Вы находитесь на странице: 1из 19

Seismic attribute-based

characterization of coalbed
methane reservoirs: An example
from the Fruitland Formation,
San Juan basin, New Mexico
Ivan Dimitri Marroqun and Bruce S. Hart

ABSTRACT
The Fruitland Formation of the San Juan basin is the largest producer of coalbed methane in the world. Production patterns vary
from one well to another throughout the basin, reflecting factors
such as coal thickness and fracture and cleat density. In this study,
we integrated conventional P-wave three-dimensional (3-D) seismic and well data to investigate geological controls on production
from a thick, continuous coal seam in the lower part of the Fruitland
Formation. Our objective was to show the potential of using 3-D
seismic data to predict coal thickness, as well as the distribution and
orientation of subtle structures that may be associated with enhanced permeability zones. To do this, we first derived a seismic
attribute-based model that predicts coal thickness. We then used
curvature attributes derived from seismic horizons to detect subtle
structural features that might be associated with zones of enhanced
permeability. Production data show that the best producing wells
are associated with seismically definable structural features and thick
coal. Although other factors (e.g., completion practices and coal type)
affect coalbed methane production, our results suggest that conventional 3-D seismic data, integrated with wire-line logs and production data, are useful for characterizing coalbed methane reservoirs.

INTRODUCTION
The San Juan basin is the worlds largest producer of coalbed methane. Most coalbed methane in this basin is produced from coal
seams of the Upper Cretaceous Fruitland Formation. Production
began in 1951 with the Ignacio Blanco-Fruitland gas field at Ignacio,

Copyright #2004. The American Association of Petroleum Geologists. All rights reserved.
Manuscript received August 25, 2003; provisional acceptance December 31, 2003; revised manuscript
received May 3, 2004; final acceptance May 19, 2004.
DOI:10.1306/05190403088

AAPG Bulletin, v. 88, no. 11 (November 2004), pp. 1603 1621

1603

AUTHORS
Ivan Dimitri Marroqun  Earth and
Planetary Sciences, McGill University, 3450
University, Montreal, Quebec, Canada H3A 2A7
Ivan Dimitri Marroqun is a Ph.D. student of
geophysics at McGill University, Canada. He
received his B.Sc. combined degree in physics
and geology from the Universite de Montreal
in 1994 and his M.Sc. degree in geophysics
from Ecole Polytechnique in 1998. His current
research includes characterization of conglomerate and coalbed reservoirs by 3-D seismicbased means.
Bruce S. Hart  Earth and Planetary
Sciences, McGill University, 3450 University,
Montreal, Quebec, Canada H3A 2A7;
hart@eps.mcgill.ca
Bruce Hart holds a Ph.D from the University
of Western Ontario. He held positions with the
Geological Survey of Canada, Penn State, and
the New Mexico Bureau of Mines and Mineral
Resources prior to joining McGill University
in 2000. His research interests focus on the
integration of 3-D seismic data with other data
types for reservoir characterization programs.
He has been an associate editor of the AAPG
Bulletin since 2000.
ACKNOWLEDGEMENTS
Principal funding for this project was provided
by a Department of Energyfunded investigation of optimizing infill drilling in naturally
fractured tight-gas reservoirs (contract number
DE-FC26-98FT40486; Larry Teufel, principal investigator). Portions of this project were funded
by a Natural Science and Engineering Research
Council award (238411-01) to Hart. We thank
Williams Energy (in particular, Ralph Hawks) and
Tom Engler (New Mexico Tech) for supplying
data and insights into Rosa field. Software used
in this study was graciously provided by Landmark Graphics Corporation (GeoGraphix) and
Hampson-Russell. We thank both companies
for their ongoing support of our research and
former AAPG editor John Lorenz and two anonymous reviewers for their helpful comments
that helped improve the focus of this paper.

