You are on page 1of 34

The role of hydromechanical coupling in fractured rock engineering

Jonny Rutqvist · Ove Stephansson

Abstract This paper provides a review of hydromechan- erties, the HM properties of fractured rock masses are
ical (HM) couplings in fractured rock, with special em- best characterized in situ.
phasis on HM interactions as a result of, or directly con-
nected with human activities. In the early 1960s, the cou- Résumé Ce papier passe en revue les couplages hydro-
pling between hydraulic and mechanical processes in mécaniques (HM) dans les roches fracturées, en mettant
fractured rock started to receive wide attention. A series tout spécialement l’accent sur les interactions HM résul-
of events including dam failures, landslides, and injec- tant de ou directement connectées à des activités humai-
tion-induced earthquakes were believed to result from nes. Au début des années soixante, on a commencé à
HM interaction. Moreover, the advent of the computer vraiment s’intéresser au couplage entre les processus hy-
technology in the 1970s made possible the integration of drauliques et mécaniques dans les roches fracturées. Une
nonlinear processes such as stress–permeability coupling série d’évènements, dont des ruptures de barrages, des
and rock mass failure into coupled HM analysis. Cou- glissements de terrain et des séismes induits par des in-
pled HM analysis is currently being applied to many jections, a été envisagée comme la conséquence d’inter-
geological engineering practices. One key parameter in actions HM. En outre, l’émergence de la technologie des
such analyses is a good estimate of the relationship be- ordinateurs dans les années soixante-dix a rendu possible
tween stress and permeability. Based on available labora- l’intégration de processus non linéaires tels que le coupla-
tory and field data, it was found that the permeability of ge contrainte–perméabilité et la rupture d’un massif ro-
fractured rock masses tends to be most sensitive to stress cheux dans l’analyse HM couplée. L’analyse HM couplée
changes at shallow depth (low stress) and in areas of low est couramment réalisée dans de nombreuses applications
in-situ permeability. In highly permeable, fractured rock en géologie de l’ingénieur. Un paramètre clé dans une tel-
sections, fluid flow may take place in clusters of con- le analyse est une bonne estimation de la relation entre la
nected fractures which are locked open as a result of pre- contrainte et la perméabilité. À partir de données disponi-
vious shear dislocation or partial cementation of hard bles de laboratoire et de terrain, on a trouvé que la per-
mineral filling. Such locked-open fractures tend to be méabilité de massifs de roches fracturées tend à être plus
relatively insensitive to stress and may therefore be con- sensible aux changements de contrainte à faible profon-
ductive at great depths. Because of the great variability deur (faible contrainte) et dans les régions de faible per-
of HM properties in fractured rock, and the difficulties in méabilité in situ. Dans les sections très perméables des
using laboratory data for deriving in-situ material prop- roches fracturées, l’écoulement du fluide peut prendre
place dans des zones de fractures connectées qui sont
maintenues ouvertes du fait de la dislocation initiale par
Received: 29 October 2002 / Accepted: 27 November 2002 cisaillement ou de la cimentation partielle du remplissage
Published online: 17 January 2003 minéral induré. De telles fractures maintenues ouvertes
© Springer-Verlag 2003
tendent à être relativement insensibles à la contrainte et
peuvent alors être conductives à grandes profondeurs. Ce-
pendant, ce papier met en avant la grande variabilité des
J. Rutqvist (✉) propriétés HM en roches fracturées et les difficultés à uti-
Lawrence Berkeley National Laboratory, liser les données de laboratoire pour en déduire les pro-
Earth Sciences Division, priétés in situ du matériau. À cause des difficultés telles
MS 90-1116, Berkeley, CA , 94720 USA
e-mail: jrutqvist@lbl.gov que les propriétés dépendant de la dimension, des désor-
Tel.: +1-510-4865432, Fax: +1-510-4865686 dres dans l’échantillon et un échantillonnage non repré-
O. Stephansson
sentatif, les propriétés HM des massifs de roches frac-
Royal Institute of Technology, turées sont mieux caractérisées in situ.
Department of Land and Water Resources Engineering,
100-44 Stockholm, Sweden Resumen El presente artículo revisa los acoplamientos
O. Stephansson hidromecánicos en rocas fracturadas, haciendo énfasis en
GeoForschungsZentrum, 14473 Potsdam, Germany las interacciones hidromecánicas resultantes o directa-

Hydrogeology Journal (2003) 11:7–40 DOI 10.1007/s10040-002-0241-5


8
Fig. 1a–f Schematic overview
of a fractured geological medi-
um composed of an intact po-
rous rock matrix and macro-
fractures

mente relacionadas con las actividades antrópicas. A co- Introduction


mienzos de los años sesenta, el acoplamiento entre los
procesos hidráulicos y los mecánicos en rocas fractura- In earth sciences, the term “hydromechanical (HM) cou-
das comenzó a ser tenido en cuenta de forma generaliza- pling” refers to the physical interaction between hydraulic
da. Una serie de sucesos, incluyendo roturas de presas, and mechanical processes. HM interactions are common in
deslizamientos de terreno y terremotos inducidos por in- geological media (e.g., soils and rocks) because such me-
yección, fueron considerados como una consecuencia de dia contain pores and fractures which can be fluid filled
las interacciones hidromecánicas. Después, la introduc- and deformable. In general, a fluid-saturated, porous medi-
ción de los computadores hacia los setenta permitió la um or rock fracture can deform as a result either of change
integración de procesos no-lineales, tales como el aco- in the external load or of change in the internal pore-fluid
plamiento esfuerzos-permeabilidad y el fallo de la matriz pressure (Fig. 1). An increased compressive external load
rocosa, en análisis hidromecánicos acoplados. Este tipo (or stress) means that the porous medium as a whole will
de análisis se aplica actualmente a muchos estudios de be compressed to both a smaller bulk volume and a smaller
ingeniería geológica. Un parámetro clave para ello es la pore volume. If the external load is applied “rapidly”, the
correcta estimación de la relación entre esfuerzos y per- reduced pore volume will tend to compress pore fluid and
meabilidad. Basándose en datos disponibles de laborato- thereby increase pore-fluid pressure, because the fluid has
rio y de campo, se ha deducido que la permeabilidad de no time to escape. This is the so-called undrained HM re-
las rocas fracturadas es fundamentalmente sensible a sponse. If, on the other hand, the external load is applied
cambios de la tensión a profundidades someras (baja ten- “slowly”, the fluid has time to escape the compressing vol-
sión) y en áreas de permeabilidad baja. En secciones de ume, in which case the fluid pressure shows almost no in-
rocas fracturadas muy permeables, el flujo de fluidos crease at all. This is the so-called drained HM response.
puede tener lugar en grupos de fracturas interconectadas Likewise, a reduction in fluid pressure or fluid mass can
que están abiertas como resultado de cizallas previas por cause a settlement of the porous media, with accompany-
dislocación o por cementación parcial del relleno mine- ing reduction in bulk and pore volume.
ral. Estas fracturas abiertas tienden a ser relativamente In general, the HM couplings discussed above can be
insensibles a la tensión y pueden por tanto ser conducti- described as “direct” HM couplings, or couplings that oc-
vas a grandes profundidades. Sin embargo, este artículo cur through deformation and pore-fluid interactions. Direct
pretende destacar la gran variabilidad de las propiedades HM couplings include two basic phenomena (Wang 2000):
hidromecánicas en rocas fracturadas, y las dificultades
(i) a solid-to-fluid coupling that occurs when a change
asociadas al uso de datos de laboratorio para estimar pro-
in applied stress produces a change in fluid pressure
piedades de campo de los materiales. Debido a proble-
or fluid mass;
mas como la dependencia de las variables con el tamaño,
(ii) a fluid-to-solid coupling that occurs when a change
las alteraciones de muestras y el muestreo no representa-
in fluid pressure or fluid mass produces a change in
tivo, las propiedades hidromecánicas de las rocas fractu-
the volume of the porous medium.
radas se caracterizan mejor in situ.
and are labeled (i) and (ii). In any of these two cases,
Keywords Fractured rocks · Mechanical · the reduction in pore volume leads to a reduction of the
Hydromechanical coupling · Stress · Permeability cross-sectional area and a reduction in fluid flow capac-

Hydrogeology Journal (2003) 11:7–40 DOI 10.1007/s10040-002-0241-5


9

pling, especially the coupling between stress and perme-


ability, which is most relevant in fractured rock. The pre-
sentation is structured in the sequence of observations,
analysis, and applications. After providing a brief back-
ground, the basic HM behavior of porous intact rock and
fractured porous rock derived from laboratory experi-
ments and theoretical studies is introduced. Thereafter,
in-situ HM behavior of fractured rock masses is intro-
duced. This includes current HM conditions in fractured
rock masses and observations of in-situ coupled HM in-
teractions triggered by human activities. Finally, model-
ing of HM coupling is briefly reviewed, current applica-
tions of HM analysis are presented for a number of engi-
neering practices.

Fig. 2 Hydromechanical couplings in geogical media. (i) and (ii) Background


are direct couplings through pore volume interactions, whereas
(iii) and (iv) are indirect couplings through changes in material
properties Direct HM interactions in the earth’s crust have been
recognized since the late 1800s (Wang 2000). The earli-
est observations included the response of water levels in
ity. Furthermore, reduction of pore volume may result in wells to ocean tidal loading and passing trains (solid-to-
a stiffer material, as more contacts are created between fluid couplings), as well as subsidence of the land sur-
neighboring grains. These changes imply that mechani- face caused by extraction of water, oil, and gas (a fluid-
cal and hydraulic processes can affect each other to-solid coupling). Also, Meinzer (1928) observed large
through changes in material properties, which can be yields of groundwater from the Dakota Sandstone, which
considered “indirect” HM couplings. Accordingly, two could not be explained without the extra storage provid-
basic phenomena of indirect HM coupling may be con- ed by HM changes in pore volume. Observations such as
sidered: these lead to Terzaghi’s (1923) classical contributions to
the HM field, for problems related to settlement analysis,
(iii) a solid-to-fluid coupling that occurs when an applied and Theis (1935) analysis of pumped wells.
stress produces a change in hydraulic properties; Terzaghi defined effective (intergranular) stress, σ′zz,
(iv) a fluid-to-solid coupling that occurs when a change in a geological medium, as the total vertical stress, σzz,
in fluid pressure produces a change in mechanical less the pore-fluid pressure, p:
properties
(1)
These phenomena are labeled (iii) and (iv) in Fig. 2. Equation (1) is the most fundamental equation for de-
Both direct and indirect coupled processes may be fully scribing coupling between hydraulic and mechanical
reversible. However, inelastic responses, including processes in geological media. Terzaghi’s effective stress
yielding, fracturing, or fault slip, cause irreversible law was initially applied to problems related to settle-
changes in porous and fractured media. Although direct ment analysis, dam and slope stability, and petroleum
HM couplings occur in all types of geological media, technology for well stimulation. Terzaghi also derived
they tend to be most important in relatively soft and the first coupled equation for consolidation, which is the
low-permeability rocks and soils. Indirect HM cou- diffusion equation for excess (greater than hydrostatic)
plings tend to be most important in fractured rock or in- water pressure pex:
tact rock with flat intergrain micropores, where changes
in permeability caused by fracture or pore dilation can (2)
be dramatic.
This paper focuses on coupled HM processes trig- where c is a type of diffusivity term known as the con-
gered by human activities (such as underground injection solidation coefficient.
and underground construction), leaving out those in- Jacob (1940) developed the Theis concept of aquifer
duced by natural geological processes, which are thor- storage into a mathematical definition for the coefficient
oughly reviewed in an accompanying article of this spe- of storage or storativity, S:
cial issue (Neuzil 2003, this volume). This paper is fo-
cused on the deeper, water-saturated zone of the bedrock, (3)
leaving out issues related to HM behavior in shallow
soils and clays, which are also treated in an accompany- where ρf is the density of water, A is the horizontal cross-
ing paper in this issue (Alonso et al. 2003, this volume). sectional area for a vertical column of aquifer, ∆mf is the
Furthermore, this paper is focused on indirect HM cou- change in water mass in the column, and ∆h is the

Hydrogeology Journal (2003) 11:7–40 DOI 10.1007/s10040-002-0241-5


10

change in head. Assuming incompressible grains in a The coefficient α, which usually ranges between 0 and 1,
compressible porous media, Jacob (1940) derived the ex- has been measured in laboratory experiments (e.g., Nur
pression for S of the type and Byerlee 1971) for a range of geological materials.
For soils, this coefficient is generally close to one,
(4) whereas in rocks it can be significantly less than one.
The original Biot (1941) formulation of linear poro-
where g is gravity, ϕ is porosity, H is aquifer thickness, elasticity was subsequently reexamined several times by
Kf is the bulk modulus of pore fluid, and K′ is the verti- Biot (1955, 1956) himself, by Geertsma (1966) for set-
cal compressibility of the aquifer. Jacob’s diffusion equa- tlement analysis, by Verruijt (1969) for aquifer behavior;
tion for two-dimensional (in the plane of the aquifer) and by Rice and Cleary (1976) for redefining the materi-
transient flow in a confined aquifer is: al constants. The theory has also been extended to more
sophisticated material models and to incremental varia-
(5) tions in nonlinear systems. However, the pioneering
work by Terzaghi, Theis, and Biot form the basic frame-
work for modern coupled HM analysis in geological me-
where T is transmissivity and t is time. Jacob (1950) de-
dia, currently applied in many disciplines of earth sci-
rived the three-dimensional diffusion equation,
ences.
Human-triggered, indirect HM interactions in the
(6) earth’s crust were probably first recognized in the 1920s
and 1930s when seismic tremors were felt near oil-pro-
where h is hydraulic head, which is the elevation head ducing fields (Pratt and Johnson 1926) and water reser-
plus the pressure head (that is h=z+p/ρfg), µf is fluid vis- voirs (Carder 1945). These HM interactions, known as
cosity, k is permeability, and Ss is specific storage de- “induced seismicity”, were the results of pressure-in-
fined as: duced irreversible failures, usually in the form of shear
slip along discontinuities. The most fundamental criteri-
(7) on for fault slip is derived from the effective stress law
and a Coulomb criterion, rewritten as:
Jacob’s diffusion equations (e.g., Eqs. 5 and 6) are “par-
(11)
tially coupled” in the sense that two- or three-dimension-
al flow is coupled to a simplified, one-dimensional me- where σsc is the critical shear stress, σsc0 is cohesion, µs
chanical model of vertical deformation. This is the basic is coefficient of shear friction, and σn is the normal stress
concept behind most groundwater flow equations and (Scholz 1990). This equation was also applied in the ear-
well-test solutions derived in geohydrology. ly 1900s to various engineering applications in soil me-
Biot (1941) extended Terzaghi’s concept of one-di- chanics, dam and slope stability, and petroleum technol-
mensional consolidation to a general theory of three-di- ogy.
mensional consolidation. Biot’s equations for isotropic Intentional hydraulic fracturing was introduced into
linear elastic porous media can be written in a “mixed petroleum industry in the late 1940s to increase produc-
stiffness form” (Wang 2000), as tion through well stimulation (Clark 1949). The most
fundamental criterion for tensile failure (or hydraulic
(8) fracturing) is that incipient fracture propagation will oc-
cur when the fluid pressure exceeds the least principal
(9) stress by an amount equal to or greater than the tensile
strength of the rock:
where σm is the total mean stress (positive for compres-
sion), K is the usual (drained) bulk modulus (Fig. 1b), εv (12)
is the volumetric strain (positive for contraction), α is
Hydraulic-fracturing theory, as described by Hubbert
the Biot-Willis’ coefficient (Biot and Willis 1957), ξ is
and Willis (1957), was suggested as a method to deter-
the increment of fluid content (positive for “gain” of flu-
mine in-situ stress in the 1960s (Fairhurst 1964). In
id), and M is Biot’s modulus. Equation (8) governs the
classical hydraulic-fracturing stress-measurement theo-
elastic responses of the pore structure; Eq. (9) governs
ry, the breakdown of the borehole wall and fracture ini-
pore-fluid responses. The two equations are coupled
tiation takes place when the least compressive tangential
through the volumetric strain and fluid pressure terms.
effective stress is equal to the tensile stress of the rock
Since the theory describes interaction between pore fluid
(Haimson and Fairhurst 1967). In parallel with the de-
and elastic responses, it has been called the theory of
velopment of techniques for stress measurement, the
poroelasticity. The introduction of the Biot-Willis coeffi-
technology for propagation and control of hydraulic
cient as a factor multiplied to fluid pressure in Eq. (8)
fractures was developed in petroleum engineering
signifies a modification and generalization of Terzaghi’s
(Perkins and Kern 1961; Geertsma and Deklerk 1969).
effective stress law to:
In the early 1960s, HM interactions in fractured rock
(10) received wider attention. A series of events – including

Hydrogeology Journal (2003) 11:7–40 DOI 10.1007/s10040-002-0241-5


11

dam failures, landslides, and injection-induced earth- computer code ROCMAS. This formulation was based
quakes – were suggested to be triggered by HM interac- on an extension of Biot’s theory of consolidation (Biot
tion. These events called for improved HM analysis, es- 1941) to include discrete fractures in addition to the po-
pecially regarding dam stability which, at that time, did rous matrix, and uses a fully implicit solution technique.
not include special treatment of fissured rocks and rock Since then, many computer codes capable of modeling
fractures. As a result, some of the earliest works on cou- coupled HM processes in fractured porous media have
pled HM behavior in fractured rocks were related to dam been developed using various numerical methods, in-
foundations (Londe and Sabarly 1966; Louis and Maini cluding finite-element, distinct-element and boundary-el-
1970). This included application of the parallel plate ement methods (Jing and Hudson 2002). At the same
flow concept to describe fracture flow (Snow 1965; time, more realistic constitutive models have been devel-
Louis and Maini 1970), which is given by oped to describe coupled HM interaction in rock frac-
tures, the ones by Barton et al. (1985) and Walsh (1981)
(13) being the most commonly applied. Efforts have also
been made to incorporate HM coupling into effective
where Qf is volume flow rate per unit plate (or fracture) medium theories (e.g., Oda 1982). As a result of these
width (w), ∇h is head gradient, and b is the fracture aper- developments, direct and indirect HM interactions have
ture or the physical separation between two smooth, par- been integrated into coupled HM numerical analysis of
allel plates. In rock mechanics, the concept of “deform- complex geological media, which includes material het-
able” fractures was introduced, where fractures can re- erogeneity, complex geometry, and material nonlinearity.
spond to changes in effective normal stress and shear During the last 20 years, most of the research and de-
stress in an elastic manner, even if no failure takes place velopment on HM coupling in fractured rocks and most
(Fig. 1e, f). This is formulated according to Goodman applications of HM coupled analysis have been conduct-
(1970) as: ed as part of oil and gas exploration, hot-dry rock geo-
thermal energy investigations, and studies for nuclear
(14) waste disposal. Relatively new applications of HM cou-
(15) pling include deep injection of solid waste and geologi-
cal sequestration of greenhouse gases. The results of
In Eq. (14) ∆un is the normal deformation of the fracture these research efforts and some applications of coupled
caused by a change in effective normal stress, ∆σ′n, with HM analysis will be discussed in more detail below.
the magnitude of opening or closure depending on the
fracture normal stiffness kn (Fig. 1e). Likewise, Eq. (15)
describes the shear displacement, ∆us, which depends on Fundamentals of HM Behavior in Intact Rock
the shear stiffness, ks, and the change in shear stress, ∆σs and Rock Fractures
(Fig. 1f). Since pioneering works by Snow (1965), Louis
and Maini (1970), and Goodman (1970), the coupled Rock masses can be considered to be composed of intact
HM behavior of rock fractures has continued to be a hot (unfractured) rock matrix and rock fractures. Although
topic, both in theoretical studies and practical applica- intact rock has fractures in the form of microcracks,
tions, including the development of more sophisticated which may be similar in behavior to macrofractures, it is
constitutive models. The empirical work on constitutive still useful to study the fundamentals of intact rock and
models by Barton and Bandis (Barton and Choubey rock fractures separately. This section presents funda-
1977; Bandis et al. 1983; Barton et al. 1985) has been es- mental HM behavior of intact rock and rock fractures, as
pecially important for practical applications. has been derived from controlled laboratory and field
Since digital computers become widely available, in tests.
the 1970s, much development of HM coupling in frac-
tured rock has been related to improving numerical
modeling and its application to complex geological sys- Fundamental HM Behavior of Intact Rock
tems. Sandhu and Wilson (1969), Noorishad (1971), The fundamental macroscopic HM behavior of intact
Ghaboussi and Wilson (1973), and Gambolati and Freeze rock can be described in terms of measurable macro-
(1973) were among the first to use numerical modeling scopic quantities for porous media (as a whole). Macro-
for coupled HM analysis. These early finite-element scopic properties for hydromechanical behavior of intact
models treated either linear hydroelastic phenomena in rock include Biot’s α, Biot’s modulus M, and the rela-
porous media with fully implicit coupling (e.g., Sandhu tionship between stress (or strain) and permeability.
and Wilson 1969) or discrete-fractured media using se- Most published HM experiments have been conducted to
quential explicit coupling between the hydraulic and me- determine Biot’s α (e.g., Nur and Byerlee 1971, and
chanical analysis (e.g., Noorishad 1971). For fractured Bernabe 1986) or permeability variation with both con-
and porous media, fully coupled HM numerical models fining pressure and fluid pressure (e.g., Brace et al.
have been available since the early 1980s, when Noori- 1968; Bernabe 1986; Neuzil 1986; Kilmer et al. 1987). A
shad et al. (1982) presented a coupled HM formulation recent review of experiments on intact rock is presented
and finite-element scheme that later evolved into the in this Special Issue by Heiland (2003, this volume).

