Anda di halaman 1dari 16

Estimating fracture trace

intensity, density, and mean


length using circular scan lines
and windows
M. B. Rohrbaugh Jr., W. M. Dunne, and M. Mauldon

ABSTRACT
Fracture characterization protocols that reduce sampling bias are
likely to yield higher quality input for exploration and development
decisions when dealing with naturally fractured reservoirs. A new
set of estimators for fracture density, intensity, and mean trace
length corrects for sampling biases and provides a useful integrated
description for bulk aspects of a fracture network. These estimators
are based on counts of intersections between fracture traces and
circular scan lines and of trace terminations in circular windows.
Application to synthetic fracture patterns with known parameters
validates the use of the new estimators, which are then applied to
natural fault trace maps from seismic volumes and joint trace maps
from rock pavements. The new estimators are distribution independent and eliminate the effects of orientation, censoring, and
length biases, which limit the effectiveness of other sampling techniques. Estimator accuracy improves as sample size increases, particularly for larger circles that exceed a fracture-defined block size.
Estimator accuracy for mean trace length improves when the sample exceeds threshold count values for fracture terminations based
on guidance from the analysis of similar synthetic patterns. These
new estimators also provide both inputs and independent checks of
predictions for fracture-generator programs used to model fracture
populations in a rock volume.

INTRODUCTION
Quantification of fracture parameters such as density, size, and intensity aids in the assessment of hydrocarbon flow and storage in
fractured reservoirs (Reiss, 1982; Nelson, 1985; Dershowitz and
LaPointe, 1994; Narr, 1996) but is complicated by the difficulty of
deciding on the best approach for incorporating fractures into a
reservoir model. Problems arise from the lack of consensus as to
Copyright 2002. The American Association of Petroleum Geologists. All rights reserved.
Manuscript received January 24, 2000; revised manuscript received January 10, 2001; final acceptance
June 15, 2002.

AAPG Bulletin, v. 86, no. 12 (December 2002), pp. 20892104

2089

AUTHORS
M. B. Rohrbaugh Jr. Tennessee
Department of Environment and
Conservation, Division of Underground
Storage Tanks, 540 McCallie Avenue, Suite
550, Chattanooga, Tennessee, 374022013
M. Bruce Rohrbaugh Jr. received his B.S.
degree in geology from West Virginia
University in 1997 and his M.S. degree in
structural geology from the University of
Tennessee, Knoxville in 2000. His research
interests include hydrogeology, application of
computers to solving geologic problems, and
structural geology. He is currently employed
as a geologist with the Tennessee Department
of Environment and Conservation.
W. M. Dunne Department of Geological
Sciences, 306 G&G Building, University of
Tennessee, Knoxville, Tennessee, 37996-1410;
wdunne@utk.edu
William M. Dunne, although born in the
United States, received his B.S. degree and
Ph.D. in geology from the University of Bristol,
England. He joined the Department of
Geological Sciences at the University of
Tennessee in 1988 as an associate professor
and is now is a professor and department
head. His research interests include fracture
characterization particularly in younger rocks,
deformation in thrust belts from large to small
scale, and deformation analysis of
sedimentary rocks.
M. Mauldon Department of Civil and
Environmental Engineering, Virginia Tech, 200
Patton Hall, Mail Code 0105, Blacksburg,
Virginia, 24061; mauldon@vt.edu
Matthew Mauldon, although born in England,
has geology (B.A.) and civil engineering (M.S.)
degrees and a Ph.D. in civil engineering from
the University of California at Berkeley. He
spent eight years on the faculty at the
University of Tennessee, where he
collaborated with Dunne and Rohrbaugh.
Mauldon is now an associate professor in the
Via Department of Civil and Environmental
Engineering at Virginia Tech, where he
teaches and conducts research in the areas of
rock mechanics, engineering geology, and
geotechnical engineering.

ACKNOWLEDGEMENTS
Acknowledgment is made to the donors of
The Petroleum Research Fund, administered
by the American Chemical Society, for partial
support of this research. Additionally, the
Geological Society of America and the Southeastern Section of the Geological Society of
America are thanked for their partial support
of the field work. We would also like to thank
M. F. Schaeffer for permission to use the fracture trace map from Rocky Creek, South Carolina, and Camilo Montes, Yen-Yit Chan, You Li,
and Jim Calcagno for programming assistance.
Steve Laubach, John Lorenz, and Bill Dershowitz are thanked for their insightful and
constructive reviews.

which parameters to quantify, the difficulties in measuring the parameters, and intrinsic sampling biases. To address these problems,
we offer a set of estimators for fracture parameters based on the
use of circular scan lines and windows where a scan line is the perimeter and the window is the interior of a circle. These estimators
reduce sampling biases and provide a useful integrated description
of bulk aspects of a fracture network. Use of these estimators requires only counts of the number (n) of fracture traces intersecting
the circumference and/or counts of the number (m) of fracture
traces terminating in the circle interior. Performance of the new
circular scan line/window estimators is evaluated for both synthetic
and natural fracture patterns to demonstrate that estimates converge on true values for a fracture trace population, to demonstrate
that the new estimators outperform or match existing estimators,
and to discuss issues of estimator performance.
Fracture Parameters
A redundant, commonly mutually inconsistent vocabulary exists to
describe the amount of fracturing in a rock. Rather than reviewing
this terminology, we define three key parameters of a fracture pattern: density, size, and intensity (Table 1).
Density
Fracture density is commonly treated as the number of observed
isolated fractures or fracture segments per unit length, area, or volume (Dershowitz and Herda, 1992; Ghosh and Daemen, 1993).
This is a scale-dependent quantity that we call apparent density.
Fracture density is defined in this article as the number of fractures
per unit length, area, or volume, enumerated in terms of unique
points, such as fracture centers (Mauldon, 1998; Mauldon and Dershowitz, 2000). Apparent density overestimates density (Kulatilake
and Wu, 1984; Mauldon et al., 2001), and the magnitude of this
overestimation increases as sample size decreases (Figure 1B). For
example, the apparent density (number of visible traces divided by
circle area) of fractures of Set 1 in Figure 1A is 0.0014 per m2 for
a circular window of radius 75 m and increases to 0.0020 per m2
for a smaller window of radius 25 m. If trace centers (dots in Figure
1A) are counted to estimate true density for Set 1, the estimates
(count divided by circle area) are 0.0010 per m2 for both the circle
of radius 25 m and the circle of radius 75 m. In practical applications, one half the number of fracture trace terminations is used as
an unbiased estimate of the number of trace centers, because centers cannot be identified unless both ends of a trace are visible
within the window (Mauldon, 1998; Mauldon et al., 2001).
Size
Fracture size is defined in one, two, or three dimensions, as fracture trace length, area, and volume, respectively. For composite
connected fractures, an investigator should decide whether to
characterize individual segments or the entire composite. Typically

