subsequently, in order to encompass a complete framework on stress measurement, it was deemed
appropriate to dedicate the first pages to the concept of “stress”, by citing its main characteristics and
classification. In Chapter 5, the importance, accuracy and uncertainty inherent to the in-situ stress
measurement are discussed. In section 5.1, special emphasis is made on the sequence to be followed, in
order to define a stress model. Chapter 6 refers to the analysis of geological structures (e.g. faults, joints,
veins), considered highly relevant, and the first step in the process leading to the assessment of the state of
stress, as an accurate interpretation of the geologic characteristics serves as reference for the planning and
execution of stress measurement tests and ancillary activities.

Chapters 7 and 8 constitute the central core of the document, in which the most commonly used methods for
evaluating the in-situ stress state for civil and mining purposes are explained. Firstly, classification criteria for
such methods are discussed, as well as general procedures, their advantages and limitations, and lastly, their
main applications are explained.

Finally, Chapter 9 addresses case studies, where the particular experience of WG2 members will be reviewed
in order to select the most representative and interesting cases where the investigation of the state of stress
was a governing parameter in the design of a particular underground excavation.

2. SCOPE

Measuring stresses in a rock mass is a rather complex task requiring the interaction of different engineering
fields, trained staff and special equipment, in all of the various techniques used. In this document, main focus
is placed on the methods and tools currently available for measuring the stress tensor in the field. The main
purpose is to illustrate, in a general manner - however delved into enough - the main characteristics of such
methods, basic theory, advantages, assumptions and limitations, along with practical cases where such
techniques have been utilized.
3. THE CONCEPT OF STRESS AND STRESS TENSOR

It is common to correlate the term “stress” with “pressure”, that is, the force per unit area; however, such
simple definition does not consider other relevant features required to for a rigorous concept: Hudson &
Harrison (1997) Hudson et al (2003) and Weijemars (2011), explain that the “force per unit area” concept is not
an adequate definition, since stress is neither a scalar nor a vector quantity, but a tensor, which involves three
components to fully define it: magnitude, direction, and the plane along which the stress is acting; hence,
Hudson & Harrison (1997) emphasize that “pressure” is a particular stress state, that will be explained later.

In order to give a more structured definition regarding the term “stress”, the forces acting on a body should be
considered. In first place, it is well known that to reach a state of equilibrium in a solid, the internal and external
forces over a selected free body must be equal in magnitude but opposite in direction; any unbalance of such
forces will result in deformations. Depending upon the free body analyzed, all the forces acting over it could be
considered as external forces.

From this basic concept related to the acting forces over a solid, and based on physics and mathematical
concepts, in order to define the complete state of stress acting at a given point, nine unknown quantities are
required be known (actually, six unknowns, given that by the laws of equilibrium, three unknowns are
numerically equal to other three). The unknowns are the normal and shear stresses, applied over a particular
point, with a specific orientation. Written in matrix form, the stress tensor is defined as ():

y
yz
yx xz
zyz zx x
(a)xy (b)

Figure 1. a) Stresses acting at a point. (Positive convention). b) Stress tensor

It is not the purpose to present herein details of the math defining the concept of stress tensor. Extensive
literature published on this subject in available, such a publications by Valliapan (1981), Cornet (1993),
Amadei & Stephansson (1997), Hudson & Harrison (1997).

Six independent variables or stress components, therefore, define the state of stress of a (rock mass) particle,
which will include a given orientation. The definition of such variables constitutes the main objective for the in-
situ measurement of stress.

The state of stress of particular interest in geomechanics is that in which the analyzed element is oriented in
such way, that no shear stresses are exerted over it, but only normal stresses. The planes along which only
normal stresses are present are known as principal planes, and the normal stresses acting on such planes are
known as principal stresses (). The relevance of defining the principal stresses and the planes along which
they act, lies on the fact that upon applying some algebra, it is possible to obtain the state of stress at any
given orientation, and vice versa. From a mathematical standpoint, the magnitude and direction of the principal
stresses correspond to the eigenvalues and the eigenvectors of the stress tensor, respectively.

3 1

2

(a) (b)

Figure 2. a) Principal stresses at a point. (Positive convention). b) Stress tensor

Finally, upon addressing back to the difference between pressure and stress, Hudson & Harrison (1997) point
out that the “pressure” should be referred only to a particular state of stress where all normal stresses are
equal and no shear stresses are present, so that the pressure at any given point can be fully expressed by a
unique value (scalar quantity), as opposed to stress, which requires six independent components in a given
coordinate system, or the magnitude and orientation of the principal stresses, to be entirely defined.

4. STRESS CLASSIFICATION FOR A ROCK MASS

Materials present in the earth’s crust, such as rock and soil, are constantly subjected to different types of
stress, the magnitude and direction of which depend on their origin or source. Several attempts have been
made to classify stresses acting on a rock mass; however, there is no compromise as such. A simple and
useful approach is given by Amadei & Stephansson (2003) who state that stresses in a rock mass may be
divided into two main groups: the in-situ stresses and the induced stresses. In-situ, or virgin stresses, are
stresses acting on a material free of any external disturbance. Induced stresses, on the other hand, are the
result of both human activity, such as drilling and excavating, and the alteration of original conditions of the
material induced by physical processes, such as drying and swelling.

In recent years, Zang and Stephansson (2010) made a proposal regarding stress types and their classification,
which follows a slightly different approach: here, a scheme separates the stresses types into three levels,
where the first classifies stress in two main groups: stresses acting on a solid rock mass, and stresses
originated by a disturbance, either man-induced, such as a tunnel, or by a geological structure exerting
influence over the manmade structure, such as a shear zone near a tunnel.

The second level is composed by the different types of in-situ stresses, classified according to their source:
in-situ stresses may be gravitational, tectonic or residual, where tectonic stresses may be classified according
to their domain, or area of influence, where stress may be considered uniform. This last distinction
corresponds to the third level within this proposal. A more detailed description of this classification is illustrated
in Figure 3.
The purpose of this document is, then, to discuss the methods and techniques and used in current practice to
determine the in-situ stresses.

Figure 3. Stress Classification scheme. After Zang and Stephansson (2010)

5. RELEVANCE, ACCURACY AND UNCERTAINTY

5.1. RELEVANCE

Large projects have a significant impact on the community - among which are those featuring an underground
excavation - and undergo a number of stages, each of which is assessed, both economically and technically,
under different levels of accuracy and complexity. As the project progresses, more and better quality
information is required. However, in tunnel engineering, regardless of the project’s stage, knowledge of the
ground properties is a key aspect for the work to be executed. In each new stage of an underground project,
the site characterization is an essential step for the development of the study, which must be fed with more
detail and additional data. Thus, in the early stages, where the lack of information is not uncommon, the main
source of knowledge regarding the in-situ stresses in the area of interest may come from either one or all the
following sources:

 Geological and geomorphological features. (i.e. faults, dikes, topography, glacial events)
 Data bases
 Boreholes

A geological model built from the above information is known as Best Estimate Stress Model (BESM), as per
Zang & Stephansson (2010), considered as a first step, among others to follow, in order to ascertain what they
have entitled a Final Rock Stress Model (FRSM). A Final Rock Stress Model implies that all available data
have been gathered, processed and evaluated, resulting in a more realistic scenario regarding the state of
stress of the rock mass, that will be ultimately used as input data for the design of a particular underground
excavation. The general outline proposed by Zang & Stephansson (2010) is shown in Figure 4.

Figure 4. Generation of the Final Rock Stress Model. After Zang and Stephansson (2010)

After defining the BESM, the next step implies the measurement of stresses, or SMM, as referred to in Figure
4. This activity may be carried out by use of various methods and technics. These methods will be discussed
ahead, as they constitute the main topic of this document; however, it is worth emphasizing the close link
between site characterization and the need for more and good quality information, for the project to evolve
smoothly from one stage to the next.

The next step implies the gathering of information from previous stages, to be validated either through
numerical analyses - by use of state-of-the-art software - or by means of a particular algorithms, as referenced
in Zang and Stephansson, (2015). Once this Integrated Stress Determination (ISD) has been conducted, the
Final Rock Stress Model may be defined.

Based on the above, regardless of the purpose for which a tunnel is built (as part of the civil infrastructure or in
mining operations) the knowledge of the state of stress condition is of outmost relevance. In this respect, it is
well established in underground excavations, that in-situ stresses play a governing role in defining geometry
and shape, excavation sequence, and orientation of an underground excavation (Amadei and Stephansson,
1997).

Any engineering work carried out in a rock mass necessarily implies modifying the original state of stress; any
such change may lead to an unstable condition of the opening, even to the point of failure. Hence, it is of
special interest to evaluate if there may be a stress concentration where the rock strength is exceeded, leading
to failure or phenomena such as slabbing, overbreaking or popping. In the traditional practice of tunnel design
and construction, more attention is usually placed on the crown instead of the walls. However, in the case of
large caverns, the opposite occurs, as such structures usually have high and long walls, as compared to its
width, wherefore horizontal stresses could induce large deformations here rather than in the crown.

Particularly relevant is the case of deep underground caverns in hydropower projects, which stability and
potential failure mechanisms are strongly related to, and dependent on, the orientation and magnitude of the
principal stresses, as discussed by Lu et al (2010), and Gonzalez de Vallejo et al (2004). Therefore, important
design decisions, such as orientation, excavation and support installation sequences depend on the state of
stress, where inadequate decisions could imply major changes to general project layout, leading to extensions
in time for completion and possible overrun costs. Accordingly, the longitudinal axis of large underground
caverns should be oriented parallel the major principal in order to reduce high deformations in the walls and
therefore additional support resulting from redistribution of stresses caused by the excavation. Nilsen &
Thidemann, (1993) presented detailed studies, referenced by Lu et al (2010), which indicate that for rock with
high and anisotropic in-situ stresses, the optimum orientation for a cavern is obtained when the longitudinal
axis is at an angle varying between 15° and 30° with respect to the horizontal projection of 1.

5.2. ACCURACY AND UNCERTAINTY

In spite of the latest technology, standard procedures and experience gained from a number of projects, the
accuracy obtained in the measurement of in-situ stresses has its limitations (Amadei and Stephansson,1997).
There is no mechanism to compare the values obtained from a single measurement of the state of stress. The
alternative has been to report results in terms of intervals, and a number of tests performed, in order to
compare results. The uncertainty in stress measurement may stem from three different sources: the
anisotropic nature of stress, as stress may vary from point to point; uncertainty, inherent to the measurement
procedures; and uncertainty in the analysis of data and an adequate interpretation of results.

Of these three, the natural uncertainty is that of greatest concern, given the intrinsic variability of parameters,
such as strength and deformation, of a material such as rock. Hudson & Feng (2015) have identified a series
of stress perturbation factors responsible for the variation of the in-situ stresses, such as heterogeneity,
anisotropy, fractures, and the influence of a free surface. However, the theory proposed by Anderson,
endorsed by Hudson & Feng (2015), asserts there are no shear or normal stresses acting on this free face,
meaning that a free surface is a principal plane. This free surface can be either natural or man-induced; in an
excavation carried out for engineering purposes (tunnels, caverns or mines), a significant alteration of the pre-
existing in-situ stresses around the excavation occurs, along with deformations and possible water inflow.
Hudson et al (2009) describe the stress state for such case, as shown in the figure below

Figure 5. Stress tensor at a free surface, after Hudson & Feng (2015)

Amadei & Stephansson (1997) provide measures to reduce uncertainties associated with the measurement of
stress, as listed below:

- Discard obvious mistakes in measurements (e.g. values outside of range).


- Compare results of several tests through different methods in adjacent areas.
- Use statistical techniques to evaluate results.
- Monitor as many test conditions as possible. Monitored variables should include air temperature,
humidity, drill fluid characteristics, etc.
- Analyze if the scatter of data could be due to topographic, geologic or anisotropic effects.
- Lab test instruments and equipment under similar conditions to be utilized in the field.

5.3. INFLUENCE OF GEOLOGICAL FEATURES ON THE STATE OF STRESS

As discussed in Chapter 5, available geological, geomorphological and borehole core data should be
analyzed, prior to any activity regarding the measurement of stress, in order to formulate the Best Estimate
Stress Model (BESM), as defined by Zang & Stephansson (2010). A propos, Zoback, Mercier & Vergely (2002)
give a number of indications the geological environment provides, regarding the state of stress state at a point
of particular interest, which allows the engineer to prepare a plan and make decisions regarding field
arrangement of equipment to be used for measuring the state of stress. Zoback (2011) mentions two types of
geological indicators of stress: slickenside or gouge material in faults and discontinuities, and the orientation of
igneous dikes. In general, analysis of rock mass fractures should be a first step towards the formulation of a
stress model.

Fractures in a rock mass may be used as an indication of the orientation of stress: in the case of tensile
cracks, the minor principal stress acts perpendicularly to the plane of largest deformations. However, to make
an adequate assessment, the axial direction of the cracks require be known, which may be somewhat
complicated. In Figure 6 the aforementioned concept is schematically illustrated.

Figure 6. Orientation of various fracture types with respect to the principal stresses.

In spite of the usefulness in the analysis of cracks, the analysis of geological faults in the area of interest is
perhaps the most useful tool for determining stress orientation. Studies undertaken by Anderson E.M. in the
mid 20th Century are quite relevant, as they provide a useful framework in order to identify the directions of the
principal stresses. Thus, in a normal failure, the major principal stress is, in essence, vertical, whereas in a
reverse fault, such stress acts over a horizontal plane, as illustrated in Figure 7. Table 1 summarizes the stress
state according to the failure type. It is of note that the classification system for geological faults presented
herein is based on the actual slip of the blocks separated by the fault.

In the case of igneous dikes for instance, propagation of fractures is controlled by the orientation of the minor
principal stress (3), that is, 3 is oriented perpendicularly to the direction of propagation of the fracture, which
is the basis of the hydraulic fracturing technique.

Table 1. Relative stress magnitudes and faulting regimes. After Zoback (2011)
Regime Stress
  
Normal fault v hmax hmin
Strike-slip fault hmax v hmin
Reverse fault hmax hmin v

Figure 7. Faults types and principal stresses direction. Modified after Mercier & Vergely (2002)

Amadei and Stephansson (1993) point out an often error is incurred in when the vertical stress due to
overloading is assumed as one of the principal stresses. In mountainous terrains close to valleys - as shown
Figure 8 - the principal stresses tend to be parallel and normal to the surface, although, to the extent where the
analysis is performed to a greater depth and a state of stress similar to that obtained for a horizontal surface is
observed.

Figure 8. Topographic influence in the orientation of the in-situ stresses state


The implications this topographic effect has on the design of underground works have been widely studied and
verified by several researchers: Broch (1984) asserts that modifications to the state of stress due to
topography plays a governing role in the design of tunnels and shafts, on issues such as alignment and final
lining requirements. Research has been carried out in the development of analytical models to determine the
state of stress in steep reliefs: perturbation methods and conformal mapping are two examples. However,
current practice tends to use numerical modeling for such analyses, in which finite elements, boundary
elements, or finite differences software are available.

6. METHODS FOR MEASUREMENT OF IN-SITU STRESS

An important issue related to the measurement of stresses, either in-situ or in the laboratory, lies on the fact
that stresses cannot be measured directly, for which their determination and evaluation depend on the
interpretation of displacements or strains, or by the induced fractures when a pressure is applied (Wittcke
2013). Accordingly, a series of methods have been developed in order to assess the state of stress, at least in
an approximate manner, given its relevance in the design of underground excavations. As discussed later,
there are a significant number of techniques related to the in-situ stress evaluation; nevertheless, only those
related to a direct measuring and relevant within the tunnel industry will be considered herein.

6.1. CLASSIFICATION

There are different criteria for the classification of in-situ stress measurement methods: one is related to the
location where the measurement is conducted, according to the operational type, as per Ljunggren et al
(2003). The classification based on this criterion is summarized in Table 2. This table includes the volume of
rock involved in the measurement.

