You are on page 1of 5

Anisotropy and beyond: Geologic perspectives

on geophysical prospecting for natural fractures


RANDALL MARRETT, STEPHEN E. LAUBACH, and JON E. OLSON, The University of Texas at Austin, USA

Recent geologic research on natural


fractures challenges assumptions fre-
quently made by geophysicists. Open
fractures are not necessarily oriented
parallel to the maximum horizontal
stress, and fractures do not necessar-
ily close when the fluid pressure
within them is reduced. Even in the
most mechanically favorable envi-
ronment, precipitated cements can
prop fractures open or seal fractures
of any orientation. Fracture sets typi-
cally show dispersion in strike, and
multiple sets of open fractures can
coexist. More importantly, fractures
comprise populations that commonly
range over orders of magnitude in
aperture and length and that occur in
nonuniform clusters. Instead of iso-
lated, regularly spaced, large, equally Figure 1. Rose diagrams of maximum compressive stress (SHmax, upper row, blue) and open frac-
compliant fractures, the Earth presents ture strike (lower row, red) for study areas in Texas and Wyoming. (a) SHmax and (b) fracture
complex fractal clustering of fractures strike, East Texas, average ENE fracture strike is similar to SHmax trends, but open fractures have
having a wide range of sizes and vari- a spread of 130°; (c) SHmax and (d) fracture strike, four West Texas wells. (e) SHmax and (f) frac-
able compliance dictated by natural ture strike, western Green River Basin; (g) SHmax and (h) fracture strike southern Powder River
Basin, horizontal well image log data (after Laubach et al., 2004).
cements in the fractures and the rock
mass. Going beyond anisotropy to
document these essential fracture attributes in the interwell discarded. Among seismically important attributes, frac-
space is a key challenge for geophysicists. ture openness can determine the extent of mechanical cou-
pling across fractures, fracture orientation may control the
Motivation for seismic characterization of fractures. direction of velocity anisotropy, and fracture sizes and abun-
Fractures are notoriously challenging to study in the sub- dance can control the magnitude of seismic signature
surface, where they can have dramatic effects on fluid flow. (Marrett, 1997). Open fracture length may be related to the
Direct study of subsurface fractures, using logs or core from magnitude of velocity anisotropy, and could also affect dif-
boreholes, is hampered by several sampling problems. fraction patterns.
Fractures commonly are nearly vertical, so vertical wellbores Here we outline some recent core, outcrop, and model-
are unlikely to intersect many fractures. Sampling proba- based findings on natural fracture populations that suggest
bility also is poor because the spacing between conductive that subsurface fracture patterns are highly heterogeneous
fractures typically is large in comparison with borehole on a range of scales and in a variety of ways. Thus, there
diameters. Additionally, heterogeneity of fractures com- are first-order implications for the expected seismic response
monly occurs on length scales that are a fraction of well spac- of fractures that are typically not accounted for in current
ing, so important lateral changes in fracture attributes can geophysical approaches.
remain undiscovered. Heterogeneity can also be manifest
as significant variation of fractures from one layer to the next. Fracture geology. Outcrops commonly contain sets of large,
Because of the limitations in wellbore-based observations, more-or-less planar, mostly evenly spaced, barren (no min-
seismic detection and characterization of fractures poten- eral fill), opening-mode fractures. Known since the early
tially valuable tools for subsurface prospecting. days of geology (Pollard and Aydin, 1988; Cosgrove and
The presence of fracture signal in seismic reflection data Engelder, 2004), these joints undoubtedly influence common
is detected by studying anisotropic behavior of seismic conceptions of fractures. Yet inherently sparse information
velocities and amplitudes (Queen and Rizer, 1990; about subsurface fractures suggests they differ from joints
Schoenberg and Sayers, 1995). However, anisotropic attrib- in some important ways. Cores and well logs, particularly
utes are challenging to measure, and their interpretation may image logs, provide direct samples of subsurface fractures,
rely on assumptions about fracture orientations, shapes, but fracture sampling using wells is notoriously incomplete.
openness, sizes, and spatial arrangement that are challeng- For example, prior to the advent of horizontal wells, sub-
ing to verify independently. Based on the geologic literature, surface fracture size and spatial distributions were mostly
geophysicists frequently assume that fractures are oriented conjectural. Consequently, geologists have long utilized out-
parallel to the maximum horizontal stress, that fractures crops containing exhumed fractures as a proxy source of
close when the fluid pressure within the fractures is reduced, information on subsurface fractures. Outcrops, however,
and that there is a single set of parallel, evenly spaced, open are subject to uplift and weathering-induced fractures that
fractures. Many of these assumptions need to be revised or are nonrepresentative of the subsurface.