Hydrogeology Journal (2003) 11:7–40 DOI 10.1007/s10040-002-0241-5


12

stress can explain the relatively insensitive permeability


in these media at high stress.
Quite a few empirical models have been used to
match observed permeability, k, versus confining pres-
sure, P, and fluid pressure, p, in intact rock. Among
them are exponential functions (e.g., Louis et al. 1977),
(18)
power functions (e.g., Kranz et al. 1979),
(19)
and the logarithmic function proposed by Jones and
Ovens (1980):
(20)
In Eqs. (18), (19) and (20), ko is permeability at some
reference effective stress, which is 0 in the case of
Eq. (18), and a1, a2, a3, and a4 are various fitting pa-
Fig. 3 Literature data of permeability versus effective confining rameters to match experimental data.
pressure for intact rock of Pierre shale (average from Neuzil Triaxial tests are frequently used in petroleum engi-
1986), Westerly granite (Brace et al. 1968), and MWX tight sand
gas (Kilmer et al. 1987) neering to simulate reservoir conditions and to record
porosity values versus vertical effective stress (Van-Golf
Racht 1982). A stress–permeability relation obtained
Figure 3 presents literature data of permeability ver- from triaxial tests may be combined with a permeabili-
sus effective confining stress from laboratory tests on ty–porosity relation to obtain the stress–porosity rela-
shale, granite and low-permeability sandstone. Shale, tionship (e.g., Davies and Davies 1999). A permeabili-
which has the lowest permeability, also has the most ty–porosity relationship is provided by the theoretical
stress-sensitive permeability. The tight gas sandstone is Carman-Kozeny relationship (Scheidegger 1974) which,
very stress sensitive at low stress but appears to attain a for the case of packed uniform spheres, can be written as
residual permeability at higher stress.
The differences in the stress–permeability relationship (21)
for different rock types in Fig. 3 can be explained by dif-
ferences in pore shapes. The effect of pore shape can be where dg is the diameter of spheres and τCK is flow tort-
studied, treating pores as elliptical cracks in a linear elas- uosity. However, empirical relationships of the type
tic medium, according to the following equation:
(22)

(16) are widely used to match experimental data where a5


and a6 are fitting constants (Van-Golf Racht 1982).
where bc0 is the crack’s aperture under zero pressure, ν is The macroscopic Biot-Willis’ coefficient, α, can also
Poisson’s ratio, E is Young’s modulus, R is aspect ratio, be expressed in terms of grain and bulk properties:
and P is confining pressure, meaning an isotropic com-
pressive stress (Walsh 1965; Iwano 1995). This equation (23)
implies that cracks will be completely closed when the
pressure reaches where K is the drained bulk modulus of the medium, and
Ks is the bulk modulus of the solid grains. Thus, α≈1 for
(17) a medium in which rock grains have a large bulk modu-
lus (very stiff) compared with bulk modulus for the me-
Equation (17) shows that cracks subjected to a stress P dium as a whole. Similarly, the inverse of Biot’s modu-
will close more easily if they are flat (i.e., if they have a lus, M, can be expressed in micromechanical parameters
small aspect ratio, R). Cracks with small aspect ratios are as
generally found in less-permeable rocks, such as shale
and granite, which explains their relatively sensitive (24)
stress–permeability relationship. Also, in low-permeabil-
ity gas sand, the basic cause of unusually stress-sensitive Equation (24) is valid for an ideal porous medium char-
permeability at low stress has been ascribed to high-as- acterized as a fully connected pore space in a microscop-
pect-ratio sheet pores that are commonly observed be- ically homogenous and isotropic matrix (Detournay and
tween grain boundaries (Brower and Morrow 1985). Iso- Cheng 1993). Equation (24) indicates that 1/M varies
tropic pores in the sandstone that are more resistant to with effective stress, since ϕ and α both vary with stress.

Hydrogeology Journal (2003) 11:7–40 DOI 10.1007/s10040-002-0241-5


13

Wang (2000) lists values of M and other poroelastic con-


stants for various rock types. (28)
Some laboratory experiments have been carried out to
study permeability changes caused by deviatoric stresses where bv is the void aperture defined as the volume ac-
(e.g., Zoback and Byerlee 1975; Wang and Park 2002). In cessible for water per unit area of the fracture, which is
these tests, the axial load on the core sample is increased, equal to the average physical fracture aperture b- over A.
and the axial permeability is measured as a function of
differential stress (axial stress less confining stress). Test
results on granite by Takahashi et al. (1995), Lee and Mechanical normal closure behavior
Chang (1995), and Souley et al. (2001) show that the per- When normal stress is applied on joints, the normal de-
meability first decreases about 1 order of magnitude, until formation is typically nonlinear, as shown in Fig. 4a. The
the axial stress has reached about half of the maximum rate of deformation is greatest at low values of normal
stress (stress at rock failure). This reduction in permeabil- stress, indicating that fracture stiffness increases as nor-
ity results from closure of existing pores and microfrac- mal stress increases. A common feature of fracture de-
tures. At higher stress, that is, values over half of the rock formation is a hysteresis effect during stress loading and
strength, the permeability increases, with increasing axial unloading (not included in Fig. 4a), which is caused by
stress causing the onset of unstable crack growth. The processes arising from surface mismatch, sampling dis-
permeability increase near the peak stress can be dramatic turbances and crushing of asperities (Barton et al. 1985).
and is related to macroscopic failure by the coalescence In Fig. 4a, a size effect on fracture normal closure is in-
of microcracks (Souley et al. 2001). dicated by the experimental results of Yoshinaka et al.
(1993), which showed that the maximum closure, δmax,
increases with sample size. Several empirical models
Fundamental HM Behavior of Rock Fractures have been developed and applied for normal closure be-
In a fracture, fluid flow takes place in the void between havior. The first nonlinear joint model is Goodman’s
two rock surfaces that are irregularly shaped and partly (1974) hyperbolic form which can be written as:
in contact with each other. The HM behavior of rock
fractures has been studied in experiments using various (29)
techniques, such as uniaxial normal compression tests,
triaxial cell tests, and various biaxial compression/shear
tests. Most experiments have been conducted to study where ∆un is joint normal displacement, and kni and σ′ni
HM behavior under normal closure; only a few have are normal stiffness and effective normal stress at an ini-
been conducted to study the effects of shear on perme- tial reference stage (Fig. 4a). However, the most com-
ability. monly applied joint model today is Bandis’ hyperbolic
function (Bandis et al. 1983):

Hydraulic fluid flow (30)


Hydraulic behavior in a rock fracture can be modeled
analogously to that in a confined aquifer using a diffu- where δ is current normal closure, δmax is maximum nor-
sion equation of the form mal closure, and kn0 is normal stiffness at the zero stress
intercept (Fig. 4a). The basic parameters kn0 and δmax can
(25) be estimated from Barton-Bandis basic joint parameters,
joint roughness coefficient (JRC0), and joint compressive
where Tf and Sf are fracture transmissivity and storati- strength (JCS0) using empirical relationships (Barton et
vity, respectively. A hydraulic aperture, bh, can be de- al. 1985).
fined as the parallel plate fracture aperture b in Eq. (13) The second most commonly used model is the logarith-
that produces the same relationship between Q and ∆h. mic type, which has mostly been applied in hot-dry rock
Hence, fracture transmissivity is given by geothermal-reservoir engineering. In one version, the log-
arithmic model can be written as (Evans et al. 1992):
(26)
(31)
The hydraulic or effective aperture bh can therefore be
back-calculated from fracture transmissivity, which can The parameter (dkn/dσ′n)−1, which Evans et al. (1992)
be determined in a flow test. In analogy with Jacob’s denoted “stiffness characteristics”, can be estimated the-
equations for storage in a confined aquifer (Eqs. 3 and oretically from Hertz’s analysis of deformation for two
4), the coefficient of storage (or storativity) of a rock fracture surfaces containing linear elastic hemispheres in
fracture can be estimated as (Rutqvist et al. 1998): contact.
Both of these models are attractive, because their be-
(27) havior can be estimated from basic fracture surface data
and they have been matched to a large number of labo-

Hydrogeology Journal (2003) 11:7–40 DOI 10.1007/s10040-002-0241-5


14
Fig. 4a–d Typical mechanical
and hydromechanical fracture
responses under normal closure
(a, b) and shear (c, d). Effects
of sample size is indicated with
the laboratory sample response
(dashed lines) compared with
in-situ fracture response (1-m2
size)

ratory data in the literature. In general, the hyperbolic portion of the stress/deformation curve corresponding to
Eq. (30) has been shown to match mated fractures better elastic deformation of the fracture, there is minimum di-
whereas the logarithmic Eq. (31) matches unmated frac- lation. The onset of rapid dilation occurs when asperities
tures better. However, as shown by Wei and Hudson begin to slide against each other. Rate of dilation (slope
(1988), Evans et al. (1992), and Zhao and Brown of ∆un curve in Fig. 4c) increases and reaches a maxi-
(1992), the logarithmic function can also match labora- mum at the peak shear stress (Barton et al. 1985).
tory data of mated fractures over an engineering stress The figure also illustrates the effect of scale. For a
range (i.e., less than about 10 MPa). At very high stress, larger sample, the peak shear stress is smaller and takes
on the other hand, the logarithmic model cannot be used place after a larger shear displacement magnitude. As a
to produce a residual aperture, since the fracture will consequence shear stiffness, representing the first steep
close completely and ultimately produce a negative ap- elastic part, would be smaller with increased sample
erture. size. Furthermore, the onset of shear dilation is delayed
in larger samples because a larger displacement is re-
quired to reach peak shear stress.
Mechanical shear behavior The peak shear strength depends on normal stress,
Mechanics of fracture shear behavior have been devel- with a higher peak stress for a higher normal stress. This
oped in the field of rock mechanics since the 1960s, pio- was examined by Byerlee (1978) for normal stresses up
neered by Patton (1966) who predicted the existence of to 100 MPa, who showed the relationship
scale effects. The influence of scale effects was quanti-
fied by Barton and Choubey (1977), and conclusive ex- (32)
periments were presented by Bandis et al. (1983). Fig-
ure 4c illustrates a typical shear stress displacement of a Barton and Choubey (1977) studied shear behavior at en-
clean, rough, dilatant fracture under constant normal gineering stress levels and developed the following rela-
stress. It is characterized by rapid increase in shear stress tionship:
up to a peak, followed by a loss in load-carrying capaci-
ty. The shear displacement is accompanied with a shear (33)
dilation, as shown in the lower curves of Fig. 4c. For a

Hydrogeology Journal (2003) 11:7–40 DOI 10.1007/s10040-002-0241-5


15

where σscmob is the mobilized shear strength, JRCmob is


the mobilized joint roughness coefficient, JCS is the
joint compressive strength, and Φr is the residual friction
angle. The term “mobilized” denotes current values of
σsc and JRC at any current shear displacement, which
can be before, at or after peak-shear stress (Barton et al.
1985). The dilation curve (∆un versus ∆us) can be calcu-
lated by the following expression:
(34)

(35)

where dmob is the mobilized dilation angle and m is a


damage coefficient, given values of 1 or 2 for shearing
under low or high normal stress, respectively (Olsson
and Barton 2001). Equations (33), (34) and (35), and the
size dependency of JRC are the basic empirical equa-
tions of the Barton-Bandis joint model for fracture shear.

Hydromechanical coupling under normal closure


Londe and Sabarly (1966) were perhaps the first to per-
form HM experiments on rock fractures during their in- Fig. 5 Fracture hydraulic conductivity versus normal stress for
vestigation of the Malpasset Dam failure in the early various sample sizes of artificial and natural tension fractures
1960s (described below). Experimental results typically (from Witherspoon et at. 1979)
show a decrease in fracture transmissivity with normal
stress (Fig. 4b), but with an apparent residual transmissi-
vity, Tr, at high stress when the fracture appears to be tribution, implying that fewer of the largest, least-fre-
mechanically compressed. The residual transmissitivy quent flow channels would be included in a smaller sam-
indicates that the fluid flow at high stress may be domi- ple. A conflicting observation regarding the size effect
nated by tube-like flow channels, which have a low as- reported by Raven and Gale (1985) may have resulted
pect ratio and therefore cannot be closed easily (see from unrepresentative sampling of the fracture surface
Eq. 16). In a few of the experiments (e.g., Kranz et al. asperities (Raven and Gale 1985).
1979, and Iwano 1995) values of Biot’s α for fracture Witherspoon et al. (1980) developed a modified cubic
flow have been determined to be less than one and to law which they validated against laboratory experiments
vary with effective normal stress, with a higher α value on artificial tension fractures in samples of granite and
at low stress. marble. They considered a general flow law
Figure 4a, c shows a size effect for normal closure.
Witherspoon et al. (1979) suggested a size effect on the (36)
experimentally determined hydraulic properties of rock
fractures. This suggestion was based on data from exper-
iments carried out on an ultra-large core (0.95 m in di- where f is a friction factor that accounts for the rough-
ameter), an in-situ block test (1 m2) reported by Pratt et ness of the fracture surface, b is an apparent physical ap-
al. (1977), and smaller laboratory samples (0.15 m in di- erture, and C is a constant depending on the flow domain
ameter) reported by Iwai (1976). Their results, presented geometry and the properties of the fluid (e.g.,
in Fig. 5, showed that, at the maximum stress level that C=ρfgw/12µf for parallel flow (see Eq. 13)). If n=3, this
could be attained (10–20 MPa), the minimum values of is a cubic law or modified cubic law, and in such case
fracture hydraulic conductivity were not the same for the apparent physical aperture is related to the hydraulic
each rock specimen but increased with specimen size. aperture as:
Barton and Bakhtar (1982) noted similar behavior in an (37)
in-situ block experiment, finding that the effective aper-
ture could not be decreased to less than 30 µm at the Witherpoon et al. (1980) interpreted the apparent physi-
maximum normal stress of 7 MPa. The effect of size on cal aperture as the residual aperture, br, plus an apparent
stress–permeability coupling of fractures was confirmed mechanical opening, according to
in theoretical studies by Neuzil and Tracy (1981) and (38)
Swan (1983), who predicted increasing permeability
with sample size. Neuzil and Tracy (1981) attributed this Using a slightly different approach, Elliot et al. (1985)
size effect to a truncation of the aperture frequency dis- parameterized the apparent physical aperture as

Hydrogeology Journal (2003) 11:7–40 DOI 10.1007/s10040-002-0241-5


16

(39) monly holds for a wide variety of fractures. A significant


number of cases where the cubic law does not hold in-
where b0 is the apparent physical aperture when δ is ze- volve cases in which the hydraulic aperture appears to be
ro. A third approach used by Rutqvist (1995a) parame- less sensitive to closure than predicted by the cubic law.
terized the apparent physical aperture as These cases can typically be fitted with a modified cubic
(40) law (i.e., Eq. 36 with n=3 and f≠1.0). In many cases, the
modified cubic law breaks down as closure is increased,
where bi and δi are the apparent physical aperture and with these cases typically exhibiting a progressive de-
normal closure at initial effective stress. In the latter case crease in the sensitivity of hydraulic aperture to further
the current hydraulic aperture can be calculated conve- closure. Boinott (1991) concluded that this breakdown
niently as frequently occurs when the hydraulic aperture is less
than about 50 µm. During the 1990s, these observations
(41)
have been confirmed in several studies (e.g., Zhao and
Brown 1992; Iwano 1995). This shows that residual ap-
where f′=1/f1/3, ∆un is the normal deformation of the erture is an important parameter for the coupled stress-
fracture from initial conditions (Fig. 4a), and bhi is the flow behavior of rock fractures.
hydraulic aperture at initial effective stress. A complete relationship between fracture transmissi-
Barton et al. (1985) criticized the use of an apparent vity and effective normal stress can be derived by com-
physical aperture and proposed an empirical relationship: bining a fracture flow law with an equation for fracture
normal closure. For example, Rutqvist (1995a) combines
(42) Goodman’s (1974) model in Eq. (29) with Eqs. (26) and
(41) to derive:
where bE is the “real” physical aperture (also called “me-
chanical aperture” and usually having the symbol E in (43)
the literature). Here the “real” physical aperture corre-
sponds to the arithmetic mean of the separation of the where κni was denoted as the initial “hydraulic normal
two fracture surfaces, which can be measured directly stiffness”, since it reflects how hydraulic aperture chang-
with a feeler gauge. Equation (42) implies a nonlinear es with stress. The hydraulic stiffness is related to me-
relationship between hydraulic and mechanical aper- chanical stiffness by
tures, and hence a nonlinear relationship between ∆bh
and ∆un. This is different from the approaches by With- (44)
erspoon et al. (1980), Elliot et al. (1985), and Rutqvist
(1995a), which all have a linear relationship between ∆bh
and ∆un (or ∆δ). Intuitively, Barton’s Eq. (42) is more which shows that the hydraulic stiffness is usually equal
correct. However, as demonstrated by Wei and Hudson to or less than the mechanical stiffness. Similarly, Alm
(1988) and Zhao and Brown (1992), over an engineering (1999) derived a relationship between transmissivity and
stress range of σ′n<10 MPa, the linear relationship seems normal stress by combining Eqs. (26) and (31), leading
to match experimental data as well as Barton’s equation. to
There has been conflicting laboratory evidence re-
garding the validity of the cubic law and the effects of (45)
contact area for fluid flow in rock fractures. As men-
tioned above, Iwai (1976) and Witherspoon et al. (1980) In addition to the above models, there are a number of
validated the cubic law against laboratory experiments empirical and theoretical models to describe the relation-
on tension fractures in granite and marble. They found ship between normal stress across a fracture and its
that the cube of the aperture is proportional to the flow if transmissivity. Models for macroscopic hydromechanical
a correction is made for a residual aperture at high stress. behavior of fractures have been developed from theories
By contrast, a number of studies during the 1980s and of microscopic behavior of rough fractures. These in-
1990s have shown a dramatic deviation from the cubic clude the beds-of-nail model by Gangi (1978), the aper-
law (for example, Engelder and Scholz 1981, Raven and ture frequency model by Neuzil and Tracy (1981), the
Gale 1985, and Pyrak-Nolte et al. 1987). quasi two-dimensional aperture-void model by Tsang
Boitnott (1991) carefully reviewed the conflicting ev- and Witherspoon (1981), the Hertzian contact models by
idence on the cubic law. He pointed out that much of the Walsh (1981) and Swan (1983), and stochastic models,
contradictory observations result from different data sets e.g., Brown and Scholz (1986) and Brown (1987).
having been analyzed differently; the term “cubic law” Walsh’s (1981) model, which is the most well known,
does not have a consistent definition in the published lit- can be written as:
erature. He used the general formulation of the flow law
in Eq. (36) and re-evaluated the conflicting data from 16
references of published data. He found that the predic- (46)
tions by the cubic law (i.e., n=3 and f=1 in Eq. 36) com-