2090

Fracture Characterization Using Circular Scanlines and Windows

Table 1. Vocabulary for Fracture Parameters*


Parameter

Definition

Density

Linear
Areal (q)
Volumetric

Size

Linear (l)
Areal
Volumetric
Linear
Areal (I)
Volumetric

Intensity

Estimator

Number of fractures per unit length


Number of fractures per unit area
Number of fractures per unit volume (in all cases count is conducted using
unique points such as fracture centers or half of the fracture terminations)
Mean fracture trace length
Mean fracture area
Mean fracture volume
Number of fracture per unit length (L0 /L1 L1)
Fracture length per unit area (L1 /L2 L1)
Fracture area per unit volume (L2 /L3 L1)

q m/2pr 2

l (pr/2)(n/m)

I n/4r

*Where L is a dimension of length and r is radius. See Figure 3 for illustration of m and n.

for fracture studies on exposed surfaces, trace length


or aperture is measured because fracture areas and
volumes are commonly not directly measurable (Der-

showitz and Herda, 1992; Marrett et al., 1999; Ortega


and Marrett, 2000). For the two-dimensional fracture
patterns discussed in this article, we use mean trace
length as our size parameter (Table 1; Figure 1).
Intensity
Fracture intensity is a pattern characteristic that incorporates both density and size (Dershowitz and
Herda, 1992; Mauldon and Dershowitz, 2000). Intensity is defined as number of fractures per unit sample
length, fracture length per unit surface area, or fracture area per unit rock volume, in one, two, or three
dimensions, respectively (Table 1). Consequently, intensity has the same dimensions whether calculated
linearly, areally, or volumetrically. In this article, we
examine two-dimensional samples and, therefore, use
the areal intensity: fracture length per unit area.
Current Measurement Methods

Figure 1. Problems due to censoring and length bias when


sampling fracture traces. (A) Fracture pattern with two sets sampled by three progressively larger circles (dots trace centers;
r radius). (B) Decreasing density overestimates for increasing
sample (circle) size. (C) Decreasing mean trace-length underestimates for increasing sample size. (D) Undersampling of joint
trace lengths due to censoring as shown by the probability distribution function (pdf). (E) Oversampling of longer traces due
to length bias as shown by a pdf.

Two common sampling methods are used for estimating fracture parameters: straight scan lines and
areal sampling (Figure 2) (LaPointe and Hudson,
1985; Priest, 1993; Wu and Pollard, 1995, Becker and
Gross, 1996; Marrett et al., 1999; Ortega and Marrett,
2000). Straight scan lines sample the fractures they
intersect and are used to systematically record fracture
characteristics, such as number, orientation, aperture,
and so on. (Priest and Hudson, 1981). Areal sampling
involves mapping the fracture trace pattern and recording desired fracture characteristics at locations in
the map area (e.g., Priest, 1993; Wu and Pollard,
1995).
Rohrbaugh et al.

2091

Figure 2. Circular scan line/window (dotted circle), areal (irregular white window), and straight scan line (dotted line) sampling of a fracture trace population. Solid lines represent visible
fracture traces, and dashed lines represent covered fracture
traces.

Sampling Biases
Orientation bias occurs where scan lines or the long
axes of inequant sampling areas are not perpendicular
to a fracture set. In both cases, intensity is underestimated (Terzaghi, 1965; Priest, 1993; Mauldon and
Mauldon, 1997). Scan line estimates are corrected by
dividing by the cosine of the angle between the scan
line and the normal to the fracture set (Terzaghi, 1965;
Peacock et al., in press). However, as this angle approaches 90, the cosine approaches zero, and corrected estimates approach infinity, significantly overestimating intensity (Priest, 1993). No procedures are
currently available for correcting orientation bias from
inequant sampling areas.
Censored fracture traces extend beyond the exposure or seismic coverage, so that one or both ends
are not visible (Figure 1A) (e.g., Cruden, 1977;
Baecher and Lanney, 1978; Einstein and Baecher,
1983; Kulatilake and Wu, 1984; LaPointe and Hudson,
1985; Pickering et al., 1995; Mauldon, 1998; Marrett
et al., 1999). Such traces are referred to as singly or
doubly censored, respectively. Using censored traces to
estimate density and size directly, rather than using estimators such as those in Table 1, leads to overestimates of density and underestimates of size (Figure 1).
For example, a count of all visible trace segments in a
circle of radius 25 m (4 segments vs. 2 centers) of Set
1 (Figure 1A) overestimates density by a factor of two
(0.002 vs. 0.001 per m2). Similarly, using the censored
2092