1106 THE LEADING EDGE SEPTEMBER 2007


environment, reactive fracture sur-
faces are susceptible to accumulating
cement deposits. Core studies show
(Figure 2) and modeling studies pre-
dict that in otherwise open fractures,
isolated deposits of cement are com-
mon. Laboratory tests show that par-
tial mineral fill can make fracture
aperture insensitive to static changes
in effective stress. The prevalence of
strong, spatially isolated mineral
bridges that resist fracture closure is
not widely appreciated.
Cement precipitation in the host
rock during or immediately after frac-
ture formation is another mechanism
that can increase the resistance of nat-
ural fractures to closing. This process,
which is probably widespread, essen-
Figure 2. Subsurface and outcrop data on open and sealed fractures. (a) Rocks may have many tially freezes fractures open. Lander et
sets of open fractures or none (modified from Laubach and Ward, 2006). (b) Even in the most al. (2002) showed that, without seal-
mechanically favorable environment, precipitated cements can seal any orientation fracture (mod-
ified from Laubach, 2003).
ing fractures, as much as 20% whole
rock volume of quartz cement can pre-
cipitate in a rock’s pore space after
fractures form. Thus, fractures do not
necessarily close when the fluid pres-
sure within them is reduced, even if
they lack cement bridges (Figure 2). A
simple calculation demonstrates how
host rock stiffening can affect fracture
aperture compressibility (Olson et al.,
2007). Assuming linear elasticity and
plane strain, the compliance of frac-
ture aperture (opening per unit dri-
ving stress) for a fracture of a given
total length, L, can be written as

where v is Poisson’s ratio, Δσ is the dri-


ving stress, and E is Young’s modulus.
If after initial fracture opening (Figure
3, A to B), diagenetic cementation in
the host rock increased Young’s mod-
ulus by a factor of 5, for instance, frac-
Figure 3. Plot of fracture opening (aperture) versus driving stress. When driving stress is zero ture compliance would be reduced by
(A), the fracture is closed. Increasing driving stress by increased pore pressure or reducing crack- that same factor and it would take five
normal compression causes the crack to open (A to B). If diagenesis occurs at time B, crack clos- times the stress change to close the
ing takes more stress because of host rock stiffening effect (after Olson et al., 2007). fracture (Figure 3, B to C) as it took to
open it.
We do not believe that a consensus exists in the geologic Of course, fracture aperture can be closed without any
community about the general attributes of subsurface frac- kinematic aperture change by the diagenetic process of
tures. Nevertheless, geologic observations do test the valid- cement precipitation, which is insensitive to fracture com-
ity of assumptions about fractures that geophysicists accept pliance or fracture-stress orientation relationships. Empirical
as generally true. For example, Figure 1 shows that open evidence shows that heterogeneous patterns of infilling of
fractures are not necessarily oriented parallel to current-day large fractures by cements is a common occurrence. Core
maximum horizontal stress. In these examples, stress ori- demonstrates that sealed and open fractures having identi-
entation data from reliable measurements are consistent cal strike can be interspersed over vertical distances that
with regional stress maps. Yet observations from extensive range from a few meters or less to decimeters and over lat-
core collections show that open fractures can have arbitrary eral distances of m to km. Production and core data demon-
strike relative to SHmax (maximum horizontal compression). strate that it is the degree of cement fill in fractures rather
Production data show that these fractures also govern fluid than fracture orientation that limits fluid flow. Flow occurs
flow. only where fractures are not sealed with cement. In the
Chemical processes can account for resistance of frac- absence of reliable measurements of both open fracture
tures to closure. Fractures at depth in sedimentary basins strike and SHmax, these features should not be presumed to
are exposed to hot (>80°C), mineral-laden water. In this be parallel. Even in the most mechanically favorable envi-