Hydrogeology Journal (2003) 11:7–40 DOI 10.1007/s10040-002-0241-5


17

where T0 and b0 are joint transmissivity, and aperture at in hydraulic aperture. This has been attributed to asperity
some reference effective stress, σ′no, he is the standard damage under increasing shear deformation with accom-
deviation of the asperity height distribution and τw is a panying gouge production. As a result, Olsson and
tortuosity factor that depends on the normal stress. This Barton (2001) found that Eq. (42), relating hydraulic and
equation is derived using Hertzian theory, assuming that real physical apertures, breaks down after peak shear dis-
the fracture consists of two uncorrelated rough surfaces, placement. They presented a new tentative relationship
each with random topography of exponential distribu- expressed as
tion. Walsh (1981) studied experimental data of Kranz et
(48)
al. (1979) and found that the tortuosity factor can be ne-
glected for fluid flow, since aperture is raised to the third which is valid after 75% of the shear displacement at
power and overwhelms the tortuosity term, which is only peak stress. This relationship matched their laboratory
raised to the first power. If the tortuosity factor is ne- data quite well.
glected, and b0 is assumed equal to bh0, then Eq. (46) re-
duces to Eq. (45), which is also equivalent to the empiri-
cal model developed in petroleum engineering by Jones In-Situ HM Conditions in Fractured Rock Masses
(1975). The equivalence of Eqs. (45) and (46) in this
case becomes obvious if (bh0)3 is broken out of the This section describes the in-situ HM conditions found
square bracket in Eq. (45) with the resulting (bh0)3 C=T0, in fractured rock masses. They are related to the current
and (dkn/dσ′n)−1=√2 he distribution of in-situ stress and permeability in the
Equivalent logarithmic models have also been applied Earth’s crust. In turn, the current in-situ stress and per-
by Zhao and Brown (1992), and Evans et al. (1992) for meability distributions are results of past and ongoing
hydromechanical normal closure behavior, and by geological processes. However, in an engineering time
Bandis et al. (1983) for mechanical normal closure be- frame they may be taken as initial quasi-static condi-
havior of unmated rock joints. Therefore, it can be said tions. A correlation between in-situ stress and permeabil-
that the theoretical logarithmic normal closure model by ity can indicate a coupling between the modern stress
Walsh (1981) has been widely applied and that a loga- field and permeability. For example, a decreasing perme-
rithmic function seems to fit laboratory data, at least at ability with depth, or a correlation between anisotropy of
engineering stress levels. principal stress and permeability fields may indicate that
permeability is strongly dependent on stress. If, on the
other hand, permeability is correlated to maximum shear
Hydromechanical behavior during shear stress direction, it will indicate that fractures are prone to
Laboratory measurements of hydromechanical behavior further shear slip following any human-induced distur-
during shear have been rare because of a lack of special- bance (for example, during fluid injection or under-
ized test equipment. The first comprehensive, experi- ground excavation).
mental study of permeability changes caused by shear
was conducted by Makurat et al. (1990), who developed
a coupled shear-flow test apparatus that could apply bi- Conceptual Model of Crystalline Fractured Rock
axial stress. They concluded that whether the conductivi- Fractured rock masses consist of discontinuities of vari-
ty increases or decreases with shear depends on both the ous scales, from decimeter-size single joints to kilome-
joint and rock properties, as well as the exact nature of ter-size major fault zones. Therefore, mechanical and
the stress applied. Makurat et al. (1990) determined that hydraulic properties can be, and typically are strongly
decreases in hydraulic conductivity during shearing were heterogeneous. As an example of a fractured rock mass,
a result of gouge production, which tended to block flow Fig. 6 shows a conceptual model for the U.S. Geological
paths. The experiments were modeled using a Barton- Survey’s fractured rock research site near Mirror Lake,
Bandis joint model, which predicted an increasing aper- New Hampshire. The investigations at Mirror Lake sug-
ture with shear dilation, according to gest that the bedrock contains a small number of highly
bE = bEi + ∆un (47) transmissive fractures within a larger network of less-
transmissive fractures. The highly transmissive fractures
where ∆un is the dilation of the fracture obtained from appear to form local clusters. Each fracture cluster occu-
Eq. (34), and bEi is the initial aperture before shear. This pies a near-horizontal, tabular-shaped volume which is
equation overpredicted increases in fracture permeability several meters thick and 10–40 m across (National
during shear because it does not correct for the formation Research Council 1996). The domination of a few
of gouge material in the fractures. connected fractures has been observed at many frac-
Recent tests on granite reported by Esaki et al. (1999) tured crystalline rock sites, for example, Stripa (Olsson
and Olsson and Barton (2001) show that the joint trans- 1992), Forsmark (Carlsson and Olsson 1977), Laxemar
missivity starts to increase after about 1 mm of shear and (Rutqvist et al. 1997), and Äspö in Sweden (Rhén et al.
then increases rapidly by about 1–2 orders of magnitude 1997). A clustered or compartmentalized system of
up to 5 mm shear. These tests also indicated that the me- highly conductive fracture networks has been reported
chanical dilation of the fracture is greater than changes from Kamaishi Mine, Japan (Doe 1999) and Sellafield,

Hydrogeology Journal (2003) 11:7–40 DOI 10.1007/s10040-002-0241-5


18
Fig. 6 Vertical cross section
and conceptual model of the
US Geological Survey’s frac-
tured rock research site near
Mirror Lake, New Hampshire.
Four clusters of highly perme-
able fractures labeled A–D oc-
cur in the less-permeable frac-
tured rocks. Borehole packers
are shown as closed sections
(extracted from National Re-
search Council 1996)

Fig. 8 Conceptual three-dimensional view of a highly water-con-


ductive minor shear fault located at the Äspö Hard Rock Laborato-
ry, Sweden (Hakami 1995). The aperture variation is simplified in
the model into open areas (b>0.1 mm) and contact areas (closed;
b≤0.1 mm). The elliptically shaped contact areas reflect the anisot-
ropy in correlation distances obtained from aperture measurements
Fig. 7 Inflow features in the drift system at the storage site for with d1 and d2 typically 8 and 16 cm, respectively
low- and intermediate-level nuclear waste, Forsmark, Sweden (ex-
tracted from Carlsson and Olsson 1977). The depicted inflow fea-
tures can be described as (1) outflow in open channels in a major Figure 8 provides a three-dimensional view of a high-
fracture, (2) evenly distributed flow along subhorizontal tension
joints, (3) flow focused to the termination of a steep fracture at a ly permeable fracture. Hakami (1995) conducted exten-
subhorizontal fracture, (4) enhanced flow at fracture intersections, sive fracture aperture measurements of a minor fault in-
and (5) diffuse flow through unfractured rock tersecting a drift at Äspö Hard Rock Laboratory, Swe-
den. The fault belongs to a group of steep, dipping frac-
tures zones, typically 10–30 m long, of a type found to
U.K. (Sutton et al. 1996). At the Forsmark site for low- be the most conductive in the area. It is located at a
and intermediate radioactive nuclear waste disposal in depth of about 200 m where the minimum principal
Sweden, a detailed observation of inflows was classified stress is about 5 MPa. The fault has been reactivated sev-
into five main types, as shown in Fig. 7. Carlsson and eral times and contains fracture filling materials such as
Olsson (1977) found that most water inflow was related calcite, chlorite, epidote, and quartz. Measurements
to features labeled 1 and 4, which are associated with showed that the average aperture of the fracture (outside
fractures opened by large-scale fracture surface undula- contact zones) was about 2 mm, with the coefficient of
tion and fracture intersections. variation 130%. The contact area, defined as the area

Hydrogeology Journal (2003) 11:7–40 DOI 10.1007/s10040-002-0241-5


19

with aperture less than 0.1 mm, covers about 40% of the
total area of the fracture. The conceptual model in Fig. 8
shows that the fault is undulating and “locked open” by a
shear slip of about 4 cm. This appears to belong to flow
feature 1 in Fig. 7.

In-situ HM Behavior from Injection Tests


In-situ HM properties can be inferred from a so-called
hydraulic jacking test (also called a step-rate test) con-
ducted in single boreholes. The technique was first ap-
plied by Londe and Sabarly (1966) and Louis et al.
(1977) to study pressure-sensitive permeability under
dam foundations. Rutqvist (1995b) and Rutqvist et al.
(1997) used hydraulic jacking tests, combined with cou-
pled numerical modeling, for determining the in-situ HM
properties of fractures in crystalline rocks. Hydraulic
jacking tests were conducted by a step-wise increase of
the fluid pressure. At each step, the well pressure was
kept constant for a few minutes until the flow was steady
(Fig. 9a). The numerical analysis of these injection tests
shows that the flow rate at each pressure step is strongly
dependent on the aperture and normal stiffness of the
fracture in the vicinity of the borehole, where the pres-
sure changes the most (Fig. 9b). Figure 9c shows field
test results of a hydraulic jacking test on a fracture at
267-m depth, in a crystalline fractured rock at Laxemar
in southeastern Sweden (Rutqvist et al. 1997). At the
first cycle of step-wise increasing pressures, the flow
rate increases as a nonlinear function of pressure. A tem-
poral peak pressure is obtained at a flow rate of
1.3 l min−1 before the pressure begins to decrease with
an increasing flow rate. A shear-slip analysis of the par-
ticular fracture, which was inclined to the principal in-
situ stresses, indicated that these irreversible fracture re-
sponses could be caused by a shear slip, as the fluid
pressure reduced the shear strength of the fracture. The
subsequent step-pressure cycle took a different path be-
cause of the change in hydromechanical properties re-
sulting from shearing and fracturing.
The overall results from the hydraulic jacking tests
conducted at Laxemar showed that the pressure sensitivi-
ty of the fractures are strongly dependent on the initial
hydraulic permeability. The permeability of the most-
conductive fractures is relatively insensitive to injection
pressure, whereas the permeability of the least-conduc- Fig. 9a–c In-situ determination of normal stress versus trans-
tive fractures can be strongly dependent on the injection missivity relationship using a combination of pulse, constant head
pressure. From the borehole-televiewer image, the most and hydraulic jacking tests. a Schematic representation of pressure
conductive fractures appear to be open fractures that are and flow versus time, b the radius of influence in each test, and
c results of hydraulic jacking test (Rutqvist et al. 1997)
incompletely cemented, indicating flow channels in a
fracture that are “locked open” by shear dislocation or
mineral filling (Rutqvist et al. 1997). ite. The fracture has a dip of about 45° relative to the hor-
There are a few examples where hydraulic jacking izontal plane, and the normal stress across the fracture
tests have been conducted with simultaneous mechanical was estimated to be 4.4 MPa. Figure 10 presents hydrau-
measurements of fracture deformations (Gale 1975; Jung lic and physical apertures determined at one intersecting
1989; Martin et al. 1990; Myer 1991). Jung (1989) mea- boreholes as a function of fluid pressure in the fracture.
sured fracture hydraulic and mechanical widths during The figure shows that, at the initial hydrostatic fluid pres-
hydraulic injection into a large hydraulic fracture (over sure (corresponding to an effective normal stress of
100 m in diameter) located at about 250-m depth in gran- 1.9 MPa), the hydraulic conducting aperture is about

Hydrogeology Journal (2003) 11:7–40 DOI 10.1007/s10040-002-0241-5


20

Fig. 10 In-situ measurements of hydraulic aperture and apparent


physical aperture during hydraulic jacking tests into an artificial
fracture in granite at a depth of about 250 m (data extracted from
Jung 1989)

Fig. 11 Permeability measured in short-interval well tests in frac-


0.5 mm whereas the apparent physical aperture is about tured crystalline rocks at Gideå, Sweden (data points from Wladis
1 mm. At fracture extension pressure (corresponding to et al. 1997). Effects of shear dislocation and mineral precipita-
0.3 MPa above the effective normal stress), both the hy- tion/dissolution processes obscure the dependency of permeability
on depth (stress). The permeability values on the left-hand side re-
draulic and physical apertures are estimated to be around present intact rock granite, or flow feature 5 in Fig. 7, whereas the
2.0 mm (Fig. 10). The very large residual aperture indi- permeability values on the right-hand side represent highly con-
cated in Fig. 10 seems to agree with a general size effect ductive fractures, possibly flow feature 1 in Fig. 7
discussed above, and may also have been caused by a
fracture shear slip during the hydraulic injection.
if fractures are intersecting, they are either completely
cemented with minerals or isolated from a conducting
Correlation Between Permeability and Depth fracture network. The maximum permeability values on
(Stress) the right-hand side of Fig. 11 represent the transmissivity
Figure 11 presents a profile of short-interval (3-m packer of at least one intersecting fracture, which is highly con-
separation) well tests performed at a typical granitic rock ductive and connected to a larger network of conducting
for the site investigation of the Swedish radioactive fractures. As discussed above, these highly conductive
waste disposal program, located at Gideå, Sweden. The fractures are likely “locked open” either by bridges of
figure may show a general decrease in permeability with hard mineral filling or by large shear dislocation. The
depth, but the depth dependency is obscured by a very fluid flow in these “locked-open” fractures takes place in
large permeability spatial variation (up to 6 orders of channels between bridges, and is likely to be relatively
magnitude). A general observation from Fig. 11, and insensitive to stress. The phenomenon of “locked-open”
from many fractured crystalline rock sites, is that the fractures is well known in hydrocarbon reservoirs, espe-
most pronounced depth dependency can be found in the cially in carbonate rocks where localized dissolution of
upper 100–300 m of the bedrock. It has been suggested the wall rock can create highly conductive pathways of
that this depends on near-surface fracturing owing to several millimeters, even at high in-situ stress (Dyke
stress relief caused by erosion and isostatic rebound after 1995). In crystalline rocks, the “locked-open” fracture
glaciation, enhanced by weathering and solution of frac- category could occur in mineral-filled or shear-dislocat-
ture minerals (Wladis et al. 1997). However, the more ed fractures throughout the rock mass. However, large
pronounced depth dependency in shallow areas can also channels are more likely to occur in fault zones where
be explained by the nonlinear normal stress–aperture re- large movements allow substantial shear dislocation and
lationship of single extension joints (Fig. 4b). Such a de- extension of oblique fractures (National Research Coun-
crease in hydraulic aperture with depth has been ob- cil 1996). It has also been observed that highly perme-
served by Snow (1968), who used detailed pumping tests able fracture zones are relatively insensitive to stress
and fracture statistics to determine fracture apertures at compared to competent fractured host-rock media
numerous dam foundations. His results showed that the (Carlsson and Olsson 1986; Rutqvist et al. 1997).
hydraulic aperture decreases from about 200 µm at the
ground surface to about 50 µm at 60-m depth.
The permeability in Fig. 11 varies from 5×10−19 to Correlation Between Anisotropy of In-situ Stress
4×10−13 m2. The minimum values on the left-hand side and Permeability
of Fig. 11 represent the permeability of the rock matrix, An in-situ stress-dependent permeability has been in-
indicating that no fracture intersected the test interval or, ferred from observed in-situ stress and permeability anis-

Hydrogeology Journal (2003) 11:7–40 DOI 10.1007/s10040-002-0241-5


21

otropies. One example is a study at an underground hy- dependency of permeability appeared only for the upper
dropower plant in Juctan, Sweden (Carlsson and Olsson 100–300 m, whereas at greater depth fractures are closed
1979). The study, which was conducted at about 250-m to a residual permeability value. This residual permeabil-
depth in a mountain of granitic rock, showed that the an- ity value is very high for “locked-open” fractures, and
isotropic permeability of the fractured rock was correlat- their stiffness is high because of hard mineral bridges.
ed with the anisotropic in-situ stress field. An anisotropy Hence, at these depths, bulk permeability cannot be ex-
factor of 3 (kH=1.9×10-16 m2 and kh=0.6×10-16 m2) was pected to be especially stress sensitive.
obtained for the permeability in a stress field that
showed a local anisotropy factor of 4.3 (σH=15 MPa and
σh=3.5 MPa). One reason to believe that this anisotropy Correlation Between Maximum Shear Stress
of permeability was in fact stress induced is that the and Permeability
measured in-situ stress field (which probably was affect- In the last decade, investigations of the active lithospher-
ed by the nearby excavation) was extremely anisotropic. ic-plate boundary in California have shown that fractures
Most importantly, the minimum horizontal stress was favorably oriented for shear slip, so-called critically
very low, which implies that some fractures are in a low stressed fractures, tend to be active groundwater flow
normal stress (low normal stiffness) regime, in which paths (Barton et al. 1995, 1998; Ferrill et al. 1999). The
permeability can be highly stress dependent. rationale for bulk permeability being dominated by criti-
Another example is provided by a study of a fractured cally stressed fractures is that most fractures in the bed-
oil reservoir in the Ekofisk Field in the North Sea (Teufel rock are cemented because of water/rock chemical reac-
and Farrel 1995). The top of the reservoir is at a depth of tions. If shear slip occurs on a critically stressed fracture,
2.9 km, and reservoir thickness is up to 305 m. The res- it can raise the permeability of the fracture through sev-
ervoir consists of two fractured chalk intervals separated eral mechanisms, including brecciation, surface rough-
by a relatively impermeable layer of unfractured chalk ness, and breakdown of seals (Barton et al. 1995). Late-
(see also Lewis et al. 2003, this volume). Fractures with ly, similar correlations have been found at the KTB Sci-
two distinct orientations were logged from boreholes, entific Drill Hole in Germany down to 7 km (Ito and
with one set far more abundant than the other. Hydraulic Zoback 2000), and at Äspö, Sweden, in Precambrian
testing showed that the direction of maximum principal rocks of the Baltic Shield deep in the Eurasian plate
permeability was aligned with the less-abundant fracture (Talbot and Sirat 2001). On the other hand, experience
set, and that its direction was parallel to the maxi- from injection experiments at the hot-dry rock geother-
mum compressive stress. An anistropy factor of 4.4 mal sites in Soultz, France, and Rosemanowes, U.K.,
(kH=1.59×10−13 and kh=3.6×10−14 m2) in permeability suggests that a pore-pressure increase of 5–6 MPa over
was obtained for a stress field with an anisotropy factor ambient is needed to stimulate significant microseismi-
of about 1.8 (calculated using Eq. 1, with σH=40 MPa, city (Evans et al. 1999). This indicates that under undis-
σh=33 MPa, and p=24 MPa) for effective stresses. In this turbed stress and pressure conditions, the fractures
case the minimum compressible stress is about 9 MPa, would not be at the verge of shear failure.
which implies that fractures should be relatively com-
pressed close to a residual aperture. Thus, it may be
questioned whether the permeability anisotropy observed Observations of Human-induced HM Interactions
by Teufel and Farrel (1995) really results from a present- in the Earth Crust
day stress field. Furthermore, it appears that the perme-
ability has remained constant at Ekofisk despite a 22- This section introduces field observations of HM interac-
MPa reduction in pore pressure over 18 years of produc- tions in fractured rock masses triggered by human activi-
tion (Dyke 1995), which indicates that the permeability ties. All these observations (e.g., land subsidence due to
at reservoir depth is not particularly stress dependent. fluid extraction, injection-induced shear failure and se-
Stowel et al. (2001) questioned whether modern-day ismicity, and excavation-induced squeezing) are evi-
stress controls the orientation of the most-conductive dence of direct HM couplings accompanied by indirect
fractures in deep reservoirs. They reviewed stress, frac- couplings. Thus, the coupled responses are triggered by
ture, and production data from low- to moderate-matrix- changes in stress or fluid pressure and are accompanied
permeability reservoirs at three sites in the western Unit- by changes in material properties (see Fig. 2).
ed States. Their comparison of stress directions and ori-
entation of flow-controlling fractures showed that open
fractures in the subsurface are not necessarily parallel to Land Subsidence due to Fluid Extraction
the maximum principal compressive stress, and that frac- Evidence of direct fluid-to-solid coupling is perhaps
tures parallel to the maximum principal stress are numer- most obvious in land subsidence caused by underground
ous. They referred to partially cemented “locked-open” fluid removal. In this case, land subsidence is caused by
fractures as the most important flow features at depth. a reduction in fluid pressure that increases the effective
The findings by Stowel et al. (2001) are not surprising, stress, which in turn results in a contraction of the geo-
especially since their data came from depths of logical media. However, the land subsidence process can
2,400–6,400 m. As noted in the section above, the depth be quite complex, and the amount of subsidence during

