You are on page 1of 12

SELF-ORGANIZATION OF NATURAL MODE-I FRACTURE

APERTURES INTO POWER-LAW DISTRIBUTIONS

Julia Gale
Bureau of Economic Geology, Jackson School of Geosciences, The University of Texas at Austin, Austin, TX 78713

ABSTRACT: Opening-mode fractures in siliciclastic and carbonate host rocks from different geologic settings consistently show
power-law aperture-size distributions when measured along one-dimensional scanlines. Exponents (slopes) and pre-exponential
coefficients (intensities) of the power laws, were compared in order to constrain the range of fracture intensities of various aperture
sizes over the observation scales. The coefficients range over two orders of magnitude, from 0.1 to 62.4, and the exponents range
from -0.45 to -1.36. On a compilation plot of power laws a broad wedge-shaped envelope defines the natural range of frequencies
of fractures for aperture sizes from approximately 1000 to 0.01 mm. There is an inverse correlation between the coefficient and
exponent for most of the data sets. Power laws having steep slopes tend to have low intensities, whereas power laws having
shallow slopes tend to have higher intensities. There are, however, a few exceptions to this correlation. The reasons for the
correlation and the departures from it are considered in the light of geologic and mechanical properties at the time of fracture
formation for each case. The subcritical crack index, mechanical layer thickness, total strain and strain rate are all likely controls
over the relative proportion of narrow to wide fractures in a population.

which is the slope of the curves and is negative. The


1. INTRODUCTION exponent reflects the relative proportions of wide
and narrow fractures in the population. A small
Empirical studies have shown that the aperture sizes exponent (shallow slope) indicates a fracture
of opening-mode fractures within a single set are population with few narrow fractures relative to
self-organized into power-law distributions [1-3]. wide fractures, whereas a large exponent (steep
Power-law aperture-size distributions ranging over slope) indicates a population with a large number of
five orders of magnitude have been reported in narrow fractures relative to wide fractures. The
sandstone and limestone [1] and over three orders of values of the exponent and coefficient are different
magnitude in dolomites [4]. The range over which for different fracture sets but the range and reasons
power-law distributions apply is typically that for the variation have not been addressed
observed at the field outcrop scale down to previously.
microfractures; on the order of 1 m to 1 μm. Small
and large opening-mode fractures that have the The aim of this contribution is to present a
same orientation may be different size fractions of compilation of opening-mode fracture aperture-size
the same fracture set. data sets so that the aperture-size distributions from
different rocks may be compared and the range of
Power-law descriptions of fracture aperture intensities of naturally occurring fractures of
populations take the form: F = ab-c, where F is different aperture sizes may be constrained. The
cumulative fracture frequency, a is the coefficient, b size-range of fractures studied in each set includes
is the fracture aperture, and c is the exponent of the small sealed fractures and larger open fractures.
power-law relationship. At any given aperture size, Possible geologic and mechanical reasons for
the intensity of fractures is defined by the pre- variation in the aperture-size distributions are then
exponential coefficient, which governs the position considered in the light of properties in effect at the
of the curves on the ordinate, and the exponent, time of fracturing.
2. GEOLOGIC SETTING OF FRACTURE SETS continuous for about 300 m, normal to the ENE
fracture trend, but the total outcrop height is limited
The fracture sets examined in this paper are from a
to 2 m. The outcrop is approximately 10 m wide
wide variety of geologic settings. For all data sets a
across the creek bed, parallel to the fracture trend.
brief description of the geology is given but in the
Fracture-aperture and spacing-size distributions
interest of brevity, geological maps and sections are
were obtained using a scanline in the lower chalk
not included. The main variables for the fracture
layer [5]. Mechanical thickness of the fractured
data sets are compared (Table 1). Some fracture sets
layer is difficult to establish because the lower part
occur in outcrops, others are from the subsurface
of the layer is not exposed. The upper boundary to
cores and there is a mix of siliciclastic and
the layer is also problematic because some fractures
carbonate host rock lithologies. Where data have
truncate at the marl layer, whereas others propagate
been extracted from previously published work,
through the marl into chalk above [6]. Typically,
references for these studies are indicated (Table 1).
marl-chalk cycles near the top of the Austin Chalk
are of the order of 1 m thick [7] and this would
2.1. Austin Chalk, Grove Creek
represent a minimum mechanical thickness if
The Grove Creek outcrop is near Waxahachie, north
fractures were truncated at both bounding marl
Central Texas and is stratigraphically at the top of
layers. Otherwise, mechanical thickness could be
the Upper Austin Chalk, just below the overlying
many 10s of meters.
Ozan Formation. The outcrop comprises two chalk
horizons separated by a marl layer. The outcrop is
Table 1. Scanline and power-law parameters for all compared data sets.