lengths, the estimate of mean trace length for Set 1


(total visible trace length divided by number of traces)
is 21.3 m, in contrast to the true mean trace length of
40 m. In both cases, increasing the size of the sampling
circle relative to fracture size decreases censoring and,
consequently, error. For example, increasing circle radius from 25 to 75 m (Figure 1A) decreases apparent
density from 0.0020 to 0.0014 per m2 (true density
0.001 per m2) and increases apparent mean trace
length from 21.3 to 27.4 m (true value 40 m).
Length bias occurs because longer fracture traces
have a greater probability of being sampled than
shorter traces (Baecher and Lanney, 1978; Einstein and
Baecher, 1983; LaPointe and Hudson, 1985; Mauldon
1998). Consequently, mean trace-length estimates and
trace-length distributions are skewed toward longer
fractures (Figure 1E). For example in Figure 1A, Sets
1 and 2 have the same fracture density, but Set 1 has
longer traces. As a result, the largest window (75 m
radius) samples 24 of the longer Set 1 fractures vs. 17
of Set 2. If these data are taken at face value,
mean trace length for the whole pattern will be
overestimated.
Pattern heterogeneity refers to a change in fracture
parameters with a change in position. This sampling
bias can occur, for example, when sampling rock volumes with localized fracture development near faults.
Heterogeneity is partly a function of scale and may be
especially pronounced for borehole sampling, which
provides limited samples of subsurface fracture networks. Use of multiple subdomains with homogeneous
parameters minimizes the effects of heterogeneity
(Turner and Weiss, 1963; Whitten, 1966; LaPointe
and Hudson, 1985; Kulatilake et al., 1997). Alternatively, geostatistical methods may be employed to determine the magnitude and rate of change of a parameter as a function of distance and direction (LaPointe
and Hudson, 1985; LaPointe, 1993; Priest, 1993; Jian
et al., 1996).
Advantages/Disadvantages of Current Estimation Methods
Straight scan lines (Figure 2) provide rapid estimates
of fracture intensity (Priest and Hudson, 1981; LaPointe and Hudson, 1985; Becker and Gross, 1996).
Unprocessed straight scan line data, however, are subject to orientation bias, length bias, censoring, and pattern heterogeneity (Terzaghi, 1965; Baecher and Lanney, 1978; Priest and Hudson, 1981; LaPointe and
Hudson, 1985; Priest, 1993; Mauldon and Mauldon,
1997; Peacock et al., in press).

Fracture Characterization Using Circular Scanlines and Windows

Areal sampling (Figure 2) reduces censoring and


length bias as compared to scan line sampling but (1)
is subject to orientation bias in the plane for anything
except a circular sampling area, (2) is more time consuming than use of scan lines, (3) may disguise pattern
heterogeneities, and (4) may introduce censoring and/
or length biases for inequant or small sampling areas
(LaPointe and Hudson, 1985; Kulatilake et al., 1993;
Wu and Pollard, 1995; Kulatilake et al., 1997).

NEW ESTIMATORS
Because neither areal sampling nor straight scan lines
are completely satisfactory in terms of efficiency, accuracy, and lack of bias, this article examines the use
of circular scan lines and windows for characterizing
intensity, density, and mean trace length (Figure 2).
Data from circular scan lines and windows are applied
to estimators (Table 1) that do not require knowledge
of fracture spacing, trace length, or orientation distributions and are, therefore, distribution independent
(Mauldon, 1998; Mauldon et al., 2001).

Figure 3. Fracture trace pattern with sampling circle. (A) Solid


dots are intersection points (n) between fractures and circle. (B)
Triangles are fracture endpoints (m) in the circular windows.

ESTIMATOR PERFORMANCE FOR


SYNTHETIC FRACTURE PATTERNS
The ability of the new estimators to correct for sampling biases was investigated by comparing estimates
of intensity, density, and mean trace length to known
values for synthetic fracture trace patterns. A new
computer program called JAWS (Joint Analysis using
Windows and Scanlines) (Rohrbaugh, 2000), was used
to generate and sample traces with uniformly distributed centers in a square region. These traces were sampled with 100 circles of known radius placed randomly
and independently in a smaller square analysis region
centered on the generation region so as to avoid edge
effects (Gilmour et al., 1986). Trace-circle intersections and trace terminations inside circular windows
were counted (Figure 3), and counts were input into
estimators (Table 1) to yield comparison values.
For example, two synthetic fracture sets with orientations of 60 30 and 150 30 were deployed
to produce orientation bias when sampled, with fracture lengths of 50 10 L (L is an arbitrary unit of
length) and 100 10 L that exceed circle radius of 20
L by factors of 2.5 and 5, respectively, to create a censoring bias, and with a difference in length between
sets of a factor of two to yield a length bias. Running

means for the estimates (Figure 4) converge on the input values for intensity, density, and mean trace length
and correspond closely to them after about 40 samples,
whereas individual estimates fluctuate about the mean.
These results show that the new estimators deal successfully with orientation, censoring, and length biases.
Similar results were found for all other synthetic cases.
Comparison to Other Estimators
Having established that the new estimators overcome
sampling biases, their accuracy was compared to that
from straight scan lines and/or areal sampling of a synthetic pattern. To achieve equivalent sampling comparison, estimates were obtained using 100 circular
scan lines/windows of radius 10 L , 100 straight scan
lines of the same length (2p (10 L) 63 L), and
areal samples consisting of 100 circle interiors.
For example, a fracture set with orientation of 060
30, mean trace length of 40 20 L, intensity of
1.95 L/L2, and density of 0.049 per L2 was generated
(Figure 5A). Intensity estimates from both circular
scan lines and areal samples accurately estimate intensity, but map construction for areal samples is likely to
Rohrbaugh et al.

2093

Based on these results, the use of the estimators based


on circular scan lines and windows is recommended.