Table 2. Methods for rock stress measurement based on operational type (Ljunggren et.al, 2003)
Category Method Volume (m³)
Hydraulic fracturing 0.5 - 50
Overcoring 10-3 – 10-2
Methods performed in boreholes HTPF(Hydraulic teats on pre-existing 1 – 10
fractures)
Borehole breakouts 10-2 - 100
Strain recovery methods 10-3
Methods using drill cores Core-discing 10-3
Acoustic Methods 10-3
Jacking methods 0.5 – 2
Methods performed on rock surfaces
Surface relief methods 1–2
Earthquake focal mechanism 109
Analysis of large scale geological structures
Fault slip analysis 108
Other Relief of large volumes (Back analysis) 102 - 103

A second classification discussed by Ljunggren et al (2003), states that the available methods for determining
the in-situ stresses could be classified in two main groups: the first comprises methods that alter the in-situ
state of stress in the rock; such alteration may be induced either by inducing deformations, or by the opening
of cracks. Examples are the overcoring and hydraulic fracturing methods, considered the most common
techniques currently used. The second group comprises methods which are based on observation of the rock
behavior itself, with no particular intervention, that is, the material is not subjected to any load or displacement.

Table 3. Methods for rock stress measurement based on the alteration - perturbation of the host rock.
Category Method
Hydraulic methods: e.g. Hydraulic fracturing, HTPF
Methods that disturb the in-situ rock conditions Borehole stress relief methods: e.g. Overcoring
Surface relief methods
Statistics of measured data (database) Strain
recovery methods
Methods that DO NOT disturb the in-situ rock Core-discing
conditions Acoustical Methods
Relief of large rock volumes (Back analysis)
Earthquake focal mechanism

On the other hand, Stephansson & Zeng (2015) establish that stress measurements techniques may be
classified from the physical mechanism, experimental technique and ultimate borehole depth in five different
categories:

 Category 1: mechanisms related to rock fracture as applied to boreholes.


 Category 2: mechanisms related to elastic strain relief due to coring.
 Category 3: mechanisms related to crack-induced strain relief in drill cores.
 Category 4: borehole seismic logging or indirect methods
 Category 5: techniques based on physical properties of pre-existing fault zones in the earth’s crust
and related earthquakes.

Methods included within each one of these categories are mentioned in the table below.

Table 4. Classification scheme for the in-situ stress measurements methods proposed by Zeng & Stephansson
(2010)
Category Methods
- HF
- HTPF
Mechanism related to rock fracture as applied to
- Sleeve fracturing
boreholes
- Drilling induce tensile fractures
- Borehole breakouts
- Surface relief methods
Mechanisms related to elastic strain relief due to
- Borehole relief methods
coring
- Relief of large rock volumes
- Inelastic strain recovery (ASR)
- Differential strain rate analysis (DRA)
Mechanism related to crack-induced strain relief in - Differential strain analysis (DSA)
drill cores - Differential wave velocity analysis (DWVA)
- Wave velocity analysis (WVA).
- Acoustic emissions (Kaiser effect)
Borehole seismic logging or indirect methods (wave - Shear-wave polarization
Category Methods
- Shear wave splitting
propagation methods) - Analysis of Stonely waves

- Fault plane solutions (FPS)


Techniques based on physical properties of pre- - Focal mechanisms of earthquakes
existing fault zones in the Earth’s crust and related • Natural seismicity (NS)
earthquakes • Induced seismicity (IS).

6.2. HYDRAULIC METHODS

As observed in Tables 1 and 3, hydraulic methods are conducted within boreholes. Zeng and Stephanson
(2015) outline the general procedure for this method: a section of the borehole is isolated by means of
inflatable packers; then, fluid pressure (usually water) is applied along the isolated section, and gradually
increased, until either existing fractures are opened, or tensile fractures are induced in the rock. The fluid
pressure required to change the initial state of fractures in rock at a given depth is measured, to finally
associate it with the in-situ state of stress. The direction of the in-situ stresses is deduced by direct
observation, or measuring the orientation of the hydraulically induced, or open fractures. The most widely used
hydraulic methods are: hydraulic fracturing or hydraulic jacking of pre-existing fractures (HTPF), as explained
below.

6.2.1. Hydraulic Fracturing (HF)

This technique, also known as hydrofracturing, or hydrofrac, and occasionally minifrac (Haimson & Cornet,
2003) was developed originally for the oil industry to increase oil production, by inducing new fractures in the
rock upon injecting water at high pressure in a sealed section of the borehole. The theoretical background of
the method goes back to studies by Hubbert and Willis in 1957, who, based on the theory of elasticity,
concluded that stress in the rock mass around the borehole was directly related to the pressure recorded
during pressurization (Wittke, 2004). Several breakthroughs have made since, the most relevant summarized
by Haimson (1993), as listed below:

 Numerous interpretation techniques for the logged pressure data and fracture traces on the borehole
wall have been developed, some being more accurate; however, the margin of error is not usually
greater than 10%.
 Different equipment and accessories are currently available, such as drillpipes, drillpipes and hoses,
wirelines and hoses, and multihoses.
 The techniques related to fracture tracing includes: oriented impression packer, geophysical imaging,
borehole televiewer, televiewer-impression packer and electric resistivity. More information on such
techniques may be found in Haimson & Cornet (2003), Feng et al (2015), Guglielmi et al (2015).
 In spite that most common stress calculation procedures are based on elastic models, alternative
methods that use concepts of fracture mechanics or poroelasticity (Yew & Weng, 2015) are available.
In this document, only hydrofracturing in open holes, drilled in relatively competent rock for civil and mining
engineering purposes has been considered; hence, data interpretation and stress computations shall refer to
this particular application, and not to hydrofracturing in the oil industry.

6.2.1.1. Assumptions, advantages and limitations of the HF

Considering the assumptions made in hydrofracturing, Haimson & Cornet (2003), Ljunggren et at (2003) and
Wittke (2004) point out:

 The classical interpretation of this kind of test is only possible if the borehole axis is aligned (i.e.
parallel) to one of the principal stresses. Likewise, the borehole axis must be contained in the induced
fracture plane. This brings about that an excessive deviation of the borehole axis with respect to a
principal stress will not allow interpreting the results through the classical method. To this respect,
within the “classical interpretation”, when both, the borehole and the induced fracture are nearly
vertical, the stress component parallel to the direction of the hole axis is assumed as a principal stess
equal to the overburden.
 If the aforementioned is fulfilled, i.e. the borehole axis coincides with a principal stress, only the
maximum and minimum stresses in a plane perpendicular to this axis may be established. Once the
fracture starts to propagate, it will follow the direction of least resistance, which implies that a
minimum stress is acting perpendicularly to this propagation direction. In turn, based on the theory of
elasticity, the direction of the maximum stress acting on the same plane coincides with the direction of
fracture propagation.
 The fracture orientation is assumed to continue beyond the borehole. Based on this hypothesis, the
principal stress directions may be obtained from the fracture’s impression on the borehole walls.
 The rock behavior is assumed linearly elastic, isotropic and homogeneous, and its tensile strength
must be ascertained.
 In flat and non-tectonic active areas, where the in-situ state of stress is assumed not to be influenced,
the vertical stress may be assumed to be a principal stress.

This method may be considered an efficient way to determine the major and minor principal stresses,
especially in flat zones at great depths. According to Ljunggren et at (2003), the HF method is often conducted
in vertical boreholes from the surface, in areas with enough, considering the size of the equipment required for
the test. One advantage of the method is that the elastic properties of the rock mass (Young’s modulus,
Poisson ratio) are not required be known. However, the tensile strength of the rock should be obtained, if
possible, although it could be somewhat difficult, as its value should be obtained at the same depth as the HF
test.

As per the method’s limitations, as mentioned earlier, in order to perform the test, a fracture-free borehole
segment is required, however not always easy to find. Despite the segment needs be only between 1.5 and
2.0 m, if possible, a longer fracture-free portion should be searched for, in order to avoid any interaction
between the new fracture being created by the test and pre-existing fractures.

A drawback of the hydrofrac method is the difficulty to be performed in inclined holes, or in vertical holes cut by
transverse fractures, or in highly fractured rock. Special attention should be paid to geological structures, such
as folds, as these features constitute weakness planes, so that during a test, the fracture occurs along such
planes, instead of creating a new fracture perpendicular to 3 (Haimson, 1993).

In summary, in this method, the in-situ state of stresses is obtained for a perpendicular plane to the borehole
axis, and such axis must be closely aligned with a principal stress.

6.2.1.2. Procedure

Regarding the procedure to follow in a hydraulic fracture test, the American Society for Testing and Materials
(ASTM) has developed Standard ASTM D-4645. The ISRM has also published a suggested method (Haimson
& Cornet, 2003) in the International Journal of Rock Mechanics & Mining Sciences. The following procedure
describes the general steps to conduct this type of test (Haimson, 1993):

 Isolate an intact segment at a target depth 0.5 m to 2.0 m long in a drillhole. The term intact implies
that in the test section, no discontinuities or fractures should be present in the borehole walls; hence,
prior tele-images of the borehole wall are recommended in selecting a suitable test section. The test
section is sealed off using two inflatable rubber packers known as straddle packers (Wittke, 2004).

 Pressure inject fluid (usually water), at a rate where the pressure is increased steadily, but relatively
fast, to induce fracture initiation under the tensile stress exerted by the fluid. The creation of new
fractures takes place as soon as the fluid pressure reaches a critical level known as the breakdown
pressure. At this point, before the presence of the new fractures, the hydraulic pressure drops slightly.
This breakdown pressure is detected on the PC screen.

 Stop pumping and keep the hydraulic line shut. At this point, a decay in pressure takes place. This fall
in pressure occurs in two stages: in the first, the drop is fast; in the second stage, the drop occurs
slower. This transition level between fast and slow pressure decay is known as the shut-in pressure
(Ps) which indicates the closure of the fracture.

 Bleed the pressure, during several minutes, in order to finish the first pressurization cycle, also called
frac cycle Wittke (2004).

 Carry out at least three additional pressurization cycles at the same flow rate and pressure increase
as in the first cycle. The peak of each cycle corresponds to the fracture re-opening pressure (Pr). The
Ps value is recorded for each of these new cycles.

During the test, fluid pressure and flow rate are continuously logged. The Ps, Pc and Pr values are used to
calculate the stress magnitudes, whereas fracture tracing techniques are the tool for determining the minor
principal stress orientation, which is useful but not essential to determine the length of the steel liner.

In order to provide a clearer idea of the recorded data during a HF test, Figure 9 illustrates a typical log. In this
figure, each of the characteristic values obtained during the test are indicated (Ps, Pc and Pr). Also, in Figure
10 a general layout of the test is presented.
Figure 9. Typical pressure-time and flow rate-time record during a hydraulic fracturing test. After Haimson
(1993)

Figure 10. General layout for the hydraulic fracturing test. After Guo et al (2006)

6.2.1.3. Test data interpretation

Figure 9 suggests that the interpretation of results obtained during HF tests is limited to recognizing peak
values and the identification of an abrupt change in slope during the pressure drop stage. Although
theoretically correct, in practice, the obtained logs are not so straightforward to analyze, as the one shown in
the Figure. In this regard, Wittke (2004) emphasizes the inherent difficulty to determine the Ps value where the
methods available for its determination have a common denominator: the need for high resolution of pressure
versus time records.

A number of procedures have been developed to obtain Ps, most in graphical form (Haimson,1993), where the
analysis is conducted using the pressure-time curves or the flow rate - pressure curves. These methods are
explained below.

Method 1: Ps value from the pressure-time curves (shut-in curves)

In this method, the shut-in pressure is obtained from the pressure decay that follows the fracture initiation or
the re-opening of the fracture. The fracture propagates until the pressure is equivalent to the normal stress
acting perpendicular to the plane in which the fracture has been created. At this point, the pressure drop
continues, but at a different rate. This continuous drop is explained by one, or all, of the following factors:
permeability of the rock, and flow toward the natural system of joints through the induced fracture and other
leakages. These leakages conceal the Ps value, for which a number of graphical methods to evaluate Ps have
been proposed. Three of the most common procedures to estimate the Ps value from the shut-in curve are: a)
the tangent divergence (Gronserth & Kry, 1983), b) the tangent intersection (Enever & Chopra, 1986) and c)
the logarithmic method (Doe et al, 1983). These methods are shown in the figure below.

Figure 11. Graphical method for determining Ps from the Pressure Vs time curve. (a) Tangent divergence
method. (b) Tangent intersection method. (c) Log Pressure Vs Time.
Method 2: Ps value from the flow rate – pressure curves

During the HF test, the fracture closure implies a significant reduction in the flow rate. If the flow rate values
are plotted against the pressure values for each cycle during the HF, a bi-linear relationship may be found.
Based on this curve, the Ps value may be determined as the intersection of the two extrapolated straight
segments of the curve flow rate versus pressure, as shown in Figure 12.

Flow rate vs pressure is also used to determine the re-opening pressure (Pr) by means of the so-called step
re-fracture test. In this test, the pressure-flow behavior of the fractured zone is determined by pumping at a
constant rate, thus obtaining the equilibrium pressure. Again, the plots that relate these two variables show a
bi-linear behavior. In this case, the Pr value corresponds to the pressure where the change in slope takes
place.

Figure 12. Graphical method for determining Ps from the Flow rate Vs. Pressure curve.

For further insight into this topic, see the procedures developed by Lee & Haimson (1989), Cornet (1993),
Rutqvist & Stephansson (1996) Guo et al (2006), on the evaluation of the Ps value. In regard to the evaluation
of the Pr parameter, Lee & Haimson (1989) present a statistical approach.

6.2.1.4. Stress computations in the HF method

The computation process, related to the evaluation of the in-situ state of stress by means of the HF method,
depends, to a great extent, on the orientation of the borehole; likewise, as several models have been proposed
to conduct the analyses, additional information to that obtained through the test may be required, in order to
apply the selected model. To this effect, Haimson (1993) mentions three models that may be used in the HF
method: the first: the elastic model, which is the most widely used due to its simplicity. The other options which
deal with the poro-elasticity theory and fracture mechanics require lab tests, special equipment and further
assumptions, and are therefore beyond the scope of this document. Hence, steps for establishing the in-situ
state of stress in a vertical borehole by using the elastic model is explained next.

Case 1. Vertical borehole and axial generated fractures – Elastic Model

In this model, the rock is assumed as a homogeneous, elastic and isotropic material that fails when its tensile
stress is exceeded by the pressure applied to it; thus, the theory of elasticity is the conceptual foundation by
which the mathematical expressions are obtained. Further considerations include: within the test interval, there
is a plain strain condition and the rock mass surrounding the zone where the test takes place is impervious
enough as to avoid any flow or fluid transport.

Along with the abovementioned assumptions, perhaps the most relevant hypothesis refers the fact that the
vertical stress (v) is assumed as a principal stress:

v = *H

In this manner, according to Hubbert & Willis (1957), the elastic solution is given by the expression below:

Pc = T + 3h - H - Uo

Solving for H:


H = T + 3h - Pc - Uo

where,

Pc = breakdown pressure
T= Tensile strength
H = Maximum horizontal stress
h = Minimum horizontal stress
Uo = Pore pressure

As stated above, the orientation of the induced fracture gives light with respect to the orientation and
magnitude of minimum and maximum horizontal stresses (h and H, respectively). If the vertical stress is
indeed a principal stress, then h and H are also principal stresses. Thus, the pressure required to keep the
fracture open should be equal or similar in value to h. In a HF test, this pressure corresponds to Ps. This
reasoning implies that:

Ps = h

Where:

Ps = shut-in pressure

In this manner, the magnitudes of the three principal stresses are obtained.
Case 2. Inclined boreholes

Even though the HF test is normally performed in vertical holes, on occasions, in geotechnical prospections,
inclined drillholes are required due to logistic and/or topographic problems (e.g. locations of difficult access),
or to identify particular features in the rock mass (e.g. vertical discontinuities). In such boreholes, the
assumptions made for the case of vertical drillholes are no longer valid, so the equations presented above
cannot be used, and a three-dimensional analysis is then required. Haimson (1993) proposes a solution for
boreholes inclined in similar direction to the maximum horizontal stress. Accordingly:

H = (3h - Pr - Uo -v*sin²cos²
H = (3h - Pc +T- Uo -v*sin²cos²

where:
 = inclination of the borehole with respect to the vertical

6.2.2. Hydraulic Tests on Pre-existing Fractures (HTPF)

This method, known as hydrojacking, was first introduced by Cornet & Valette (1984). In essence, it is an
extension of the HF method, utilizing the same equipment and measuring the same parameters. In the HTPF
method, readings are taken on an existing fracture, rather than creating a new fracture. The objective is to
apply pressure to re-open the existing fracture, in order to determine the normal stress acting over the fracture
plane (Ljunggren et at, 2003).