Hydrogeology Journal (2003) 11:7–40 DOI 10.1007/s10040-002-0241-5


22

Fig. 12a, b Schematics of land subsidence at the Wilmington Oil Fig. 13 Stress change in reservoir caused by a decrease in reser-
Field in California. a Horizontal strain causing horizontal faulting, voir pore pressure. The minimum horizontal stress was determined
and b vertical and horizontal displacement (from Segall 1989) by hydraulic fracturing (from Segall 1989)

fluid removal depends on a number of factors (such as


size and depth of the exploited formation, magnitude of
the pressure decline, compressibility, and permeability of
the rocks). One of the most dramatic examples of land
subsidence has occurred in Mexico City, caused by
groundwater withdrawal and associated aquifer compac-
tion. The maximum rate of subsidence approaches 60 cm
per year, with total subsidence during the 20th century as
great as 10 m (Galloway et al. 1999). In the United
States, more than 80% of the identified subsidence re- Fig. 14 Schematics of faulting associated with underground ex-
sults from exploitation of underground water resources, traction of oil and gas (Segall 1989)
with the agriculture-intensive San Joaquin Valley, Cali-
fornia, being the largest (Galloway et al. 1999). As of the
last geodetic survey (1970), subsidence in excess of one mal energy extraction (Segall and Fitzgerald 1998). Ex-
foot (0.3 m) affected more than 5,200 acres (21 km2), traction-induced earthquakes are triggered by poroelastic
with the maximum subsidence more than 9 m. Although stresses associated with reservoir contraction. Declining
rare, dramatic land subsidence has also been observed pore pressure causes reservoir rocks to contract slightly
during oil and gas exploration. One well-known example with associated reduction of the horizontal stress. Fig-
is the Wilmington Oil Field in California, which experi- ure 13 presents field data showing such fluid pres-
enced a subsidence of almost 9 m in the period sure–stress coupling in the Ekofisk Field and for the
1926–1967 (Fig. 12). The amount of subsidence at the McAllen range, Texas Field. Because the reservoir is elas-
Wilmington Oil Field was substantial because of a sig- tically coupled to the surrounding rocks, the contraction of
nificant reduction of the reservoir pressure over a large the reservoir and the reduction of stress in the reservoir is
vertical interval. More recently, oil and gas extractions at balanced by increased stress in the neighboring crust. This
the Ekofisk Field in the North Sea have caused a subsi- stress results in subsidence, horizontal contraction above
dence of the sea bottom of more than 10 m since the the reservoir and, in some instances, triggered seismicity
1970s, and have necessitated to repeated, large-volume (Segall 1989). Examples of faulting associated with fluid
seawater injections to jack-up the platforms. Lewis et al. withdrawal are illustrated schematically in Fig. 14. One of
(2003, this volume) presents the subsidence of the Eko- the most obvious examples of seismicity caused by fluid
fisk Field in more detail. extraction is the Wilmington Oil Field where spectacular
surface deformation was accompanied by seismicity. Six
shallow earthquakes (M 2.4–3.3) occurred within the oil-
Seismicity and Failure Associated with Deep Well field between 1947 and 1955, when the rate of subsidence
Fluid Extraction/Injection was greatest (Segall 1989). The earthquakes were generat-
Seismicity caused by fluid extraction has been observed ed by slip-on bedding planes at depths of 470–520 m,
during oil production at numerous oilfields since the which sheared off tens of oil wells over several square
1920s (Segall 1989) and, more recently, during geother- kilometers (Segall 1989). Although the earthquakes at

Hydrogeology Journal (2003) 11:7–40 DOI 10.1007/s10040-002-0241-5


23
Fig. 15a, b Injection flow rate,
injection pressure and seismic
activities during fluid injection
at Hijiori hot-dry rock site in
Japan. a A 1-day high-pressure
hydraulic fracturing operation,
and b a 1-month low-pressure
circulation experiment (modi-
fied from Sasaki 1998)

Wilmington Oil Field caused equipment damage, their oped in Fenton Hill, New Mexico, in the late 1970s. Re-
magnitude was moderate. However, oil production has sults from the monitoring of hydraulic fracturing experi-
also been attributed to more substantial seismic activity ments have been reported by Aki et al. (1982) from 3.5-
around the world (Cypser and Davis 1998). km-deep, large-volume injection at Fenton Hill (U.S.),
Seismicity caused by fluid injection has been ob- by Pine and Bachelor (1984) from 2.5-km-deep, large-
served at sites used for deep injection of hazardous volume injection done in Rosemanoes (U.K.), by Jupe
waste, during reservoir stimulation, and during forced et al. (1992) from 0.5-km-deep injection at Fjällbacka
circulation in geothermal fields. A classic example of in- (Sweden), and by Gaucher et al. (1998) from
jection-induced seismicity occurred at a deep (3,638 m) Soultz-sous-Forêts (France), and several other sites
well for liquid waste injection at the Rocky Mountain worldwide.
Arsenal near Denver, Colorado (Hsieh and Bredehoeft An illustrative example is presented in Figs. 15 and
1981). During the first months of operation in 1962, over 16 from the Hijiori hot-dry rock site in Japan. Figure 15
15,000 m3 were injected, the wellhead pressure ranging shows flow, pressure, and seismic activities associated
from 0 to 7.2 MPa. In March 1962, Denver experienced with two different experiments: a one-day, high-pressure
its first earthquake in 80 years. The injection continued, hydraulic-fracturing experiment performed in 1988, and
off and on, until Evans (1966) publicly suggested a cor- a one-month low-pressure circulation experiment per-
relation between fluid injection and seismic activity. The formed in 1989. In both, the seismic rate (Fig. 15) ap-
correlation was observed through seismic monitoring, pears to be strongly dependent on wellhead pressure.
which showed that seismic activity generally increased However, the spatial distribution of the microseismic
during injection periods and declined during periods of events in the two experiments is quite different (Fig. 16).
no injection (Hsieh and Bredehoeft 1981). However, the Microseismic events induced by the 1988 hydraulic frac-
three largest earthquakes occurred at the site more than turing migrated towards the east and distributed along a
one year after the injection was discontinued, in 1967, vertical plane with a strike nearly parallel to the direction
with magnitudes of M 4.8 and 5.3 (Cypser and Davis of the maximum principal stress. This indicated that pre-
1998). The continuous seismicity years after injection existing fractures were reopening along a zone that de-
stopped was explained by postinjection propagation of veloped in the direction of maximum principal stress, al-
built-up injection pressure throughout the reservoir, though microseismic events were caused by shear slip.
which could reach and release strain from tectonically Seismicity accompanying the 1989 circulation test with a
stressed faults (Healy et al. 1968). low injection pressure was diffuse, forming a seismic
Since the Denver Basin earthquakes, many injection- “cloud”. It was concluded that the seismic cloud accom-
induced quakes have been documented (Nicholson and panying the circulation test resulted from the permeation
Wesson 1992). However, of the thousands of injection of water into joints, which slipped when the effective
wells in the world, only a few have induced notable stress was reduced by the increased fluid pressure.
earthquakes. Such notable examples are an experimental
well in the Chevron Oil Field near Rangely, Colorado,
and sites at El Dorado, Arkansas. Seismicity and Failure Associated with Reservoir
Intentional hydraulic stimulation and microseismic Impoundment
monitoring have also been applied at fractured geother- Substantial seismicity and large-scale failure of rock
mal (hot-dry rock) sites since the first site was devel- masses have on several occasions been linked to surface

Hydrogeology Journal (2003) 11:7–40 DOI 10.1007/s10040-002-0241-5


24

1945). Since then, numerous such cases have been re-


ported, and seem to fall into two categories of mechani-
cal earth-crust response (Simpson et al. 1988). In many
cases, there is a rapid response, in which seismicity in-
creases almost immediately after reservoir impound-
ment. In other cases, there is a delayed response, in
which the main seismic activity does not occur until
some years after the reservoir has been filled and the wa-
ter level is maintained at a stable height. Cases of rapid
response tend to produce swarms of small earthquakes,
located at shallow depth, in the immediate vicinity of the
reservoir. In the delayed response case, the earthquakes
tend to be larger, at a greater depth, and are often located
at some distance from the deep part of the reservoir. An
example of delayed response, and one of the most devas-
tating earthquakes induced by reservoir impoundment,
occurred at a dam that was being built in Konya, India,
containing Shivajisagar Lake. Prior to the dam being
built, the area recorded very low seismic activity. After
being filled in 1962, seismograph records showed fre-
quent earthquakes directly beneath the lake. However,
large earthquakes did not occur until late 1967, when
magnitude 5.5 and 6.2 events occurred. The largest
earthquake, which caused considerable damage and
claimed the lives of 200 people, occurred at a depth of
5 km, some 10 km downstream from the dam. It oc-
curred just after the water level had reached its maxi-
mum value (Scholz 1990).
The best-known example of dam failure is the col-
lapse of the Malpasset Dam in France in 1959, that killed
at least 500 people. The Malpasset Dam failure took
Fig. 16a, b Spatial distribution of the microseismic events at place in a sudden movement, like an explosion, with
Hijiori hot-dry rock site in Japan. a Events during a one-day high- nothing abnormal detected on the dam a few hours be-
pressure hydraulic fracturing operation. b Events during a one- fore the failure (Londe 1987). Six years of research
month low-pressure circulation experiment (modified from Sasaki
1998). Each plane view is 2 by 2 km showed that, under the load of the dam, permeability was
reduced by a factor of approximately 0.01. As a result,
an “underground hydraulic barrier” was created, in
reservoir impoundment (Gupta 1992). The process is which the concentrated action of seepage forces contrib-
analogous to fluid injection, in that increased fluid pres- uted to failure by sliding of the dam foundation along a
sure will reduce the effective stress and shear strength fault located under the structure (Fig. 17). At the time of
along existing discontinuities, leading to shear failure or the design and even at the time of the failure, very little
fracturing. In addition, the added loads from the reser- was known about the mechanism of water seepage
voir water increases stress in surrounding rock forma- through fractured media. Foundation design procedure
tions and load dam abutments through direct elastic and did not include any study of uplift in the rock, except at
poroelastic effects. Large-scale failure of rock masses the contact between concrete and foundation (Londe
during reservoir impoundment can be devastating. Ex- 1987). Furthermore, the variation in rock permeability as
amples are landslides (e.g., Vaiont, Italy, 1963) and dam a function of applied stress was an unknown phenome-
failures (e.g., Malpasset, France, 1959). A landslide or non before the elaborate and systematic analysis under-
dam failure may be preceded by induced seismic activity taken as a result of the accident (Habib 1987). It was
that can be a warning signal for a larger failure. Howev- during the investigation of the Malpasset Dam failure
er, large failures may also take place abruptly without that the concept of fractured rock hydraulics and
forewarning and can be catastrophic. stress–permeability coupling in single fractures was ini-
In rare instances reservoir impoundment has been tiated.
linked to some of the largest and most damaging earth-
quakes induced by humans (Gupta 1992). The first real
evidence relating increased seismicity to water storage Excavation-induced Squeezing
came in 1936, when an increase of seismicity occurred Solid-to-fluid interaction can be directly observed in a
after the creation of Lake Mead, held back by the porous medium as a pressure pulse, if the porous medi-
Hoover Dam on the Arizona–Nevada border (Carder um (soil and rock) is squeezed rapidly by a mechanical

Hydrogeology Journal (2003) 11:7–40 DOI 10.1007/s10040-002-0241-5


25

Fig. 18 Excavation-induced pressure changes in rocks during tun-


nel boring at the Grimsel test site in Switzerland. The pressure is
measured in a 10-m-long packed-off section of a borehole located
about 3 m from the wall of the tunnel (see sketch of tunnel and
monitoring section below the pressure curve)
Fig. 17 Londe’s hypothesis for the 1959 failure at the Malpassat
Dam, France. The figure shows how an underground hydraulic
barrier was created under load of the dam due to extremely stress-
sensitive permeability (modified from Londe 1987)
Excavation-induced Permeability Changes
Stress-induced changes in permeability around excava-
tions have been studied since the early 1980s. These
force. For example, a rapid contraction of porous media studies have been related to research on the excavation
leads to a local pressure increase. Fluid pressure can disturbed zone (EDZ) around tunnels, which is an issue
then diminish as fluid flows out of the pressurized area. of concern for the safety assessment of nuclear waste re-
An observable pressure pulse will only occur if the positories. The effect of the EDZ on hydraulic behavior
loading is so rapid and the permeability is so small that around tunnels was first realized in an experiment at the
the fluid has no time to escape. One example of this Stripa Mine, Sweden, as a part of the Stripa Project (see
phenomenon is the response in water-well levels caused Witherspoon 2000 for a recent recapitulation). The ex-
by ground compression of passing trains, studied by periment was a large-scale permeability test, in which
Jacob (1940). Figure 18 presents another recent exam- the radial head distribution showed a peculiar drop with-
ple, which was observed in fractured crystalline rocks in 5 m of the drift wall. This head decrease could be ex-
at the Grimsel test site for geological disposal of radio- plained by a “skin”, or zone of reduced permeability
active waste, Switzerland. The figure shows fluid-pres- around the drift. This skin was thought at the time to re-
sure response in a drift wall during boring of the drift flect closure of fractures caused by tangential stress con-
(2.28 m in diameter). The fluid pressure was monitored centration around the opening. It was later found that, as
in a packed-off borehole section, about 10 m long, ori- the water pressure fell below 17 m of water head, out-
ented parallel to the future drift, at a radial distance of gassing occurred and thus the increased gradient, just
about 3 m from the drift wall (Fig. 18). The figure outside the wall, could also be attributed to effects of
shows two distinct pressure peaks, which occurred dur- two-phase conditions in the fractures.
ing two consecutive, eight-hour working shifts as the After this first observation at the Stripa Project, EDZs
face of the tunnel-boring machine (TBM) passed paral- have been investigated at many sites in various rock
lel to the monitoring section. The pressure peaks can be types. A large fraction of these studies have involved
explained by an increased mean stress at or near the lo- measurements of mechanical responses during mine-by
cation of the packed-off monitoring section. Such in- experiments or seismic velocity evaluation of a damaged
crease in mean stress is possible if the virgin stress zone due to blasting. In other studies, the mechanical and
around the drift is anisotropic. After the end of each hydraulic responses have not been measured directly, but
shift, when the excavation was stopped, the fluid pres- a skin zone around the excavation has been inferred in or-
sure declined because fluid slowly diffused away from der to explain hydraulic behavior (e.g., buffer mass test
the pressurized zone. Similar transient phenomena have and site characterization and validation test in Stripa by
also been observed during excavations in Opalinus clay Börgesson et al. 1992, and Olsson 1992). These dam-
at Mont Terri Rock Laboratory, Switzerland (Thury and aged-zone observations and the suspicion of a permeabil-
Bossart 1999), and in fractured granite at the Under- ity skin motivated deliberately designed in-situ tests to
ground Research Laboratory (URL), Canada (Read and quantify permeability changes around underground open-
Chandler 1999). ings. Such in-situ tests have been conducted since the
1990s at the Äspö hard rock laboratory, Sweden (Emsley

Hydrogeology Journal (2003) 11:7–40 DOI 10.1007/s10040-002-0241-5


26

Fig. 19 Schematic of the excavation disturbed zone (EDZ) around


a drift in fractured rock

et al. 1997), Kamaishi and Tono mines in Japan (Sugihara


et al. 1999), and at the Underground Research Laboratory
in Manitoba, Canada (Bäckblom and Martin 1999).
Findings of the above-mentioned experiments lead to
a few general observations regarding EDZs and exca-
vation-induced permeability changes (Bäckblom and
Martin 1999). First, the EDZ is the rock zone where rock
properties and conditions are changed because of pro-
cesses induced by excavation. The EDZ includes a dam- Fig. 20a, b Pre/postexcavation permeability tests conducted in
aged zone of induced rock failure and fracturing stem- fractured unsaturated tuff at Yucca Mountain, Nevada. a Geometry
ming from excavation processes, an unsaturated zone, a of tunnel and test boreholes, and b ratio of pre- and postexcava-
zone with altered stress distribution around the drift, and tion permeability plotted against pre-excavation permeability (ex-
a zone of reduced fluid pressure (Fig. 19). In the case of tracted from Wang et al. 2001)
excavation-induced damage, high transmissivity will
probably develop close to the opening. For mechanical
excavation (using no blasting) in a moderate stress envi- permeability at four niches excavated by a mechanical
ronment, the damage zone may be limited to a few centi- (alpine mining) method. Permeability was measured be-
meters thickness where a limited change in porosity and fore and after excavation in 30-cm, packed-off sections
permeability may take place. When drill-and-blast is along 10-m-long boreholes located about 0.65–1 m
used for excavation, the damage zone is more extensive above the drifts, supposedly outside the damaged zone
and therefore increased permeability is likely, especially (Fig. 20a). The results showed that measured air perme-
in the floor where the permeability can increase by 2–3 ability generally ranged from 1×10−15 to 1×10−10 m2
orders of magnitude. In high-stressed, relatively intact (Fig. 20b), which is at least 3 orders of magnitude higher
rock conditions like at the URL in Canada, induced frac- than the matrix permeability measured from core sample
turing and rock loosening at the wall of excavations can tests. This shows that at least one hydraulic conductive
cause dramatic (up to 6 orders of magnitude) permeabili- fracture intersects each 0.3-m borehole section and that
ty changes. For the rock mass outside the damage zone, the fractures, which are practically dry at all times, form
the experiments at Äspö, Tono, Kamaishi Mine, and a well-connected network for airflow. The results in
URL indicate very little excavation-induced changes in Fig. 20b show that the pre/postpermeability ratios vary
permeability. For example, at the ZEDEX experiment in in the range 1–400, with a change in arithmetic mean of
Äspö, the permeability sometimes increased and some- about 60. There is also a tendency for stronger perme-
times decreased before and after excavation, and it was ability increases in those sections where the initial per-
not possible to come to a firm conclusion about changes meability is small. Preliminary analyses by Wang and
in properties (Bäckblom and Martin 1999). Elsworth (1999) and Rutqvist and Tsang (2002) show
A recent study in unsaturated fractured tuff at Yucca that the permeability increase can be explained by open-
Mountain, Nevada, provides a comprehensive dataset ing of horizontal fractures unloaded by the drift. A rela-
pertaining to excavation-induced permeability changes tively large change in permeability around the drifts at
(Wang et al. 2001). The data include measurements of air Yucca Mountain, in relation to the other sparsely frac-