Colorado. Today, Fruitland coalbed methane wells have


a cumulative production of 9 tcf (0.25  1012 m3),
an annual production of 0.9 tcf (0.025  1012 m3),
and contain in-situ gas resources estimated at 55 tcf
(1.6  1012 m3; Dugan, 2002).
The production patterns in the Fruitland vary on
basinwide and local scales, reflecting differences in geologic settings and engineering factors (e.g., completion).
Geologic factors affecting the capacity of coal seams to
store and produce coalbed methane include coal composition and rank, gas composition, water production,
coal thickness, and degree of fracturing (e.g., Kaiser
and Ayers, 1994). An understanding of the controls
acting on these seams can be used to guide exploration
and development activities. Despite the increasing importance of coal seams as sources of methane gas, little
seismic data have been acquired to aid in characterizing
coalbed methane reservoirs (Shuck et al., 1996). Shuck
et al. (1996) and Ramos and Davis (1997) used seismic
data to help characterize a coalbed methane reservoir
in the Fruitland Formation at Cedar Hill field in the
San Juan basin (Figure 1). These studies sought to determine fracture location, direction, and density. Shuck
et al. used multicomponent seismic data, whereas Ramos
and Davis used both amplitude variation with offset
(AVO) and modeling analysis. Selected aspects of these
studies are summarized in the Discussion.
This paper aims to show how two of the primary
geologic controls on coalbed methane production, coal
thickness and subtle structures that may be associated
with enhanced permeability zones, may be predicted
from conventional (P-wave) three-dimensional (3-D)
seismic data. We use 3-D seismic attributes to generate
a model that predicts the thickness of the coal seam and
then use curvature analysis to delineate subtle structural features. We note that the best production trends
are associated with thick coal and seismically detectable
structural lineaments. This methodology should have
application in other coalbed methane plays. Issues that
arose during the course of our attribute analyses shed
insights on how to address certain technical aspects of
these studies.

GEOLOGICAL FRAMEWORK
The San Juan basin lies in the east-central part of the
Colorado Plateau in northwestern New Mexico and
southwestern Colorado. This basin is roughly circular
and covers an area of about 6700 mi2 (17,353 km2;
1604

Laubach and Tremain, 1994). During the Late Cretaceous, shoreline progradation (to the northeast) led
to deposition of the northwest-southeast striking
Pictured Cliffs Sandstone (Ayers et al., 1994). Intermittent transgressive-regressive shifts of the shoreline
resulted in the interfingering of the upper tongues of
the Pictured Cliffs Sandstone and continental facies of
the overlying Fruitland Formation. The Fruitland Formation is the primary coal-bearing formation in the San
Juan basin.
The Laramide orogeny began during the Late Cretaceous and continued into the Eocene. Various aspects of this tectonic event have been discussed by
Hamilton (1987) and Lorenz and Cooper (2003). The
San Juan basin is structurally delimited by the Hogback
monocline to the northwest, Archuleta anticlinorium
to the northeast, Nacimiento uplift to the east, Defiance uplift to the west, and the Zuni uplift to the south.
Coalification of the Fruitland began during the Laramide orogeny, and this tectonic activity may have influenced the development of cleats (the term applied to
fractures in coals), minor folds, and faults in this unit.
Based on the variability of cleat strike in Fruitland coals,
Tremian et al. (1994) noted the existence of two principal face-cleat domains, northwest-southeast strike
in the north of the basin (domain 2), and northeastsouthwest strike in the south (domain 1), separated by
an east-trending transition zone (Figure 1). They stated
that permeability in the face-cleat direction should
be greater than in the butt-cleat direction because of
greater connectedness in the former direction. Laubach
and Tremain (1994) suggested that fracture swarms,
which are associated with increasing fracture and cleat
connectivity and intensity, should be considered as exploration targets in coalbed methane reservoirs.