6.2.2.1. Assumptions, advantages and limitations of the HTPF method

In general, the main assumptions within the HTPF are summarized as follows (Haimson & Cornet, 2003):

 The method assumes that isolated pre-existing fractures, or weakness planes, are present in the rock
mass.
 These discontinuities are not all aligned within a narrow range of orientations (strikes) and inclinations
(dips).
 The pre-existing fractures are mechanically opened by hydraulic tests.

The main characteristic and advantage of this method is that the borehole’s alignment needs not be assumed
parallel to a principal stress. A major implication occurs in the case when it is carried out in a vertical borehole:
in this particular, but frequent situation, the vertical stress needs not be a principal stress, as considered within
the traditional hydrofracturing method, thus allowing higher versatility to the HTPF test.
Additionally, other relevant advantages to this method as compared with the traditional HF are mentioned by
Amadei (1997). In the HTPF method, for instance, there is no need to determine the tensile strength of the
rock, and the pore pressure exerts no effect on the test.

Regarding limitations or specific considerations for the HTPF method, special attention must be given to the
characteristics of the fracture: the re-opened fracture should be long enough as to prevent stress
concentrations on the borehole walls from affecting the normal stress acting across the fracture. Ljungreen
(2003) states that the tested fracture should comply with the following characteristics: it must be of a size such
that the normal stress may be assumed uniform, and its shape, planar, free of undulations. The HTPF method
requires that several tests be carried out in fractures of different characteristics (dip and orientation) for each
region where the state of stress is considered homogeneous; the rock mass where the test is conducted
should not be highly jointed, as each test is executed in an isolated fracture. In this regard, Haimson & Cornet
(2003) state that when the isolated interval where the test is being conducted includes several fractures, only a
single fracture requires be opened.

Lastly, with respect to the applicability, Cornet (1993) has found that the HTPF method may be used with good
results in homogeneous rock formations, but when the rock mass is stratified, it has limitations.

6.2.2.2. Procedure

For this test, the same equipment and procedure than for the traditional HF method are used for determining
stress, as per ASTM D-4645 Standard, or the method suggested by the ISRM - Haimson & Cornet (2003)
Nevertheless, Amadei (1997) points out that the main difference between these approaches falls on the
special attention that must be given to the characteristics of the opened fracture, as mentioned before. With
respect to the number of tests required for a full and reliable stress tensor determination, the procedure should
be performed out at least six times, each on a different fracture, in terms of dip and orientation; nevertheless
Haimson & Cornet (2003) suggest carrying out additional tests for verification purposes. In the figure below, a
general layout of the HTPF method compared to the HF test is illustrated.

Figure 13. Rock stress measurement in the HTPF method compared to the traditional hydraulic fracturing.
After Ljunggren et al (2003)
6.2.2.3. Test data interpretation

Given its similarity to the hydraulic fracturing method, similar data are obtained from the HTPF test. The HTPF
method is based on four parameters: test depth, shut-in pressure (Ps), and the characteristics of the fracture
with respect to its dip and orientation. In this case, the Ps value is equivalent to the normal stress (n) acting
across the fracture plane.

With respect to test interpretation, Marulanda et al (1990) mention a number of considerations to be taken into
account, especially in permeable and heavily jointed rock masses:

 Determining the Ps values may be difficult due to the high pressure rate and fluid losses. In these
materials small pieces of rock or fill material could be transported through the flow channels, creating
a permanently open joint that could cause an abrupt pressure drop, thus making the shut-in pressure
difficult to determine.
 In these types of materials, a high-capacity pump should be available, due to the high loss of water,
which could surpass a standard pump´s capacity. In extreme cases, pre-grouting could be
considered, in order to reduce the permeability of the rock mass.

In general terms, in heavily jointed rock, permeability plays an important role in the selection of adequate
equipment.

6.2.2.4. Stress computations in the HTPF method

Assume that N tests have been run over the same number of fractures. Each fracture denoted by i is
characterized by its normal ni vector, which orientation is defined by azimuth  and dip . The aim is to obtain
the components of the stress tensor (X) in the rock mass, where the normal stresses (ni) acting on every
tested fracture have been determined. The mathematical expression that relates the normal stress in each
fracture with the stress tensor is:

σ ¿=σ ( X i ) ni ∙ ni

where (Xi) is the local stress tensor present at the center X i on the ith fracture plane. Cornet (1993) has
shown that if some assumptions are considered valid, the stress field (X) may be expressed as:

σ ( X )=σ ( X 3 )=S+ x 3 ∙θ

Where, S and  are two second-order symmetrical Cartesian tensors. Cornet (1993) has proved that seven
parameters are required to determine the stress field. These parameters are:
S1 and S2 = eingenvalues of S
 = orientation of the eigenvector with respect to the north associated with the eigenvalue S1 in the
horizontal plane

1, 2 and 3 = eigenvectors of 


= angle between the directions of 1 and S1 in the horizontal plane
x3 = considered depth

Since the mathematical development of these concepts is beyond the scope of this document, the reader may
refer to Cornet (1993), who has shown that by the combination of the two above equations, for the ith normal
stress measurement (ni), the ith fracture with orientation angles  and a dip  and at depth x3, the following
equation may be written:

σ ¿=θ 3 ∙ x 3i ∙ cos βi −0.5 sen β i [ S 1+ S 2+ ( θ 1+ θ2 ) x 3 i + ( S1−S 2) ∙ cos 2 ( α −λ ) + ( θ 1−θ2 ) x 3 i ∙ cos 2 ( α−λ−η ) ]=¿ 0¿
2 2

An equation like this is obtained for each conducted test, which means that at least seven independent
measurements of normal stress (N =7) are required in order to build a system of N non-linear equations, and
seven unknowns can be constructed (Amadei, 1997) that can be solved through the use either of numerical
methods or specialized software. Once the unknowns have been calculated and for the depth of interest, the
principal components (eigenvalues) of (X) and their orientation (eigenvectors) can also be determined.

6.3. RELIEF METHODS

The theory behind these methods is based on the response of the rock mass in terms of deformation, or strain,
when subjected to a stress relief process. According to Amadei (1997), this relief may be obtained by the
following mechanisms: over- or undercoring holes, cutting slots or underexcavation. The execution of a relief
method for determining the in-situ state of stress implies isolating a rock sample from the surrounding rock
mass and measuring its reaction. In this type of methods, interpretation of results will depend on:

 Constitutive model, that is, the stress-strain relationship.


 Assessment of the rock properties from samples.
 The sensitivity of the instruments used.

In the constitutive model, the trend is to resort to linear elasticity. In addition, despite these methods were
originally developed for competent rocks, in practice, they have also been used in weak or soft materials. In
order to describe these relief methods, a classification proposed by Amadei (1997) has been followed. Thus,
the relief techniques used to measure the in-situ state of stress may be classified into three main groups: a)
borehole relief methods b) surface relief methods, either in underground or surface excavations, and c)
methods where relief occurs in large volumes of rock. Each method is explained below.

6.3.1. Borehole relief methods


Despite variations in procedures and/or equipment used in any particular method, all in-situ stress evaluations
are performed inside a borehole, thus avoiding interferences associated with the measurement of stress at the
surface, such as rock mass disturbances, stress concentrations or climatic conditions. Due to their
advantages, these methods are currently the most common in engineering practice, performed in a variety of
projects. As a result of their success, a number of variations of the originally proposed method have been
developed since their onset in the 60’s (Sjöber et al 2003). In this respect, in 1997, Amadei and Stephansson
(1997) presented a list of available techniques, including 10 types of borehole relief methods. Because of such
a significant number of techniques, there is no unified classification as such; however, Ljundgreen et al (2003)
have proposed a relatively simple category, where the borehole methods are divided into the following
subgroups:

 Overcoring of cells in pilot holes.


 Overcoring of borehole bottom cells.
 Borehole slotting.

includes an overall explanation of each subgroup, and the main technique developed for each case.

Table 5. Borehole relief methods classification proposal. After Ljunggren et.al (2003)
Sub-group Methods Characteristics
 Soft inclusion cells for example:
CSIR, CSIRO, SSPB (Borre probe),
Overcoring a pilot hole in which
Overcoring of  Deformation meters(gauges)
the measuring cell is installed.
cells in pilot measuring displacements of the wall
Nowadays these methods
holes during overcoring (USBM gauge,
dominate the market.
IST)
 Stiff/solid cells
The cell is installed at the bottom
 Doorstoppers (CSIR doorstopper and of the borehole. This bottom can
Overcoring of
DDGS cells) be either flat or curved according
borehole-bottom
 Spherical strain cells to the used cell type
cells
 Conical strain cells (CCBO cell) The conical strain cells are used
mainly in Asia
A contact strain sensor is
attached to the walls of a large
Other methods  Borehole slotting
diameter borehole

CSIR: Council of Scientific and Industrial Research developed in South Africa


CSIRO: Commonwealth Scientific and Industrial Research Organization developed in Australia
SSPB: Swedish State Power Board, known as Borre probe which is a type of CSIR cell
DDGS: Deep Doorstopper Gauge System
CCBO: Compact Conical Ended Borehole Overcoring
USBM: US Bureau of Mines
IST: Sigra In-situ Stress Tool

Explanation of each method is beyond the scope of this document, wherefore only the more relevant and
common techniques are herein explained. These techniques correspond to the methods that make use of the
CSIR, Doorstoppers and USBM cells. Other methods are referred to in documents prepared by Obara and
Sugawara (2003) on the CCBO cell; Sjöberg & Klasson,(2003) on the Bore Probe cell (SSPB); Thomson &
Chandler (2004) on the DDGS Bottom cell; and Amadei & Stephansson (1997) on other techniques
mentioned.

6.3.2. Overcoring methods – Overview

The term overcoring is applied to a borehole, usually with a diameter between 76 and 96 mm, drilled
concentrically through, and over, a previously drilled pilot hole, in order to relieve the stresses of the rock
around it. Instruments or sensors designed to measure deformations resulting from stress relief have been
previously installed in the pilot hole, fixed to the borehole walls. In this method, the corresponding strains are
measured before, during, and after the hole is overcored, and the difference in deformation values obtained
before and after overcoring are correlated to the in-situ state of stress (Sjöber et. al 2003). The relation
between strains and in-situ stresses is established based on the theory of elasticity, where a number of
assumptions regarding the material´s behavior are made, and rock properties, Young’s modulus and Poisson’s
ratio must be determined. Ank et al (2009) point out the relevance of monitoring the temperature during testing
in order to avoid modifying the properties of the grout used to fix the cell and to minimize the temperature
difference between the deformation readings before and after overcoring.

Depending on the particular overcoring method, a 2-D or a 3-D state of stress is obtained for each test.
Accordingly, methods using cells in pilot holes provide a full stress tensor from a single hole, whereas
techniques using a cell at the bottom, only the state of stress in the plane perpendicular to the borehole may
be determined. Further ahead, a general procedure for a 2-D and 3-D stress measurement by use of the
overcoring technique will be discussed.

6.3.2.1. Assumptions, advantages and limitations of the Overcoring methods

Ask et al (2009) list the main assumptions made for the overcoring methods:

 The rock material has a continuous, homogeneous, isotropic, and linear-elastic (C.H.I.L.E) behavior.
 No stress/strain variations are present along the axis of the used probe. To accomplish this, the
measuring probe is mounted far enough from the end of the borehole.
 The deformation of the overcored sample during stress relief is identical in magnitude to that
produced by the in-situ stress field, but of opposite sign.

Based on these assumptions, and given the limitations of surface relief methods, overcoring techniques are
currently the most widely used method to determine the in-situ state of stress: They provide the following:

 In most cases, the complete state of stress as determined by use of a single borehole.
 Measurements are taken in undisturbed rock, away from the influence of excavations or weathering
phenomena.
 A number of methods to choose from is available, allowing selection of the one most suitable,
according to site conditions, requirements and characteristics of the project as well as available
budget.
 Vast experience gained in several civil and mining projects worldwide.

Nevertheless, as in all in-situ stress measurement methods, there are disadvantages. One main drawback of
the overcoring method is that the reliability of the estimated stress is directly affected by the estimate of elastic
properties. Corthésy et al (2003) state that in soft rocks, relevant problems have been identified related to
stress redistribution and stress concentrations that could exceed the yield point of the rock, leading to inelastic
deformations and, consequently, assumption that rock behaves as a linearly elastic material is no longer valid.

In addition, according to Sano et al (2005), and Cai and Peng (2011), there are two issues with respect to
overcoring techniques that influence accuracy and reliability of methods:

 The use of instruments sensitive to temperature changes when measuring deformations.


 The assumptions related to the linearly elastic and isotropic behavior of the material.

Regarding the former, most of the overcoring techniques use resistance strain gauges to measure strains
induced by the relief of stress; however, since these instruments are sensitive to variations in temperature,
additional strains are recorded, thereby inducing errors in the stress values. Hence, Cai and Peng (2011)
emphasize the need for using a temperature-compensation method. Further information related to available
procedures to reduce errors caused by temperature sensibility may be found in Cai et al (2010). As for the
latter, this is a common situation to all the stress measurement methods that can be addressed either by one
or more of the following strategies (Cai et al, 1995):

 To use a deformation modulus according to the level of stress. This may be achieved by use of the
incremental elasticity theory and iterative procedures.
 Numerical modeling and laboratory testing.
 The use of biaxial loading test for determining anisotropic parameters.

Finally, Sano et al (2005) lists a number of considerations to bear in mind when tests are being conducted in
deep borings, as borehole walls tend to undergo considerable damage by drilling, which alters the mechanical
properties of the surrounding rock, stress distribution around the borehole, and permeability. Moreover,
thermal stresses could change stress distribution due to temperature differences between the rock and the
mud water used for drilling.

The abovementioned advantages, disadvantages and assumptions may be considered as general


characteristics where the distinctive features of the employed method have not been taken into account. For a
more specific description, the main characteristics, along with advantages and disadvantages for the distinct
overcoring methods, are mentioned in the tables below.

Table 6. Characteristics of the borehole relief cells. Gathered from Ljunggren et.at (2003)
No of active Continuous
Cell types Measuring depths Borehole requirements
gauges logging
38 mm pilot hole, usually 150
CSIRO cell 9-12 Up to 30 Yes (via cable)
mm drill hole.
Normally: 10–50 m, 38mm pilot hole, usually 76-9
CSIR cell 12 modified versions: up to No 0mm drillhole. Modified
1000m versions accept water
36mm pilot hole, 76mm
Bore probe Practiced to 620 m. Yes, built in
9 drillhole.
(SSPB) Tested for 1000m data-logger
Accepts water-filled holes
38 mm pilot hole, usually 150
Normally 3; Normally 10–50 m;
mm drillhole. Modified
USBM cell modified modified versions up to No
versions accept
versions 4 1000m
water
60 m (CSIR
Yes in modified
Doorstopper 3-4 doorstopper) up to 1000 60 -76 mm diameter
versions
m (DDGS)

Table 7. Advantages and disadvantages of the most common relief cells.