ESTIMATOR PERFORMANCE WITH


NATURAL FRACTURE PATTERNS

Figure 4. Estimates (jagged black lines) and running means


of estimates (small open circles) for intensity, density, and mean
trace length of the synthetic pattern with dashed lines and numbers for control values, showing that estimators successfully deal
with sampling bias issues.
be much more time intensive than is deployment of
circular scan lines. In contrast, running means for the
straight scan lines underestimate intensity by a factor
of 1.7 with the Terzaghi correction applied to the
mean set orientation (Figure 5B).
The running mean density estimate from the circular window estimator corresponds closely to the control value of 0.049 per L2, whereas the same size areal
samples overestimate density by a factor of 3.5 if trace
segments are counted (Figure 5C). Running means for
circular window estimates of mean trace length also
correspond to the control value, whereas the same size
areal samples tend to yield estimates of a little less than
4 L, which is an underestimate by a factor of 11 as
compared to the input value of 40 L (Figure 5D).
2094

Given the success of the new estimators with synthetic


fracture patterns, we investigated their applicability to
natural patterns. This comparison used new and existing trace maps of fault and/or joint patterns that range
from single sets of parallel fractures to multiset patterns to near polygonal patterns (Table 2; Figure 6).
This geometric variety was selected so as to provide a
thorough test of the estimator equations. Also, fault
trace maps from seismic reflection data (Figure 6C, D;
Table 2) were included to demonstrate that circular
scan lines may be used effectively with this common
industry data source. Trace maps were digitized and
imported into JAWS for analysis. After specifying the
number and size of the sample circles, JAWS randomly
distributed them in the sample area and only retained
intersection and termination counts for circles that
were completely contained within a map (Stoyan et al.,
1995). The new circle-based estimates for the natural
data sets were compared to estimates from areal samples of the maps.
Circle-based intensity estimates mostly match
areal estimates for both fault and joint trace patterns
(e.g., Figures 6; 7C, D), as would be expected from the
similar correspondence that was found during the analysis of synthetic fracture patterns (e.g., Figure 5B). The
circle-based density and mean trace-length estimates
mostly matched the apparent density and mean tracelength estimates where fracture/fault traces are small
relative to window size, so that censoring is minimal
(e.g., Figures 6C, D; 7E, F). In contrast, at Llantwit
Major, 86% of the master joint traces are censored, and
direct density and mean trace-length estimates from
areal samples differ by a factor of 4 from the circlebased estimates (density: 1.5 vs. 0.35 per m2; mean
trace length: 2.2 vs. 8.7 m)(Figures 6A; 7A, B). Having
already tested and demonstrated the performance of
the circle-based estimators using synthetic traces, we
interpret these circle-based estimates at Llantwit Major as being representative of true population characteristics. It follows that the direct areal estimates for
the characteristics of these greatly censored traces are
off by a factor of 4. Again, the circle-based estimators
yield superior results to previous estimators where the
potential for sampling biases exists.

Fracture Characterization Using Circular Scanlines and Windows

Figure 5. Estimator performances for a synthetic fracture trace pattern. (A) Part of a synthetic fracture pattern used to test relative
performance of new and old estimators. Pattern characteristics are described in text. (B) Intensity estimates from circular scan lines
(thick black line), circular areas (light gray line), and straight scan lines (thin line). Control value is 1.95 L/L2. (C) Density estimates
from circular windows (thick black line) and circular areal sampling of apparent density (light gray line). Control value is 0.049 per
L2. (D) Mean trace length estimates from circular windows (thick black line) and circular areal sampling of apparent trace length
(light gray line). Control value is 40.0 L. Open circles in (B), (C), and (D) are the running means for appropriate estimators.
DISCUSSION OF ESTIMATOR
PERFORMANCE
Effects of Block Size on Estimates
Block size in two dimensions is the unfractured area
bounded by fracture traces from two or more sets

(LaPointe, 1988). Block size depends on the fracture


spacing or frequency and could be an issue for the new
estimators for circles smaller than block size. Block size
effects on the intensity estimator were investigated at
Llantwit Major and the light gray region at Amroth
(Figure 6; Table 2). Block size at Llantwit Major is
about 0.71 m by 0.30.5 m, whereas at Amroth the
Rohrbaugh et al.

2095

2096

Fracture Characterization Using Circular Scanlines and Windows


Early Westphalian
(Carboniferous), Early
Westphalian coal
measures (sandstone)
Jurassic, Hugin
Formation, reservoir
sandstone
Upper Cretaceous to
Holocene sediments

Amroth, Wales, United


Kingdom,
SN17560722*
(Figure 6B)
Sleipner Vest field,
North Sea,
(Figure 6C)

Cartier Trough, Timor


Sea (unspecified)
(Figure 6D)
Telpyn Point, Wales,
United Kingdom,
SN18330731*
(Figure 6E)
Ward Lake, California,
United States, not
specified (Figure 6F)
Cretaceous, Mt. Givens
Granodiorite

Late Namurian
(Carboniferous), Upper
Sandstone Group

Early Jurassic, Porthkerry


Formation (limestone)

Rock Age and


Formation Name

Llantwit Major, Wales,


United Kingdom,
SS95756754*
(Figure 6A)

Name, Location,
Figure No.

Table 2. Geology of Fracture Trace Populations

Segall and Pollard


(1983), 100 cm,
unknown, 1835.6 m2
(exposed area), 1169.1
m2 (stippled area)

Walsh et al. (1996),


40m displacement,
unknown, 555.6 km2
This study, 40 cm, 1:25,
247.6 m2

This study, 40 cm, 1:20,


162.0 m2 (exposed
area), 104.7 m2
(stippled area)
Ottesen Ellevset et al.
(1998), 200 m,
unknown, 110.7 km2

This study, 20 cm, 1:25,


49.5 m2

Source, Truncation Limit,


Mapping Scale,
Surface Area

Joints 010020

Joints 200, 267,


290, 318

Faults WSW

Faults variety of
orientations

Joints 200, 290,


316

Joints cross075
master165

Sets

Single set

Orthogonal pattern
(200 & 290); other
younger sets

Single set in terms of


orientation

Orthogonal pattern
(200 & 290) and
one younger joint
set
Somewhat polygonal

Ladder pattern with a


bimodal trace length
distribution

Pattern Geometry

Age of joints has been reported as


either pre-Eocene or post-Pliocene
(Segall and Pollard, 1983); joint sets
are of regional extent

Same as Amroth

165 set formed during Late


CretaceousEarly Miocene Alpine
compression (Nemcock et al., 1995);
075 set formed due to relaxation or
contraction at later stage (Rawnsley
et al., 1998)
200 and 290 set formed during
alternating r2 and r3 stress
directions during Variscan thrusting
(Dunne and North, 1990)
Late JurassicEarly Cretaceous
extension, possibly related to salt
pillow formation (Ottesen Ellevset et
al., 1998)
PliocenePleistocene normal faults
(Walsh et al., 1996)

Origin

*British National Grid Reference System (Universal Transverse Mercator system).