SEPTEMBER 2007 THE LEADING EDGE 1107


velocity is decreased in all azimuths
and the net effect can be less velocity
anisotropy than from either of the frac-
ture sets alone. The case of nonorthog-
onal fracture sets might be diagnosed
if fractures have different normal and
shear compliances. That rock may con-
tain fracture sets having differing ori-
entations has long been appreciated,
but recent studies show that rocks may
Figure 4. Fracture patterns generated using randomly oriented starter cracks and a subcritical have many sets of fractures that are
propagation model for increasingly anisotropic prefracture strain states (from a to c) followed by open concurrently. And even given
isotropic biaxial extension. only one set of fractures (formed by
one deformation event), the assump-
tion that all individuals within that
set are parallel may not always be rea-
sonable. Some core data sets show
substantial dispersion in strike for
nominally coeval fractures that cannot
be ascribed to core orientation errors.
Locally, fracture sets in outcrop also
show wide strike dispersion. This vari-
able fracture orientation in some cases
can be attributed to perturbations of
stress fields caused by the presence of
large faults (Rawnsley et al., 1992) or
to nearly isotropic loading conditions
that may promote random or orthog-
onal fracture patterns to develop dur-
ing a single loading event. Geo-
mechanical modeling results (Figure
4) show how differences in strain
anisotropy for a given deformation
event can significantly affect the ori-
entation of natural, opening-mode
fracture sets.
Beyond orientation and fracture
cementation, another crucial fracture
attribute is size. Subsurface fractures
in a single set commonly have aper-
tures and lengths that range over
orders of magnitude in size, with small
fractures far more abundant than large
fractures (Marrett et al., 1999). In such
cases, fractures of different size share
orientations, kinematics, and timing
Figure 5. Abundance of fractures as a function of fracture size. Fractures commonly have aper- relative to diagenetic events, so they
tures and lengths that range over orders of magnitude in size and follow power-law frequency are most simply interpreted as differ-
distributions. Fractures from the Marble Falls Limestone, Texas, follow a common power-law ent size fractions of a single fracture
distribution of apertures, even though one subset of the data was collected at outcrop scale using set with a common genesis. One con-
a hand lens whereas other data were collected petrographically in a microscope (Marrett et al., sequence of the broad spectrum of
1999). The power-law regression to data from microfractures is extrapolated for comparison with fracture sizes is that fracture intensity,
data collected along a 60-m line of observation at outcrop scale.
the abundance of fractures in space, is
inherently scale-dependent, and varies
ronment, precipitated cements can seal any orientation frac- as a function of minimum observed fracture size. As the
ture (Laubach, 2003). Recognition of, and distinction among, threshold for counting fractures is decreased, fracture inten-
uncemented fractures, partially cemented fractures con- sity increases rapidly. Another consequence is that average
taining cement bridges, and completely cemented fractures fracture size is poorly defined and depends sensitively to
is essential in effective reservoir characterization and is a detection threshold.
challenge for seismic methods. This might be feasible if min- These aspects of fracture size distributions are prob-
eral fill within a fracture (e.g., mineral bridges) affects shear lematic, because models of seismic velocity anisotropy vari-
compliance across the fracture differently than normal com- ation with fracture characteristics (e.g., Thomsen, 1993) are
pliance (Sayers and Dean, 2001). defined in terms of average fracture size and fracture inten-
Given that open fractures exist in the subsurface, their sity. To the extent that fracture size distributions follow sys-
orientation is often surmised from velocity anisotropy. tematic patterns, theory can be modified to account for
However, multiple sets of fractures are common, and if a realistic fracture parameters (e.g., Marrett, 1997). For exam-
second set forms at a high angle with the first set, then ple, fracture apertures commonly follow power-law distri-

1108 THE LEADING EDGE SEPTEMBER 2007


butions (Figure 5), which can be char-
acterized by two parameters that
replace fracture intensity and average
fracture size in formulations of the
magnitude of velocity anisotropy.
Models of seismic velocity
anisotropy typically presume a statis-
tically uniform arrangement of frac-
tures in space (e.g., Thomsen, 1995),
but subsurface fractures commonly
are clustered (Figure 6a). Moreover,
the largest fractures tend to occur in
clusters. For example, Figure 6b shows
the autocorrelation function (Davis,
2002) using logarithmically graduated
lags for the data shown in Figure 6a.
Positive autocorrelation characterizes
the fractures for almost all lags less
than a few meters, indicating the frac-
tures occur in meter-scale clusters.
Negative autocorrelation dominates
lags of 5–15 m, and a spike of positive
autocorrelation occurs for lags of
15–25 m. This pattern of autocorrela-
tion suggests that ~15 m wide
domains of unusually low fracture
intensity lie between fracture clusters,
the centers of which are spaced about
20 m apart.
Fracture clustering may affect seis-
mic response in at least two ways.
First, because the largest fractures tend
to be clustered, the probability for
them to be connected is much higher
that it would otherwise be. Mechanical
connectivity among fractures should
magnify compliance and enhance
velocity anisotropy compared with
isolated fractures of the same size and
abundance. Second, velocity ani-
sotropy may be heterogeneous on
length scales that are long compared
with individual fractures but short
Figure 6. Arrangement of fractures in space. Instead of regularly spaced fractures of comparable compared with seismic wavelengths.
size, complex clustering of fractures having a wide range of sizes is common. (a) Fractures in the
Marble Falls Limestone, Texas, are arranged in clusters, where the largest-aperture fractures tend The extent to which such heterogene-
to lie. (b) Autocorrelation function (Davis, 2002) of indicator series (value of 1 within fractures, ity can be teased from the seismic sig-
value of 0 between fractures) documents concentration of fractures inside of clusters that are ~5 nal remains to be addressed.
m wide and spaced ~20 m apart. This is the same data set as used in Figure 5 to represent out-
crop-scale observations. Discussion. The well-known chal-
lenges of obtaining meaningful geo-