Cell types Advantages Disadvantages
CSIRO cell - It is possible to determine the 3D - Long overcoring lengths are required (30
state of stress from one single cm at least)
measurement point - Special calibration for temperature is
- Very good results in isotropic and required
homogeneous materials and - Problems in water filled holes
acceptable in non-homogeneous - Limited success under low (<10°C) and
and medium-grained rock. high temperatures (>40°C)
- Redundant measuring - High cost
- Less sensitive to rock - Curing time for the epoxy (varies from 10
heterogeneities to 20 hours)
- Continuous monitoring
CSIR cell - It is possible to determine the 3D - Long overcoring lengths are required (30 -
state of stress from one single 50 cm at least)
measurement point - High scattering in soft rocks
- Redundant measuring - Usually the cell is not recoverable
- Good results in isotropic and - It is needed to clean the pilot hole walls
homogeneous materials - Curing time for the glue/grout (varies from
- Extensive research and previous 1 to 20 hours)
experience - The results can be affected by the grain
size of the rock and discontinuities
Doorstoppers - Used in highly stressed rock - It is a 2D method so the stresses are only
masses or in very jointed determined in the perpendicular plane to
materials (better possibilities for the borehole. In order to obtain the
successful since there is no need complete stress state three (or two) non-
a pilot hole) parallel holes are required.
- Long overcoring lengths are not - The end of the borehole must be flat
required (5 cm are enough) which requires to polish the hole bottom
- It requires less time than the - If the hole is wet, cementing problems can
triaxial cells, (2–3 tests can be take place.
conducted per - Curing time for the glue/grout (varies from
- day) 1 to 20 hours)
- Continuous monitoring (in some
equipment)
Cell types Advantages Disadvantages
- Used in water-filled boreholes (in
some equipment)
- it can be placed in smaller
diameter boreholes.
USBM Cell - the gauge is recoverable and - It requires an unbroken core of at least 30
reusable cm length
- no cementing or gluing is required - the gauges can be damaged if the core
- Good previous experience and breaks
proved success - three nonparallel holes are necessary to
- Easy to install calculate the in-situ stress field
- the gauge can be regularly - The selected gauge type depends on the
calibrated to ensure accuracy minerals in contact
- Less sensitive to temperature - Bad performance in heterogeneous
changes materials
- The gauge needs to be calibrated before
and after installation.
- The results can be affected by
discontinuities, and grain size of the rock

6.3.2.2. General Procedure

As mentioned earlier, depending on the technology used (i.e. cell type) the measured state of stress may be 2-
D or 3-D.

The general procedure in an overcoring test implies two main stages:

 Drilling the overcoring hole and performing the measurement of the in-situ strain relief.
 Assessment of Young’s modulus of the rock by use of a biaxial chamber or laboratory tests.

The steps for carrying out the measurement of in-situ stresses are described next. Steps included are for both
the overcoring of borehole-bottom cells (doorstoppers) and the overcoring of cells in pilot holes (CSIR cell and
USBM gauge).

Doorstopper procedure – 2D stress measurement.

The procedure described below may be found in Ljunggren (2003) and INTERFELS (2005):

a) Preparation of borehole for insertion of the Doorstopper cell:


A diamond drilled hole (usually 76 mm outer diameter) is drilled to the target depth; the rock core is
removed and the borehole bottom is flattened and polished with a special drill bit.
b) Placing of the Doorstopper cell:
A 2D cell containing a strain gauge rosette is lowered into the hole with a special installation tool. The
cell is then fixed to the flattened bottom.
c) Removing the installation tool and initial measurement:
The cell is fixed to the bottom and the initial reading is taken. The installing tool is removed so the
doorstopper cell is ready for overcoring.
d) Overcoring:
A new core is drilled in order to generate stress relief at the bottom of the borehole, inducing
deformations that are measured with the strain gauge rosette.
e) Core retrieval and repetition of the strain measures:
The core is caught and retrieved with a special device; after the core is removed, a second reading is
taken. From the recorded strains, along with the elastic parameters determined in laboratory, the
stresses acting in a plane perpendicular to the borehole are obtained

For clearer interpretation of each of the above steps, see Figure 14.

Figure 14. Doorstopper procedure – 2D stress measurement. After Interfels (2005)

CSIR cell overcoring procedure – 3D stress measurement.

The steps for a 3D pilot hole overcoring measurements are described next:

a) Preparation of the borehole for insertion of the CSIR cell:


A diamond drilled hole (usually 76 mm outer diameter) is drilled to the target depth. In some
procedures, the bottom of the hole is flattened with a special drill bit. Next, a concentric hole of
smaller diameter (36mm) and 30 cm long, known as the pilot hole is drilled.
b) Placing of the CSIR cell:
A 3D cell containing an orientation device is placed into the hole with a special installation tool. In this
step, there are several procedures to put the cell in place; the one employed by INTERFELS (2005)
implies the use of detachable aluminum rods, to put the instrument head in place; then, compressed
air is used to expand the cell in the hole and the strain gauges are glued to the hole´s wall.
c) Removing the installation tool of the CSIR cell:
Once the cell is fixed in the pilot hole, an initial reading is taken. The installation tool is removed so
the cell is ready for overcoring.
d) Overcoring:
The pilot hole is overcored in order to generate a relief of stress, inducing strains that are measured
with the strain gauges.
e) Core retrieval and repetition of the strain measures:
The core is caught and retrieved with a special device; immediately after the core is removed, a
second reading is taken. From the recorded strains, along with the elastic parameters determined by
biaxial tests and/or laboratory tests, the in-situ stress tensor may be computed.

For the above methods, and depending on site conditions and specific equipment characteristics, the average
time required for a single measurement may be approximately one hour. In order to obtain reliable results, at
least 7 to 10 single measurements should be taken, a task which, under normal circumstances, might take up
to two days. Figure 15 illustrates the above steps in order to provide a clearer description.

Figure 15. CSIR cell overcoring procedure – 3D stress measurement. After Interfels (2005)

USBM cell overcoring procedure – 3D stress measurement.

In general, this procedure is quite similar to the abovedescribed CSIR cell procedure. Differences center on
the overcored hole, which has a larger diameter (150 mm) and in the characteristics of the gauge itself. A
detailed description of the accessories and testing procedures for this gauge may be found in the last version
of the ASTM D-4623, and in Amadei and Stephansson (1993).

For a full 3D stress tensor estimation by use of a USBM gauge, a set of three boreholes shall be drilled, two of
which should be drilled in the wall of the underground excavation at an angle of 45° between them. These
holes should be inclined upward to allow drainage. The third borehole shall be drilled either upwards or
downwards, and preferably drilled last, to avoid problems related to drilling and core retrieval. Advantages and
disadvantages were mentioned in Table 7; however, a 30 cm minimum intact rock core is required, free of
cracks, fills, or imperfections after the overcoring process.

Biaxial Test

As discussed above, the overcored samples obtained from the USBM gauge, the CSIR cell and the CSIRO
cell, have to be tested in order to establish the rock’s Young’s modulus. A biaxial test is then carried out, used
to measure the response of the rock to loading and unloading cycles, thus obtaining its deformation properties.
This test also provides data regarding the material’s behavior by establishing limits in which the rock has a
linearly elastic stress-strain relationship. The test is performed in a biaxial chamber, which consists of a
cylindrical steel jacket, a rubber membrane and seals.

According to Sjöberg et al (2003), the following considerations have to be kept in mind regarding the biaxial
test:

 The test must be carried out on site as soon as possible.


 The core sample should be tested under similar conditions as those in the borehole.
 Extreme temperature changes have to be avoided.

For more detailed information on the biaxial test equipment and capabilities, see comments on presented in
Amadei and Stephenson (1993).

6.3.2.3. Test interpretation – Overcoring methods

As explained before, overcoring tests are performed in two stages, in each during which deformation
measurements are taken. The interpretation and use of these measurements is discussed next.

Analysis of strain measurements during overcoring

A typical log for an overcoring test relates strain measurements versus time. Under normal conditions, stable
readings just before overcoring shall be taken. Once the overcoring process begins, it is common practice to
record a maximum or a minimum, before passing the strain gauges. At this point, the stress relief process
takes place, manifested as an increase in the deformation values. Then, another local maximum or minimum
is expected, when the drill core bit has passed the strain gauge position. Finally, once the overcoring has been
completed, the readings show once more a steady state.

This ideal behavior, however, will most likely show variations. Sjöber (2003) states that minor variations are
almost always recorded, but larger divergences may be due to:

 Equipment flaws (e.g. malfunctioning of the logger or the strain gauge)


 Rock imperfections (e.g. presence of previously undetected microcracks)
 Rock behavior (e.g. creep)
 Additional strains caused by temperature changes in the strain gauge

Once a critical review to the measurements has been conducted, strain differences (before and after
overcoring) are computed for each strain gauge, to be used as input data for the stress computation. Figure 16
illustrates ideal overcoring results during the strain measurement stage.
Figure 16. Idealized overcoring results

Analysis of biaxial test data

The key objective of the biaxial test is to obtain the deformation properties of the rock submitted to loading and
unloading. Thus, in order to calculate these properties, the theory for an infinitely-long, thick-walled circular
cylinder, subjected to uniform external pressure is considered (Sjöberg et al, 2003)

D2
σ T =2 p
D 2−d 2
Where:
p = applied pressure
D = outer diameter
d = inner diameter

Young’s modulus is assessed as a secant modulus by means of the following expression:

p∗d
∗D 2
σT Ui
E= =2 2 2
Δ εT D −d
where,

T: strain in a tangential direction


Ui = diametric deformation recorded the gauges

A typical result of a biaxial test in shown in the Figure 17 below.


Figure 17. Biaxial test results. After Sjöberg et al. (2003)

6.3.2.4. Stress computations in the Overcoring methods

Theoretical background related to the stress computation for different cell types and material behavior is
provided by Amadei and Stephansson (1997), and in Standard ASTM D-4623, for the particular case of the
USBM gauge; however, as an example of calculations, the general equations for computing stresses in an
isotropic material, when an USBM gauge is used in a single borehole, are shown next:

Stress calculation with a USBM gauge in a single borehole

From elasticity theory, the radial displacement (ur) induced by application of in-situ stress field in hole with
radius (a) is given by the following expression, obtained by Amadei (1983):

uro
=f 1 o σ xo +f 2o σ yo + f 3 o σ zo+ f 4 o τ yzo +f 5 o τ xzo + f 6 o τ xyo
a
Where the coeficients fio depend on the elastic properties of the rock. In this case, the change in hole diameter
(Udo) may be expressed as:
U do=2uro
When there is a plane of elastic symmetry normal to, for instance, the z-axis (see Figure 18) both f4 and f5 are
equal to zero, so a general expression for the diametric deformation is:

U do
=f 1 o σ xo + f 2 o σ yo + f 3 o σ zo +f 6 o τ xyo
2a
Figure 18. Examples of elastic symmetry with respect to the Z-axis

Now, drilling a vertical borehole in the Z-axis, the equation reduces to three unknowns. If no longitudinal strain
takes place during overcoring, the principal stresses can be determined in a plane perpendicular to the
borehole axis by means of the following expressions, as given by Amadei & Stephansson (1997)

E
σ xo = ¿
6 ( υ2−1 )
E
σ yo= ¿
6 ( υ 2−1 )
E
τ xyo = [ U 2 −U 3 ]
4 ( υ −1 ) √ 3
2

U do
In this case, Ui are the values of measured with the USBM gauge at = 0°, 60° and 120° from the X-
2a
axis (see Figure 24Figure 19)

Once the values for xo, yo and xyo have been established, the major and minor principal stresses are
obtained as follows:

√( )
2
σ xo +σ yo σ xo −σ yo 2
σ 1= + + τ xyo❑
2 4

√( ) +τ
2
σ xo +σ yo σ xo −σ yo 2
σ 3= − xyo❑
2 4
The orientation angle  between 1 and the X-axis is:
τ xyo
sin 2 Φ=

√( )
2
σ xo −σ yo 2
+τ xyo❑
4
Figure 19. Stress computation using an USBM gauge in a single borehole

6.3.3. Surface relief methods

The basic concept of these methods is based on the concept of disturbing the stress equilibrium in the rock
mass through a borehole, or small excavation, and measuring the associated deformations to such
disturbance. The general procedure is described below.

6.3.3.1. Assumptions, advantages and limitations of surface relief methods

The following setbacks may occur when the in-situ state of stress is obtained by use of a surface relief method:

 Humidity and dust may influence the performance of the gauges or pins.
 The measured displacements may be affected by factors such as the disturbance of the rock mass
(due to blasting or any other excavation method) and weathering phenomena.
 Stress concentrations in the walls of the excavation (at the surface or underground) have to be
assumed.
 The obtained stresses correspond to the principal stresses in the plane parallel to the wall where the
test is conducted.

Given the characteristics of these methods, a plane stress condition is assumed; moreover, from the
behavioral model adopted, the rock is considered as an isotropic and linearly elastic material, with Young's
modulus (E) and Poisson's ratio (.

6.3.3.2. Procedure

This method is considered to be the first technique applied in an underground excavation’s walls in order to
evaluate the state of stress, in which the general procedure implies installing sensors, such as gauges or pins,
at the rock’s surface. Then, a cut or drill in the rock wall is made, so that a change occurs in the state of stress,
which consequently induces deformations. Amadei & Stephansson (1997) summarize the method to achieve
this stress relief:

 Isolate a block of rock from the surrounding rock mass.


 Center hole drilling or undercoring.
 Monitor the deformation in the hole due to drilling of parallel and near hole.

In the figure below, a general layout for the first two methods mentioned above is shown.

(a) (b)
Figure 20. Surface relief methods. a) Rock block isolation Modified from Amadei (1997) after Merril (1964). b)
Undercoring (center hole drilling). Modified from Amadei (1997) after Duvall (1974).

Once the disturbance in the state of stress has been caused, the manner in which the material responds to the
stress relief is recorded by preinstalled gauges and pins. Finally, from the elasticity theory, the stresses may
be obtained.

6.3.3.3. Stress computations in the undercoring method

In surface relief methods, stresses are computed, by means of the elasticity theory from displacements
obtained during testing. These deformations (u) are related to the stresses xo, yo and xy acting over the
same plane where the test is carried out (in most cases, this plane is the excavation wall). From elasticity, the
radial displacement (ua) at the point (a, Q) induced by a borehole with radius (r) is:

ua
=f 1 σ xo + f 2 σ yo + f 3 τ xy
r
Where

−r
f 1= [ 1+υ+ H cos 2θ ]
2 Ea
−r
f 2= [ 1+υ−H cos 2 θ ]
2 Ea

−r
f 3= H sin 2θ
Ea
2
(1+υ) r
H =4− 2
a

If the diametrical measurements induced by the drilling of a center hole are U1, U2 and U3, and they are located
at angles 1, 2, and 3, a system of equations may be expressed as:

[]
U1

[ ][ ]
2r
f 11 f 21 f 31 σ x
U2
= f 12 f 22 f 32 ∙ σ y
2r
f 13 f 23 f 33 τ xy
U3
2r

Once the system has been solved, two principal stresses may be obtained from the same equation, by using
the USBM gauge in the overcoring method.

6.3.4. Relief of large rock volumes. Back analysis

In this method, statistical and numerical tools are used in order to find the stresses that generate the measured
displacements in several cross-sections during an underground excavation. In this method, large volumes of
rock are involved, with values varying from 10² to 10³ m³ (Ljungreen, 2003). This technique has been used in
other civil engineering works, where the unknown variable is not always the stress. For example, in slope
stability analyses, this method is used to find the strength parameters of the material.

6.3.4.1. Assumptions, advantages and limitations of back analysis

To evaluate the state of stress by this method, both displacement and deformation parameters of the rock
mass must be known. The first requirement may be obtained by means an adequate and sufficient
instrumentation system, whereas the accurate determination of the rock mass modulus may be somewhat
challenging, in spite of the available methods for this. Regarding the monitoring of deformations, Christiansson
(2006) believes installation of extensometers to be the most complete method for recording the displacements
and emphasizes the importance of having precision equipment, especially in crystalline rocks.
Regarding advantages, back analysis methods involve large volumes of material and in general, expeditious to
use; however, knowledge and experience in numerical modelling is required. In turn, Amadei & Stephansson
(1997) state another important advantage, regarding the method that may be used to modify design decisions,
such as support requirements or construction processes, while the excavation progresses.

The main weaknesses of this method are related to its applicability, since it can only be implemented during
the excavation process of an underground excavation; however, examples of the use of back analysis by
implementing analytical and/or numerical analyses in 3D, along with multivariable regression techniques to
assess the in-situ stresses state, may be found in Barton & Infanti (2006). Liu et.al (2006) and Zhu et al (2006),
and Liu & Chang (2010).