**Universal Transverse Mercator system for United States.

5000 ft grid based on Nevada State Plane coordinates, Nevada State Plane projection.

Joints 068, 274,


292, 356
Cambrian, Great Falls
Metagranite
Rocky Creek, South
Carolina, United
States, 511215.3 m
E** 3821843.5 m
N** (Figure 6H)

M. F. Schaeffer, 1998,
unpublished data, 25
cm, 1:120, 1446.6 m2
(exposed area), 922.2
m2 (exposed, stippled
area)

Multiple fracture sets;


length distribution is
independent of
orientation

Cooling joints formed first during


thermoelastic relaxation from
cooling; the tectonic joints postdate
cooling joints; based on termination
relationships, the 325 tectonic set is
oldest, followed by 359 set, with
040 set youngest. (Barton et al.,
1993)
Joints are at high angles and of
tectonic origin; greenschist facies
minerals along the joints 300 Ma;
the laumonitite mineralization
relates to Mesozoic rifting 150200
Ma (M. F. Schaeffer, 1999, personal
communication).
Orthogonal set of
cooling joints with
the 050 being
dominant; poorly
developed tectonic
joints
Joints cooling050,
320; tectonic010,
040, 325, 359
Barton et al. (1993), 20
cm, 1:50, 233.2 m2
(exposed area), 124.1
m2 (stippled area)
Miocene, Tiva Canyon
Tuff (upper lithophysal
zone)
P100, Yucca Mt.,
Nevada, United
States, 562051.8 ft
E 763407.7 ft N
(Figure 6G)

block size for the two dominant sets is about 2.5 m by


1 m. At each pavement, 100 circles with diameters
smaller than the block size and 100 circles with diameters larger than the block size (0.2 and 1.5 m at
Llantwit; 0.5 and 3.0 m at Amroth) were used to estimate intensity. The smaller circles yield estimates
with much greater variability than the larger circles
(Figure 8). Therefore, to reduce the variability of estimates, circles larger than mean block size are recommended for estimating fracture parameters. Similar
conclusions are applicable to the other estimators and
to fault networks interpreted from seismic cubes. For
example, the greatest estimator error occurs at Sleipner Vest and Cartier Trough for the smallest sampling
circles (Figure 7E, F).
Although these estimators are effective with samples from two-dimensional surfaces such as rock pavements and maps, they cannot be used effectively with
borehole imagery data because terminations inside the
borehole cannot be uniquely constrained and, hence,
m counts are subject to interpretation. Thus, none of
the existing estimation approaches are particularly effective with borehole data sets at present.
Furthermore, inferring subsurface joint patterns
from boreholes is, in general, problematic, because
boreholes are likely to be smaller than block size in
naturally fractured rocks. Perhaps this issue can be resolved by considering microfractures in core specimens
(Laubach, 1997; Ortega and Marrett, 2000), but implementation of the circle-based estimators on microfractures inside a solid core is beyond the scope of this
article.
Effects of Pattern Heterogeneity
To consider the effect of pattern heterogeneities and
circle size on estimator variance, intensity estimates for
fractures from the rock pavements were analyzed as a
function of circle size (Tables 2, 3; Figure 6). Fault
trace maps could not be included in the analysis because their greater size as compared to the joint patterns precluded the use of sampling circles with the
same size.
The fracture patterns that were selected for this
analysis encapsulate a wide range of likely fracture geometries and settings (Table 2; Figure 6). They include
a ladder geometry where cross-joint spacing is controlled by the spacing of master joints (Figure 6A); an
orthogonal pattern of large fractures overprinted by
younger, less persistent joints (Figure 6B); patterns
with clustered fractures due to fault development or
Rohrbaugh et al.

2097

Figure 6. Fracture trace maps. (A) Llantwit Major. (B) Amroth. (C) Sleipner Vest (modified from Ottesen Ellevset et al., 1998). (D)
Cartier Trough (modified from Walsh et al., 1996). Continued.

igneous processes (Figure 6E, G); and patterns in crystalline rocks that vary from a single fracture set in a
granite (Figure 6F) to a complex array of fractures that
2098

accumulated during a long uplift and erosion history


(Figure 6H). Characteristics of these patterns include
spatial heterogeneity due to different ages of joint sets,

Fracture Characterization Using Circular Scanlines and Windows

Figure 6. Continued. (E) Telpyn Point. (F) Ward Lake (modified from Segall and Pollard, 1983). (G) P100 (modified from Barton et
al., 1993). (H) Rocky Creek (modified from M. F. Schaeffer, 1998, unpublished data). Light-gray regions in (B), (F), (G), and (H)
represent visually identified homogeneous subdomains, and black regions in (F) and (H) are unexposed. See Table 2 for geologic
information for trace maps.

Figure 8. Variability of estimates as a function of circle size


with respect to block size (small circles open diamonds; large
circles black diamonds; see text for size details). Variability
is represented by the coefficient of variation (standard deviation
divided by the mean) for 100 circle counts for each sample.