Well/outcrop Aperture size Strain


Formation Lithology name Power Law Scanline parameters (mm) %
no. length
coefficient exponent fractures (mm) max min
Carbonates
Austin Chalk
[5] chalk Grove Creek 0.1052 -0.558 136 246160 100.0 0.05 0.15
Cupido dolostone Las Palmas 11 11.99 -1.055 575 2487.14 17.50 0.075 7.62
Cupido dolostone Las Palmas 9 32.1 -0.978 221 3533.29 15.87 0.05 12.31
Cupido dolostone La Escalera 3 20.11 -0.886 150 2444.92 55.00 0.50 12.80
Cupido dolostone La Escalera 6 31.63 -1.078 206 5381.85 104.0 0.50 13.16
Pedernales
Marble Falls limestone Falls 0.86 -1.19 830 55489.6 18.00 0.0049 0.47
Clear Fork Apache
[13] dolostone Canyon 2.47 -0.891 98 5484.3 10.00 0.08 0.55
Knox dolostone Well X 0.43 -1.36 31 5.955 0.013 0.0003 1.25
Ellenburger dolostone Barnhart field 30 -0.777 60 14.099 0.179 0.0005 6.14

Siliciclastics
Deerlodge _
Weber [17] sandstone North 0.45 -0.68 - - ~1000 ~0.1
Blakeney
Ozona [1] sandstone Kruger 19.3 -0.769 235 18.609 0.360 0.0002 4.77
Homer L.
Pottsville[19] sandstone Jenkins 62.38 -0.446 27 22.5 0.153 0.0004 2.53
San Juan
Dakota sandstone basin 7259 ft 0.199 -1.310 124 49.59 0.011 0.0002 0.46
San Juan
Dakota sandstone basin 7181 ft 0.309 -1.217 86 47.01 0.381 0.0003 1.18
Piceance
Cozzette sandstone basin, SHCT 7.67 -0.813 56 16.45 0.627 0.0003 5.13
2.2. Cupido Formation, NE Mexico microcrystalline calcite fill, which seals narrow
Cupido Formation shallow water platform fractures (<1 mm) and lines and bridges wider
carbonates [8] are exposed in a series of Laramide- fractures. A later dark-colored sparry calcite fills the
age, isoclinal folds with wavelengths on the order of remaining pore space in the wide fractures. The
kilometers in the Sierra Madre Oriental to the south mechanical thickness of layers containing the
and west of Monterrey, NE Mexico [9]. Data were fractures is not known, but exceeds the height of the
collected from an exposed section of Cupido obvious beds exposed at the falls, which are
Formation in the forelimb of the San Blas anticline approximately 1 to 3 meters thick. and are bounded
in Cañón Boquilla Corral de Palmas and in Cañón by stylolites that cut the filled fractures. The age of
La Escalera. Fractures are preferentially developed the fractures is uncertain; they may have formed
in dolomitized units that occur near the top of during the late Paleozoic Ouachita-Marathon-
depositional cycles [10]. These fractures also occur Appalachian orogeny.
in displaced blocks in early, pre-folding, solution-
collapse breccias, and are cut by bedding-parallel, 2.4. Clear Fork Group dolostones, Apache
compaction-related stylolites and therefore pre-date Canyon, Texas
the Laramide folding event [10]. The fractures Leonardian Clear Fork Group dolostones are well
developed very early during the diagenetic exposed in the area of Apache Canyon in the Sierra
sequence, when adjacent limestones were not fully Diablo in Hudspeth and Culberson Counties, Texas.
lithified [10]. Mechanical layer thickness for them The Sierra Diablo is situated on the western edge of
was consistently small (less than 1 m in all cases). the Delaware Basin along a major structural high.
Fracture sets in the different beds are broadly Fractures are steeply inclined in flat-lying beds and
synchronous. They consist of three most fractures have height and length dimensions of
contemporaneous subsets which have been analyzed centimeters. Both open and sealed fractures are
together: dominant opening-mode fractures normal present. Cements in fractures include dolomite and
to bedding that bisect hybrid extension–shear calcite. At least some of the calcite in appears to be
fractures. a near-surface, late pedogenic deposit and thus
Scanlines normal to the dominant fracture set were unrelated to fracture attributes to be expected in the
constructed along single mechanical units. Palmas subsurface [13,14].
11 comprises two adjacent beds, which when
2.5. Knox Group dolostones, Mississippi
treated together range from 21 to 30 cm thick. Core samples of Lower Ordovician Knox
Palmas 9 and the data from Cañón La Escalera are dolostones from approximately 14,500 ft were
from individual beds. Fractures are lined and examined using SEM-based cathodoluminescence.
bridged with synkinematic dolomite cement, and The host rocks are dolomitized carbonate
remaining porosity has been mostly filled by mudstones from a shallow water platform
postkinematic calcite [11]. A very small amount of depositional environment. Local dessication and
porosity is preserved in a few of the largest collapse structures are present. Two fracture sets
fractures. were identified with synkinematic dolomite and
later calcite filling each set. Data presented here are
2.3. Marble Falls Limestone, Pedernales Falls from the later fracture set, which trends ENE. Some
State Park, central Texas fractures stop at bed boundaries in the core, while
Pennsylvanian Marble Falls Limestone is exposed others terminate within beds or cross bed
in the Pedernales river-bed in Pedernales Falls State boundaries, giving a wide range of fracture heights.
Park, central Texas. The Marble Falls Limestone is
2.6. Ellenburger Group dolostones, West Texas
a shallow marine wackestone-packstone with Microfracture data were collected from a core
abundant crinoids, which dips approximately 14º to sample at 9079 ft from the Unit No. 3 well in
the southeast. Some beds contain chert nodules and Barnhart field, an Ellenburger hydrocarbon
chertified burrows [12]. East-trending filled reservoir located in southeast Reagan County, West
fractures cut through multiple beds and have en Texas. The Ellenburger Group in the Barnhart area
echelon arrangements both along strike and down comprises shallow-water, lower Ordovician
dip, with both left- and right- stepping patterns. The carbonates containing both dolostone and limestone
fractures contain an early light-colored
[15]. In the Unit No. 3 well these are brecciated Alabama were sampled with 20 drilled sidewall
throughout. The breccias have been interpreted as cores. The well is in the southern Black Warrior
being due to palaeocave collapse [16]. The fractures Basin, a north-trending basin that lies between the
in this study post-date the brecciation and an early Appalachian and Ouachita orogenic belts. The
dolomitization event and are sealed with dolomite Lower Pennsylvanian Pottsville Sandstone was
[4]. Crack-seal texture in the dolomite is common. deposited during uplift of the Appalachian-Ouachita
There is no clear mechanical stratigraphy because Mountains. Pottsville sandstones vary from dark,
brecciation has thoroughly disrupted the sequence. low porosity (< 7%) litharenite to light quartzose
sandstone with higher porosity (~7 to 8 %) [19].
2.7. Weber Sandstone, Deerlodge North,
Fracture aperture data were collected from SEM-
Colorado
based CL and SEI images [19].
The Deerlodge North outcrop in NW Colorado was
studied by Ortega [17] as an analog for Rangely 2.10. Dakota Sandstone, San Juan Basin, New
Field. However, it was found that the structural and Mexico
diagenetic history and mechanical stratigraphy of Two samples from a Dakota Sandstone core from
the outcrops was so different from that of the Weber the San Juan basin in NE New Mexico were imaged
Fm in Rangely Field at approximately 1760 m using SEM-based CL. Microfracture aperture data
(5775 ft) that it did not make a good analog for were collected from scanlines in each thin section.
fractures [17]; Ortega’s outcrop data set is included Both samples, from depths of 7181 ft and 7259 ft,
here. In outcrop the mechanical layer thickness is contain natural fractures, but the fracture cements
very large at 250 m. This has allowed very tall, are different. Fractures in the sample from 7181 ft