6.4. JACKING METHODS

The term Jacking has been given to those methods intended to obtain the in-situ state of stress at, or near, the
surface (no deeper than 7 m), where an altered equilibrium in the rock is restored through the use of a device,
such as a jack. These techniques where initially developed to evaluate the deformation characteristics of the
rock mass (Young’s modulus mainly) and are thus also known as stress compensating methods (Amadei and
Stephansson,1997) the most widely used is the flatjack method.

6.4.1. Procedure

In general, the following procedure is briefly described. The stress equilibrium in the rock mass is altered by
cutting slots on the surface of rock excavations (underground or surface). As in other methods, this alteration
is evidenced by deformations, which can be measured with a strain gauge, or reference pins, located in the
slots. As mentioned before, the stress equilibrium is altered, and stress is applied by means of a jack, in order
to restore it, that is, to return to the zero deformation state.

As the flat jack is the most common within the jacking methods, this technique only is discussed in this
document. Standard guidelines have been developed by the ASTM, in its ASTM D 4729 Standard, as well as
the document edited by Ulusay & Hudson (2007) for the ISRM.

According to Amadei and Stephansson (1997) a flat jack is composed of two metal plates welded together,
which are inserted into the slot, then grouted in place and pressurized until the deformation readings have
returned to their original position, that is, to the state before equilibrium was distorted. The required pressure to
annul the deformations is known as cancellation pressure (pc), interpreted as the tangential stress normal to
the jack (). The main characteristics of the equipment and general set-up for the test are described below:

 The plates are usually square or rectangular, of width not less than 0.6 m (2 ft). The plates can bear
up to tens of megapascals of pressure.
 There are two ways to create the slots: by overlapping holes, or by using a large diamond disk saw.
To overlapping holes is more suitable for deep slots (deeper than 1.5 m), whereas the saw is used for
shallower slots.
 Either mortar or resin may be used as grouting material; however, grout should be designed to have
as similar strength and deformability as the surrounding rock, which is not always possible.
 Hydraulic pumps are used to apply pressure. Pressure should remain constant over a period of 5
minutes.
 To increase the rock volume involved during the test, several coplanar flat jacks may be installed in
the slot.
 Deformation shall be measured in the vicinity of the slot. According to the ASTM, the points at which
measurement are taken, should be located within a distance of L/2 of the flat jack slot, where L is the
flat jack width (ASTM, 2008)

Since each flat jack test only yields one component of the in-situ stress tensor, a total of six tests are required
to obtain the complete in-situ state of stress.

6.4.2. Assumptions, limitations and advantages of the flat jack

The main assumptions, weaknesses and advantages regarding the flat jack test have been discussed by a
number of authors; however, the compilation prepared by Amadei and Stephanssonn (1997) is perhaps the
most complete source of information on this subject. Within the assumptions for this method, are the following:

 The rock mass behavior is elastic (either liner or non-linear) and isotropic.
 During the pressurization stage, the rock mass is under compression in the direction perpendicular to
the jack surface, with an efficiency of 100%.
 The applied stress across the jacks is uniform.
 Creep phenomena are neglected (which can be a problem in weak rocks such as shales).
 By means of this test, it is not possible to measure shear stresses (

On the advantages of this method, Amadei and Stephansson (1997) point out the following:

 The rock mass deformation properties need not be known. In fact, this method may be used to
evaluate such properties (Goodman, 1989).
 The equipment used for the test is sturdy and stable.
 Compared to another methods (for instance borehole relief and HF methods) relatively large rock
volumes may be involved in the tests.

In spite of the abovementioned advantages, currently this method is not commonly used as a method for
measuring the in-situ state of stress, mainly due to its limitations regarding the following aspects:

 The stress measurement may only be performed near the surface of an opening, which implies that
inherent disturbance caused by the excavation processes could influence the results.
 The presence of major discontinuities in the rock mass could affect the test.
 Results may be affected by environmental conditions, such as temperature or dust.
 Due to the geometrical characteristics of the flat jack, the contact area could change during testing
wherefore the applied pressure will not be uniformly distributed, so the applied pressure will differ
from the actual pressure acting against the rock.
 In underground excavations, the shape of the opening is of utmost relevance, as the stress
concentration is required be known in order to correlate the measured cancellation pressure (p c) to
the virgin stresses. Thus, the stress concentration factors need be established by use of numerical
models, or other available methods found in the literature.
 In soft and weak rocks, several drawbacks may occur, such as rock softening due to the water used
for the grouting or by creep phenomena.

6.4.3. Test interpretation – Flat jack

During a flat jack test the process related to the alteration of the stress state and pressurization is illustrated in
Figure 21, where elastic behavior is assumed. As mentioned before, during testing, the rock is assumed to be
under compression, perpendicular to the jack surface. At the onset of the test, the distance between two
reference pins is d0 and the unknown normal stress is defined is sq (Point A). Once the slot has been cut, the
surface of the slot becomes a “free surface”; therefore, the normal stress acting on it is reduced from sq to
zero, and the distance between the pins is reduced by an amount 2d (Point B). When the pressurization
stage takes place, the pins return to their original position when the applied stress is equal to the cancellation
pressure (pc).

Figure 21. Mechanical interpretation of the flat jack test. After Amadei & Stephansson (1997)

6.4.4. Stress computations in the flat jack method

In each flat jack test, only one value of the cancellation pressure (pc) is obtained, which is assumed equivalent
to the tangential stress () acting over the plane where the test is conducted Figure 22), assuming that the
stresses y, x and xy act in a bi-dimensional field, where three different flat jack tests were conducted, in a
plane parallel to the axis of an underground opening, in order to measure and Assuming the rock
mass as an isotropic and homogeneous material, the relationship between these stresses is given by the
following matrix:
[ ][ ][ ]
σθ 1 f 11 f 12 f 13 σ x
σ θ 2 = f 21 f 22 f 23 ∙ σ y
σθ 3 f 31 f 32 f 33 τ xy

Where parameters fij depend upon the geometry of the excavation.

Figure 22. Stress state in a flat jack test carried out at a circular excavation. After Amadei and Stephansson
(1997)

When the slots are cut at an angle with respect to the axes of the underground excavation, each p c value may
be related to the six components of the in-situ stress tensor. To this end, the orientation of each slot and the
stress concentration factors associated with the shape of the excavation are required.

6.5. CORE-BASED METHODS

These methods are based on the analysis of cores retrieved from drilling. Their advantage lies in that no
additional field procedures are required to evaluate the state of stress, other than the drilling itself. Within these
techniques are the acoustic emission (AE) method and the strain recovery methods. In general, in spite of the
advances during the past decades with respect to the gauging instruments and material models, the core-
based methods are still being developed, reason for which their reliability in tunneling and underground
excavations is still limited. Hence, only a general overview of these methods will be discussed herein.

6.5.1. Strain recovery methods

The main principle of these methods is based on the fact that once a core is extracted from the rock mass, it
undergoes a stress relief due to the lack of confinement, expressed in terms of an "expansion”. From the
theoretical point of view, this expansion is greater in the direction of the maximum stress direction and smaller
in the direction of the minimum stress (Ljungreen et.all, 1993). Applying the theory of elasticity, this relaxation
has two components: one purely elastic, which is instantaneous; the other, time-dependent anelastic strain.
There are two main methods that use this concept:
 Anelastic strain recovery (ASR)
 Differential strain curve analysis (DSCA)

6.5.2. Anelastic strain recovery (ASR)

This method is founded on the analysis of microcraks that occur during the relaxation caused by core
extraction from a borehole. Field measurements have shown that the opening and propagation of microcracks
are linked to strain recovery. Thus, if this anelastic strain recovery is measured, it is possible to obtain the
orientation of the principal strains. Upon knowledge of this characteristic of the principal strains, the orientation
of the principal stresses may also be obtained.

The ASR technique is used in deep and oriented holes where neither the HF nor overcoring methods are
feasible or are too expensive to use. However, regarding its applicability, the following aspects are significant
limitations to be considered:
 Temperature changes creating thermal strains,
 Water loss (dehydration) of core samples,
 Pore fluid pressure diffusion,
 Drilling mudrock interaction,
 Core recovery time
 Accuracy of core orientation.

The above circumstances entail that this method is rarely used in practice. Besides, Amadei & Stephansson
(1997) mention results obtained in Japan by Matsuki and Sakaguchi (1995), where a poor correlation between
the ASR and overcoring measurements was ascribed to problems with the gauging devices, thus exposing the
need for more research on this matter.

6.5.3. Differential strain curve analysis (DSCA)

Anamdei & Stephansson (1993) explain the concept behind this technique, where the stress history of a rock
core may be evaluated by meticulous monitoring of the strain behavior after re-loading. Once the core has
undergone expansion due to the lack of confining pressure, microcracks are assumed to have had enough
time to align themselves in the direction of the original in-situ stresses. Thus, if a hydrostatic load is applied to
the rock sample, then the required strain to close the microcracks may be calculated from strain gauge
readings.

In regard to the assumptions in the DSCA method, the orientations of the strains required to close microcracks
are considered principal - and proportional in volume - to the in-situ stress magnitude in order to associate
them to the orientations of the principal stresses, and the ratios between the principal strains are related to the
ratios of the principal stresses.

As with the ASR method, the DSCA may be used in deep boreholes, as long as a core sample is recovered.
Regarding its limitations: 1) the method is only two-dimensional; 2) required measurements are difficult to
perform, and 3) the poor correlation with other techniques, such as overcoring, thus questioning its reliability
(Ljungreen et.all, 1993). Figure 23 is a schematic representation of the principles behind the two
aforementioned methods.

Figure 23. Basic concepts for the strain recovery methods. a) ASR method and b) DSCA method

6.5.4. Acoustic Emissions (AE) Method

The acoustic emissions may be defined as transient elastic waves created by a sudden release of energy by a
local source, or sources, within a material (Lehtonen & Särkkä, 2006). In a rock mass, this phenomenon is an
indication of fracturing and deformation phenomena. In turn, the AE method resorts to the Kaiser effect to
estimate the in-situ state of stress, where the Kaiser effect may be defined as a phenomenon in which the
acoustic emissions’ rate undergoes a rapid change, immediately after the initial stress level has been
exceeded in a repeated loading cycle.

The basis of this method is relatively similar to that applied in the strain recovery approaches, since in the AE
technique the rock specimen undergoes a stress relief due to core sampling, and, subsequently in the lab,
subjected to a re-stressing process. In this loading process, the in-situ state of stress is computed by means of
acoustic emissions that take place under uniaxial compression. The goal is to find the point where the Kaiser
effect occurs. This point is considered as the normal component of triaxial in-situ stress, in the direction of
loading, reason for which a sufficient number of rock cores sub-specimens taken in several directions from the
main core sample have to be tested, in order to find the stress tensor. The assessment of this Kaiser effect
point is carried out from cumulative event graphs drawn as a function of stress, which is not a straightforward
task. In Figure 24 the graph, along with the Kaiser effect point, is explained.
(a) (b)
Figure 24. Acoustic emission (AE) method. a) Cumulative acoustic emission curve b) “Kaiser effect point”
location. Taken from Lehtonen & Särkkä (2006),after Boyce et al,( 1981).

Regarding the advantages of this method, Ljunggren et al (2003) assert:

 A good correlation has been found with results obtained from HF and overcoring tests.
 In general, equipment required for its implementation is simple.
 It is possible to find the 3D state of stress

In spite of these benefits, there are still doubts concerning its reliability, and the fact that the acoustic
emissions is measured from the uniaxial compression test, is still controversial.

7. APPLICATIONS

Projects, such as underground mines, tunnels in hydropower, road and railway, nuclear waste storage facilities
and other types of works, are usually built in rock, being the constituent material of excavations. In the early
stages of an engineering design process, a clear understanding of the in-situ stress condition of the rock is
required to define the overall layout, the support system and construction method of the project, as such
structures will be subjected to long-term stresses. Because the field of application for the various techniques
mentioned so far is somewhat extensive, involving a significant number of engineering fields, such a civil,
mining and petroleum engineering and geophysics, only the applications related to the different types of
tunnels, mining activities and hydropower developments have been herein considered.

7.1. MINING

The mining industry has acknowledged the importance of knowing the in-situ state of stress, not only from the
technical standpoint, but from the economical. Therefore, measurement of the in-situ state of stress plays an
important role during the different stages of a mining project, as summarized below.
7.1.1. Mine design and long-term planning

It is common practice within the mining industry to have well-established programs oriented to evaluate the in-
situ state of stress, before and during mining operations. Before operations, specifically in the design of
underground mine projects, knowing the in-situ stress is of utmost relevance for the dimensioning of rock
pillars and chambers, as for the general layout of the project. Mining schedules are thus proposed, on the
basis of geological as well as global ore market conditions.

7.1.2. Increase production

For the case of mines in operation, a noteworthy example where the knowledge of the state of stress has
technical and economic implications is the dimensioning of the support rock pillars: measuring the state of
stress may lead to re-dimensioning or even removal of valuable ore-containing pillars, thereby increasing
production and profit, which were perhaps not initially considered.

7.2. TUNNELS

In these structures, the in-situ stress methods have played a significant role, given the relevance stress has in
the design of such works. Some of the more important applications in this area are mentioned below.

7.2.1. Numerical analysis

Numerical analysis is currently of great use in the in-situ measurement of rock stress, as the data obtained
have become the essential input for numerical modeling, which, in turn, has become a widespread method in
the design of underground excavations. In numerical analysis, the in-situ stresses are considered as a
boundary condition, based on which the stress analyses are carried out. From stress analyses, key aspects in
the design of a tunnel may be determined, such as:

 Stress redistribution around the opening.


 Stress concentrations.

Acting stresses identifying their nature (tension or compression) and magnitude, in order to identify the
feasibility of different phenomena, such as rockbursts, spalling or squeezing.

Stress analysis is a quite helpful tool to determine the extent of the zone where stresses in the rock mass have
been disturbed by adjacent excavations, or to dimension the support requirements, such as bolt length and
density, or the need for steel sets.
7.2.2. Selection of the optimal shape

In an underground excavation, the closer its shape to a circle, the better, in order to minimize stress
concentrations; theoretically, an ideal cross section is an ellipse, in which the ratio of horizontal to vertical axes
matches the ratio of horizontal to vertical principal stresses (Cai & Peng, 2011); however, such shape is not
always possible to implement due to space requirements, costs or other considerations.

A good example on how the shape of a tunnel may be modified based on in-situ stress measurements in order
to overcome stability problems, is given by Cai & Peng (2011), who cited a case of the largest nickel mine in
China. The mine began construction in the mid 60’s; however, it took almost a decade to start operations, due
to large deformations and instabilities in the tunnels. Opon measuring the in-situ stresses, it was found that the
horizontal stress was extremely larger than the vertical stress. The cross section was changed from the
traditional horseshoe shape with vertical walls, to an egg-shaped geometry. This measure, along with a
change in the support system, stabilized the underground openings and the mine was able to begin operating.

7.2.3. Optimization of excavation sequences

In normal practice, the excavation sequence has been based on suggestions and previous experiences of the
designer; in many cases, similar procedures are applied regardless of the natural variability and complexity of
the different characteristics of the rock mass. Advances in numerical modeling provide a new way to deal with
this, providing a rational support for design decisions, based on mathematical calculations. Nevertheless, the
quality and reliability of these models depends exclusively on the input data, where reliable values of the stress
tensor, regarding magnitude and orientation, are essential.

7.2.4. Support and lining design

Checking the state of stress in the periphery of tunnels is sometimes required, in cases in which the in-situ
stress methods can be used to evaluate the tangential and axial stresses in the roof of the excavation, and the
load on the rock pillar between excavations, thus providing essential data on its stability. With respect to tunnel
support, the doorstopper method has been used to measure the acting stress in shotcrete and concrete.