Figure 7. Estimates of natural fracture characteristics using 10


circular scan lines/windows (diamonds) and areal samples
(dashed lines). (A) Density at Llantwit Major. (B) Mean trace
length at Llantwit Major (MJ master joints; CJ cross joints).
(C) Intensity at Amroth. (D) Intensity at Sleipner Vest. (E) Mean
trace length at Sleipner Vest. (F) Density at Cartier Trough. Estimates are in meters for Llantwit Major and Amroth and in
kilometers for Sleipner Vest and Cartier Trough.

varying degrees of persistence, clustering, and the local


heterogeneities arising from the mechanics of joint
origin.
Spatially homogeneous fracture domains were established initially by visual inspection. However, a
first-pass visual inspection to eliminate obvious heterogeneities was not sufficient to identify homogeneous
domains in all pavements. This insufficiency means
that the circle-based estimators could be tested both
for their performance in homogeneous regions and for
their ability to detect pattern heterogeneity in regions
of proposed homogeneity.
Intensity was sampled for the entire pavements at
Telpyn Point and Llantwit Major and for subdomains
at Amroth, Ward Lake, Yucca Mountain, and Rocky
2100

Creek (Figure 6). To simplify comparison among the


different fracture geometries and sampling schemes,
circle-based intensity estimates within 15% of the areal
value were considered accurate and are marked with Y
(for yes) in Table 3. Ranges of low (1 fracture per
m), moderate (11.99 per m), and high (2 per m)
were adopted for intensity magnitude to facilitate comparison (Table 3). Circle radii of 12 m were used to
achieve a high likelihood of intersections between fracture traces and circular scan lines as a function of block
size or fracture spacing, while not exceeding pavement
size. Ten circles were used to balance the need for a
large sample with the need to be able to do work in a
timely manner.
The intensity of seven single fracture sets and two
entire fracture patterns was evaluated (Table 3). Four
fracture sets and two fracture patterns yielded accurate
circle-based intensity estimates as compared to the estimate for the entire pavement. We interpret the
agreement between the two methods as indicating that
these patterns or fracture sets are spatially homogeneous. Three fracture sets, however, yielded inaccurate
results. Inspection of these pavements indicated that
the inaccuracy results from spatial heterogeneity. For
example, the 316 set at Amroth is more abundant at
the eastern end of the pavement subdomain (Figure 6),
so, unlike the 200 set, it is heterogeneously distributed. The cross joints at Llantwit Major are less intense
at the western end of the pavement, which is where

Fracture Characterization Using Circular Scanlines and Windows

Table 3. Accuracy of Intensity Estimates Using Ten Circular Scanlines for a Variety of Natural Fracture Patterns
Estimate Accuracy, Scanline Radius
Natural Fracture Patterns

1.0 m

1.5 m

2.0 m

Intensity** (m/m2)

Llantwit Major, cross joint set

Y
14.2
Y
5.4
N
49.5
Y
9.9
Y
15
Y
9.0
N
50.7
Y
7.6
Y
6.0

Y
8.8
Y
12.6
N
63.1
Y
10.6
Y
13.3
Y
14.0
Y
2.9
Y
9.7
Y
4.8

N
18.6
Y
7.7
N
35.9
Y
0.1
Y
2.5
Y
3.4
N
23.4
Y
8.2
Y
13.7

High
2.00
High
3.40
Low
0.55
Mod
1.80
Low
0.50
High
3.60
Mod
1.00
High
2.80
High
3.30

Llantwit Major, master joint set


Amroth, 316 set
Amroth, 200 set
Telpyn Point, 290 set
Telpyn Point, 200 set
Ward Lake, 015 set
Yucca Mountain, all fracture sets
Rocky Creek, all fracture sets

*Y(yes)/N(no) indicates whether estimate is/is not within 15% of the areal value; / denotes an over/underestimate.
**Actual percent error for the intensity estimate are shown in the second row for each fracture system. Low/moderate/high intensity indicates values in the ranges
(0:1), (1:2), or (2), respectively.

larger sampling circles are restricted because of pavement width (Figure 6). The subdomain at Ward Lake
eliminates a low intensity region for the fracture set
but contains heterogeneities due to fracture clusters
and covered areas that prevent random circle placement. Thus, the geometry of older joints, joint persistence, and joint clustering, as expected, generate spatial heterogeneity. More important, based on these
results, we believe that a sampling strategy of 10 circles
with a size that exceeds the block size or fracture spacing but is substantially smaller than the minimum dimension of a sample area will yield an intensity estimate within 15% of the actual intensity for a
homogeneous fracture pattern.
Minimum Count of m for Reliable Tracelength Estimates
Of the three estimators discussed here, the mean tracelength estimator exhibits the greatest variability. Low
trace densities, large fracture trace lengths relative to
circle size, or small total sample area can lead to small
counts of fracture endpoints (m). Such small counts
may be an issue for the petroleum industry be-

cause boreholes are commonly small with respect to


fractures. These small counts of m can lead to significant errors in the mean trace-length estimator (Table
1), because as m tends to zero in the denominator, the
estimate tends to infinity. Thus, it is important to ensure that a small m count is an adequate sample of a
fracture pattern with large trace lengths and not a sampling artifact due to insufficient sample size, thereby
producing an overestimate of mean trace length.
To evaluate this issue, five simple synthetic trace
patterns with a single fracture set of intensity 2 m/m2
were examined. The five cases were for trace patterns
with individual traces of length 10, 16, 20, 25, or 50
m, respectively. For each case, a suite of 35 sampling
strategies (combinations of circle size and number) was
used, with circle radii ranging from 1 to 20 m and number of circles ranging from 2 to 10. This combination
of cases and strategies yielded a total of 175 samples in
most of which trace lengths exceeded the circle diameters, consistent with the petroleum industry situation of fractures larger than borehole diameter.
The accuracy of circle-based estimates of mean
trace length, as determined by comparison with the
Rohrbaugh et al.