1110 THE LEADING EDGE SEPTEMBER 2007


logic samples of fractures using boreholes (see, for exam- Structural Geology, 2004). “Diagenesis in porosity evolution of
ple, Narr, 1996; Mauldon and Mauldon, 2005) underline the opening-mode fractures, Middle Triassic to Lower Jurassic la
need for seismic information on hard-to-measure attributes. Boca Formation, NE Mexico” by Laubach and Ward
Unfortunately, most seismic data analysis techniques cur- (Tectonophysics, 2006). “Permeability, porosity, and shear-wave
rently practiced are based on equivalent or effective media the- anisotropy from scaling of open-fracture populations” by Marrett
ories which assume unrealistic fracture geometries and (RMAG Fractured Reservoirs: Characterization and Modeling
distributions. Nonetheless, seismic methods offer the hope of Guidebook, 1997). “Extent of power-law scaling for natural frac-
measuring key fracture attributes between boreholes and on tures in rock” by Marrett et al. (Geology, 1999). “Fracture sampling
length scales that are most meaningful to fluid-flow simula- on a cylinder: From scanlines to boreholes to tunnels” by Mauldon
tion. Appreciating the complexity of fracture systems as illus- and Mauldon (Rock Mechanics and Rock Engineering, 1997).
trated by outcrop and subsurface geologic investigations will “Estimating average fracture spacing in subsurface rock” by Narr
be essential for creating new seismic analysis tools and (AAPG Bulletin, 1996). Combining Diagenesis and Mechanics in
improved processing and analysis of seismic data. Natural Fracture Network Characterization by Olson et al. (Geological
Society of London Special Publication 270, 2007). “Progress in
Suggested reading. The Initiation, Propagation, and Arrest of Joints understanding jointing over the past century” by Pollard and
and other Fractures by Cosgrove and Engelder (Geological Society Aydin (GSA Bulletin, 1988). “An integrated study of seismic
of London Special Publication 231, 2004). Statistics and Data anisotropy and the natural fracture system at the Conoco bore-
Analysis in Geology by Davis (Wiley, 2002). Predicting and hole test facility, Kay County, Oklahoma” by Queen and Rizer
Characterizing Fractures in Dolostone Reservoirs: Using the Link (Journal of Geophysical Research, 1990). “Joint development in per-
between Diagenesis and Fracturing” by Gale et al. (Geological Society turbed stress fields near faults” by Rawnsley et al. (Journal of
of London Special Publication 235, 2004). “Feasibility of seismic Structural Geology, 1992). “Azimuth-dependent AVO in reservoirs
characterization of multiple fracture sets” by Grechka and containing nonorthogonal fracture sets” by Sayers and Dean
Tsvankin (GEOPHYSICS, 2002). “Lithologic and structural controls (Geophysical Prospecting, 2001). “Seismic anisotropy of fractured
on natural fracture distribution and behavior within the Lisburne rock” by Schoenberg and Sayers (GEOPHYSICS, 1995). “Weak elas-
Group, northeastern Brooks Range and North Slope subsurface, tic anisotropy” by Thomsen (GEOPHYSICS, 1993). “Elastic anisotropy
Alaska” by Hanks et al. (AAPG Bulletin, 1997). “Interaction due to aligned cracks in porous rock” by Thomsen (Geophysical
between quartz cementation and fracturing in sandstone” by Prospecting, 1995). TLE
Lander et al. (AAPG Annual Convention, 2002). “Practical
approaches to identifying sealed and open fractures” by Laubach Acknowledgments: We gratefully acknowledge comments from S. Fomel and
(AAPG Bulletin, 2003). “Are open fractures necessarily aligned support by the U.S. Department of Energy Office of Basic Energy Sciences
with maximum horizontal stress?” by Laubach et al. (Earth and and the Fracture Research and Application Consortium.
Planetary Science Letters, 2004). “Coevolution of crack-seal texture
and fracture porosity in sedimentary rocks: Cathodoluminescence Corresponding author: jolson@mail.utexas.edu
observations of regional fractures” by Laubach et al. (Journal of

SEPTEMBER 2007 THE LEADING EDGE 1111

You might also like