A well-known application in hydropower projects is related to the determination of the minor principal stress by
means of the hydrofracturing method in headrace pressure tunnels or pressure shafts. In this case, the minor
principal stress in the rock mass is compared to the water head in order to establish the need for a steel liner
and its length: a steel liner will be required if the internal water pressure is higher than the minimum principal
stress. In order to preclude water leaks due to hydraulic fracturing, the following criterion must be met:

whw < 3
7.2.5. Prediction of rockbursts

Rockburst phenomena may occur in deep excavations in competent brittle rock masses. It is, in essence, a
dynamic process of prior energy accumulation and its sudden relief during excavation. Within the evaluation of
rock bursts, in order to quantitatively compute the magnitude and distribution of underground energy
accumulation, a clear understanding of the state of in-situ stresses is of utmost importance, in order to predict
the probability of such phenomena occurring. From knowledge of stresses and strains, the magnitude and
distribution of energy build-up around the excavation face may be obtained, by the use of numerical models
and earthquake theory.

7.3. CAVERNS

Caverns are usually built for hydropower projects, storage facilities, sports halls or metro stations. In many
cases, the main target is to obtain the major principal stress in order to face the cavern as parallel as possible
to this principal stress. In other cases, sufficient horizontal stress has been a prerequisite for stable large span
underground caverns. As in the aforementioned applications, the results from 2D or 3D stress measurements
may be used as input for numerical modeling, in order to carry the stress analysis that it will serve for design of
the structure.

8. CASE STUDIES

This section will cover practical cases in which some of the techniques herein described were used for the
determination of the in situ stress state, whose results served as basis for decision making.

8.1. CASE HISTORY – PORCE III HYDROPOWER PLANT – HYDRAULIC FRACTURING


AND OVERCORING TESTS

8.1.1. General Characteristics

Main Project data (General project layout: see Figure 10.1):

 Location: Northwestern Colombia


 Generation power: 660 MW
 Hydrostatic head: 365m
 Year of operation: 2010
 Powerhouse: underground (270 m overburden), housing 4 units.
Pressure conduit:

 Upper headrace tunnel 12,300m long; 10.2 m diameter.


 Upper surge shaft: 149m long; 15 m diameter, located at K12+125 of headrace tunnel
 Lower connection shaft: 102m; 8m diameter that connects to the headrace tunnel through a lower
40m long gallery, 10.2 m diameter
 Lower headrace tunnel: 323m long, connects to the distributor and finally to the turbines in the
powerhouse
 Tailrace tunnel: 827 m long; 10 m diameter

Figure 10.1. General Project Layout

8.1.2. Hydraulic Fracturing

Hydraulic fracturing will occur when the in-situ minor principal stress is lower than the internal hydraulic
pressure of the conduit, wherefore aperture of discontinuities or tensile failure of the rock mass takes place,
thus creating unwanted water leaks towards the ground surface. Accordingly, the pressure tunnels were
initially located as far in the mountain as possible in order to preclude such phenomena from occurring and at
the same time, thus minimizing the need or length of a steel liner, as was the case in the lower headrace
tunnel.

During the conceptual and initial design stages, confinement of the pressure conduits (surge shaft, pressure
shaft, and lower headrace tunnel) were reviewed, upon applying empirical overburden rules in order to locate
such pressure conduits, at least in a preliminary manner, however always considering that a minimum factor of
safety of 1,3 is required in the measures to be implemented in order to counteract such phenomenon, among
which were criteria by Broch and Dann.

However, in order to verify these overburden analyses and minimum confinement stresses for the pressure
conduits, between the surge shaft and the downstream sector, down to the powerhouse, the conceptual and
initial designs had foreseen conducting, at a later timely stage, hydraulic fracture, as well as overcoring tests
from inside a 630-m long exploration gallery that had been excavated to reach the vicinity of the underground
powerhouse, in order to measure the minor principal stress in this particular area.

Consequently, 22 hydrofracture tests were conducted in the powerhouse exploration gallery, distributed in two
niches excavated to provide room for hydrofracturing equipment: the first, in the farthermost upstream end of a
gallery´s branch, the second, 15 m downstream of this location. This amount of tests was conceived to have
an ample number of results, including tests where eventually no hydraulic fracture of the rock, or inconclusive
or incoherent results were obtained, and on the other hand, having coverage of all possibilities in which the
minor principal stress was found.

Geologically, the project is located within Paleozoic-age metamorphic rocks, composed of quarzitic feldspar
gneiss in the lower portion, gneissic schist in the mid portion, and schist in the upper portion. Rocks are mainly
folded in the schist sequence in a general north-south direction, as consequence of multiple seismic events
that have affected Colombia´s Western Cordillera.

Unconfined strength of intact rock (gneiss) ranged between 25 and 140 MPa, with an average of 80 MPa.
According to Deere and Miller (1966), the rock in the area of the underground powerhouse may be classified
as medium strength rock.

Seventy-six mm cored drillholes were conducted, in order to examine cores and thus best locate the packers
and have either the 0,5m- or the 1.0 m-long pressure chamber’s flute coincide with the rock’s discontinuities.
Typical drillhole lengths were 15m. Drillholes were oriented in different directions to increase the probability of
covering all discontinuities of the rock mass, as well as to finding the actual orientation of the minor principal
stress. Thus, holes were drilled: one horizontal, two inclined 45° downward and 45° forward, respectively, and
two vertical, downwards.

Typical test behavior may be divided into three categories:

8.1.2.1. Effective hydrojacking

Aperture of the rock fracture occurs during the first cycle of four, upon gradual increase of pressure (0.5 -
1.0 MPa/s) at the moment when the rock’s tensile strength, σt, is exceeded, as evidenced by a sudden
increase in flow and an asymptotic drop in the pressure curve. A second cycle, in which the initial aperture
pressure was reached, flow was kept steady for some time to allow the propagation of the fissure, then
shutting the flow instantly to obtain the shut-in pressure value, Ps; in a third cycle, such procedure is repeated,
performing the shut-in maneuver once the fissure aperture pressure is reached, in order to corroborate the
shut-in pressure value; and a fourth and final cycle, upon applying a stepped pressure increase followed by a
sudden flow shut off, before purging the air out of the system, in order to check, once more, the shut-in
pressure value, Ps. In this type of behavior, the pump´s pressure and maximum flow sufficed to overcome the
fracture pressure, Pf, of the rock, and hence the minor principal stress vale was established σ3, or shut-in
pressure, Ps, Eight tests were performed to obtain such results.

8.1.2.2. Incipient hydrojacking

In this case, once the gradual pressure increase has been performed, upon reaching the critical failure value,
normally 8.0 MPa to 9.0 MPa in this type material, the onset of the sudden increase in flow was observed,
however the pump pressure was insufficient to continue opening the fissure and hence, allowing an increase in
flow to the point of stabilizing itself at a given value, so that the proper conditions to perform the shut-in
maneuver were not attained. Nonetheless, a second and third cycle were performed, similar to the first, in
order to discard the possibility of a clogging taking place –with aft washing and aperture- of the fissure. This
result continues to be of value, as the fracture pressure, Pf, and hence the shut-in pressure, Ps, were found to
be greater than those obtained in the hydrofracturing tests where hydrojacking was obtained (P s). Six of the
22 tests showed this type of behavior.

8.1.2.3. No hydrojacking

Characteristic of this behavior was the fact that once and the maximum pump pressure reached values of
about 9.0 MPa, no further increase in flow could be obtained, even though the pump’s capacity could reach 14
MPa, due to the presence of open discontinuities that caused the flow to fall short. As in the previous case, this
result is meaningful as the fracture pressure, P f, and hence the shut-in pressure, Ps, were greater than those
obtained in the tests where hydrojacking was attained. Seven out of the 22 tests showed this type of behavior.

8.1.3. Results

Figure 10.2, below, illustrates a typical hydrofracture test preformed represented in the pressure decay curve,
or shut-in curve, of the last cycle, where a value of 3.95 MPa is observed.
Pressure [Kg/cm2]

Time

Figure 10.2 – Minor principal stress from shut-in pressure

Test results indicate the following:

 There were several tests in which the rock was not fractured or there was no aperture of
discontinuities, even under a pressure of 9.4 MPa.

 At Location 1, the lowest value of the minor principal stress was 3.95 MPa, according to the shut-in
curve, and 3.54 MPa as per the Q vs P curve for the same test (N1-I3 between 11.2 m and 11.70 m
depth.

 At Location 2, the lowest value of the minor principal stress was 3.6 MPa, according to the shut-in
curve, and 3.56 MPa according to the Q vs P curve for the same test (N2-V1, between 13.85 m and
14.35 m).

 In the remaining tests where there was re-aperture of discontinuities, values of the minor principal
stress greater than 4.2 MPa were found, interpreted both based on the shut-in curve and the Q vs P
curve.

Figure 10.3, below, graphically illustrates values of the minor principal stress as deduced from the tests, with
respect to the depth or location of the test. The same figure, the variation of the theoretical stress, estimated
as the vertical rock overburden, times the rock’s unit weight (2.7 ton/m3), as well as the empirical criterion of
the minor stress that may be deduced according to the equation by Stephansson.
Elevation masl

Measured minimum Principal stress

Figure 10.3 - Minor principal stress versus depth of test

From the above results, the following may be deduced:

 It is of note that in the powerhouse area, a stress relaxation phenomenon exists that could be as high
as 50% of the stress as deduced based on rock overburden and the unit weight of the rock. This
reduction in the in-situ stress is mainly due to topographical effects, as there is a steep slope in the
area with topographical protrusions responsible for such phenomenon within the rock mass.

 With respect to Stephansson’s empirical criterion (1993), the measured minor principal stress is about
35% to 25% less than that deduced empirically.

8.1.4. Conclusions

8.1.4.1. Hydraulic Fracturing

 Hydrofracturing tests performed in the powerhouse’s exploration gallery allowed estimating the in-situ
minor principal stress. However, results obtained from the tests, indicate that in this particular area, at
the onset of the steel lining in the headrace tunnel, there are stresses lower than those deduced from
empirical overburden criteria. Consequently, the factor of safety against hydraulic fracture at this point
was 1.04, which is unacceptably below the required 1.3.
 The reduction in the measured in-situ minor principal stress is due mainly to the presence of
topographic protrusions in the area. In general, topographic protrusions, and steep valley slopes will
tend to induce relaxation of stresses in the rock mass, and hence, a decrease in confining minor
principal stress values.

 To counteract the hydraulic fracture phenomenon, with a minimum safety factor of 1.3, the initially
designed steel liner length requires be extended by 135 m. In general, the initially estimated
theoretical steel liner length should be checked, and adjusted, if required, based on hydrofracturing
tests performed along the newly constructed portion, or portions, of a pressure tunnel, starting at the
upstream end of the design stage steel liner, and continue testing upstream, until test results indicate
that a factor of 1.3 has been attained.

 Orogeny, or formation process of mountain ranges, where typically one plate subducts another plate
lifting and folding rock, as is the case of the Andean Range, also provides interesting clues regarding
the relation and orientation of the main (three) principal stresses, as well as between the relation
between the horizontal/vertical stresses (K factor), and the phenomena that have consequently
developed, which values may be estimated quite reliably, based on hydraulic fracture, as well as
overcoring tests. The values of the three main principal stresses, or stress tensor, constitutes a
significant piece of information for the optimum orientation of the underground powerhouse, whereas
the minor stress obtained from such tests, may corroborate the values of the minor stress obtained
from hydraulic fracture or hydrojacking tests. In the particular case of Porce III, the minimum principal
stress obtained from previously conducted overcoring tests of 3.92 MPa, correlate quite well with the
minor stresses obtained from hydrojacking tests at Locations 1 and 2, of 395 MPa and 3.6 MPa,
respectively.

 The lowest measured minor principal stress value was 3.95 MPa at Location 3, with an azimuth of
227° and inclination with respect to the horizontal plane of 20°; this minor stress value is quite similar
in magnitude to the minor stress values of 3.95 MPa 3.6 MPa interpreted from the hydraulic fracture
tests performed at nearby Locations 1 and 2, respectively of the exploration gallery.

 The pump used in the hydrofracturing tests can reach pressures of up to 2000 psi (140 bar
approximately), the upper value in the pressure scale corresponding to the upper 20-mA value in the
transducer’s scale (zero pressure corresponds to 4 mA). However, maximum pump flow reached only
about 30 l/min, which although in many cases may sufficed to crack open a fissure in the rock mass,
or re-open an existing one, could, on occasions, fall short of the flow required to maintain open and
propagate the fissure under a steady flow condition to perform the shut-in maneuver, as discussed
above (see No Hydrojacking),. This may occur in fissured and/or folded, rather than intact, rock,
where the pumped pressure water will tend to quickly find one or more open discontinuities and
scatter the flow among them, rendering the flow insufficient to conduct the test adequately. Hence, an
adequate pump, not only from the pressure but the flow standpoint, should be implemented from the
start, with a larger flow capacity, according to a previous analysis of discontinuities.

 The minor principal stress orientation and value requires delving, in more detail, into the pressure
versus flow curves, for adequate interpretation, as the initially pumped water could flow more easily
into a fissure that is not normal to the minimum stress, rendering a false value of such, a test that
should ultimately be discarded. This, amongst other misleading causes, should encourage conducting
a significant buy sufficient number of tests in order to end up with an adequate value, which is
fundamental for decision-making regarding the length of the steel liner.

8.1.4.2. Overcoring

 In addition to the Hydrofracturing tests, in-situ stress measurements were conducted using the USBM
overcoring cell technic at three locations along the powerhouse exploration gallery: 13 tests at
Location 1, 9 tests at Location 2, and 11 tests at Location 3, for a total 33 tests.

 An equal number of lab biaxial chamber tests were performed, except for two cores at Locations 1
and 2, that broke during retrieval and hence no modulus of elasticity data could be obtained.
However, the stress tensor at these locations could be obtained based on all the remaining data.

 Based on the biaxial chamber results, the rock has a relatively high degree of anisotropy, which
variation in moduli between axes varies between 22% and 64%.

 Results obtained at Location 2 show the highest correlation value (0.92), considered to be quite
acceptable. Consequently, tests conducted at Location 2, seem to be the most representative and
hence the most appropriate to be used for the analysis of redistribution of stresses and strains in the
underground powerhouse.

 According to stress tensor data at Location 2, the stress ratio K for the orientation of the powerhouse
of (N27°E), is about 1.83, with a horizontal and vertical stresses of 16.07 MPa and 8.76 MPa,
respectively. 

 The minor principal stress (3) at Location 2, is about 7.76 MPa - which is equivalent to an
overburden of 282 m - and inclined 56° with respect to the horizontal plane, with an azimuth of 208°.

 The lowest measured minor principal stress value was 3.92 MPa at Location 3, with an azimuth of
227° and inclination with respect to the horizontal plane of 20°; this minor stress value is quite similar
in magnitude to the minor stress values of 3.95 MPa 3.6 MPa interpreted from the hydraulic fracture
tests performed at nearby Locations 1 and 2, respectively of the exploration gallery.

 The magnitude of this stress indicates that there may exist a significant in-situ relaxation of stresses
in the powerhouse area with respect to those deduced from empirical rules, such as discussed in the
Hydraulic Fracturing Report - Analysis and Interpretation of Tests, included above, which reinforces
the need to extend the steel lining, as suggested therein.

 Regarding the orientation of the major and intermediate stresses as measured on site at Locations 2
and 3, with respect to orientation of the powerhouse (N27°E), these are considered to be unfavorable
with respect to the design hypothesis, as they are acting in simultaneous form on the caverns’ walls,
which eventually would require additional support, due to possible rock blocks or wedges generated
by discontinuities along the excavated branch off the exploration gallery.
 The main and most frequent impasse to adequately perform overcoring tests has fallen on the too-
often difficulty of retrieving intact minimum 30 cm long rock cores, free of any significant discontinuity
that will most likely induce failure of the core sample, thus rendering the biaxial lab test impossible to
perform, wherefore either additional overcoring holes need to be drilled, or the existing ones
extended, until intact cores can be retrieved.

 However, problems may emerge not only during core retrieval, but during the overcoring test itself,
whenever the segment of rock being overcored ruptures and causes the connection cable to gyrate in
an uncontrolled manner, along with the overcoring bit, until a full stop of the equipment is achieved,
by cutting of power. On one occasion, at a different project, this mishap caused rupture of the cable
as well. To prevent this accident from happening, the inner 1.5 “ drillhole rock cores should be
inspected in detail to look for unwanted discontinuities that may induce this phenomenon when the
inner hole is being overcored.