2101

input mean trace-length values, is graphed as a function of m counts in Figure 9. For this graph , an accurate
result was defined as within 15% of the input value.
Using this criterion, virtually all sampling strategies
that yielded m counts of 30 or greater produced accurate estimates. These results (Figure 9) suggest that
for a specified fracture intensity, a threshold value of
m counts exists for which accurate estimates are a near
certainty.
This analysis could not be extended directly to natural fracture patterns, because, due to problems of censoring and length bias, no direct technique exists to
accurately estimate mean trace length for comparison
to the results of the circle-based estimator. Although
threshold values of m counts for accurate estimates
could be determined for our synthetic patterns, our
models had Poissonian trace center distributions,
which is unlikely in nature. Still, results from the synthetic patterns should provide guidance, such as in the
case considered here, where a fracture pattern with
large fractures and an intensity of 2 m/m2 that is being
sampled by small circles (or small-diameter boreholes)
needs an m count greater than 30 to yield an accurate
estimate at the 15% level.

USING NEW ESTIMATORS WITH


FRACTURE GENERATORS
One way to evaluate fractures as factors in hydrocarbon transport and storage in reservoirs is to use computer programs to generate representative fracture networks in synthetic rock volumes (Dershowitz and
LaPointe, 1994; Swaby and Rawnsley, 1996; Renshaw,
2000). These synthetic networks are then modeled for
fluid flow and storage. The primary contribution that
the circle-based estimators make to such an approach
is providing more accurate estimates for program inputs and validating outputs. Intensity estimates can, for
example, substitute for dimensionally equivalent density or frequency input values in programs. In addition,
these estimates of density, size, and intensity provide
tests for determining whether synthetic networks are
truly representative.

Figure 9. Graph illustrating the performance of the mean


trace-length estimator as a function of the total number of m
counts in terms of the percentage of accurate estimates for a
given total m count.

2.

3.

4.

5.

CONCLUSIONS
6.
1. New circle-based estimators for fracture intensity,
density, and mean trace length virtually eliminate
orientation, censoring, and length biases, which se2102

verely limit the effectiveness of the straight scan line


and area methods.
Estimator accuracy is improved and variability reduced by using circles larger than mean fracture
block size or fracture spacing for a single set.
A sampling strategy of 10 circles with a diameter
that exceeds block size or fracture spacing but is
significantly less than the minimum dimension of a
sample region yields an intensity estimate within
15% of the actual intensity for a homogeneous fracture pattern.
When using small circles to estimate mean trace
length of large fractures, care should be taken to use
sampling strategies that gather a sufficient m count
to eliminate spurious overestimates of fracture size
from small samples. Guidance about the necessary
number of m counts can be gained from analyzing
synthetic patterns with similar characteristics.
The new estimators provide both inputs and independent checks of predictions for fracture-generator programs that model fracture populations in a
rock volume.
Fracture characterization protocols that deal with
sampling biases, such as the ones presented in this
article, are likely to yield improved input for exploration and development decisions.

Fracture Characterization Using Circular Scanlines and Windows

REFERENCES CITED
Baecher, G. B., and N. A. Lanney, 1978, Trace length biases in joint
surveys: Proceedings of the 19th U.S. Symposium on Rock Mechanics, v. 1, p. 5665.
Barton, C. C., E. Larsen, W. R. Page, and T. M. Howard, 1993,
Characterizing fractured rock for fluid-flow, geomechanical,
and paleostress modeling: methods and preliminary results from
Yucca Mountain, Nevada: United States Geological Society
Open-File Report 93-269, 62 p.
Becker, A., and M. R. Gross, 1996, Mechanism for joint saturation
in mechanically layered rocks: an example from southern Israel:
Tectonophysics, v. 257, p. 223237.
Cruden, D. M., 1977, Describing the size of discontinuities: International Journal of Rock Mechanics and Mining Sciences and
Geomechanical Abstracts, v. 14, p. 133137.
Dershowitz, W. S., and H. H. Herda, 1992, Interpretation of fracture
spacing and intensity, in J. R. Tillerson and W. R. Wawersik,
eds., Proceedings of the 33rd U.S. Symposium on Rock Mechanics: Rotterdam, Balkema, p. 757766.
Dershowitz, W. S., and P. R. LaPointe, 1994, Discrete fracture approaches for oil and gas applications, in R. A. Nelson and S.
Laubach, eds., Rock mechanics: Rotterdam, Balkema, p. 19
30.
Dunne, W. M., and C. P. North, 1990, Orthogonal fracture systems
at the limits of thrusting: an example from southwestern Wales:
Journal of Structural Geology, v. 12, p. 207215.
Einstein, H. H., and G. B. Baecher, 1983, Probablistic and statistical
methods in engineering geology, specific methods and examples, part 1: exploration: Rock Mechanics and Rock Engineering, v. 16, p. 3972.
Ghosh, A., and J. J. K. Daemen, 1993, Fractal characteristics of rock
discontinuities: Engineering Geology, v. 34, p. 19.
Gilmour, H. M. P., D. Billaux, and J. C. S. Long, 1986, Models for
calculating fluid flow in randomly generated three-dimensional
networks of disk-shaped fractures: theory and design of
FMG3D, DISCEL and DIMES: Berkeley, California, Lawrence
Berkeley Laboratory, Earth Sciences Division, LBL-19515,
144 p.
Jian, X., R. Olea, and Y. Yu, 1996, Semivariogram modeling by
weighted least square: Computers and Geosciences, v. 22,
p. 387397.
Kulatilake, P. H. S. W., and T. H. Wu, 1984, The density of discontinuity traces in sampling windows (technical note): International Journal of Rock Mechanics and Mining Sciences and
Geomechanical Abstracts, v. 21, p. 345347.
Kulatilake, P. H. S. W, D. N. Wathugala, and O. Stephansson, 1993,
Joint network modeling with a validation exercise in Stripa
Mine, Sweden: International Journal of Rock Mechanics and
Mining Sciences and Geomechanical Abstracts, v. 30, p. 503
526.
Kulatilake, P. H. S. W., R. Fiedler, and B. B. Panda, 1997, Box fractal
dimension as a measure of statistical homogeneity of jointed
rock masses: Engineering Geology, v. 48, p. 217229.
LaPointe, P. R., 1988, A method to characterize fracture density and
connectivity through fractal geometry: International Journal of
Rock Mechanics and Mining Sciences and Geomechanical Abstracts, v. 24, p. 421429.
LaPointe, P. R., 1993, Pattern analysis and simulation of joints for
rock engineering, in J. A. Hudson, ed., Comprehensive rock
engineering, volume 3rock testing and site characterization:
New York, Pergamon Press, p. 215239.
LaPointe, P. R., and J. A. Hudson, 1985, Characterization and interpretation of rock mass joint patterns: Geological Society of
America Special Paper 199, 37 p.