Hydrogeology Journal (2003) 11:7–40 DOI 10.1007/s10040-002-0241-5


27

tured sites mentioned above, is possibly caused by a reservoirs. These are irregularly distributed,and, because
well-connected fracture network that can create consis- of their scarcity, do not belong to any common statistical
tent changes in bulk permeability. grouping. Dyke (1995) concluded that the most reliable
way characterizing the response of a fracture subset is
through back-analysis of in-situ reservoir behavior.
Pressure-sensitive Permeability During Injection/
Withdrawal into Fractured Reservoirs
Pressure-dependent permeability for injection pressures Modeling of HM Coupling in Fractured Rocks
below lithostatic stress has been confirmed at hot-dry
rock sites (e.g., Fenton Hill, U.S. (Brown 1993), The advancement of computer technology in recent de-
Rosemanowes, U.K. (Richards et al. 1994), Fjällbacka, cades has made it possible to analyze problems using
Sweden (Eliasson et al. 1990), Higashihachimantai, Japan fully coupled HM analysis (including nonlinear, inelastic
(Hayashi and Abe 1989), Falkenberg, Germany (Jung processes and heterogeneous geological media). This is
1989), and Soultz, France (Baria et al. 1999a). At a geo- of great benefit for practical applications. Jing and
thermal site, transmissitivity is one of the main factors Hudson (2002) provide a comprehensive review of nu-
governing production of heat and, consequently, the eco- merical methods in rock mechanics, including finite-ele-
nomics of the operation. The hydraulic properties of a ment (FEM), boundary-element (BEM) and distinct-ele-
hot-dry rock reservoir are usually evaluated using hy- ment (DEM) methods. For more details on FEM model-
draulic jacking tests to assess pressure dependency of ing related to consolidation of porous media, a textbook
transmissivity. Unlike the hydraulic jacking test men- by Lewis and Schrefler (1998) is recommended. For ana-
tioned above, where one fracture is isolated and pressur- lytical solutions of HM coupling, refer to Neuzil (2003,
ized, injectivity tests in geothermal reservoirs are large- this volume) regarding geological processes, and Wang
scale and involve jacking of a large volume of fractured (2000) who presents useful analytical solutions for a
rock. Secondly, a circulation test may be conducted be- large number of practical problems. This section briefly
tween two deep wells to determine the “flow impedance”, introduces a coupled HM numerical model that uses a
which is defined as the pressure difference between the FEM technique and discusses some conceptual ap-
injection and production well divided by the injection rate proaches for using equivalent continuum medium to re-
(Hayashi and Abe 1989). It has been shown that flow im- present HM behavior in fractured rocks.
pedance can improve if a back-pressure is applied to the
production well, as a result of pressure-dependent perme-
Governing Hydromechanical Equations for Porous
ability (Eliasson et al.1990; Brown 1993).
Deformable Medium
Examples of how permeability can depend on reser-
The macroscopic hydromechanical behavior of intact po-
voir pressure have been recorded in fractured, sedimen-
rous rock can be described as governed by a static force
tary “tight sand” gas reservoirs. One prime example is
equilibrium:
the U.S. Department of Energy-supported Multiwell Ex-
periment (MWX) at the Piceance Basin in Colorado. The ∇ ⋅ σ + ρm g = 0 (49)
site was developed to perform detailed experiments on and by Biot’s constitutive equation for mechanical be-
all aspects of low-permeability natural gas reservoir havior, written in incremental form as
evaluation, stimulation, and production (National Re-
search Council 1996). One dominating set of extensional σ = D : ε − αIp (50)
fractures, some of which are incompletely cemented, is where I is the identity tensor and α is the Biot-Willis co-
the primary production sources for these tight sands. In- efficient. This equation is more general then Eq. (8) in
terference tests showed that permeability anisotropy was that the medium does not necessarily have to be isotropic
on the order of 1:100, owing to the unidirectional nature elastic. To accommodate for nonlinear or plastic me-
of the fracture system. Production tests showed that the chanical behavior, Eq. (50) can be written in incremental
natural fracture system is highly stress sensitive. By de- form as
creasing the reservoir pressure below a critical value
(typically about 6.9 MPa), the well production could be dσ = D : dε − αIdp (51)
almost totally stopped because the decrease in pressure where D is the tensor of incremental elastic constants.
created higher effective stresses that physically closed For a linear elastic isotropic medium, the D can be ex-
the fractures (National Research Council 1996). panded in Cartesian coordinates as:
Pressure-independent permeability in fractured oil
and gas reservoirs has been reported in rock having high-
ly conductive, “locked-open” fractures. Dyke (1995) list-
ed six naturally fractured hydrocarbon reservoirs, which
have been producing for a considerable time under large (52)
drawdown pressure, without any noticeable reduction in
permeability. Dyke (1995) argued that highly permeable
macrofractures appear to control permeability in many

Hydrogeology Journal (2003) 11:7–40 DOI 10.1007/s10040-002-0241-5


28

Equations (49) and (51) are combined to obtain the stress- where
compatibility equation, which for a quasi-static coupled
(59)
system can be written in a general incremental form as:
∂p  ∂F and and are average values over the integration
∇ ⋅  D : ∂ε − Iα = (53) interval. In Eq. (59), ς is the integration parameter with a
∂t ∂t  ∂t
value between 0 and 1. In differential form the final
This equation basically assures that the internal net force equation is:
(left-hand side) in all directions is in balance with exter-
nal forces and gravity effects (right-hand side) working ˆ ˆ
{ }
C C  ∆uˆ t  Fˆ t 
in the same direction.  uu uP
 ∆pˆ t =  ˆ ˆ t −1
The governing fluid flow equation can be written as: ˆ ˆ ˆ k p ∆ t + ˆ (t )∆t 
Q
CPu −ζ∆tkpp − Cpp   pp 

(54) (60)
The time marching solution technique for linear hydro-
which is a combination of Eq. (9) and Darcy’s law. elasticity is unconditionally stable for ς≥0.5, although an
early-time solution may exhibit a decaying oscillatory
behavior for arbitrary, initial time step sizes. To reduce
Finite-element formulation the oscillatory behavior, a predictor-corrector scheme is
The following presents a numerical formulation for cou- used in which a solution is sought at t1+θ ∆t and linearly
pled hydromechanical analysis, as it was developed in interpolated to t1+∆t=t2. For θ=1 and ς=0.5, the Crank-
the HM version of the ROCMAS code (Noorishad et al. Nicholson approximation is obtained. The solution of
1982, 1992). Eq. (60) is obtained by a Newton-Raphson scheme to al-
The finite-element formulation for the matrix rock is low for material nonlinearity in the stiffness and fluid
expressed in terms of strain vector rather than the strain conductivity matrices (Noorishad et al. 1992).
tensor, leading to the following two equations:
∂ {ε}  ∂p
∇ ⋅  D − ∇ ⋅  αI  = ∂F (55) Extension of Formulation for Fractured Media
∂t  ∂t ∂t Macroscopic hydromechanical behavior of a fracture un-
der normal closure is governed by a condition of total
(56) stress balance between the upper and lower fracture sur-
faces, and Biot’s equations. Biot’s equation extended to
where {εε} is a column vector containing normal and mechanical behavior of a joint can be written as:
shear strains (e.g., in Cartesian coordinates {εε}={εxx, εyy,
εzz, εxy, εyz, εzx}tr where tr denotes transpose). In (61)
Eq. (56) the effect of gravity flow is included in the
source term Q. which in a compacted form is
Using standard finite-element procedures, Eqs. (55)
and (56) are discretized in terms of unknown nodal dis- dσ f = k f du f − α1 f dp (62)
placements, pressure, forces and flow through appropri- Note that in Eq. (61) kn∆un is the effective normal stress,
ate shape (interpolation) functions as: and fluid pressure plays no role in the elastic incremental
shear behavior. Similarly to Eq. (52), a fracture stress-
(57) compatibility equation can be derived from Eq. (62) and
a cross-fracture total stress balance requirement as:

where the coefficient matrices , etc., are coeffi- (63)


cient matrices that contain material parameters such as
permeability, Biot’s constants, appropriate shape and Biot’s fluid flow equation for a fracture can be written as
transformation functions, and element volume. For the
exact expressions of these coefficient matrices, refer to (64)
Noorishad (1982, 1992).
Equation (56) is then integrated between time t1 and where fracture permeability is defined as
t2=t1+∆t to obtain:
(65)

and the volumetric strain is approximated as


(58) (66)

Hydrogeology Journal (2003) 11:7–40 DOI 10.1007/s10040-002-0241-5


29

In fractured porous media, fractures may be represented parameter that can be related to the empirical rock mass
explicitly in a discrete representation using special joint quality index RMR (Liu et al. 1999).
elements (e.g., Goodman 1976) whereas the matrix rock For nonorthogonal fracture sets, coupled HM behav-
between fractures is represented by solid elements. The ior can be modeled using a multilaminate model by
contribution from both solid elements and joint elements Zienkiewics and Pande (1977), in which nonlinearities
are superimposed in the element assembly process. In the such as sliding and separation of joints can be consid-
case of a joint element, the coefficient matrices in Eq. (60) ered. Using such a model, a permeability tensor may be
includes the material parameters given in Eqs. (63) and derived according to Snow (1969).
(64) and appropriate interpolation functions for joint ele- If fractures are not grouped in sets but rather random-
ments. However, for analysis of a large-scale fractured ly distributed, the equivalent properties may be calculat-
rock mass, meshing of discrete fractures is cumbersome ed by the concept of the “fabric” or “crack” tensor pro-
and an equivalent-continuum representation of the frac- posed by Oda (1982). The crack tensor is a unique mea-
tured rock mass is often the most practical approach. Vari- sure combining four significant aspects of hydraulic and
ous approaches for equivalent continuum representation of mechanical behavior in fractured rocks: volume, frac-
fractured porous rocks are presented below. ture size, fracture orientation, and fracture aperture.
However, as in any continuum model, a sufficient num-
ber of fractures must be included in the volume in which
Equivalent Continuum Models of HM Behavior the average properties are calculated to make the de-
in Fractured Rocks rived equivalent properties meaningful (Stietel et al.
Equivalent-continuum models for HM behavior of frac- 1996).
tured rock masses can be derived using the existing ap- In many cases in reservoir engineering, the storage of
proaches for equivalent representation in hydrogeology fluids is dominated by storage in the matrix whereas per-
or reservoir engineering and in rock mechanics. In rock meability is dominated by fluid flow in fractures. In such
mechanics, ubiquitously fractured models have been de- a case, so-called dual-porosity or overlapping-continua
veloped in which ideal sets of fractures are considered models are applicable. Dual-porosity models were first
with their spacing, orientation, and deformation proper- developed in reservoir engineering in the 1960s (e.g.,
ties. One of the simplest approaches is to assume three Barenblatt et al. 1960). Models incorporating both Biot’s
orthogonal sets of fractures in a “sugar cube” medium poroelastic theory and Bareblatt et al.’s dual-porosity
model. Using such an approach, simple equations for the concept have been studied by several authors since the
fluid flow and elastic properties can be derived. For ex- 1970s. Recent developments include Chen and Teufel
ample, consider an ideal medium with three sets of paral- (1997) and Bai et al. (1999).
lel fractures in which all sets have the same spacing and
the same hydraulic aperture. The permeability along the
direction of each set is (Snow 1965) Applications

(67) This section surveys applications where HM coupling is


considered significant and HM analysis is performed. The
where s is fracture spacing. From this equation, it is clear three main fields in which coupled HM analysis is used to-
that the fracture aperture is much more important than day are oil and gas explorations, analyses of nuclear waste
the fracture spacing for permeability. Likewise, an disposal, and geothermal energy (hot-dry rock) extraction.
equivalent compliance for coupling normal stress and Other applications where HM coupling is important in-
normal strain in each direction can be derived as (Gerr- clude coal mining and coal methane extraction, deep well
ard 1982) injection of liquid, and underground storage of natural gas.
New applications include deep injection of solid waste and
(68) geological sequestration of greenhouse gases (CO2). This
section provides a brief overview of each of these applica-
where ERM and ER are the moduli of deformation for the tions with regard to the role of HM coupling.
rock mass and the rock matrix, respectively, and kn is the
fracture normal stiffness. Models for orthogonal aniso-
tropic solids (orthotropy) can be used to derive the coef- Hydrogeology and Well-test Analysis
ficients in the D matrix of Eq. (50) according to Gerrard Coupled HM effects are generally not considered in con-
(1982). Coupling between fracture deformation and fluid ventional geohydrology and well-test analysis except
flow can be provided using Biot’s equations extended to through the concept of specific storage. It is well known
fractures, as in Eqs. (61) and (64). Liu et al. (1999) ap- that the classical solutions by Jacob (1940) and the con-
plied the conceptual model of three orthogonal fracture cept of coefficient of storage in Eq. (4) above are based
sets to calculate changes in permeability from normal on a one-dimensional mechanical model and a two- or
and shear strains. The sensitivity of permeability to three-dimensional flow geometry. The partially coupled
stress in their model is controlled by the initial perme- Jacob diffusion equation can be derived from Biot’s fully
ability and the modulus of reduction ratio, Rm=ERM/ER, a coupled equations. First, a complete flow equation for a

Hydrogeology Journal (2003) 11:7–40 DOI 10.1007/s10040-002-0241-5


30
Fig. 21 Flow and pressure re-
sponse during a constant pres-
sure injection into a subhori-
zontal fracture located at 70-m
depth. The injection pressure is
1.2 MPa and the vertical stress
normal to the fracture is esti-
mated to be 2.0 MPa. Note that
the injection rate increases dur-
ing the first 60 min, despite a
reducing pressure gradient.
This can be explained by an in-
creased permeability as a result
of increased fluid pressure
within the fracture (data ex-
tracted from Alm 1999)

fully coupled system can be derived by combining normal stress across the horizontal fracture, estimated at
Darcy’s law with Eqs. (9) and (24), leading to: 2.0 MPa. The flow rate first stays roughly constant for
about 15 min. Thereafter, the fracture starts to jack open,
(69) and the flow increases with time to reach a maximum in-
jection rate after 60 min. Later, the flow rate decreases
again, probably because it has been completely filled
This is a fully coupled equation, since fluid flow de- with pressurized water, as shown in the lower part of
pends on mechanical deformation, through the volumet- Fig. 21. This exemplifies that permeability is not con-
ric strain. In an uncoupled (Jacob’s) approach, the volu- stant but rather changes with pressure inside the fracture.
metric strain in Eq. (69) is calculated with the assump- However, the test is conducted at shallow depths and the
tion of no lateral strain and a constant total vertical fluid pressure is only 0.8 MPa below the lithostatic
stress, according to stress. In most cases, a fluid pressure well below 20% of
the lithostatic stress is sufficiently low to avoid an in-
(70) creasing permeability. A hydraulic jacking test (as de-
scribed above) can be conducted to investigate whether
and the uncoupled flow equation can be written as the injection pressure is sufficiently low to avoid signifi-
cant changes in permeability.
(71) Coupled HM analysis of injection tests can be used to
constrain subsurface hydraulic properties. The mechani-
cal responses of fractures dominate the storativity during
where Kv is uniaxial (vertical) deformation modulus. The a hydraulic injection test in low-permeability fractured
coefficient rocks. Local and regional stress balance in a rock mass
and its mechanical stiffness provide a confinement that
(72) restricts mechanical movements within certain limits.
This means that the mechanics of the rock and rock frac-
is the specific storage which, for incompressible grains tures provide practical bounds for fracture storativity and
(α= 1, and Ks=∞), reduces to the usual Jacob specific specific storage. As demonstrated by Rutqvist (1995c),
storage coefficient, Ss, defined in Eq. (7). Then, by sub- this mechanism can be used to constrain fracture storati-
stituting p=(h−z)ρg into Eq. (71), the exact Jacob diffu- vity when interpreting well tests.
sion Eq. (6) can be derived.
Most well-test formulas are based on a partially cou-
pled diffusion equation, with the assumption of a con- Oil and Gas Extraction
stant vertical stress for the derivation of specific storage, HM coupling is important for a multitude of problems re-
and of a constant permeability that does not change with lated to oil and gas exploration and extraction, including
injection pressure. The assumption of constant perme- poroelastic stress redistribution around a borehole, reser-
ability could present difficulties in fractured rocks, voir compaction and subsidence, as well as the mechanics
where the permeability can be very sensitive to injection of hydraulic fracturing (Geertsma 1957, 1966, 1973).
pressure. Figure 21 presents a recent example from a test Since the 1970s, hydraulic fracturing technology has been
on fractures in granite located at 70-m depth. The injec- the most important topic for HM coupling in oil and gas
tion pressure is 1.2 MPa, which is well below the total exploration. Recently, other areas have attracted interest:

Hydrogeology Journal (2003) 11:7–40 DOI 10.1007/s10040-002-0241-5


31

borehole stability and shale mechanics, reservoir compac-


tion and subsidence, enhanced oil recovery (Dusseault
1997), and stress-sensitive permeability changes during
production (Davies and Davies 1999). Reservoir compac-
tion is nowadays generally analyzed using partially or
fully coupled numerical modeling approaches (see Lewis
et al. 2003, this volume). In a partially coupled approach,
the pressure pattern obtained from reservoir simulators is
used in a stress-strain analysis to compute the compaction
of the reservoir and the subsidence of the land surface
(Gambolati and Freeze 1973; Gambolati et al. 1991).
However, fully coupled models based on Biot’s theory of
consolidation have also been used (e.g., Gutierrez et al.
1995; Lewis and Schrefler 1998). Recently, several fully
coupled reservoir simulation models, with the capability Fig. 22a–c Development of the HDR reservoir concepts in the
of simulating stress-dependent permeability, have been last 25 years. a Penny-shaped fracture, b shear on natural joints,
and c graben concept (after Baria et al. 1999a)
developed for oil and gas production simulations (e.g.,
Lewis and Sukirman 1993; Koutsabeloulis and Hope
1998; Gutierrez and Lewis 1998; Osoiro et al. 1998). with the orientation of the natural joints and in-situ stress
Such models may become increasingly important as con- conditions. The events were located between about 500 m
ventional oil supplies are depleted. Extraction of uncon- above and 2 km below the point of hydraulic injection,
ventional heavy oil requires thermal recovery methods or and they were used for the location of the HWR reservoir.
other techniques that involve massive stress changes cou- Later stimulations using gel and water gave improved
pled with fluid flow and evolution of mechanical proper- connectivity between the boreholes drilled into the reser-
ties (Dusseault 1997). Thus, analysis of coupled thermo- voir and reduced circulation impedance.
hydromechanical (THM) processes will be increasingly In the 1980s the term “hot-wet rock” (HWR) system
important for oil and gas exploration. was proposed for geothermal extraction systems in which
hydraulic fracturing and injection were used for perme-
ability enhancements and for sustaining production in
Geothermal Energy naturally fractured hot rock masses that exhibit only mi-
Several large projects, directed at establishing the tech- nor flow. In Europe the HWR system is sometimes called
nology for developing commercial engineered geother- the “graben concept” as it is associated with areas of gra-
mal systems, are now underway in various parts of the ben structure where stresses tend to be low, which allows
world. At the time of the first “hot-dry rock” (HDR) pro- stimulation and circulation to be carried out at moderate
ject at Fenton Hill in the U.S.A., discussion was mainly pressures (Baria et al. 1999a). The concepts that have
focused on a subsurface reservoir system consisting of evolved in the last 25 years for HDR/HWR are shown in
one or more simple, artificial, penny-shaped fractures Fig. 22. A summary of the most important HDR/HWR
propagating in the direction of the maximum principal field experiments and their target temperatures, regional
stress (Fig. 22a). The term “dry” in HDR came from the and local geology, and geochemistry was presented by
fact that the intact basement granites where the tests Nakatsuka (1999), and the relevant stress and rock me-
were performed were of very low permeability. Once chanics issues by Evans et al. (1999).
HDR reservoirs were recognized as essentially manmade In the last eight years or so the concept of HDR/WDR
variants of natural hydrothermal reservoirs, the similari- has shifted again to take greater benefit of the geological
ties between HDR hydrothermal resources and extrac- environment. Hence, at the European Community project
tions from natural systems became apparent. at Soultz-sous-Forêts, the reservoir is located in a typical
In the early 1980s, Pine and Batchelor (1984) con- graben structure, where high temperature can be encoun-
firmed that the creation of new fractures was not the tered at shallow depth at relatively low minimum stress
dominant process during the injection of water into the gradients, in which faults or fault zones are already par-
rock mass at great depth. Far more important was the tially open for flow. At Soultz, a system consisting of two
shearing of natural joints and, in particular, those aligned production wells and one injection well is planned, in
with the principal stresses of the local stress field, as which the three boreholes are likely to be aligned parallel
shown in Fig. 22b. As discussed above, the joints fail in with the direction of the maximum horizontal stress, with
shear because the fluid injection reduces the normal stress a smaller normal stress on fractures connecting the wells
across them and allows frictional slippage to occur before (Baria et al. 1999b). In Sweden, the concepts will be ap-
jacking. This was first demonstrated for the Cornwall plied to areas with low heat flow. In the county of Scania
HDR project in the Carnmenellis granite where injection in southernmost Sweden, at the edge of the Baltic Shield,
was conducted at depths greater than 2 km below ground a well is being drilled to target one of the major fault
level. Microseismic events detected during the high-flow zones in Precambrian granites and gneisses at about 4-km
rate stimulations indicated strike-slip shear consistent depth. Warm water (estimated at 70 °C) will be extracted