 Figure 25, below, graphically illustrates the results of both, the hydrofracturing as well as the
overcoring tests. The minimum measured stress level profile, falls well below the ground surface
profile and served as basis to: 1) extend the lower headrace tunnel steel liner upstream, from Station
K12+622.77 to Station K12+576.22, and 2) moving the pressure shaft 50 m upstream, from Station
K12+447.20 to Station K12+397.20, deeper into the mountain, in order to gain additional overburden
to preclude hydrofracture phenomena, both cases in compliance with a minimum factor of safety of
1.3, as shown in the upper Factor of Safety graph.
Figure 25. Minimum stress profile measured by hydraulic fracture and stress tensor measured by overcoring

8.2. CASE HISTORY – IN SITU STRESS MEASUREMENTS AT THE UNDERGROUND


RESEARCH LABORATORY (URL), MANITOBA, CANADA.

8.2.1. General Characteristics and geological framework

In this case history, the carried out activities regarding in situ stress measurement at the Atomic Energy of
Canada Limited’s (AECL’s) Underground Research Laboratories (URL) are briefly described. The URL
construction began in 1982 and was designed to address the needs of the Canadian nuclear fuel waste
management program. This project comprises one of the most comprehensive programs of geologic and
geotechnical characterisation ever developed in Canada, so this case study aims at summarizing some of the
procedures and selected results from those extensive studies. The reader who wants to obtain more
information is directed towards the original papers that are presented as references in the text.

The project location and general layout are shown in Figure 26 and Figure 27.
Figure 26. Underground Research Laboratories (URL) location, from Chandler (2003).

Figure 27. Underground Research Laboratories (URL) layout. Adapted from Chandler (2003)

Regarding the geological framework, the URL is allocated in the western part of the Canadian shield,
specifically in Pinawa (Manitoba) at 433,0 m depth. The required excavations for these facilities are located in
a granitic batholith called the Lac du Bonnet granite which belongs to the Late Kenoran age within the
Winnipeg River plutonic complex of the English river gneiss belt of the Western Superior Province. The rocks
are undifferentiated massive porphyritic granite-granodiorite. The rock mass is massive, with medium to
coarse grained with local gneissic banding.

With respect to the main geological structures, two major thrust faults were intersected by the excavations with
dips of around 25° to 30° southeast. Those faults are named fracture zone 2 and 3 (FZ2 and FZ3) as
presented on the Figure 28. Site geology at URL. Adapted from Thompson and Chandler (2004).Figure 28.
Figure 28. Site geology at URL. Adapted from Thompson and Chandler (2004).

8.2.2. In situ stress measurement

Among many other things, AECL’s in situ stress program was designed to evaluate techniques for determining
the in-situ stresses in rock and to investigate the factors which influence stress magnitude and variability
(Chandler et al., 1996) in order to achieve this a total of over 350 far-field triaxial overcoring tests and
approximately 80 hydraulic fracturing measurements were carried at the URL on an estimated volume of rock
of around 100 m X 100 m X 500 m. 

Moreover, during excavation of the circular shaft and tunnels, convergence measurements were made
according to the face advances. The convergence measurements using pre-installed extensometers were later
used to compute back analysis of the in-situ stress tensors (Chandler et al., 1996).

The main characteristics of these tests and procedures  (i.e. overcoring, hydraulic fracturing, and back
analyses) are summarized below.

8.2.2.1. Hydraulic fracturing tests at the URL - Canada.

Hydraulic fracturing tests were carried in vertical boreholes at depths from 50 m to 600 m, before the shaft
excavation took place in 1981 and 1982. Tests done below 300 m produced subhorizontal fractures or no
fracture at all (because of the equipment limitations). The horizontal measured stress magnitude above 300 m
was later confirmed by overcoring tests.
Further tests were carried below 300 m, where only subhorizontal fractures were created. The method
proposed by Hefny and Lo (1992) was used to calculate the maximum horizontal stress. However, the stress
derived from this method is sensitive to the strength parameters chosen, particularly when the ratio of
maximum to minimum horizontal stress is close to unity (Chandler et al., 1996). Given that the method
produced very high horizontal stresses that were not correlated with the observation of failure around the URL
openings, it was concluded by the authors that this methodology needed further refinements.

In addition, 14 hydraulic jacking tests of existing or created fractures were carried at URL at a depth between
445 m and 512 m.

8.2.2.2. Overcoring tests

Several overcoring tests were carried at URL using the USBM gauges, the triaxial strain cells, or the 2D
doorstopper cell. A total of around 1000 overcoring tests were carried at URL, among which 350 were far field
triaxial stress measurements. A great deal of research and development was done at URL using overcoring
methods to examine the influence of geological features, comparing different techniques, assessing the effect
of scale effect, developing deep hole overcoring techniques and improving the methods of analysis and
interpretation. The greatest limitation for overcoring tests at URL was the high horizontal stress which has
often produced disking, making elastic interpretation difficult.

In order to overcome this limitation, the doorstopper method was used at the 420 Level. To this respect, it is
worth mentioning that the smaller solid cores on which the doorstopper cell is attached were not affected by
disking of the core, while in many tests, the larger hollow cores from overcoring with Leeman cells were
subjected to discing (Chandler et al., 1996).

8.2.2.3. Shaft convergence

The shaft at URL was excavated in two distinct phases. The first phase was done from surface to 255 m as a
rectangular section (2.8 m X 4.9 m), and phase 2 was done from 255 m to 443 m deep as a 4.6 m diameter
circular shaft. The drill and blast method was used for shaft excavation. 

Failure observations were used to identify the direction and approximate magnitude of the horizontal stresses,
and in some cases providing an indication of in situ rock mass strength. Moreover, the convergence of the
circular shaft was measured at 26 locations between 267 m and 430 m. Convergence measurements were
used in back analysis to compute the in situ stress. The total convergence measured in the shaft varied from
1.0 mm to 3.0 mm, and total convergence was attained when the excavation face had advanced two diameter
(± 10 m). In situ stresses were calculated using the Kirsch equations for stresses and displacement around a
circular opening. A least square solution was fit to convergence data in the shaft.

8.2.3. Results

The Figure 29 shows the maximum stress magnitude and orientation versus depth obtained from the different
kinds of tests that were conducted for the URL. 
Based on the results compiled and shown in the above figure, it is possible to identify a good agreement
between the outcomes of the different types of tests performed at a depth above 300 m regarding maximum
horizontal stress. At this 300 m depth, the main finding can be seen in the same Figure 29, this is the sharp
transition from moderate stress to high stress values, which it was explained over the hypothesis that the
fracture FZ 2 was acting as a stress domain boundary (Chandler et al., 1996).
 
In spite of tests result of hydraulic fracturing and overcoring carried above 300 m show a good agreement with
each other, the horizontal stress orientation obtained by the firsts ones were showing differences from 30° to
60° compared to the latest ones (See Figure 29). The reason for that difference was not well established,
however, it was attribute to either the natural variation of preferred joint direction leading to a variation in the
direction of induced fracture or to the existence of a specific kind of mineral (i.e magnetite) in the granitic rocks
at URL which may have affected the compass direction on the impression packers (Chandler et al., 1996).

The stress tensor obtained from hydraulic jacking tests of existing fractures were in good agreement with
stress tensors calculated from convergence and back-analysis of displacements on the 420 level. 

Regarding stress orientation, the results show that the maximum horizontal stress rotates about 90° from
above 300 m to below 300 m, and the stress is aligned with the dip direction of FZ 2. It is noted that below FZ
2, the stress direction concurs with the principal stress direction for the Canadian Shield. However, the stress
magnitude just below 300 m is more than the double of the average stress given by academic literature
(Herget,1980).

Figure 29. Maximum horizontal stress (magnitude and direction versus depth). Taken and adapted from
Chandler et al.(1996.)
Martin and Chandler (1993) have proposed an explanation to explain the stress relief above 300 m and the
stress orientation shift below 300 m. These authors have assumed that since considerable reverse slip
movements have occurred across the FZ 2 zone, the movement has possibly created stress relief in the dip
direction of the overlying rock. Also, it is mentioned that the deformation modulus of the fractured rock mass
above 200 m was around 30 to 40 GPa, while the deformation modulus of the more massive rock mass below
200 m was around 60 GPa. This difference in modulus along with the proximity of the fault FZ 2, would be the
probable cause of the high stress measured at depths below 200 m. 

Horizontal in situ stresses calculated from convergence measurements have shown without doubts the clear
discontinuity at FZ 2, both in magnitude and direction. It was however very important to determine the
percentage of convergence that has occurred before the installation of convergence pins. Numerical models
result suggests that a large percentage of the displacement have already occurred. In situ measurements of
displacements ahead of the face suggested that 55% to 60% of the total radial displacements occurred ahead
of the face. Thus, a correction factor of 60% was used for convergence measurements in the shaft.

8.2.4. Conclusions

The URL performed a very large number of in situ stress measurements in order to establish boundary
conditions for the operating phase experiments at the site. The lesson learned from the stress measurements
at this project is summarized by Chandler et al.(1996) who established that the determination of a stress
tensor is an iterative process which continuously builds upon successive stages of stress information
collection, each stage improving the level of details. Good information may not be obtained from a single hole,
confidence is obtained from a variety of test techniques.

It was shown at the URL that even in a relatively homogeneous plutonic rock mass, the stress magnitude and
orientation can be highly variable. This is the case especially near regional geological structures. Therefore,
the knowledge of geological structures is required to correctly interpret stress measurement results (Martin and
Chandler, 1993).

8.3. CASE HISTORY - IN SITU STRESS MEASUREMENTS FOR UNLINED PRESSURE


TUNNEL DESIGN IN QUEBEC, CANADA

8.3.1. General Characteristics and geological framework

The case study 2 deals with the activities carried out by Hydro-Quebec which is the Governmental Company
that is the main electrical producer in the Province of Quebec (Canada) with a total installed production
capacity of more than 37,000 MW of hydro-electrical power. Regarding in situ stress measurement, since
1997, Hydro-Québec has prescribed the use of hydraulic jacking tests for all their projects that include design
for unlined pressure tunnels (Quirion and Tournier, 2010). 

Bearing this in mind, this case study summarises the Hydro-Quebec experience presented in Quirion and
Tournier (2010) while adding two more recent private project results and interpretations that were presented in
conferences (Rancourt et al. (2013) and Rancourt et al. (2016). As in case study 1, the reader who wants to
obtain more detailed information can resort to the original papers that are presented as references in this
document.
As for the geological  characteristics, all projects in this case review are located in a precambrian craton so-
called Canadian Shield. In the Quebec Province, the north shore of the Saint-Laurent River forms a part of this
shield. The rocks are part of the Grenville or the Superior geological provinces. Both provinces contain large
varieties of plutonic rocks (granite, monzonite, syenite, gabbro, anorthosite) and metamorphic rocks (gneisses,
migmatites, charnockite, metasedimentary, among others). The rock mass is typically cut by dykes, sills, joints
and fault zones. The intact rocks are generally hard and sound, but the rock mass quality can vary rapidly on a
short distance from excellent to very poor.  

The region is largely covered by an irregular moraine layer, that is overlain by more recent deposits. The
moraine is the result of the Wisconsin glaciation that covered the region with ice up to 4 km thick. The shape of
the topography is largely the result of the ice erosion, carving softer zones into valleys and leaving hard
unfractured rock masses forming the bulk of the higher altitude Laurentian Mountains. Also, of most
importance on the structural geology of the region, is the glacier retreat isostatic effect, which has created near
surface sub-horizontal decompression joints from the gradual uplift of the earth crust. 

8.3.2. In situ stress measurement campaign

8.3.2.1. Hydraulic jacking tests

These tests are conducted by Hydro-Quebec under the premise that the stress state defined in preliminary
stages by analytic procedures like the suggested one in Broch (1982) have to be verified through in situ stress
measurements in the final design of unlined pressure tunnels.

The equipment used is according to standards (e.g. ASTM, 1989; ISRM, 2003) with single or double packer
systems, pumps, pressure sensors and high frequency data acquisition systems. In addition, borehole
televiewer results are used to locate existing joints and open joints. The tests are done to verify the following
equation with a safety factor generally above 1,5.

σ min
FS=
ΔH w γ w
 
Where,
𝜎 : minMeasured minimum stress from hydraulic jacking (MPa)
ΔH : Static water pressure from the reservoir at the level of the test (m)
w

γw : Water unit weight (9.81 kN/m ). 3

The main reasons for choosing the hydraulic jacking method in an unlined pressure tunnel, are based on two
premises:

 For these excavations it is more important to determine the effect of the increase of water pressure in
the rock mass than to determine the exact stress tensor (Broch et al., 1997). 
 Hydraulic jacking tests are covering a bigger volume than for instance the overcoring methods, which
fit with the scale of a tunnel.

The Table 8 below presents a summary of the projects that have included a hydraulic jacking testing program
for the design of the steel liner length.

Table 8. Hydropower projects owned by Hydro-Quebec that have used hydraulic jacking testing. Modified from
Quirion and Tournier (2010)
Test Depth (max) Reservoir pressure (P)
Project Number of Tests Boreholes
Year
(m) (MPa)
Toulnustouc 2003; 83 18 145 1,6
2004
Peribonka 2003; 43 6 100 0,9
2004
EM-1 2004 14 4 60 0,8
EM-1A 2004 12 3 80 0,8
Romaine-1 2004 19 2 85 0,6
Romaine-4 2004 9 1 60 1,1
Romaine-2 2008; 32 6 150 1,7
2009
Sheldrake 2012 14 1 70 0,7
Saint-Joachim 2015 10 1 64 0,7

8.3.2.1. Special test – Constant pressure hydraulic jacking test

Special tests were carried at the Romaine-2 site as presented in Noel et al. (2011). The special test is a
constant pressure test that was proposed by Fernandez and Alvarez (1994). This test could be used as a
validation test when measured minimum stresses are marginal. These authors are suggesting that the jacking
pressure measured in prolonged tests provide a better evaluation of the jacking behavior than the short
conventional jacking test. 

In Romaine-2 project, only one packer was used to test a 30 m long hole section with constant pressure at 1.7
times the static pressure in the tunnel. The test duration was set at 8 consecutive hours in order to observe
flow rate variation. The test pressure was set in order to create jacking conditions of the tested section.
 

8.3.3. Results

For all the mentioned projects in , the hydraulic jacking programs generally aimed at testing the most possible
joint families in the tunnel area in order to find the minimum hydraulic jacking stress and if possible, identify in
which joint family it was located.

presents statistical results and the ratio of minimum stress to static pressure (FS) for the studied projects. For
cases where the ratio was below 1,5, the steel liner has been extended to reach the point along the axis of the
tunnel where the ratio is above 1,5. For the cases where the ratio is above 1,5, the steel liner was
exceptionally shortened to reach the point where the ratio is at 1,5.

Table 9. Hydropower projects owned by Hydro-Quebec that have used hydraulic jacking testing. Tests results.
modified from Quirion and Tournier (2010)
Interpreted minimum stress (MPa)
Project Min Max Avg Std. Dev. σmin/P
Toulnustouc 0,6 8,2 2,6 1,7 0,4
Peribonka 0,8 3,1 1,8 0,6 0,9
EM-1 1,1 2,7 1,6 0,4 1,4
EM-1A 1,1 2,1 1,4 0,3 1,4
Romaine-1 1,1 3,7 2,5 0,8 1,8
Romaine-4 1,0 4,2 2,5 0,8 0,9
Romaine-2 1,2 3,1 2,1 0,5 0,7
Sheldrake 1,5 3,6 2,5 0,7 2,1
Saint-Joachim 0,8 2,1 1,7 0,4 1,1

It is interesting to mention that in some projects like Toulnustouc and Romaine-2, the tests identified several
low stress areas, with minimum stresses much lower than the predicted gravity stresses estimations. While
other projects like Romaine-1 and Sheldrake showed minimum stresses higher than the gravity predictions.