Laubach, S. E., 1997, A method to detect natural fracture strike in


sandstones: AAPG Bulletin, v. 81, p. 604623.
Marrett, R., O. J. Ortega, and C. M. Kelsey, 1999, Extent of powerlaw scaling for natural fractures in rocks: Geology, v. 27, p. 799
802.
Mauldon, M., 1998, Estimating mean fracture trace length and density from observations in convex windows: Rock Mechanics and
Rock Engineering, v. 31, p. 201216.
Mauldon, M., and W. Dershowitz, 2000, A multi-dimensional system of fracture abundance measures: Geological Society of
America Abstracts with Programs, v. 32, no. 7, p. A474.
Mauldon, M., and J. G. Mauldon, 1997, Fracture sampling on a cylinder: from scan lines to boreholes and tunnels: Rock Mechanics
and Rock Engineering, v. 30, p. 129144.
Mauldon, M., W. M. Dunne, and M. B. Rohrbaugh Jr., 2001, Circular scan lines and circular windows: new tools for characterizing the geometry of fracture traces: Journal of Structural Geology, v. 23, p. 247258.
Narr, W., 1996, Estimating average fracture spacing in subsurface
rock: AAPG Bulletin, v. 80, p. 15651586.
Nelson, R. A., 1985, Geologic analysis of naturally fractured reservoirs: Houston, Texas, Gulf Publishing, 320 p.
Nemcock, M., R. Bayer, and M. Miliorizos, 1995, Structural analysis
of the inverted Bristol Channel Basin: implications for the geometry and timing of fracture porosity, in J. G. Buchanan and
P. G. Buchanan, eds., Basin inversion: Geological Society Special Publication 88, p. 355392.
Ortega, O. J., and R. Marrett, 2000, Prediction of macrofracture
properties using microfracture information, Mesaverde Group
sandstones, San Juan basin, New Mexico: Journal of Structural
Geology, v. 22, p. 571587.
Ottesen Ellevset, S., R. Knipe, T. Svava Olsen, Q. J. Fisher, and G.
Jones, 1998, Fault controlled communication in the Sleipner
Vest field, Norwegian continental shelf: detailed, quantitative
input for reservoir simulation and well planning, in G. Jones,
Q. J. Fisher, and R. J. Knipe, eds., Faulting, fault sealing and
fluid flow in hydrocarbon reservoirs: Geological Society Special
Publication 147, p. 283297.
Peacock, D. C. P., S. J., Harris, and M. Mauldon, in press, Use of
curved scan lines and boreholes to predict fracture frequencies:
Journal of Structural Geology, v. 25, p. 109119.
Pickering, G., J. M. Bull, and D. J. Sanderson, 1995, Sampling
power-law distributions: Tectonophysics, v. 248, p. 120.
Priest, S. D., 1993, Discontinuity analysis for rock engineering: New
York, Chapman and Hall, p. 5054.
Priest, S. D., and J. A. Hudson, 1981, Estimation of discontinuity
spacing and trace length using scan line surveys: International
Journal of Rock Mechanics and Mining Sciences and Geomechanics Abstracts, v. 18, p. 183197.
Rawnsley, K. D., D. C. P. Peacock, T. Rives, and J.-P. Petit, 1998,
Joints in the Mesozoic sediments around the Bristol Channel
Basin: Journal of Structural Geology, v. 20, p. 1641
1661.
Reiss, L. H., 1982, The reservoir engineering aspects of fractured
formations: Houston, Texas, Gulf Publishing, 186 p.
Renshaw, C. E., 2000, An example of the use of geological and mechanical constraints to develop a conceptual model of flow in
fractured, granitic rock: Geological Society of America Abstracts with Program, v. 32, no. 7, p. A64.
Rohrbaugh Jr., M. B., 2000, Estimating joint intensity, density, and
mean trace length using circular scan lines and circular windows: M.S. thesis, University of Tennessee, Knoxville, 96 p., 1
CD-ROM.
Segall, P., and D. D. Pollard, 1983, Joint formation in granitic rock
of the Sierra Nevada: Geological Society of America Bulletin,
v. 94, p. 563575.

Rohrbaugh et al.

2103

Stoyan, D., W. S. Kendall, and J. Mecke, 1995, Stochastic geometry


and its applications, 2d ed.: New York, John Wiley and Sons,
p. 286296.
Swaby, P. A., and K. D. Rawnsley, 1996, An interactive 3D fracture
modeling environment: Society of Petroleum Engineers, SPE
36004, p. 177187.
Terzaghi, R. D., 1965, Sources of error in joint surveys: Geotechnique, v. 15, p. 287304.
Turner, F. J., and L. E. Weiss, 1963, Structural analysis of metamorphic tectonites: New York, McGraw-Hill, 545 p.

2104

Walsh, J. J., J. Watterson, C. Childs, and A. Nicol, 1996, Ductile


strain effects in the analysis of seismic interpretations of normal
fault systems, in P. G. Buchanan and D. Nieuwland, eds., Modern developments in structural interpretation, validation, and
modelling: Geological Society Special Publication 99, p. 2740.
Whitten, E. H. T., 1966, Structural geology of folded rocks: Chicago,
Rand McNally, 678 p.
Wu, H., and D. D. Pollard, 1995, An experimental study of the
relationship between joint spacing and layer thickness: Journal
of Structural Geology, v. 17, p. 887905.

Fracture Characterization Using Circular Scanlines and Windows

Anda mungkin juga menyukai