Hydrogeology Journal (2003) 11:7–40 DOI 10.1007/s10040-002-0241-5


32

from the fault and later used for heating in the city of tests (BMTs) and test cases (TCs) of actual experiments.
Lund. Another project at Björkö in central Sweden is lo- The tasks have involved coupled THM problems from the
cated in a 10-km-diameter meteorite impact crater of Pre- scale of laboratory samples to kilometers. A number of
cambrium age about 20 km west of Stockholm (Henkel major field experiments have been TCs in DECOV-
2002). The impact has caused severe fracturing of the ALEX, e.g., Fanay Augères in France (Rejeb 1996), Ka-
bedrock, with fractures later healed from calcite, chlorite maishi Mine, Japan (Rutqvist et al. 2002), FEBEX at
and laumontite precipitation. These are low-strength min- Grimsel in Switzerland, and the drift scale test (DST) at
erals, and the present concept is to stimulate and reopen Yucca Mountain, Nevada. HM coupling is part of all
existing fractures and create a highly permeable, low- these experiments. However, direct measurements of per-
temperature geothermal reservoir. After heat exchange at meability changes during heating have only been con-
the production hole, the warm water will be fed into the ducted in the recent drift scale heater test at Yucca Moun-
existing district heating system of Stockholm. tain. This ongoing experiment indicates that permeability
decreases up to 1 order of magnitude during heating be-
cause of thermal stress and associated closure of rock
Nuclear Waste Disposal fractures. Such data were used by Rutqvist and Tsang
In general, HM coupling is part of the coupled thermo- (2002) to back-calculate in-situ HM properties of the
hydromechanico-chemical (THMC) processes that are fractured rock at Yucca Mountain. These in-situ calibrat-
being studied as part of the performance assessment for ed properties were then used to predict coupled THM ef-
nuclear waste repositories. Coupled thermo-hydrome- fects at a potential repository at Yucca Mountain, and to
chanical (THM) processes are important because the estimate the impact of HM coupling on the long-term
heat released from the spent fuel will heat the surround- performance of the repository (Rutqvist and Tsang 2002).
ing rock mass, creating thermal stresses. Depending on
repository design and waste characteristics, the tempera-
ture may reach 50–150 °C, with the greatest thermome- Deep Well Injection of Liquid Waste
chanical impact occurring around 100–1,000 years after The possibility of induced earthquakes was mentioned
waste emplacement. There are two main areas in which above in connection with the waste injection operation at
HM coupling could have a significant impact on a repos- the Rocky Mountain Arsenal near Denver, in the early
itory performance: (1) the mechanical integrity of the re- 1960s. However, although the injection-induced seismi-
pository, and (2) possible changes in radionuclide trans- city is of great concern, it is a rare phenomenon that does
port properties. HM coupling affects the mechanical in- not pose a problem for the vast majority of active injec-
tegrity and stability of the repository vaults through the tion wells. Instead, it is the potential for contaminating
effective stress law. This effect must be predicted otherwise usable groundwater that is of most concern
throughout the repository lifetime because the fluid pres- (Apps and Tsang 1996). A breech in the confining inter-
sure (in saturated rock formations) will generally in- val present above the injection zone is one possible ave-
crease over time because of restoration of hydrostatic nue of fluid migration from the injection zone. Such flu-
fluid pressure that was drained during the construction id migration can be enhanced if the confining layer is in-
phase. Transport properties may change as a result of the tersected by natural faults and fractures, or by an in-
excavation in the EDZ. Furthermore, during heating of duced fracture. However, since the 1980s, more stringent
the rock mass, thermal stresses can change fracture aper- regulation for deep well injection in the U.S. requires
tures through thermoelastic closure, opening or shear comprehensive HM characterization of injection zones
slip, leading to permeability changes that may or may and confining layers before and during injection opera-
not be irreversible after the period of heating. tions. Regulations in different states usually require the
Prediction of coupled THM processes associated with maximum injection pressure to be set below the fracture
nuclear waste repositories thousands of years in the future pressure of the injection zone, to ensure that the confin-
requires robust models and these models must be sup- ing zone does not fracture. During an injection opera-
plied with realistic input data. These issues were first dis- tion, injection pressure, injection volume, and flow rate
cussed in a series of workshops and international sympo- must be continuously monitored, because any change in
siums at the Lawrence Berkeley National Laboratory, the relationship between the variables could indicate
USA in the early and mid-1980s (Tsang 1987). In 1991, downhole problems. In addition, yearly falloff pressure
an international collaboration project DECOVALEX tests are required to investigate possible changes in the
(DEvelopment of COupled MOdels and their VALidation geometry of stimulated fractures in the injection zone
against EXperiments in nuclear waste isolation) was (Brasier and Kobelsky 1996).
launched. The project, and its follow-ups DECOV-
ALEX II and III, have been sponsored by nuclear waste
agencies in nine countries and have involved some 20 re- Deep Well Injection of Solid Waste
search organizations (Stephansson et al. 1996). Within The use of hydraulic fracturing for disposal of drill cut-
the project, the different research teams are assigned to tings and other wastes, or so-called slurry fracture injec-
use their respective numerical models to predict coupled tion (SFI), has become more widely used in the last
THM responses in well-defined, hypothetical bench-mark decade as an alternative to traditional landfill disposal

Hydrogeology Journal (2003) 11:7–40 DOI 10.1007/s10040-002-0241-5


33

for oilfield wastes. Large-scale operations have taken analysis of a coal methane extraction operation seems to
place in the Gulf of Mexico (Reed et al. 2001), Canada be essential for optimizing the production.
(Dusseault et al. 1996), Alaska (Schmidt et al. 1999),
California (Hainey et al. 1997), and the North Sea. SFI is
conducted by mixing a slurry of solids and fresh or pro- Geological Storage of Natural Gas
duced water and injecting it at high pressure into suitable Underground gas storage in reservoirs and rock caverns is
sand formations contained within layers of low-perme- a mature technology that has been practiced for decades.
ability shales. The low-permeability shales should act as In North America, it is a typical practice to operate gas
a fracture barrier to blunt upward fracture growth and storage reservoirs at or below the original reservoir pres-
prevent vertical fluid communication. Whether the low- sure, out of concerns about caprock integrity, fracturing,
permeability formation adjacent to the injection zone faulting and gas loss. As pointed out by Bruno et al.
will act as a fracture barrier is strongly dependent on the (1998), the maximum safe operating pressure depends on
stress contrast between the layers (Nolte 1991). During a several geomechanical factors, including in-situ stresses,
slurry injection, the carrying fluid bleeds off rapidly, stresses induced by local and global changes in the reser-
leaving behind solid waste within the fracture. An injec- voir, and the mechanical and hydraulic properties of the
tion operation over 24 h frequently involves an injection reservoir and overburden. Nagelhout and Roest (1997)
episode and a relaxation episode. The relaxation period presented numerical modeling of a generic underground
is necessary to relax fluid pressure and poroelastic stress natural gas storage facility to investigate the possibilities
buildup in the rock mass surrounding the fracture, which of fault slip in some of the new gas storage facilities in
otherwise would close the fracture and reduce injectivity. The Netherlands. The result of Nagelhout and Roest’s
Because the injection takes place at a pressure above (1997) simulation showed that fault slip of up to 5 cm oc-
fracturing pressure, a careful hydromechanical character- curred during the initial depletion of fluid pressure from
ization of the injection interval and the caprock, as well 300 to 115 bar. On the other hand, no inelastic fault slip
as monitoring of the fracture geometry are essential to occurred during the following simulated gas storage oper-
ensure that the fracture does not break though the overly- ations. However, the fluid pressure did not exceed the ini-
ing cap. During operation, fracture extension can some- tial reservoir pressure by more than 14%.
times be tracked by monitoring of acoustic emission and
surface tiltmeters (Withers et al. 1996).
Geological Sequestration of CO2
Recently, underground storage of carbon dioxide (CO2)
Coal Mining and Coal Methane Extraction into permeable aquifers, such as deep saline aquifers, de-
Coal is generally characterized by a dual porosity, con- pleted oil and gas reservoirs, and coal seams, has been
taining both micropore and macropore systems. The mac- suggested as an important potential method for reducing
ropore system consists of a naturally occurring network the emission of greenhouse gases to the atmosphere
of fractures called the cleat system, with apertures vary- (DOE 1999). The injection would take place at a depth
ing in the range 100–0.1 nm (Harpalani and Chen 1995). below 800 m so that the CO2 would be within the tem-
Because of the cleat structure, the permeability of coal perature and pressure range of a supercritical fluid. As a
has shown to be extremely sensitive to stress changes and supercritical fluid, CO2 behaves like a gas with low vis-
gas pressure. In addition, the release of methane gas from cosity but with a liquid-like density of 200–900 kg m−3,
coal induces shrinkage in the matrix, which also strongly depending on pressure and temperature. Because super-
affects the permeability of the cleat system. These HM critical CO2 is less dense than water, deep underground
coupled processes can have important consequences dur- disposal requires a sufficiently impermeable seal (cap-
ing underground mining of coal (Liu and Elsworth 1997; rock) above an underground storage zone to trap the in-
Xiao and Xu 2000) and during extraction of methane jected CO2. A caprock may be discontinuous and may
from coal seams. To recover a large percentage of meth- contain imperfections such as faults and fractures of var-
ane gas in coalbeds, reservoir pressure must be reduced ious sizes (from small, meter-scale fractures to kilome-
significantly (Harpalani and Schraufnagel 1990). Two ter-scale faults). Furthermore, the hydraulic properties of
distinct effects are associated with the reduction of pres- a fault may change with injection-induced hydraulic
sure during methane extraction: (1) the release of gas, and pressure. Thus, the performance assessment of CO2 stor-
(2) an increase in effective stress. The increase in effec- age in an underground brine formation requires coupled
tive stress causes a decrease in the permeability of the HM analysis, including multiphase flow and failure anal-
coal owing to mechanical closure of flow paths. Howev- ysis. A first study of coupled HM numerical modeling of
er, an opposing effect of coal-matrix shrinkage can occur, the effects during CO2 injection into a brine formation
one that is attributed to a desorbtion phenomenon. This has recently been conducted by Rutqvist and Tsang
shrinkage widens the fractures that are primarily respon- (2002). The simulation is conducted as a linked
sible for gas flow in coalbed reservoirs, and thus increas- TOUGH2–FLAC3D simulation, in which TOUGH2 is a
es permeability. These effects occur simultaneously and, reservoir simulator for multiphase flow and heat trans-
depending on in-situ conditions, the permeability can de- port, and FLAC3D is a rock mechanics code (Rutqvist et
crease or increase during production. A coupled HM al. 2002).

Hydrogeology Journal (2003) 11:7–40 DOI 10.1007/s10040-002-0241-5


34

Concluding Remarks
Based on laboratory and field data reviewed in this arti-
cle, a few general observations can be made. First of all,
indirect coupled processes (for example, stress-depen-
dent permeability) tend to be most pronounced in intact
rocks containing flat microcracks, such as shale and
granite, and in rocks containing macrofractures. The sen-
sitivity of permeability to stress in fractured rock de-
pends on hydraulic properties, such as fracture perme-
ability and interconnectivity of the conducting fracture
network, and on mechanical parameters, such as fracture
normal stiffness and fracture shear strength. In general,
the most important parameters for the stress–permeabili-
ty relationship of a single connected fracture are
– initial permeability,
– initial fracture normal stiffness, and
– fracture shear strength.
Fig. 23 Schematic representation of possible permeability chang-
All three of these parameters depend on the initial effec- es at shallow and deep locations in fractured bedrock. The solid
tive stress normal to the fracture. However, whereas ini- lines represent the depth- (or stress)-permeability function for in-
tact rock, clean tension joint and highly conductive and locked-
tial fracture normal stiffness and fracture shear strength open fractures
can be well correlated to the virgin in-situ stress, no such
clear correlation between permeability and the virgin
stress exists. This is because permeability is sensitive to has already been sheared past its peak shear strengths,
the effects of mineral filling (see Fig. 11). At one ex- which means that the shear dilation curves are flat and
treme, mineral filling can completely block the fluid that any further shear may give relatively small changes
flow in a fracture. At the other extreme, hard mineral in permeability (see Fig. 4d). However, if the initial per-
filling can keep the fracture open, creating large flow meability is small, it can indicate that it is a mated joint
channels between rock bridges in the fracture plane. without previous shear, or it could be a fault that is al-
Many field experiments have shown that, in general, most completely mineral filled. In the case of a mineral-
fractures with a large initial permeability are less sensi- filled fracture, any shearing could induce extreme chang-
tive to changes in effective normal stress than fractures es in permeability through breakdown of seals.
with a small initial permeability. Hydraulic borehole tests Fractured rock permeability at depth is likely to be
show that relatively large open-flow channels can exist at less sensitive to a disturbance (e.g., underground con-
depths of several thousand meters. These fractures are so struction) than fractured rocks at shallow depths (as illus-
conductive that they are likely to be locked open by large trated in Fig. 23). This has been confirmed in large-scale
shear displacement and/or hard mineral filling. These field tests in fractured rocks. For example, at the Yucca
highly permeable fractures will tend to be relatively stress Mountain site, Nevada, U.S.A., the initial horizontal
insensitive because their stiffness is relatively high. The stresses are very low, on the order of 2 MPa. This implies
fact that these features stay open indicates that they con- that fractures are open much more than their residual val-
sist of low-aspect-ratio features which, according to ues, and that the initial fracture normal stiffness is rela-
Eq. (16), could stay open in very high-stress environ- tively low. Results from field tests at Yucca Mountain in-
ments. Another extreme is fractures that are shallow and dicate that rock mass permeability can be reduced 1 order
partially filled with soft mineral filling. Shallow mineral- of magnitude by stress increases during large-scale heater
filled fractures have a relatively low fracture normal stiff- tests, before reaching a residual permeability value. Tests
ness, and at the same time a small initial hydraulic aper- of excavation-disturbed zones at other sites where the ini-
ture because the fracture surfaces are coated with mineral tial stress is higher indicate little change in permeability
filling. Such fractures can be extremely sensitive to outside the damaged zone. One reason may be that per-
changes in stress and fluid pressure. meability of fractures tends to be close to their residual
A fracture with a large initial permeability is also like- values at depths, which implies that any further stress in-
ly to be less sensitive to shear displacements. There are crease will not produce any further fracture closure
several reasons for this. First, a large initial aperture is (Fig. 23). On the other hand, a reduced stress, for exam-
likely found in large (long) fractures (see Fig. 5) which, ple, from unloading normal to a drift wall, can produce an
for a given shear displacement, do not dilate as much as opening of fractures at any depth.
small fractures (see Fig. 4c, d). Second, a large initial ap- Pressure-induced changes during injection are also
erture is likely to be found in a fracture that has already likely to be more pronounced at shallow depth. One ex-
been sheared. For example, the fracture depicted in Fig. 8 ample is the injection test presented in Fig. 21, showing
has experienced a shear displacement of centimeters. It an increasing fracture transmissivity with fracture fluid

Hydrogeology Journal (2003) 11:7–40 DOI 10.1007/s10040-002-0241-5


35

pressure. In this case an injection pressure of 1 MPa Appendix


above the initial pressure gives significant changes in
fracture permeability. In the example of the hydraulic Abbreviations
jacking test in Fig. 9c, conducted at a depth of about Most symbols are defined the first time they occur. The
300 m, an injection pressure of 1 MPa over the initial following list contains an explanation of the symbols
gives no significant changes in permeability. Injection- that need further explanation and symbols that are most
induced fracture shear also tends to be more likely in frequently used. Boldface letters represent matrix or vec-
shallow areas, because the in-situ stress field can be tor quantities, and scalars are shown in italic.
more anisotropic, and at the same time the shear strength
A Area (m2)
is small because of the relatively small effective stress.
b Physical fracture aperture (m)
Figure 23 shows a generally reduced stress dependen-
bE Barton et al.’s (1985) “real” physical aperture
cy for permeability with depth. However, fracture per-
(Eq. 42) (m)
meabilities at great depth can be highly stress dependent
bh Hydraulic conducting aperture (Eq. 26) (m)
if their initial values are low in relation to their initial
C Flow geometry constant (= ρfgw/12µffor parallel
normal stiffness. As seen in the conceptual model for
flow, Eq. 13) (-)
fractured rocks in Fig. 6, there are clusters with highly
Ĉuu FEM incremental stiffness matrix (Pa)
permeable fractures, which are connected by sections
ĈuP FEM coupling matrix (-)
containing much less-permeable fractures. These low-
tr
permeability bridges of more competent rock may con- ĈPu ĈuP
trol the overall kilometer-scale flow through the rock ĈPP FEM fluid storage matrix
mass. Since these bridges have a lower permeability, D Tensor of elastic constants (Pa)
they are also likely to have a more stress-dependent per- E Young’s modulus (Pa)
meability than the more-permeable fracture clusters. f A friction factor for fluid flow in a rough frac-
To be able to estimate the range of possible HM re- ture (Eq. 36) (-)
sponses to a human-induced disturbance (e.g., massive in- F Mechanical force vector (={Fx, Fy, Fz}tr in Carte-
jection or underground construction), an in-situ testing pro- sian coordinates)
gram is recommended. If HM interactions are significant F̂ FEM nodal force vector
for a certain application (i.e., geological sequestration of g Acceleration of gravity (m s−2)
CO2) then in-situ hydromechanical characterization should G Shear modulus (Pa)
be carried out routinely, just as hydraulic permeability tests h Total hydraulic head (=p/ρfg+z) (m)
are done. This includes a complete determination of the in- I Identity tensor (all components 0 except diago-
situ stress field, followed by in-situ tests to back-analyze nals which are 1) (-)
the stress–permeability relationship of the fractured media. JCS Joint compressive strength (measured by a ham-
Although coupled numerical modeling has advanced mer rebound test) (Pa)
greatly over the past 30 years, applying numerical mod- JRC Joint roughness coefficient (a measure of frac-
eling to coupled HM analysis can lead to great uncertain- ture surface roughness) (-)
ties. Many of these uncertainties stem from the inability JRCmob Mobilized JRC(a function of shear displacement) (-)
to characterize heterogeneous fractured rock masses and k Permeability (m2)
to capture all critical material features in a numerical k0 Permeability at a reference stress (m2)
model. In fractured rock, it may be possible to predict kn Fracture normal stiffness (Fig. 1e) (Pa m−1)
the general responses but, on a local scale, it can be diffi- kni Fracture normal stiffness at an initial effective
cult to predict responses other than probabilistically. In a stress (Fig. 4a) (Pa m−1)
sparsely fractured rock mass, one critical discrete frac- kn0 Fracture normal stiffness at zero normal stress
ture, connected to a network of conducting fractures, can (Fig. 4a) (Pa m−1)
have substantial effects on permeability, depending on its ks Fracture shear stiffness (Fig. 1f) (Pa)
location and orientation. Such a situation is difficult to k Permeability tensor (m2)
predict, because the precise locations of such fractures –
k̂pp FEM incremental conductance matrix
are generally not known in advance. K Drained bulk modulus (Pa)
Kf Bulk modulus of pore fluid (Pa)
Acknowledgements Technical review and comments by Dr. K′ Modulus of drained uniaxial (or vertical) porous
Christopher E. Neuzil, US Geological Survey, Dr. Chin-Fu Tsang, medium deformation (Pa)
Lawrence Berkeley National Laboratory, and Tech. Lic. Ki-Bok Ks Bulk modulus of solid grains (Pa)
Min, Royal Institute of Technology, Sweden are much appreciat-
ed. The following organizations are gratefully acknowledged for M Biot’s effective isothermal storage constant
their financial support: the Director, Office of Science, Office of (Eqs. 9and 24) (Pa)
Basic Energy Sciences, Division of Chemical Sciences, Geosci- p Fluid pressure (Pa)
ences and Biological Sciences, of the US Department of Energy, P Total isotropic pressure (= σxx+σyy+σzz, compres-
under contract no. DE-AC03-76-SF00098; the DECOVALEX Pro-
ject through the Swedish Nuclear Power Inspectorate; and the Eu- sive positive) (Pa)
ropean Commission through the BENCHPAR project under con- P′ Effective isotropic pressure (= σ′xx+σ′yy+σ′zz,
tract FIKW-CT-2000-00066. compressive positive) (Pa)