From the aforementioned results the key aspects identified from their assessment are summarized as follows:

 These results are showing the high heterogeneity and high spatial variability of the stress tensors
within fractured crystalline rock masses that have been subjected to decompression. Besides, they
show the difficulties of trying to estimate stresses using gravity assumptions.

 Quirion and Tournier (2010) have observed that the ratio of the peak pressures P max/P (see
c r3 max

reference Figure 30 obtained for the conducted tests, are relatively constant with values between 1,4
and 1,8 for all the project sites. This could indicate that in crystalline rocks masses, the ratio
represents a typical signature of the fracture initiation/reopening process. It is noted that those peak
values are not influenced by data interpretation because they come from raw data. 

 Those zones identified especially on Toulnustouc and Romaine-2 projects where in many tests, the
measured minimum hydraulic jacking stress was much lower than the gravity stress predictions from
the depth of rock cover even when using the topographic corrections. These results can be due to the
effect of open joints systems that somehow are forming possible arches below which stresses are
relaxed. This may also worsen by regional fault, shear zones, and dykes structural effects. 

 Those zones identified especially on Romaine-1 and Sheldrake projects where in several tests, the
measured minimum hydraulic jacking stress was higher than the gravity stress predictions from the
depth of rock cover. These results are showing a rock mass that has been, locally at least, kept
compressed. This could possibly be explained by topographic variations, old tectonism or remnant
stresses. But for the pressure tunnel design purpose, those results indicate high rock mass
confinement, so the liner does not need to be extended, it can stay as planned or may possibly be
shortened. In this sense, the liner length to be reduced in the Sheldrake project was investigated by
Rancourt et al., (2013) using a simple numerical model to study the minimum stress variations along
the liner according to a calibrated model for the case of a surface. Figure 30 shows the numerical
model geometry and minimum stress results). The steps of the carried process are described in the
following lines: 

o In first place, the model was calibrated to fit with the in place measured hydraulic jacking
stress. This was done by increasing the rock density in the model parameters, until the
minimum stress isocontours fit with test results at the test locations. The bold line in the
Figure 30Figure 30. Numerical model used to determine the final length of the liner for intake
tunnel of the Sheldrake Project. represents the stress 1.5 times above the static pressure
(900 kPa). 

o Secondly, the excavation of the powerhouse was added to the model to see its effect on the
minimum stress distribution. The minimum stress after the powerhouse excavation is
presented in a dotted line on the figure below. It can be observed that the effect of the
excavation pushes the minimum stress around 15 m further upstream. Thus, the liner length
that was preliminarily designed at 55,0 m, was adjusted down to 43,0 m to fit the dotted line
that takes into consideration the effect of the excavation. The powerhouse operates
successfully with no leakage since its commissioning in 2013.   

Figure 30. Numerical model used to determine the final length of the liner for intake tunnel of the Sheldrake
Project.

As for the constant pressure test, the results are presented on Figure 31 which shows P (pressure) and Q
(flow rate) curves for the test. The flow rate shows rapid fluctuations that may be associated with very small
pressure variations indicating that the fractures are fully open. But the most important result is that the flow
rate does not increase during the test showing that no uncontrolled fracture extensions were occurring within
the tested zone. 
Figure 31. P and Q curves for the constant pressure test. Adapted from Noel et al (2011).

8.3.4. Conclusions

Based on the results of the hydraulic jacking tests that were carried for the Hydro-Quebec projects mentioned
in Table 9, the following conclusions were made:
 For an unlined pressure tunnel, it is more important to determine the effect of the increase of water
pressure in the rock mass than to determine the exact stress tensor. For the design of pressure
tunnels, hydraulic jacking tests is the method chosen by Hydro-Quebec because it reproduces at a
smaller scale, the effect of a pressure tunnel saturating a rock mass with pressurized water.

 Old crystalline rocks that have been subjected to glaciation are showing decompression joints,
faulting and complex topography that strongly influence the local stresses distribution.

 Near surface low minimum stress zones were interpreted as decompressed rock masses with open
jointing. In general, these results in the liner being extended to reach sufficient minimum stress.

 Near surface high minimum stress zones were linked with locally highly compressed rocks,
topographic variations, or locked in stresses. When the minimum stress is high the liner length stays
as planned or can even be shortened.

 Constant pressure test was used to validate steel liner length in the case of a project where hydraulic
jacking results were showing very slow minimum stress increase with the increase in rock cover. The
result of the constant pressure test has confirmed the preliminary liner length design. 
9. REFERENCES
Amadei, B. 1983. Rock Anisotropy and the Theory of Stress Measurements.
Amadei, B. & Stephansson, O., 1997. Rock Stress and its Measurement First Edit., Springer.
Ask, D. et al., 2009. Rock stress, rock stress measurements, and the integrated stress determination
method(ISDM). Rock Mechanics and Rock Engineering, 42(4), pp.559–584.
Barton N. 2007 Rock Quality, Seismic Velocity, Attenuation and Anisotropy // Taylor & Francis Group.-.
London, UK.
Boyce, G. M., 1981. Acoustic Emission Signatures of Various Rock Types in Unconfined Compression.
Brekke, T.L. and Ripley, B.D., 1987, Design Guidelines for Pressure Tunnels and Shafts, EPRI, Research
Project 1745-17, AP-5273, Final report, 160 pages.
Brock, E., Dahlo, T.S., Hansen, S.E., 1997, Hydraulic jacking tests for unlined pressure tunnels, Hydropower
1997, Balkema, pp. 581-585.
Broch, E., 1982, The Development of unlined pressure shafts and tunnels in Norway, Rock Mechanics: Cavern
and Pressure Shafts, Wittke ed., Balkema, pp. 545 – 554.
Broch, E., 1984, Unlined High Pressure Tunnels in Areas of Complex Topography, Water Power & Dam
Construction, Nov, pp.21-23.
Cai, M., Qiao, L., Yu, J., 1995. Study and test of techniques for increasing overcoring stress measurement
accuracy.
Cai, M. & Peng, H., 2011. Advance of in-situ stress measurement in China. Journal of Rock Mechanics and
Geotechnical Engineering, 3(4), pp.373–384.
Chandler, N.A., Read, R.S., Martin, C.D., 1996, In Situ Stress Measurement for Nuclear Fuel Waste
Repository Design, Proc. 2nd American Rock Mechanics Symposium, Montreal, Balkeman, Rotterdam, pp
929-936.
Chandler, N.A., 2003, Twenty Years of Underground Research at Canada’s URL, WM’03 Conference, Tucson,
AZ, February.
Chang, C., Lee, J., 2010. Interaction between regional stress state and faults: Complementary analysis of
borehole in situ stress and earthquake focal mechanism in southeastern Korea.
Christiansson, R. & Hudson, J., 2003. ISRM Suggested Methods for rock stress estimation—Part 4: Quality
control of rock stress estimation. International Journal of Rock Mechanics and Mining Sciences, 40(7-8),
pp.1021–1025.
Cornet, F., Valette, B., 1984. In situ stress determination from hydraulic injection test data. Journal of
Geophysical Research 89.
Cornet, F., 1993, The HTPF and the Integrated Stress Determination Methods, Rock Testing and Site
Characterization, Comprehensive Rock Engineering- Science Direct, pp 413-432.
Corthésy, R., Leite, M., Gill, D., Gaudin, B., 2003. Stress measurements in soft rocks. Volume 69, Issues 3–4.
Pages 381-397.
Doe, B. R., 1983. The possible bearing of the granite of the UPH Deep Drill Holes, Northern Illinois, on the
origin of Mississippi Valley ore deposits.
Deere D., Miller, R., 1966. Engineering classification and index properties for intact rock.
Duvall, W.I. (1974) Stress relief by center hole. Appendix in US Bureau of Mines Report of Investigation RI
7894.
Enever, J.R., Chopra, P.N., 1986. Experience with hydraulic fracture stress measurements in granite.
Fairhurst, C., 2003. Stress estimation in rock: a brief history and review. International Journal of Rock
Mechanics and Mining Sciences, 40(7-8), pp.957–973.
Feng, X., 2015. Stochastic soil water balance under seasonal climates. Proceedings of the Royal Society.
Goodman R.E., Kulhawy, F.H., 1987. Foundations in rock. Chapter 15, Butterworths, London, UK.
Gonzalez de Vallejo, Luis I., 2004. Ingeniería Geological. Pearson Educación. Madrid, España.
Gronseth, J. M. and Kry, P. R., 1981, Instantaneous shut-in pressure and its relationship to the minimum
stress. Proc. Workshop on Hydraulic Fracture Stress Measurements, Monterey, CA, pp. 55-60.
Guglielmi, P., 2015. Mechanical characterization of oligo (ethylene glycol) - based hydrogels by dynamic
nanoindentation experiments.
Guo TK., 2016. Development of miniature hydraulic fracturing simulation teaching experimental device [J].
Exp. Technology Management.
Haimson, B.C., 1993, The Hydraulic Fracturing Method of Stress Measurement: Theory and Practice, Stress
and Stress Measurements Methods, Chap 14, Comprehensive Rock Engineering, pp.395-412, Pergamon,
Oxford.
Haimson, B.C. & Cornet, F.H., 2003. ISRM Suggested Methods for rock stress estimation—Part 3: hydraulic
fracturing (HF) and/or hydraulic testing of pre-existing fractures (HTPF). International Journal of Rock
Mechanics and Mining Sciences, 40(7-8), pp.1011–1020.
Herget, G., 1980, Regional stresses in the Canadian Shield, in Proc. Of 13 th Canadian Rock Mechanics
Symposium, CIM 22, pp 9-16.
Hubbert, M., Willis, D., 1957. Mechanics of Hydraulic Fracturing. Trans AIME.
Hudson, J.A. and J. P. Harrison, 1997, Engineering Rock Mechanics, An Introduction to the principles, Elsevier
Science.
Hudson, J.A., Cornet, F.H. & Christiansson, R., 2003. ISRM Suggested Methods for rock stress estimation—
Part 1: Strategy for rock stress estimation. International Journal of Rock Mechanics and Mining Sciences,
40(7-8), pp.991–998.
Hudson, J.A., and X.T., Feng, 2015, Rock Engineering Risk, Taylor & Francis Group, 594 pages.
Interfels GmbH, 2005. Instrumentación Geotécnica de Túneles. Consultor Geotécnico.
Lee, M.Y., Haimson B.C, 1989. Statistical evaluation of hydraulic fracturing stress measurement parameters.
Lehtonen, A., Särkkä, P., 2006. Evaluation of rock stress estimation by the Kaiser effect.
Liu, Y. H., Cen, Z. Z., Xu, B. Y., 1995. A numerical method for plastic limit analysis of 3-D structures.
Ljunggren, C. et al., 2003. An overview of rock stress measurement methods. International Journal of Rock
Mechanics and Mining Sciences, 40(7-8), pp.975–989
Lu B., J.M. Wang, X.L. Ding, A.Q. Wu, S.H. Duan, S.L. Huang, 2010, Study of deformation and cracking
mechanism of surrounding rock of Jinping I underground powerhouse, Chinese Journal of Rock Mechanics
and Engineering, 29 (12), pp. 2429-2441.
Martin, C.D., and N. A. Chandler, 1993, Stress Heterogeneity and Geological Structures, Int. J. Rock Mech.
Sci. & Geomech. Abstr., Vol. 30, No. 7, pp 993-999.
Marulanda, C., 1990. Hydraulic Fracturing as a decisive element in the design optimization of pressure
tunnels. Bogotá, Colombia.
Marulanda, A., Gutiérrez, R.,García, J.P.,2003. Hydraulic Fracturing Tests - Execution, Analysis and
Interpretation. Porce III Hydropower Project. Bogotá, Colombia.
Matsuki, K., Sakaguchi, K., 1999. Applicability of Core-based in-situ Stress Measuring Methods to Limestone
Evaluated by Comparison with Conical-Ended Borehole Method. Japan.
Mercier, J., Vergely, P., 2002. The Paikon Massif Revisited, Comments on the Late Cretaceous - Paleogene
Geodynamics of the Axios - Vardar Zone. How Many Jurassic Ophiolitic Basins? (Hellenides, Macedonia,
Greece).
Merrill, R.H., 1964. In situ determination of stress by relief techniques, in Proc. Int. Conf. State of Stress in the
Earth’s Crust, Santa Monica, Elsevier, New York, pp. 343–369.
Nilsen, B., Thidemann, A., 1993. Rock Engineering, vol 9. Norwegian Institute of Technology. Division of
Hydraulic Engineering.
Noel, J.F, Claisse, M., Rancourt, A.J. and D. Babin, 2011, Introduction of constant pressure hydrojacking tests
to validate a pressure tunnel’s steel liner length, 1st Int. Congress on Tunnels and Underground Structures in
South-East Europe, Dubrovnik, Croatia, April 7-9.
Obara, Y., Sugawara, K., 2003. Rock Stress. Proceedings of the Third International Symposium on Rock
Stress. Kumamoto, Japan.
Quirion, M and J.P. Tournier, 2010, Hydraulic jacking tests in crystalline rocks for hydroelectric projects in
Québec, Canada, Rock Stress and Earthquakes – Xie (ed.), Taylor & Francis Group, London, pp. 513-518.
Rancourt, A.J., Gourdeau, S. and Camiré, G., 2013, The design and construction of the Courbe-du- Sault
hydroelectric project, Sheldrake River, Québec, Canada, GeoMontreal2013, Canadian Geotechnical Society
Conf.
Rancourt, A.J., Claisse, E.B., Gourdeau, S., 2016, The design and construction of the Hydro-Canyon Saint-
Joachim hydroelectric tunnel and related deep excavations, Sainte-Anne du Nord River, Québec, Canada,
Tunnelling Association of Canada 2016 Annual Conference, Ottawa.
Rutqvist, J., Stephansson, O., 1998. Determination of fracture storativity in hard rocks using high-pressure
injection testing.
Sano, T., 2003. An Overview of Rock Stress Measurement Methods.
Sjöberg, J., Christiansson, R. & Hudson, J.A., 2003. ISRM Suggested Methods for rock stress estimation—
Part 2: overcoring methods. International Journal of Rock Mechanics and Mining Sciences, 40(7-8), pp.999–
1010.
Sjöberg, J., Klasson, H., 2003. Stress measurements in deep boreholes using the Borre (SSPB) probe.
Available at: https://www.sciencedirect.com/science/article/abs/pii/S1365160903001151.

Thompson, P.M. and Chandler, N.A., 2004, In Situ rock stress determinations in deep boreholes at the
Underground Research Laboratory, Int. J. Rock Mech. Sci. & Geomech. Abstr., 41, pp 1305-1316.
Ulusay R, Hudson JA (eds) (2007) The complete ISRM suggested methods for rock characterization, testing
and monitoring: 1974–2006. ISRM Turkish National Group, Ankara, p 628.
Ulusay, R., 2014. The ISRM Suggested Methods for Rock Characterization , Testing and Monitoring.
Valliapan, S., 1981, Continuum Mechanics Fundamentals, A.A. Balkema, Rotterdam, 228 pages.
Weijemars, R., 2011. Analytical Stress Functions Applied to Hydraulic Fracturing: Scaling the interaction of
tectonic stress and unbalanced boreholes pressures, US Rock Mech., Geomech. Symp., San Francisco.
Wittke, G, 2004. Computation of strains and pressures for tunnels in swelling rocks. Tunnelling and
Underground Space Technology
Wittke, J., 2013. Evidence for deposition of 10 million tonnes of impact spherules across four continents
12.800 years ago.
Yew, C.H. and Weng, X. 2015. Mechanics of Hydraulic Fracturing, Second Edition, Elsevier.
Zang, A. and O. Stephansson, 2010. Stress Field of the Earth's Crust, Springer Link, 978-1-4020-8443-0.
Zang, A. and O. Stephansson, 2015. Numerical investigation on optimized stimulation of intact and naturally
fractured deep geothermal reservoirs using hydro-mechanical coupled discrete particles joints model.
Zoback, M., 2011. Reservoir Geomechanics. Stanford University, California, United States.

You might also like