DATABASE
Our database consisted of a time-migrated 3-D seismic
data volume and wire-line logs from 59 wells (Figure 1)
within and around the survey area (27 of which were
in the seismic area). Records of gas production from
36 wells in the seismic area were used to analyze spatial
trends in coalbed methane production. Unfortunately,
we did not have wire-line logs for all of these wells. The
3-D seismic grid covers an area of about 17 mi2 (45 km2)
with a bin size of 150  150 ft (45  45 m). The seismic
data were acquired with a northeast-southwest orientation and then processed to have crosslines oriented

Seismic Attribute-Based Characterization of Coalbed Methane Reservoirs

Figure 1. Location map showing cleat strike domains of the Fruitland Formation coal seams in the San Juan basin and major
structural elements bounding the basin (modified from Tremain et al., 1994). The study area is in a boundary region between a
region (domain 1) dominated by northeast-southwest striking face cleats and a region (domain 2) dominated by northwestsoutheaststriking face cleats. The inset at the upper right shows the 3-D seismic area with well locations. Line AA0 is the location of
the log cross section shown in Figure 2.

north-south and inlines running east-west. Seismic


data quality is variable but is uniformly poor around
the margins of our data set.
Various combinations of gamma-ray, caliper, density, and resistivity logs were available for wells; sonic
logs were not available for any of the wells in our study
area. As such, we generated pseudosonic logs from den-

sity logs using the Gardner equation (Gardner et al.,


1974). This equation provided unreasonably low velocities values for coal because coals fall off the main
trend used to derive the equation. Accordingly, we
assumed a coal velocity of 8695 ft/s (2650 m/s) based
on handbooks published by geophysical logging companies. Sonic logs were used for modeling purposes
Marroqun and Hart

1605

Figure 2. Log cross section through the Fruitland Formation. The logs, from left to right, are the gamma ray (GR) and bulk density
(RHOB). Note the presence of the thick coal seam overlain by a succession of thin coal seams and the general thickening trend of this
seam to the southeast (right). See Figure 1 for location.

and to make well ties. No checkshot surveys or other


sources of velocity information were available. Image
logs, oriented core, or other independent means of detecting fracture orientation were not available to us.

RESULTS
Geologic Analysis
We identified coal seams and established the vertical
and lateral extents of the Pictured Cliffs Sandstone and
Fruitland Formation using gamma-ray and density logs.
Figure 2 shows a sample log cross section through the
Fruitland Formation in the seismic area. The top of
the Pictured Cliffs Sandstone was identified by high
gamma-ray and low-density log values. This hot shale
signature is thought to be a marine flooding surface.
Upper Pictured Cliffs tongues have a blocky well-log
response and are considered to be stratigraphically located between the Picture Cliffs Sandstone and lowermost Fruitland coal seam (e.g., Ayers et al., 1994).
These strata were found to have an average thickness of
135 ft (41 m). Coal is identified on logs by a combination of low-density and low gamma-ray values. Following Ayers et al. (1994), we identified coal using
density and gamma-ray cutoff values 2.0 g/cm3 and
75j API, respectively. Two coal-bearing intervals were
delineated: a thick, continuous coal seam in the lower
part of the Fruitland Formation (with an average thick1606

ness of 26 ft [8 m]) and an upper succession of thin,


laterally less continuous coal seams interbedded with
clastic strata. The thick accumulation of the lower Fruitland coal seam is believed to have been deposited
during periods of shoreline stillstands (cf. Ayers et al.,
1994) and produces most of the coalbed methane in
the study area. The overlying succession of thin, discontinuous coal seams is interpreted to have been
deposited in unstable flood-plain settings related to the
renewed progradation of the shoreline. An irregular
low thickness value (2.89 ft [0.88 m]) was observed for
the lower continuous seam in one well at the extreme
west margin of the seismic area. We were not able to
determine the cause of this anomalous value (e.g., faulting); the well is located in an area of poor seismic data
quality, and so it was disregarded from further analyses.
We hand drew the isopach map of the thickness of
the lower coal seam in and around the 3-D seismic
survey area (Figure 3). Our intent was only to generate
a map that shows the principal trends in the data. We
note that the coal is thicker in the southeastern part of
the seismic area and shows a northwest-southeast trend
that is approximately parallel to the inferred Pictured
Cliffs shoreline orientation. The coal seam generally
thins toward the north part of the seismic area.
Seismic Character of Coal Seams and Related Rocks
We generated seismic models by convolving a Butterworth zero-phase wavelet of bandwidth 5 15 55
65 Hz (picked to approximate frequency characteristics