wide fractures to develop, the widest fracture being are moderately occluded by late carbonate and
2 m. This is in contrast with the Weber in the anhydrite cements that postdate the fracture event,
subsurface, where mechanical layer thickness is whereas large fractures in the sample from 7259 ft
probably only 50 m and the maximum fracture are mostly open. Microfractures in both samples are
width was just 15 cm. In outcrop the synkinematic sealed by synkinematic quartz.
cement is quartz and the postkinematic cement is
2.11. Cozzette Sandstone, Piceance Basin,
calcite. In the subsurface the synkinematic cement
Colorado
is again quartz but the postkinematic cement is These data were obtained from core of the Upper
ankerite [17]. Cretaceous Cozzette Sandstone Member of the Iles
2.8. Ozona Sandstone, West Texas Formation (Mesaverde Group) from the Piceance
The Ozona Blakeney Kruger No 1 (OBK-1) well in Basin of northwestern Colorado. The Cozzette
Crockett County, West Texas provided core through Sandstone has been interpreted to be a marginal
slope- and basin-floor submarine-fan Ozona marine sandstone, possibly shoreface or offshore-
Canyon sandstones [17, 18]. This core is unusual in bar facies grading upward into barrier or strandplain
that it a vertical core has been drilled through facies [20, 21 and references therein].
steeply dipping beds, providing a bed-parallel core. Cozzette sandstones are typically sublitharenites
The nature of the hydrocarbon trap is not well and consist primarily of quartz, with minor amounts
understood, but a large fault intercepted by the well of chert and rock fragments [22]. Where mica and
suggests that the trap might be partly structural. detrital dolomite content are high, the sandstones
Ortega [17] collected aperture size data from thin are classified as litharenites. Texture is very fine to
sections and then compared them with medium sandstone having detrital silt and clay, and
macrofracture measurements in the core to test it is typically poorly sorted.
whether microfracture population measurements
could be used to predict macro-fracture populations, The core sample is from 9034.7 ft measured depth
which was shown to be the case [1]. The power law from the Slant-Hole Completion Test (SHCT-1)
presented here is for the thin-section data. well, which was drilled as part of a DOE project to
evaluate tight gas sandstones in the Piceance Basin.
2.9. Pottsville Sandstone, Black Warrior Basin, The surface location of the SHCT-1 well is 700 ft
Alabama. south of the DOE Multiwell Experiment (MWX)
Pottsville Sandstone from the Homer L. Jenkins 21- site in section 34, T6S, R94W, in Garfield County,
16 No. 1 well, Sneads Creek field, Pickens County, Colorado [23, 24]. These data were part of a larger
study looking at fracture intensity in the Cozzette On each CL mosaic, fractures were identified and
sandstone [25]. classified according to their shapes and crosscutting
relations with respect to pore-filling, grain
overgrowth, and fracture-filling cements. We
3. METHODOLOGY further classified all microfractures into those that
Single fracture sets were identified on the basis of extend beyond individual grain boundaries
relative timing, orientation and fracture fill. (transgranular) and those that are confined within
Aperture-size distribution data were collected by individual grains (intragranular). Orientation and
constructing a scanline normal to the fracture set, size were mapped electronically using
typically along a bedding plane, but in some cases commercially available software by defining four
along a plane normal to bedding. Fracture apertures points: the two fracture tips and two opposite points
were measured in the field, at the point where the on the fracture walls at the widest aperture. Fracture
scanline crossed each fracture. Fracture apertures attributes were measured and compiled using in-
less than 5 mm were measured using a comparator, house software that uses the digitized parts of the
where pre-drawn lines of given thickness were four points to calculate length, aperture, and
compared with the fracture width, thus binning the orientation [17]. The area of the CL mosaic is also
aperture sizes recorded [17]. The kinematic fracture calculated using image-processing software.
aperture is the wall-to-wall distance across the
fracture and includes both filled and open portions In any quantitative study where a subset of a
of the fracture. It is a measure of the amount of population is collected and compared with another
opening of the fracture, and therefore reflects strain data set, questions arise over whether the data are
due to fracturing. All apertures recorded here are representative of the population as a whole, and
kinematic apertures. For each data set as many over the quality of fracture observation and
fractures as possible were measured in a continuous measurement. In addition, departure from power-
scanline. Numbers of fractures in field data sets are law distributions for reasons of truncation and
limited by the continuity of the outcrop and in censoring, sampling topology and rock
microfracture data sets by the size of the thin heterogeneity are recognized. A full discussion of
section. this subject is outside the scope of this paper, and
has been previously addressed by other workers [2,
Microstructures were mostly imaged on polished 17, 26-28]. In this study every effort was made to
thin sections cut parallel to bedding using scanning collect high quality data, by ensuring all fractures of
electron microscope (SEM)-based cathodolumines- the given size range were recorded. In the field this
cence (scanned CL). The detectors and processing was done by selecting outcrops of continuous clean
used for these images record CL emissions in the exposure and by selecting a lower aperture-size
range of ultraviolet through visible into near cutoff for data collection. In most cases the lower
infrared and convert them to gray-scale intensity cutoff was 0.05 mm but for some data sets where
values. All images were acquired using an Oxford fracture intensity was very high a wider cutoff of
Instruments MonoCL2 system attached to a Philips 0.5 or 0.95 mm was selected so that a longer
XL30 SEM operating at 15 kV. Scanned CL traverse could be collected in the time available.
photographs were taken in traverses several
millimeters in length, and stitched electronically 3.1. Scaling Methods
into mosaics. Typically, a mosaic of 30 to 40 Aperture data are presented as cumulative
individual images at a scale of 1:150 is required to frequency plots of fracture apertures, normalized to
record a continuous CL image along the short side scanline length. Extrapolation of power-laws from
of a 1-by-2-inch thin section. To increase the one scale length to another has been previously
likelihood of intersecting microfractures genetically verified for several of the data sets included here [1,
related and parallel to macrofractures, mosaics were 4, 17, 25]. In these cases data were collected at two
oriented perpendicular to known macrofracture or three scales and compared. Power-law curves
strike [1]. Microstructures in the Ellenburger coincide when extrapolated, demonstrating that
dolostone sample were imaged using light- truncation and censoring departures from the
microscope based CL. power-law distribution are scale specific.
Truncation bias at the small-scale end of the plot is
produced by limits in either imaging or recognizing
and recording progressively narrower fractures.
Cupido dolostone, Las Palmas 11
Censoring bias of large fractures occurs because
these fractures are inadequately sampled along the 1000
length of the scanline chosen. Where a scanline is -1.055
extremely short, but passes through a cluster of F = 11.99 b