Hydrogeology Journal (2003) 11:7–40 DOI 10.1007/s10040-002-0241-5


36
p̂ FEM vector of nodal fluid pressure (Pa) Special symbols
Q Volume flow rate (m3 s−1) Atr Transpose of a vector
Qf Volume flow rate per unit fracture width (m3 s−1) ∇ · A Divergence of a vector (=div A)
Q̂ Nodal flow rate vector (m3 s−1) ∇A Gradient of a scalar (= grad A)
R Aspect ratio of an elliptical crack (= bc0/“crack ∇A Gradient of a vector
length”, Eq. 16) (-) A:B Contracted product of two tensors
s Fracture spacing (m)
S, Sf Sf Jacob storage coefficient or storativity (Eqs. 3
and 25) (-) References
Ss Jacob specific storage (Eq. 7) (m−1)
t Time (s) Aki K, Fehler M, Aamodt RL, Albright JN, Potter RM, Pearson
tr See special symbols below CM, Tester JW (1982) Interpretation of sesimic data from hy-
T Fluid transmissivity (m2 s−1) draulic fracturing experiments at the Fenton, Hill New Mexi-
co, Hot Dry Rock Geothermal site. J Geophys Res 87:936–951
Tr Residual transmissivity at high compressive Alm P (1999) Hydro-mechanical behavior of a pressurized single
stress (Fig. 4b) (m2 s−1) fracture: an in situ experiment. PhD Thesis, Chalmers Univer-
u Displacement (vector) [m] sity, Sweden
û FEM nodal displacement vector Alonso EE, Gens A, Delahaye CH (2003) Influence of rainfall
on the deformation and stability of a slope in overconsolidat-
un Fracture normal displacement (Fig. 4a) (m) ed clays. A case study. Hydrogeol J (in press) DOI
us Fracture shear displacement (Fig. 4c) (m) 10.1007/s10040-002-0245-1
w Width of a fracture plane (m) Apps J, Tsang C-F (1996) Preface. In: Apps J, Tsang C-F (eds)
z z-coordinate or elevation (m) Deep injection of disposal of hazardous industrial waste. Sci-
entific and engineering aspects. Academic Press, San Diego,
California
Bäckblom G, Martin CD (1999) Recent experiments in hard rock
Greek symbols to study the excavation response: Implication of the perfor-
α Biot-Willis’ coefficient (Eqs. 10and 23) (-) mance of a nuclear waste geological repository. Tunnel Under-
δ Fracture normal closure (Fig. 4a) (m) ground Space Technol 14:377–394
Bai M, Feng F, Elsworth D, Roegiers J-C (1999) Analysis of
δmax Maximum fracture normal closure (Fig. 4a) (m) stress-dependent permeability in nonorthogonal flow and de-
ε Total strain tensor (-) formation. Rock Mech Rock Eng 32:195–219
εv Volumetric strain (=εxx+εyy+εzz, positive for con- Bandis S, Lumsden AC, Barton NR (1983) Fundamentals of rock
traction) (-) joint deformation. Int J Rock Mech Mining Sci Geomech
Abstr 20:249–268
ϕ Porosity (-) Barenblatt GI, Zheltov IP, Kochina (1960) Basic concepts in the
τCK, τW Flow tortuosity factors (Eqs. 21and 46) (-) theory of homogenous liquids in fissured rocks. Prikl Mat
µf Dynamic fluid viscosity (Pa s) Mekh 24:852–864
ν Poisson’s ratio (-) Baria R, Baumgärtner J, Rummel F, Pine RJ, Sato Y (1999a)
ρf Fluid density (kg m−3) HDR/HWR reservoirs: concepts, understanding and creation.
Geothermics 28:533–552
ρm Average density of the mixture (kg m−3) Baria R, Baumgärtner J, Gérard A, Jung R, Garnish J (1999b) Eu-
σ Total stress (compression positive) (Pa) ropean HDR research programme at Soultz-sous-Forets
σ′ Effective stress (compression positive, see Eq. 1) (France) 1987–1996. Geothermics 28:655–669
(Pa) Barton NR, Choubey (1977) The shear strength of rock joints in
theory and practice. Rock Mech 10:1–54
σn Total fracture normal stress (Fig. 1d) (Pa) Barton NR, Bakhtar K (1982) Rock joint description and modeling
σ′n Effective fracture normal stress (Fig. 1d) (Pa) for the hydrothermomechanical design of nuclear waste reposito-
σ′ni Effective fracture normal stress at initial condi- ries. TerraTek Engineering, Salt Lake City, Utah, Tech Rep 83–10
tions (Fig. 4a) (Pa) Barton NR, Bandis S, Bakhtar K (1985) Strength, deformation and
conductivity coupling of rock joints. Int J Rock Mech Mining
σ′n0 A reference effective fracture normal stress in Sci Geomech Abstr 22:121–140
Eq. (31) (Pa) Barton CA, Zoback MD, Moos D (1995) Fluid flow along poten-
σm Total mean stress (=1/3(σx+σy+σz), compression tially active faults in crystalline rock. Geology 23:683–686
positive) (Pa) Barton CA, Hickman SH, Morin R, Zooback MD (1998) Reser-
σ′m Effective mean stress (= 1/3(σ′x+σ′y+σ′z), com- voir-scale fracture permeability in the Dixie Valley, Nevada,
geothermal fields. Soc Petrol Eng SPE Pap 47371
pression positive) (Pa) Bernabe Y (1986) The effective pressure law for permeability in
σs Shear stress (Fig. 1c, f) (Pa) Chelmford granite and Barre granite. Int J Rock Mech Mining
σscPeak Peak shear stress (strength) (Fig. 4c) (Pa) Sci Geomech Abstr 23:267–275
σ Macroscopic total stress tensor (Pa) Biot MA (1941) A general theory of three dimensional consolida-
tion. J Appl Phys 12:155–164
σf Macroscopic total stress vector (normal and Biot MA (1955) Theory of elasticity and consolidation for a po-
shear stress) for fractures (Pa) rous anisotropic solid. J Appl Phys 26:182–185
σ′ Macroscopic effective stress tensor (Pa) Biot MA (1956) General solutions of the equation of elasticity and
ξ Fluid volumetric strain (= ∆mf/ρf0, positive for consolidation for a porous material. J Appl Phys 27:240–253
Biot MA, Willis DG (1957) The elastic coefficients of the theory
“gain” of fluid) (-) of consolidation. J Appl Mech 24:594–601
Boitnott GN (1991) Mechanical and transport properties of joints.
PhD Thesis, Columbia University, New York

Hydrogeology Journal (2003) 11:7–40 DOI 10.1007/s10040-002-0241-5


37
Börgesson L, Pusch R, Fedriksson A, Hökmark H, Karnland O, NGW, Goodman D, Tsang C-F (eds) Fracture and jointed rock
Sandén S (1992) Final report of the zones disturbed by blast- masses. Balkema, Rotterdam, pp 81–88
ing and stress release. Swedish Nuclear Fuel and Waste Man- Eliasson TM, Sundquist U, Wallroth T (1990) Circulation experi-
agement Co (SKB), Stockholm, Stripa Proj Tech Rep 92-08 ments in 1989 at Fjällbacka HDR test site. Dept Geology,
Brace WF, Walsh JB, Frangos WT (1968) Permeability of granite Chalmers University, Rep Fj-10
under high pressure. J Geophys Res 73:2225–2236 Elliot GM, Brown ET, Boodt PI, Hudson JA (1985) Hydrome-
Brasier FM, Kobelsky BJ (1996) Injection of industrial wastes in chanical behavior of joint in the Carmenelis granite, SW En-
the United States. In: Apps J, Tsang C-F (eds) Deep injection of gland. In: Stephansson O (ed) Proc Int Symp Fundamentals of
disposal of hazardous industrial waste. Scientific and engineer- Rock Joints, 15–20 September, Börkliden, Sweden, Centek,
ing aspects. Academic Press, San Diego, California, pp 1–8 Luleå, Sweden, pp 249–258
Brower KR, Morrow NR (1985) Fluid flow in cracks as related to Emsley S, Olsson O, Stenberg L, Alheid H-J, Fall S (1997)
low permeability gas sands. Soc Petrol Eng J 25:191–201 ZEDEX – a study of damage and disturbance from tunnel ex-
Brown SR (1987) Fluid flow through rock joints: the effect of sur- cavation by blasting and tunnel boring. Swedish Nuclear Fuel
face roughness. J Geophys Res 92:1337–1347 Waste Management Co (SKB), Stockholm, Tech Rep 97-30
Brown DW (1993) Recent flow testing of the HDR reservoir at Engelder T, Scholz CH (1981) Fluid flow along very smooth
Fenton Hill, New Mexico. Geotherm Resour Council Bull joints at effective pressures up to 200 megapascals. In: Carter
22:208–214 NL (ed) Mechanical behavior of crustal rocks. Am Geophys
Brown SR, Scholz CH (1986) Closure of rock joints. J Geophys Union, Monograph 24, Washington, DC, pp 147–152
Res 91:4939–4948 Esaki T, Du S, Mitani Y, Ikusada K, Jing L (1999) Development
Bruno MS, Srinivasa MM, Danyluk PM, Hejl KA, Anduisa JP of a shear-flow test apparatus and determination of coupled
(1998) Fracture injection of solid oilfield wastes at the West properties for a single rock joint. Int J Rock Mech Mining Sci
Coyote field in California. In: Proc Symp EUROCK ’98 Rock Geomech Abstr 36:641–650
Mechanics in Petroleum Engineering, 8–10 July 1998, Trond- Evans D (1966) The Denver area earthquakes and the Rocky
heim, Norway. Soc Petrol Eng SPE Pap 47371 Mountain Arsenal Disposal Well. Mtn Geol 3:23–36
Byerlee JD (1978) Friction in rocks. J Pure Appl Geophys Evans KF, Kohl T, Hopkirk RJ, Rybach L (1992) Modeling of en-
73:615–626 ergy production from hot dry rock systems. Proj Rep Eidgenös-
Carder DS (1945) Seismic investigation in the Boulder Dam area, sische Technische Hochschule (ETH), Zürich, Switzerland
1940–1945, and influence of reservoir loading on earthquake Evans KF, Cornet FH, Hashida T, Hayashi K, Ito T, Matsuki K,
activity. Bull Seismol Soc Am 35:175–192 Wallroth T (1999) Stress and rock mechanics issues of relevance
Carlsson A, Olsson T (1977) Hydraulic properties of the Swedish to HDR/WDR engineered geothermal systems: review of devel-
crystalline rocks – hydraulic conductivity and its relationship opments during the past 15 years. Geothermics 28:455–474
to depth. Bull Geol Inst Univ Uppsala 7:71–84 Fairhurst C (1964) Measurements of in situ rock stresses with par-
Carlsson A, Olsson T (1979) Hydraulic conductivity and its stress ticular reference to hydraulic fracturing. Rock Mech Eng Geol
dependency. In: Proc Organization Econ Coop Dev-Nucl En- 2:129–147
ergy Agency OECD-NEA Int Atom Energy Agency IAEA Ferrill D, Winterle J, Wittmeyer G, Sims D, Colton S, Armstrong
Worksh Low-flow, Low-permeability Measurements in Large- A (1999) Stressed rock strains groundwater at Yucca Moun-
ly Impermeable Rocks, 19–20 March, Paris. OECD, Paris, pp tain, Nevada. Geol Soc Am, GSA Today, May 1999
249–260 Gale J (1975) A numerical, field and laboratory study of flow in
Carlsson A, Olsson T (1986) Large scale in-situ test on stress and rock with deformable fractures. PhD Thesis, University of
water flow relationships in fractured rock. Swedish State Pow- California, Berkeley
er Board Tech Rep Galloway D, Jones DR, Ingebritsen SE (eds) (1999) Land subsi-
Chen H-Y, Teufel LW (1997) Coupling fluid flow and geomechan- dence in the United States. US Geol Surv Circ 182
ics in dual-porosity modeling of naturally fractured reservoirs. Gambolati GP, Freeze A (1973) Mathematical simulation of the
Soc Petrol Eng SPE Pap 47371 subsidence of Venice. I. Theory. Water Resour Res 9:721–
Clark JB (1949) A hydraulic process for increasing the productivi- 733
ty of wells. Petrol Trans AIME 186:1–8 Gambolati G, Ricceri G, Bertoni W, Brighenti G, Vuillermin E
Cypser AC, Davis SD (1998) Induced seismicity and the potential (1991) Mathematical simulation of the subsidence of Ravenna.
for liability under U.S. law. Tectonophysics 289:239–255 Water Resour Res 27:2899–2918
Davies JP, Davies JK (1999) Stress-dependent permeability: char- Gangi AF (1978) Variation of whole and fractured porous rock
acterization and modeling. Soc Petrol Eng SPE Pap 56813 permeability with confining pressure. Int J Rock Mech Mining
Detournay E, Cheng H-D (1993) Fundamentals of poroelasticity. Sci Geomech Abstr 15:249–257
In: Hudson JA (ed) Comprehensive rock engineering. Perga- Gaucher E, Cornet FH, Bernard P (1998) Induced seismicity anal-
mon, Oxford, vol 1, pp 395–412 ysis for structure identification and stress field determination.
Doe TW (1999) Evaluating fracture network geometry from hy- In: Proc Symp EUROCK ’98 Rock Mechanics in Petroleum
draulic field data at underground test facilities. In: Faybishenko Engineering, 8–10 July 1998, Trondheim, Norway. Soc Petrol
B (ed) Proc Int Symp Dynamics of Fluids in Fractured Rocks. Eng SPE Pap 47324(1), pp 545–554
Lawrence Berkeley National Laboratory, Berkeley, California, Geertsma J (1957) The effect of fluid-pressure decline on volu-
pp 65–70 metric changes of porous rocks. Trans AIME 210:331–339
DOE (1999) Carbon sequestration research and development. In: Geertsma J (1966) Problems of rock mechanics in petroleum pro-
Reichle D et al. (eds) US Dept Energy, Rep DOE/SC/FE-1, duction engineering. In: Proc 1st Int Congr Rock Mechanics,
Washington, DC 25 September–1 October 1966, Lisbon. Lab Nac Eng Civil,
Dusseault M (1997) Geomechanics in oil and gas exploration and Lisbon, vol I, pp 585–594
drilling. In: Program 1st Asian Rock Mechanics Symp, 13–15 Geertsma J (1973) Land subsidence above compacting oil and gas
October, Seoul, pp 100–114 reservoirs. J Petrol Tech 25:734–744
Dusseault MB, Bilak RA, Bruno MS, Rothenburg L (1996) Dis- Geertsma J, Deklerk F (1969) A rapid method of predicting width
posal of granular solid wastes in the western Canadian sedi- and extent of hydraulically induced fractures. J Petrol Tech
mentary basic by slurry injection. In: Apps J, Tsang C-F (eds) 21:1571–1581
Deep injection of disposal of hazardous industrial waste. Sci- Gerrard CM (1982) Elastic models of rock masses having one,
entific and engineering aspects. Academic Press, San Diego, two and three sets of joints. Int J Rock Mech Mining Sci Ge-
California, pp 725–742 omech Abstr 19:15–23
Dyke C (1995) How sensitive is natural fracture permeability at Ghaboussi J, Wison EL (1973) Flow of compressible fluids in po-
depth to variation of in effective stress? In: Myer L, Cook rous elastic media. Int J Numer Methods Eng 5:419–442