Seismic Attribute-Based Characterization of Coalbed Methane Reservoirs

Figure 3. Hand-drawn isopach


map of the lower Fruitland coal
seam with 5-ft (1.5-m) contour
intervals showing general thickness trends. Symbols show data
control points. The coal seam
is thickest in the southeastern
part of the study area.

of our 3-D seismic data; see below) with bulk density


and our pseudosonic logs from wells in the seismic area.
An example of a seismic model is shown in Figure 4.
From this example, we note that the top of the Fruitland Formation corresponds to a zero crossing, although
for ease of picking, we interpreted the Fruitland horizon
as a moderate-amplitude continuous peak in the 3-D

seismic data. A moderate-amplitude trough observed


in the seismic model, between the Fruitland and the
top of the thick coal seam, represents the succession
of thin coal seams. We interpreted this horizon in the
3-D seismic data as a continuous and moderate- to
weak-amplitude trough. This reflection is caused by
the interference effects from the individual reflections

Figure 4. Seismic model showing the


predicted response of the Fruitland coal
seams and related strata. Line generated
using wells shown in Figure 2. Seismic
horizons: FRLD = top of Fruitland Formation,
TopTC = top thick coal seam, BottomTC =
bottom of thick coal seam, MainPC = top
of the Pictured Cliffs Sandstone.

Marroqun and Hart

1607

Figure 5. Arbitrary seismic transect


showing the interpreted seismic horizons.
Transect corresponds to seismic model
shown in Figure 4. Seismic horizons:
FRLD = top of Fruitland Formation,
TopTC = top thick coal seam, BottomTC =
bottom of thick coal seam, MainPC =
top of Pictured Cliffs Sandstone.

associated with the succession of thin coal seams.


Interference results from short-path multiples, such
that their overlapping waveforms superpose and
outweigh the amplitude of the initial primary reflection (Gochioco, 1991, 1992). The top of the thick coal
seam, on both seismic model and 3-D seismic data, is a
robust, continuous, high-amplitude trough. The bottom
of the thick coal seam, on the seismic model, corresponds to an intermediate position on the waveform
that is close to the reflection peak, and therefore, we
interpreted this horizon on the 3-D seismic data as the
underlying robust, high-amplitude peak. Moreover,
the bottom of this coal defines the top of the underlying
upper Pictured Cliffs tongues. The top of the Pictured
Cliffs Sandstone, in the seismic model, corresponds to a
trough of weaker amplitude. The corresponding horizon on the 3-D seismic data is a fairly low-amplitude
trough of poor quality. The interpreted horizons in the
3-D seismic data are shown in Figure 5 on an arbitrary
northeast-southwest seismic line trough the survey area.
A statistical wavelet was extracted from the 3-D
seismic data to determine the frequency content of the
data in a time window that covers the horizons of
interest. We observed a predominant peak frequency
( f p) of 40 Hz. By assuming an interval velocity (V ) of
8695 ft/s (2650 m/s ) for coal seams (see above), we
calculate the dominant wavelength (l) to be
l V =fp 8695 ft=s = 40 Hz  217 ft 66 m 1
1608

yielding a tuning thickness (l/4) of 54 ft (16 m). This


value is at least twice the average thickness of the lower
Fruitland coal seam in our study area. Therefore, we
conclude that this seam is a seismic thin bed, for which
amplitude variations have been used to estimate the
true coalbed thickness (Gochioco, 1991; Brown, 1999).