cumulative frequency, F
2
fractures over- sampling of large fractures occurs. 100 R = 0.997
This bias was avoided by making sure that, for

(fractures/m)
obviously clustered fracture sets, scanlines were 10 data
long enough to pass through more than one fracture es tim ate 1
es tim ate 2
cluster. In addition, only data sets with an aperture- es tim ate 3
es tim ate 4
1
size range greater than an order of magnitude were es tim ate 5
es tim ate 6
included in the study. This limit was set to eliminate power law

data sets where the censoring effect occurs over a 0.1


relatively large range of aperture sizes. 0.01 0.1 1 10 100
kinematic aperture, b (mm)

A regression model was used to obtain the power-


law distributions (straight segment on a log-log
graph) for sampled microfractures. The model fits Cupido dolostone, Las Palmas 9
all data by recursive calculations (R. Marrett,
written communication, 2000) so that artifacts and 100
an underlying power-law equation are honored. F = 32.1 b -0.978
cumulative frequency, F

data R2 = 0.993

4. RESULTS 10 estim ate 1


(fractures/m)

estim ate 2
For each fracture set a cumulative frequency plot is estim ate 3
presented showing the data points measured, the estim ate 4
1
recursive calculations and the underlying power-law estim ate 5
fit to the straight-line segment of the distribution estim ate 6
(Fig. 1). The power-law equation and correlation power law
coefficient for the regression model are given. The 0.1
power-law coefficients range over two orders of 0.001 0.1 10 1000
magnitude, from 0.1 (Austin Chalk) to 62.4 kinematic aperture, b (mm)
(Pottsville Sandstone), and the exponents range
from –0.45 (Pottsville Sandstone) to -1.36 (Cupido
dolostone).
Cupido dolostone, La Escalera 3

Grove Creek, Austin Chalk 100


F = 20.11 b -0.886
1 R 2 = 0.970
cumulative frequency, F
cumulative frequency, F (fracs/m)

-0.557
F = 0.1052b
10
(fractures/m)

2
R = 0.979
0.1
data
estim ate 1
estim ate 2
0.01 1 estim ate 3
estim ate 4
estim ate 5
estim ate 6
0.001 power law
0.01 0.1 1 10 100 1000
0.1
kinematic aperture, b (mm)
0.1 1 10 100
kinematic aperture, b (mm)
Ellenburger - Barnhart UL-U3-9079.6H
Cupido dolostone, La Escalera 6 10

F = 3.E-02 b -0.777

C u m u lative freq u en cy, F (fracs/m m )