Hydrogeology Journal (2003) 11:7–40 DOI 10.1007/s10040-002-0241-5


38
Goodman RE (1970) Deformability of joints. In: Proc Symp De- Kilmer NH, Morrow NR, Pitman JK (1987) Pressure sensitivity of
termination of the In-situ Modulus of Deformation of Rock, low permeability sandstones. J Petrol Sci Eng 1:65–81
2–7 February, Denver, Colorado. ASTM Spec Tech Publ 477, Koutsabeloulis NC, Hope SA (1998) “Coupled” stress/fluid/ther-
pp 174–196 mal multi-phase reservoir simulation studies incorporating
Goodman RE (1974) The mechanical properties of joints. In: Proc rock mechanics. In: Proc Symp EUROCK ’98 Rock Mechan-
3rd Int Congr International Society of Rock Mechanics, 1–7 ics in Petroleum Engineering, 8–10 July 1998, Trondheim,
September 1974, Denver, Colorado. National Academy of Sci- Norway. Soc Petrol Eng SPE Pap 47392, pp 449–454
ences, Washington, DC, vol I, pp 127–140 Kranz RL, Frankel AD, Engelder T, Scholz CH (1979) The perme-
Goodman RE (1976) Methods of geological engineering in dis- ability of whole and jointed barre granite. Int J Rock Mech
continuous rock. West Publishing, New York Mining Sci Geomech Abstr 16:225–234
Gupta HK (1992) Reservoir induced earthquakes. Elsevier, New Lee C-I, Chang K-M (1995) Analysis of permeability change and
York groundwater flow around underground oil storage cavern in Ko-
Gutierrez M, Lewis RW (1998) The role of geomechanics in reser- rea. In: Fujii T (ed) Proc 8th Int Congr Rock Mechanics, 25–29
voir simulation. In: Proc Symp EUROCK ’98 Rock Mechan- September 1995, Tokyo. Balkema, Rotterdam, pp 779–782
ics in Petroleum Engineering, 8–10 July 1998, Trondheim, Lewis RW, Schrefler BA (1998) The finite element method in the
Norway. Soc Petrol Eng SPE Pap 47392, pp 439–448 static and dynamic deformation and consolidation of porous
Gutierrez M, Barton N, Makurat A (1995) Compaction and subsi- media, 2nd edn. Wiley, New York
dence in North Sea hydrocarbon fields. In: Yoshinaka R, Lewis RW, Sukirman Y (1993) Finite element modeling of three-
Kikuchi K (eds) Proc Int Worksh Rock Foundation, 30 Sep- phase flow in deforming saturated oil reservoirs. Int J Numer
tember 1995, Tokyo, Japan. Balkema, Rotterdam, pp 57–64 Anal Methods Geomech 17:577–598
Habib P (1987) The Malpasset dam failure. Eng Geol 24:331-338 Lewis RW, Pao WKS, Makurat A (2003) Fully coupled modeling
Haimson BC, Fairhurst C (1967) Initiation and extension of hy- of seabed subsidence and reservoir compaction of North Sea
draulic fractures in rocks. Soc Petrol Eng J Sept:310–318 oil fields. Hydrogeol J (in press) DOI 10.1007/s10040-002-
Hainey BW, Keck RG, Smith MB, Lynch KW, Barth JW (1997) 0239-z
On-site disposal of oilfield waste solids in Wilmington Field, Liu J, Elsworth D (1997) Three-dimensional effects of hydraulic
Long Beach Unite, CA. Soc Petrol Eng SPE Pap 47371 conductivity enhancement and desaturation around mined pan-
Hakami E (1995) Aperture distribution of rock fractures. PhD els. Int J Rock Mech Mining Sci 34:1139–1152
Thesis, Royal Institute of Technology, Sweden Liu J, Elsworth D, Brady BH (1999) Linking stress-dependent ef-
Harpalani S, Chen G (1995) Estimation of changes in fracture po- fective porosity and hydraulic conductivity fields to RMR. Int
rosity of coal with gas emission. Fuel 74:1491–1498 J Rock Mech Mining Sci 36:581–596
Harpalani S, Schraufnagel RA (1990) Shrinkage of coal matrix Londe P (1987) The Malpasset dam failure. Eng Geol 24:295–
with release of gas and its impact on permeability of coal. Fuel 329
199:551–556 Londe P, Sabarly F (1966) La distribution des perméabilités dans
Hayashi K, Abé H (1989) Evaluation of hydraulic properties of la fondation des barrages voûtés en fonction du champ de con-
the artificial subsurface system in Higashihachimantai geo- trainte. In: Proc 1st Int Congr Rock Mechanics, 25 Septem-
thermal model field. J Geotherm Res Soc Jpn 11:203–215 ber–1 October 1966, Lisbon. Lab Nac Eng Civil, Lisbon, vol
Healy JH, Rubey WW, Griggs DT, Raleigh CB (1968) The Denver II, pp 517–522
earthquakes. Science 161:1301–1310 Louis C, Maini Y (1970) Determination of in situ hydraulic pa-
Heiland J (2003) Laboratory testing of coupled hydro-mechanical rameters in jointed rock. In: Proc 2nd Int Congr Rock Me-
processes during rock deformation. Hydrogeol J (in press) chanics, 21–26 September 1970, Belgrade. Inst Dev Water Re-
DOI 10.1007/s10040-002-0236-2 sour, Belgrade, vol I, pp 235–245
Henkel H (2002) The Bjökö geothermal energy project. Norges Louis C, Dessenne J-L, Feuga B (1977) Interaction between water
Geol Undersök Bull 439:45–50 flow phenomena and the mechanical behavior of soil or rock
Hsieh PA, Bredehoeft JD (1981) A reservoir analysis of the Den- masses. In: Gudehus G (ed) Finite elements in geomechanics.
ver earthquakes: a case of induced seismicity. J Geophys Res Wiley, New York, pp 479–511
86:903–925 Makurat A, Barton N, Rad NS (1990) Joint conductivity variation
Hubbert MK, Willis DG (1957) Mechanics of hydraulic fractur- due to normal and shear deformation. In: Barton N, Stephans-
ing. J Petrol Tech 9:153–168 son O (eds) Rock joints. Balkema, Rotterdam, pp 535–540
Ito T, Zoback MD (2000) Fracture permeability and in situ stress Martin CD, Davison CC, Kozak ET (1990) Characterizing normal
to 7 km depth in the KTB scientific drillhole. Geophys Res stiffness and hydraulic conductivity of a major shear zone in
Lett 27:1045–1048 granite. In: Barton N, Stephansson O (eds) Rock joints. Balk-
Iwai K (1976) Fundamental studies of fluid flow through a single ema, Rotterdam, pp 549–556
fracture. PhD Thesis, University of California, Berkeley Meinzer OE (1928) Compressibility and elasticity of artesian
Iwano M (1995) Hydromechanical characteristics of a single rock aquifers. Econ Geol 23:263–291
joint. PhD Thesis, Massachusetts Institute of Technology Myer LR (1991) Hydromechanical and seismic properties of frac-
Jacob CE (1940) On the flow of ground water in a elastic artesian tures. In: Wittke W (ed) Proc 7th Int Congr Rock Mechanics,
aquifer. Trans Am Geophys Union 22:783–787 Aachen, Germany. Balkema, Rotterdam, pp 397–404
Jacob CE (1950) Flow of groundwater. In: Engineering hydraulics. Nagelhout ACG, Roest JPA (1997) Investigating fault slip in a
Wiley, New York, pp 321–386 model of underground gas storage facility. Int J Rock Mech
Jing L, Hudson JA (2002) Numerical methods in rock mechanics. Mining Sci Geomech Abstr 34, Pap no 212
Int J Rock Mech Mining Sci 39:409–427 Nakatsuka K (1999) Field characterization for HDR/WDR: a re-
Jones FO Jr (1975) A laboratory study of the effects of confining view. Geothermics 28:519–531
pressure on fracture flow and storage capacity in carbonate National Research Council (1996) Rock Fractures and fluid flow.
rocks. J Petrol Tech 27:21–27 National Academy Press, Washington, DC
Jones FO, Ovens WW (1980) A laboratory study of low perme- Neuzil CE (1986) Ground water flow in low-permeability environ-
ability gas sands. J Petrol Eng 32:1631–1640 ments. Water Resour Res 22:1163–1195
Jung R (1989) Hydraulic in situ investigation of an artificial frac- Neuzil CE (2003) Hydromechanical coupling in geologic process-
ture in the Falkenberg granite. Int J Rock Mech Mining Sci es. Hydrogeol J (in press) DOI 10.1007/s10040-002-0230-8>
Geomech Abstr 26:301–308 Neuzil CE, Tracy JV(1981) Flow through fractures. Water Resour
Jupe AJ, Green ASP, Wallroth (1992) Induced microseismicity and Res 17:191–199
reservoir growth at the Fjällbacka Hot Dry Rock Project, Swe- Nicholson C, Wesson RL (1992) Triggered earthquakes and deep
den. Int J Rock Mech Mining Sci Geomech Abstr 29:343–354 well activities. Pure Appl Geophys 139:561–578

Hydrogeology Journal (2003) 11:7–40 DOI 10.1007/s10040-002-0241-5


39
Nolte KG (1991) Fracturing-pressure analysis for nonideal behav- Richards HG, Parker RH, Green ASP, Jones RH, Nicholls JDM,
ior. J Petrol Tech Feb 1991:210–218 Nichol DAC, Randall MM, Stewart RC, Willis-Richards J
Noorishad J (1971) Finite element analysis of rock mass behavior (1994) The performance characteristics of the experimental
under coupled action of body forces, fluid flow, and external Hot Dry Rock Geothermal Reservoir at Rosemanowes, Corn-
loads. PhD Thesis, University of California, Berkeley wall (1985–1988). Geothermics 23:73–109
Noorishad J, Ayatollahi MS, Witherspoon PA (1982) A finite ele- Rutqvist J (1995a) Coupled stress-flow properties of rock joints
ment method for coupled stress and fluid flow analysis of frac- from hydraulic field testing. PhD Thesis, Royal Institute of
tured rocks. Int J Rock Mech Mining Sci Geomech Abstr Technology, Sweden
19:185–193 Rutqvist J (1995b) Determination of hydraulic normal stiffness of
Noorishad J, Tsang C-F, Witherspoon PA (1992) Theoretical and fractures in hard rock from hydraulic well testing. Int J Rock
field studies of coupled hydromechanical behavior of fracture Mech Mining Sci Geomech Abstr 32:513–523
rocks. 1. Development and verification of a numerical simula- Rutqvist (1995c) Hydraulic pulse testing of single fractures in po-
tor. Int J Rock Mech Mining Sci Geomech Abstr 29:401– rous and deformable rocks. Q J Eng Geol 29:181–192
409 Rutqvist J, Tsang C-F (2002) Analysis of thermal-hydrologic-me-
Nur A, Byerlee JD (1971) An exact effective stress law for elastic chanical behavior near an emplacement drift at Yucca Moun-
deformation of rock with fluids. J Geophys Res 76:6414–6419 tain. J Contaminant Hydrol (in press)
Oda M (1982) Fabric tensor for discontinuous geological materi- Rutqvist J, Tsang C-F, Ekman D, Stephansson O (1997) Evalua-
als. Soils Found 22:96–108 tion of in situ hydromechanical properties of rock fractures at
Olsson O (1992) Site characterization and validation – final re- Laxemar in Sweden. In: Lee H-K, Yang H-S, Chung S-K (eds)
port: Stripa Project. Swedish Nuclear Fuel Waste Management Proc 1st Asian Rock Mechanics Symp ARMS 97, 13–15 Octo-
Co, Stockholm, Rep TR 92-22 ber 1997, Seoul, Korea. Balkema, Rotterdam, pp 619–624
Olsson R, Barton N (2001) An improved model for hydromechan- Rutqvist J, Tsang C-F, Stephansson O (1998) Determination of
ical coupling during shearing of rock joints. Int J Rock Mech fracture storativity in hard rocks using high pressure testing.
Mining Sci 38:317–329 Water Resour Res 34:2551–2560
Osoiro JG, Chen H-Y, Teufel LW, Schaffer S (1998) A two-do- Rutqvist J, Wu Y-S, Tsang C-F, Bodvarsson G (2002) A modeling
main, 3D, fully coupled fluid flow/geomechanical simulation approach for analysis of coupled multiphase fluid flow, heat
model for reservoirs with stress sensitive mechanical and flu- transfer, and deformation in fractured porous rock. Int J Rock
id-flow properties. In: Proc Symp EUROCK ’98 Rock Me- Mech Mining Sci 39:429–442
chanics in Petroleum Engineering, 8–10 July 1998, Trond- Sandhu RS, Wilson EL (1969) Finite-element analysis of seepage
heim, Norway. Soc Petrol Eng SPE Pap 47397, pp 455–464 in elastic media. J Eng Mech Div ASCE 95:641–652
Patton FD (1966) Multiple modes of shear failure in rock. In: Proc Sasaki S (1998) Characteristics of microseismic events induced
1st Int Congr Rock Mechanics, 25 September–1 October 1966, during hydraulic fracturing experiments at the Hijiori hot dry
Lisbon. Lab Nac Eng Civil, Lisbon, vol I, pp 509–513 rock geothermal energy site, Yamagata, Japan. Tectonophysics
Perkins TK, Kern LR (1961) Widths of hydraulic fractures. J Pet- 289:171–188
rol Tech Sept 1961:937–949 Scheidegger AE (1974) The physics of flow through porous me-
Pine RJ, Batchelor AS (1984) Downward migration of shearing in dia. University of Toronto Press, Toronto
jointed rock during hydraulic injections. Int J Rock Mech Schmidt JH, Friar WL, Bill ML, Cooper GD (1999) Large scale
Mining Sci 21:249–263 injection of North Slope drilling cuttings. Soc Petrol Eng SPE
Pratt WE, Johnson DW (1926) Local subsidence of the Goose Pap 52738
Creek oil field. J Geol 34:577–590 Scholz CH (1990) The mechanics of earthquakes and faulting.
Pratt HR, Swolfs HS, Brace WF, Black AD, Handin JW (1977) Cambridge University Press, New York
Elastic and transport properties of an in situ jointed granite. Int Segall P (1989) Earthquakes triggered by fluid extraction. Geolo-
J Rock Mech Mining Sci Geomech Abstr 14:35–45 gy 17:942–946
Pyrak-Nolte L, Myer LJ, Cook NGW, Witherspoon PA (1987) Hy- Segall P, Fitzgerald SD (1998) A note on induced stress changes
draulic and mechanical properties of natural fractures in low in hydrocarbon and geothermal reservoirs. Tectonophysics
permeability rock. In: Herget G, Vongpaisal S (eds) Proc 6th 289:17–128
Congr International Society of Rock Mechanics, Montreal. Simpson DW, Leith WS, Scholz CH (1988) Two types of reser-
Balkema, Rotterdam, vol 1, pp 225–231 voir-induced seismicity. Bull Seismol Soc Am 78:2025–2040
Raven KG, Gale JE (1985) Water flow in a natural rock fracture as Snow DT (1965) A parallel plate model of fractured permeable
a function of stress and sample size. Int J Rock Mech Mining media. PhD Thesis, University of California, Berkeley
Sci Geomech Abstr 22:251–261 Snow DT (1968) Rock fracture spacings, openings, and porosities.
Read RS, Chandler NA (1999) Excavation damage and stability J Soil Mech Found Div ASCE 73-91
studies at the URL. In: Amadei B, Kranz RL, Scott GA, Snow DT (1969) Anisotropic permeability of fractured media.
Smeallie PH (eds) Proc 37th US Rock Mechanics Symp, 6–9 Water Resour Res 5:1273–1289
June 1999, Vail, Colorado. Balkema, Rotterdam, vol 2, pp Souley M, Homand F, Pepa S, Hoxha D (2001) Damage-induced
861–868 permeability changes in granite: a case example at the URL in
Reed AC, Mathews JL, Bruno MS, Olmstead SE (2001) Chevron Canada. Int J Rock Mech Mining Sci 38:297–310
safely disposes one million barrels of NORM in Louisiana Stephansson O, Jing L, Tsang C F (eds) (1996) Coupled thermo-
through slurry fracture injection. Soc Petrol Eng SPE Pap hydro-mechanical processes of fractured media. Elsevier,
47371 Amsterdam
Rejeb A (1996) Mathematical simulation of coupled THM pro- Stietel A, Millard A, Treille E, Vuillod E, Thoravel A, Ababou R
cesses of Fanay-Augéres field test by distinct element and dis- (1996) Continuum representation of coupled hydromechanical
crete finite element methods. In: Stephansson O, Jing L, Tsang processes of fractured media. In: Stephansson O, Jing L, Tsang
C-F (eds) Coupled thermo-hydro-mechanical processes of C-F (eds) Coupled thermo-hydro-mechanical processes of
fractured media. Elsevier, Amsterdam, pp 341–369 fractured media. Elsevier, Amsterdam, pp 135–164
Rhén I, Gustafson G, Stanford R, Wikberg P (1997) Äspö HRL – Stowel JFW, Laubch SE, Olson JE (2001) Effect of modern state
geoscientific evaluation 1997-5: models based on site charac- of stress on flow-controlling fractures: a misleading paradigm
terization 1986-1995. Swedish Nuclear Fuel Waste Manage- in need of revision. In: Elsworth D, Tinucci JP, Heasley KA
ment Co (SKB) Tech Rep 97-06 (eds) Proc 38th US Rock Mechanics Symp, 7–10 July 2001,
Rice JR, Cleary MP (1976) Some basic stress diffusion solutions Washington, DC. Balkema, Rotterdam, vol I, pp 691–697
of fluid saturation elastic porous media with compressible Sugihara K, Matsu H, Sato T (1999) In situ experiments on rock
constituents. Rev Geophys Space Phys 14:227–241 stress conditions and excavation disturbance in JNC’s geosci-

Hydrogeology Journal (2003) 11:7–40 DOI 10.1007/s10040-002-0241-5


40
entific research program in Japan. In: Saeb S, Francke C (eds) Wang HF (2000) Theory of linear poroelasticity. Princeton Uni-
Proc Int Worksh Rock Mechanics of Nuclear Waste Reposito- versity Press, 287 p
ries, 5–6 June, Veil, Colorado, pp 159–183 Wang JSY, Elsworth D (1999) Permeability changes induced by
Sutton S, Soley R, Jeffery C, Chackrabarty C, McLeod R (1996) excavation in fractured tuff. In: Amadei B, Kranz RL, Scott
Cross hole hydraulic testing between deep boreholes 2 and 4 GA, Smeallie PH (eds) Proc 37th US Rock Mechanics Symp,
at Sellafield. European Commission, Brussels, Nuclear Sci 6–9 June 1999, Vail, Colorado. Balkema, Rotterdam, vol 2, pp
Technol Topical Rep EUR 16967 751–758
Swan G (1983) Determination of stiffness and other properties from Wang J-A, Park HD (2002) Fluid permeability of sedimentary
roughness measurements. Rock Mech Rock Eng 16:19–38 rocks in a complete stress-strain process. Eng Geol 63:291–
Takahashi M, Koide H, Xue Z-Q (1995) Correlation between per- 300
meability and deformational characteristics on the Shirahama Wang J, Cook P, Trautz R, Flexser S, Hu Q, Salve R, Hudson D,
sandstone and Inada granite. In: Fujii T (ed) Proc 8th Int Conrad M, Tsang Y, Williams K, Soll W, Turin J (2001) In-situ
Congr Rock Mechanics, 25–29 September 1995, Tokyo. field testing of processes ANL-NBS-HS-000005 REV01. Be-
Balkema, Rotterdam, pp 417–420 chtel SAIC Co, Las Vegas, Nevada
Talbot CJ, Sirat M (2001) Stress control of hydraulic conductivity Wei Z-Q, Hudson JA (1988) Permeability of jointed rock masses.
in fractured-saturated Swedish bedrock. Eng Geol 61:145–153 In: Romana M (ed) Rock mechanics and power plants. Balk-
Terzaghi K (1923) Die Berechnung der Durchlässigkeitziffer des ema, Rotterdam, pp 613–626
Tones aus dem Verlauf der hydrodynamischen Spannungs- Withers RJ, Perkins TK, Keck RG (1996) A field demonstration
erscheinungen. Akad Wissensch Wien Sitzungsber Math- of hydraulic fracturing for solid waste disposal. In: Apps J,
naturwissensch Klasse IIa 142(3/4), pp 125–138 Tsang C-F (eds) Deep injection of disposal of hazardous in-
Teufel LW, Farrel HE (1995) Interrelationship between in situ dustrial waste. Scientific and engineering aspects. Academic
stress, natural fractures, and reservoir permeability anisotropy Press, San Diego, California, pp 705–724
– A case study of the Ekofisk Field, North Sea. In: Myer L, Witherspoon PA (2000) The Stripa project. Int J Rock Mech Min-
Cook NGW, Goodman D, Tsang C-F (eds) Fracture and joint- ing Sci 37:385–396
ed rock masses. Balkema, Rotterdam, pp 573–578 Witherspoon PA, Amick CH, Gale JE, Iwai K (1979) Observations
Theis CV (1935) The relationship between the lowering of the of a potential size effect in experimental determination of the hy-
piesometric surface and rate and duration of the discharge of a draulic properties of fractures. Water Resour Res 15:1142–1146
well using groundwater storage. Trans Am Geophys Union Witherspoon PA, Wang JSY, Iwai K, Gale JE (1980) Validity of
2:519–524 cubic law for fluid flow in a deformable rock fracture. Water
Thury M, Bossart P (1999) The Mont Terri rock laboratory, a new Resour Res 16:1016–1024
international research project in a Mesozoic shale formation, Wladis D, Jönsson P, Wallroth T (1997) Regional characterization
in Switzerland. Eng Geol 52:347–359 of hydraulic properties or rock using well test data. Swedish
Tsang C-F (1987) Introduction to coupled processes. In: Tsang Nuclear Fuel Waste Management Co (SKB) Tech Rep 97-29
C-F (ed) Coupled processes associated with nuclear waste re- Xiao HT, Xu HY (2000) In situ permeability measurements to es-
positories. Academic Press, Orlando, pp 1–6 tablish the influence of slice mining on floor rock. Int J Rock
Tsang YW, Witherspoon PA (1981) Hydromechanical behavior of Mech Mining Sci 37:855–860
a deformable rock fracture subject to normal stress. J Geophys Yoshinaka R, Yoshida J, Arai H, Arisaka S (1993) Scale effects on
Res 86:9287–9298 shear and deformability of rock joints. In: Pinto da Chunha A
Van-Golf Racht TD (1982) Fundamentals of fractured reservoir (ed) Proc Symp Scale Effects in Rock Masses 93, 25 June
engineering. Elsevier, Amsterdam 1993, Lisbon, Portugal. Balkema, Rotterdam, pp 143–149
Verruijt (1969) Elastic storage of aquifers. In: DeWiest RJM (ed) Zhao J, Brown ET (1992) Hydro-thermo-mechanical properties of
Flow through porous media. Academic Press, New York, pp joints in the Crarnmenellis granite. Q J Eng Geol 25:279–290
331–376 Zienkiewics OC, Pande GN (1977) Time-dependent multilaminate
Walsh JB (1965) The effect of cracks on the compressibility of model of rocks-a numerical study of deformation and failure
rock. J Geophys Res 70:381–389 of rock masses. Numer Anal Methods Geomech 1:219–247
Walsh JB (1981) Effects of pore pressure and confining pressure Zoback MD, Byerlee JD (1975) The effect of microcrack dilatan-
on fracture permeability. Int J Rock Mech Mining Sci Geo- cy on the permeability of westerly granite. J Geophys Res
mech Abstr 18:429–435 80:752–755

Hydrogeology Journal (2003) 11:7–40 DOI 10.1007/s10040-002-0241-5

You might also like