100
F = 31.63 b -1.078 R 2 = 0.9856
cumulative frequency, F

R 2 = 0.987 1
10 data
(fractures/m)

estimate 1
estimate 2
1 data estimate 3
es tim ate 1
es tim ate 2 0.1 estimate 4
es tim ate 3 estimate 5
0.1 es tim ate 4
es tim ate 5 estimate 6
es tim ate 6 power law
power law
0.01 0.01
0.1 1 10 100 1000 0.0001 0.001 0.01 0.1 1

kinematic aperture, b (mm) Kinematic aperture, b (mm)

Knox dolostone, Mississippi

100
cumulative frequency, F (fracs/mm)

-1.361
F = 0.43b
10 2
R = 0.965

1
data
es tim ate 1
es tim ate 2
es tim ate 3
0.1
es tim ate 4
es tim ate 5
es tim ate 6
power law
0.01
0.0001 0.001 0.01 0.1

kinematic aperture, b (mm)

Pottsville Homer L. Jenkins 4240 ft


10
Marble Falls limestone, Pedernales Falls
cumulative frequency, F (fracs/mm)

SP -0.446
1000 F = 0.062 b
microfractures 2
R = 0.67
macrofractures 1
100
cumulative frequency, F

power-law
regression data
10 es tim ate 1
(fracs/m)

es tim ate 2
0.1 es tim ate 3
es tim ate 4
es tim ate 5
1 -1.1952
F = 0.8624 b es tim ate 6
2 power law
R = 0.9655
0.01
0.1
0.0001 0.001 0.01 0.1 1
kinematic aperture, b (mm)
0.01
0.001 0.01 0.1 1 10 100
kinematic aperture, b (mm)
Dakota Sst 7259 ft

10000
-1.31
F = 0.199 b
cumulative frequency, F (fracs/m)

R 2 = 0.966
1000

100 data
es tim ate 1
es tim ate 2
es tim ate 3
10 es tim ate 4
es tim ate 5
es tim ate 6
power law
1
0.0001 0.001 0.01 0.1
kine matic ape rture , b (mm)

Dakota sst. 7181 ft

10000
Fig. 1 Aperture-size cumulative frequency distributions for
F = 0.309 b -1.217
data sets listed in Table 1. Power laws and correlation
cumulative frequency, F (fracs/m)

1000 R 2 = 0.865 coefficients for the best fit function are shown. Ozona
sandstone plot from [1], Weber sandstone plot from [17].

data
100
es tim ate 1
es tim ate 2 Power-law and scanline parameters are compiled in
es tim ate 3
es tim ate 4
Table 1 and the power-law curves are extrapolated
10
es tim ate 5 over the range of kinematic apertures in a
es tim ate 6
power law
compilation plot (Fig. 2). On the compilation plot a
1 broad wedge-shaped envelope defines the natural
0.0001 0.001 0.01 0.1 range of frequencies of fractures for aperture sizes
kinematic aperture, b (mm)
from 100 to 0.01 mm. There is an inverse
correlation between the coefficient and exponent for
Cozzette SHCT 9034.7 ft most of the data sets (Fig. 3). Power laws having
steep slopes tend to have low intensities, whereas
10
power laws having shallow slopes tend to have
higher intensities. The Dakota, Knox and Clear
cumulative frequency (fracs/mm)

F = 8E-03 b -0.813
Fork data sets have large negative exponents and
R 2 = 0.8946
low coefficients, which imply relatively low overall
1
intensity but a very high number of microfractures
and few large fractures. At the other end of the
data spectrum the Pottsville Sandstone, has a small
0.1
es tim ate 1
es tim ate 2
negative exponent and a high coefficient, implying
es tim ate 3 high overall intensity but a relatively larger
es tim ate 4
es tim ate 5 proportion of large fractures. There are two groups
es tim ate 6
power law
of data that depart from this trend, the Austin Chalk
0.01 and Weber Sandstone, and the Cupido dolostone.
0.0001 0.001 0.01 0.1 1 The Austin Chalk and Weber Sandstone have low
Kinematic Aperture (mm) overall intensities but a low slope.
A simple interpretation of variation in power-law
1.00E+05 coefficients might be that they reflect a variation in
strain: the larger the strain, the more fractures
1.00E+04 present (Fig. 4). The data presented here show a
2
weak trend in support of this interpretation, with on
1.00E+03
8
6
outlying point, the Pottsville Sandstone.
4
cumulative frequency, F (fracs/m)

1.00E+02
70
10 Cupido dolostones
9
11
Austin Chalk
1.00E+01 60
7 All other data
12

50
1 5

power law coefficient


1.00E+00
40
1-Austin Chalk outcrop
1.00E-01 2-Cupido Fm. dolomite
30
3-Marble Falls Lst
4-Clear Fork dol 1

1.00E-02
5-Knox dolomite 20
6-Ellenburger 9079.6 3
7-Weber Deerlodge outcrop
8-Ozona Sst. Core, W. Texas 10
9-Dakota 7529
1.00E-03
10-Dakota 7181
11-Pottsville 4240 0
12-Cozzette 9034.7 0.1 1 10 100
1.00E-04 strain %
0.001 0.01 0.1 1 10 100
kinematic aperture, b (mm) Fig. 4. Plot of aperture-size distribution power-law coefficient
against strain.
Fig. 2. Composite plot of power law aperture-size
distributions. Disregarding the Austin Chalk, Weber and 5. DISCUSSION
Cupido curves, the other curves define a broad wedge-shaped
envelope, opening to the right (gray shaded area). Opening-mode fractures in the subsurface grow by
subcritical growth [29, 30]. Olson [31, 32] has
Measurements from the Cupido dolostone, from shown that the local stress field and subcritical
two different, but stratigraphically and structurally crack index of the rock govern which fractures
equivalent areas, show some variation, but they propagate, how fast they propagate, and what effect
form a cluster beneath the curve in Figure 3, they have on neighboring fractures. The subcritical
departing from the general trend, but forming a index has a profound effect on the number of
consistent pattern for this fracture group. The steep fractures in a population, independent of strain.
slope and high coefficient indicate high overall Consequently the size distribution of fractures will
intensity but a small proportion of large fractures. reflect both the amount of deformation and the
subcritical crack index. Moreover, the subcritical
0
index affects the relative numbers of narrow and
-0.2
Cupido dolostones
All other data y = 0.1376Ln(x) - 1.1223
wide fractures and thereby the exponent.
W eber sst. R2 = 0.9152
-0.4 Austin Chalk
Fractures may open in a single event or may
power law exponent

-0.6
repeatedly open and seal [33]. During an opening
-0.8
event the rate of opening competes with the fastest
-1 rates of precipitation to determine if the fracture
-1.2 will seal before the next strain increment. Small
-1.4 fractures completely seal with cement precipitated
-1.6 synchronously with opening, whereas large
0.01 0.1 1 10 100
fractures may retain some porosity. The aperture
pow er law coefficient
size at which porosity is preserved varies, and it is
Fig. 3. Variation of power-law coefficient and exponent. Line controlled by the temperature of the fracture fluid,
is fit to data sets (squares) other than the Cupido (diamonds), the composition and texture of the host rock and
Austin Chalk (triangle) and Weber (cross). The equation and
correlation coefficient for the best fit are shown.
precipitating mineral, and the length of time the
fracture wall is exposed to mineral precipitation, evidence of repeated opening and is common in
which is dependent on burial history and fracture bridges of large fractures but is absent in very
timing. If the widest fractures are not completely narrow microfractures [33]. This observation
sealed before the next strain increment, they may indicates that narrow microfractures are formed
act as planes of weakness, causing strain to during a single opening event, whereas larger
progressively partition into fewer fractures, which fractures grow by multiple discrete increments. It is
will grow wider. The extent to which this process not clear whether these approximate to a constant
happens should partly govern the exponent in the strain rate or whether strain rate is variable. Rapid
power-law distribution, with higher strain being fracture opening could allow even very small
linked to shallower slopes (Fig. 5). fractures to remain open because cementation
0
cannot keep up with opening, thereby allowing
Cupido dolostones these fractures to continue growth. Conversely, a
Austin Chalk
-0.2
All other data
slow strain rate may allow many small fractures to
-0.4 become sealed and stop opening.
power law exponent

-0.6
Mechanical layer thickness will also affect the
-0.8
aperture-size distribution for a given fracture set
-1 because it controls the intensity of the larger
-1.2
fractures that cross the mechanical layer for a given
strain [31]. Determination of mechanical layer
-1.4
thickness is challenging because, in addition to
-1.6 observational difficulties in outcrop and core, the
0.1 1 10 100
mechanical properties of the rock package at the
strain %
time of fracturing are required, which in turn will
Fig. 5. Plot of aperture-size distribution power-law exponent depend on the relative amount and type of
against strain. cementation in different layers at that time. In this
study, unequivocal mechanical layer thickness was
This clearly does not apply for the Austin Chalk indeterminate for many of the data sets.
relative to other rock types, although differences in
other mechanical rock properties probably make In the Austin Chalk the relatively large fractures
direct comparison between different rock types occur in clusters, with cluster spacing on the order
meaningless. Within a single rock type, three of the of 50 m and lengths along the scanline completely
four Cupido dolostone data sets follow a trend devoid of any fractures [5]. Like the Austin Chalk,
where higher strain is linked to smaller exponents, fractures in the Cupido dolostone are also clustered,
but the dolostone at La Escalera 6 departs from this with cluster spacing on the order of 1.5 m. Unlike
trend. Power-law distributions, however, are the Austin Chalk, the lengths of scanline between
continuous across the size-range from narrow, the clusters are filled with narrow fractures. The
sealed microfractures to large fractures with timing of fracturing in the Austin Chalk is not
porosity. This suggests that although strain might known, although it appears to predate some normal
focus on larger fractures as total strain increases, fault displacement [6].
smaller fractures continue to nucleate and grow.
Variations in fault displacement distributions with Subcritical crack indices for several rock types have
strain have previously been reported [34] and a been measured and are very large for the Cupido
decrease in exponent with increasing strain has been dolostone and Austin Chalk, (100 +) and are more
interpreted as the result of progressive partitioning moderate (~40 to 60) for the Dakota Sandstone
onto fewer but larger faults [35] and by fault [32]. These observations are in keeping with the
linkage [34]. highly clustered nature of the former. However,
these measurements were made on samples in their
Although no quantitative data are available, the rate current diagenetic state, and may not reflect the
at which fractures widen, relative to the rate of state of the rock at the time of fracturing.
cement precipitation could influence which
fractures continue to grow. Crack-seal texture is
6. CONCLUSIONS 3. Cowie, P. A., R. J. Knipe and I. G.Main, eds., 1996.
Scaling laws for fault and fracture populations—Analyses
Apertures of mode-I fracture sets follow power-law and applications: Journal of Structural Geology 18: 135–
distributions for aperture sizes ranging from 383.
approximately 1 μm to 0.1 m and can be considered 4. Gale, J. F. W., S. E. Laubach, R. A. Marrett, J. E. Olson,
J. Holder, and R. M. Reed. 2002. Predicting and
self-organized. Fracture sets described here with characterizing fractures in dolomite reservoirs: using the
power-law aperture-size distributions come from link between diagenesis and fracturing, In The geometry
very different geologic settings, suggesting that and petrogenesis of dolomite hydrocarbon reservoirs:
self-organization of opening-mode fractures is the Petroleum Studies Group, The Geological Society of
norm. A power law forms a straight line on a log London, Burlington House, unpaginated.
cumulative frequency (fractures/meter) versus log 5. Gale, J. F. W. 2002. Specifying lengths of horizontal
aperture size (millimeter) plot. The slope of the line wells in fractured reservoirs. Society of Petroleum
is the power-law exponent, reflecting the relative Engineers Reservoir Evaluation and Engineering, SPE
78600: 266-272.
number of narrow and wide fractures in the set. The
pre-exponential coefficient reflects the overall 6. Stowell, J. F. W., 2001. Characterization of opening-
mode fracture systems in the Austin Chalk. Gulf Coast
fracture intensity.
Association of Geological Societies Transactions 51: 313-
319.
Coefficients range from 0.1 to 62.4, and exponents
7. Hovorka, S. D. 1998. Facies and diagenesis of the Austin
range from -0.45 to -1.36. There is a negative Chalk and controls on fracture intensity – a case study
correlation between log coefficient and exponent for from North Central Texas. Geological Circular 98-2,
most of the data sets. On a compilation plot of Bureau of Economic Geology, The University of Texas at
power laws, a broad wedge-shaped envelope Austin.
defines the natural range of frequencies of fractures 8. Goldhammer, R. G., P.J. Lehman, R.G. Todd, J.L.
for given aperture sizes. Power laws having steep Wilson, W.C. Ward and C.R. Johnson. 1991. Sequence
slopes tend to have low intensities, whereas power stratigraphy and cyclostratigraphy of the Mesozoic of the
laws having shallow slopes tend to have higher Sierra Madre Oriental, northeast Mexico. Gulf Coast
Section, Society of Economic Paleontologists and
intensities. There are, however, a few exceptions to Mineralogists.
this correlation, in particular the Cupido dolostones,
9. Padilla y Sanchez, R., 1985. Las estructuras de la
Weber Sandstone and the Austin Chalk. Cupido
curvatura de Monterrey, Estados de Coahuila, Nuevo
dolostones have high intensities and steep slopes, Leon, Zacatecas y San Luis Potosi. Revista del Instituto
and Weber Sandstone and Austin Chalk have lower de Geologica de la Universidad Autonoma de Mexico 6:
intensity and a shallow slope. 1-20.
10. Ortega, O. J. and R. A. Marrett. 2001. Stratigraphic
There is a weak correlation between strain and controls on fracture intensity in Barremian-Aptian
power-law coefficients. Other influences on the carbonates, northeastern Mexico. In Genesis and controls
power-law distributions will be mechanical rock of reservoir-scale carbonate deformation, Monterrey
Salient, Mexico, ed. R. A. Marrett, 57-82. Bureau of
properties, most notably subcritical crack index, Economic Geology, The University of Texas at Austin,
strain rate and mechanical layer thickness. These Guidebook 28, Austin, Texas.
parameters are difficult to determine for the rock at
11. Monroy-Santiago, F., S. E. Laubach and R. A. Marrett.
the time of fracturing, and their roles could not be 2001. Preliminary diagenetic and stable isotope analyses
assessed quantitatively here. of fractures in the Cupido Formation, Sierra Madre
Oriental. In Genesis and controls of reservoir-scale
carbonate deformation, Monterrey Salient, Mexico, ed. R.
A. Marrett, 83-107. Bureau of Economic Geology, The
REFERENCES University of Texas at Austin, Guidebook 28, Austin,
1. Marrett, R. A., O. J. Ortega and C. M. Kelsey. 1999. Texas.
Extent of power- law scaling for natural fractures in rock. 12. Wermund E. G. and V. E. Barnes. 2003. Down to Earth at
Geology 27: 799-802. Pedernales Falls State Park, Texas. Bureau of Economic
2. Gillespie, P.A., C. B. Howard, J. J. Walsh and J. J. Geology, The University of Texas at Austin, Down to
Watterson. 1993. Measurement and characterization of Earth Series Guidebook, Austin, Texas.
spatial distributions of fractures. Tectonophysics 226: 13. Gale, J. F. W., S. E. Laubach, R. M. Reed, J. G. Moros
113-141. Otero and L. A. Gomez. 2002. Fracture analysis of Clear
Fork outcrops in Apache Canyon and cores from South
Wasson Clear Fork field, In Integrated outcrop and 24. Lorenz, J. C. and R. Hill. 1991. Subsurface fracture
subsurface studies of the interwell environment of spacing: comparison of inferences from slant/horizontal
carbonate reservoirs: Clear Fork (Leonardian-age) core and vertical core in Mesaverde reservoirs in the
reservoirs, West Texas and New Mexico, ed. F. J. Lucia, Piceance Basin: SPE Rocky Mountain Regional
191–226. The University of Texas at Austin, Bureau of Meeting/Low-Permeability Reservoir Symposium, 15-17
Economic Geology, final technical report prepared for April 1991, SPE No. 21877, 705–716. Denver: SPE.
DOE under contract no. DE-AC26-98BC15105.
25. Gomez, L., J. F. W. Gale, S. E. Laubach and S. Cumella.
14. Moros Otero, J. G. 1999. Relationship between fracture 2003. Quantifying fracture intensity: an example from the
aperture and length in sedimentary rocks. Unpublished Piceance Basin. In Piceance Basin 2003 Guidebook, eds.
Masters thesis, University of Texas at Austin. K. M. Peterson, T. M. Olson and D. S. Anderson, Chapter
6, 96-113. Rocky Mountain Association of Geologists:
15. Holtz, M. H. and C. Kerans. 1992. Characterization and
Denver, Colorado.
classification of West Texas Ellenburger reservoirs. In
Paleokarst, karst related diagenesis and reservoir 26. Marrett, R. and R. Allmendinger. 1990. Kinematic
development: examples from Ordovician–Devonian age analysis of fault slip data. Journal of Structural Geology
strata of West Texas and the Midcontinent, eds. M. P. 12: 973-986.
Candelaria, and C. L. Reed, 45–54 Permian Basin SEPM
27. Wojtal, S. F. 1996. Changes in fault displacement
Publication 92-33.
populations correlated to linkage between faults. Journal
16. Loucks, R. G. 2003. Understanding the development of of Structural Geology 18: 265-280.
breccias and fractures in Ordovician carbonate reservoirs.
28. Barton, C. A. and M. D. Zoback. 1992. Self-similar
In The Permian Basin: back to basics, eds. T. J. Hunt, and
distribution and properties of macroscopic fractures at
P. H. Lufholm, 231–252. West Texas Geological Society
depth in crystalline rock in the Cajon Pass Scientific Drill
Fall Symposium: West Texas Geological Society
Hole. Journal of Geophysical Research, B, 97: 5181-
Publication #03-112.
5200.
17. Ortega, O. 2002. Fracture-size scaling and stratigraphic
29. Atkinson, B. K. and P. G. Meredith. 1987. The theory of
controls on fracture intensity. Ph.D. dissertation, The
subcritical crack growth with applications to minerals and
University of Texas at Austin.
rocks. In Fracture mechanics of rock. ed. B. K. Atkinson
18. Hamlin, H. S. 1999. Syn-orogenic slope and basin San Diego, California: Academic Press.
depositional systems, Ozona Sandstone, Val Verde Basin,
30. Holder, J., J. E. Olson and Z. Philip. 2001. Experimental
southwest Texas. Ph.D. dissertation, The University of
determination of subcritical crack growth parameters in
Texas at Austin.
sedimentary rock. Geophysical Research Letters 28: 599-
19. Laubach, S. E., R. M. Reed, J. F. W. Gale, O. Ortega, and 602.
E. Doherty. 2002. Fracture characterization based on
31. Olson, J. E. 1993. Joint pattern development: effects of
microfracture surrogates, Pottsville Sandstone, Black
subcritical crack-growth and mechanical crack
Warrior Basin, Alabama: Gulf Coast Association of
interaction. Journal of Geophysical Research 98: 12,251-
Geological Societies Transactions 52: 585–596.
12,265.
20. Lorenz, J. C. 1983. Lateral variability in the Corcoran and
32. Olson, J. E., Y. Qiu, J. Holder, and P. Rijken. 2001.
Cozzette blanket sandstones and associated Mesaverde
Constraining the spatial distribution of fracture networks
rocks, Piceance Creek basin, northwestern Colorado:
in naturally fractured reservoirs using fracture mechanics
Society of Petroleum Engineers, SPE/DOE Paper 11608:
and core measurements: 2001 SPE Annual Technical
81–86.
Conference and Exhibition.
21. Dutton, S. P., S. J. Clift, D. S. Hamilton, H. S. Hamlin, T.
33. Laubach, S.E., R.M. Reed, J.E. Olson, R.H. Lander and
F. Hentz, W. E., Howard, M. S. Akhter, and S. E.
L.M. Bonnell. 2004. Coevolutionof crack-seal texture and
Laubach. 1993. Major low-permeability sandstone gas
fracture porosity in sedimentary rocks:
reservoirs in the Continental United States: The
cathodoluminescence observations of regional fractures.
University of Texas at Austin Bureau of Economic
Journal of Structural Geology 26: 967-982.
Geology Report of Investigations No. 211.
34. Cladouhos, T. T. and R. Marrett. 1996. Are fault growth
22. Hansley, P. L., and R. C. Johnson. 1980. Mineralogy and
and linkage models consistent with power-law
diagenesis of low-permeability sandstones of Late
distributions of lengths? Journal of Structural Geology,
Cretaceous age, Piceance Creek basin, northwestern
18: 281–293.
Colorado: The Mountain Geologist 17: 88–129.
35. Moriya , S., C. C. Childs, T. Manzocchi and J. J. Walsh.
23. Pitman, J. K., and E. S. Sprunt. 1986. Origin and
in press. Fault populations, straindistribution and
distribution of fractures in Lower Tertiary and Upper
basement reactivation in the East Pennines Coalfield,
Cretaceous Rocks, Piceance basin, Colorado, and their
U.K. Journal of Structural Geology.
relation to the occurrence of hydrocarbons. In Geology of
tight gas reservoirs: AAPG Studies in Geology 24 eds. C.
W. Spencer and R. F. Mast, 221–234. AAPG.

You might also like