SEISMIC ATTRIBUTE ANALYSIS


Seismic attribute studies are useful for integrating
wire-line log and 3-D seismic data to make predictions
of physical properties (e.g., porosity and thickness)
throughout a 3-D seismic survey area (e.g., Schultz
et al., 1994; Hampson et al., 2001). We used a windowbased approach (Chen and Sidney, 1997) to derive a
statistical relationship between coal thickness and seismic attributes (Taner et al., 1979; Brown, 1999; Chen
and Sidney, 1997). The time window for seismic attribute extraction was defined by the picks for the
top and base of the thick coal seam. The thickness of
this seam in each well was determined from the bulk
density log. The multiattribute analysis was based on
27 geophysical well logs in the 3-D seismic survey and
35 amplitude, frequency, and phase attributes. This
list was further increased by applying the following
nonlinear transforms to each attribute: natural log, exponential, square, inverse, and square root.

Seismic Attribute-Based Characterization of Coalbed Methane Reservoirs

Figure 6. Validation error


and prediction error plotted
against the number of attributes used in the stepwise
linear regression. The validation
error indicates that a combination of three attributes best
models predicted coal thickness
(cf. Hampson et al., 2001).

The choice of attributes to use in our regression


expression was made using statistical methods described by Hampson et al. (2001). The attributes were
first analyzed to determine which single attribute correlated the most strongly with coal thickness (both
values defined at well locations). This attribute was then
paired with all the remaining attributes to determine
which combination of attributes (e.g., two, three, or
more attributes) best predicted coal thickness. As more
attributes are added into the regression model, the prediction error will continuously decrease and consequently improve the fit to the data. This could lead to
overfitting of the data, however, so cross-validation
was used to determine the optimum number of attributes to keep (Hampson et al., 2001). This method consists of dividing the data into two subsets (e.g., the
training data and validation data). Although the training data set is used to train the model, the validation
data set is used to estimate the validation error. The
validation error is a measure of the quality of the fit to
the validation data set. Therefore, the point at which
adding new attributes starts to increase the validation
error determines the optimum number of attributes to
use in the analysis (Hampson et al., 2001).
Figure 6 shows the total validation error and prediction error against the number of attributes used in
the stepwise linear regression. We note that the validation error curve begins to increase after the third attribute, defining this as the optimum number of attributes
to be used. The analysis yielded the following linear
relationship:
Thickness 3:922 2:005e
8X1 3:183e9 1=X2
6:812e18 1=X3

where
X1 is maximum absolute amplitude; X2 is integrated
trace; and X 3 is total energy;
with a correlation coefficient of R 2 = 0.87 and average
error of 3.2%. Figure 7 shows the crossplot of the predicted thickness vs. the actual thickness values.
The attributes composing the regression expression are defined as follows:
The maximum absolute amplitude is the maximum
value in the analysis window interval. Theoretically,
below tuning thickness (e.g., the range of measured
coalbed thicknesses in the present study) amplitude
is directly proportional to bed thickness because of
the effects of destructive interference from reflections at the top and base of the bed. This trend is
observed in our data (Figure 8a). This attribute is
the single best predictor of bed thickness, with a
correlation coefficient of 0.84.
 The integrated trace is essentially a band-limited
(recursive) inversion, with low impedance being
represented by negative numbers. The value we used
is the average value of the integrated trace in the
analysis window. For a thinning wedge of constant
acoustic impedance, the inversion result will best
estimate the impedance of thick beds, but will not be
able to reproduce the impedance of thinner beds,
such as the coal bed studied in this paper. Thick coals
in our data are associated with strong negative values
of integrated trace, and thin coals have weaker
negative values (Figure 8b). As such, there is a strong
negative correlation between integrated trace and
thickness. Strangely, equation 2 suggests an inverse,


Marroqun and Hart

1609

poor input data quality, although it is also possible that


the model performs poorly when extrapolating beyond
the range of the input data. We thus tested the performance of the model by searching for hidden extrapolation points (Montgomery and Peck, 1992). This
procedure consists of defining the smallest possible
ellipsoid enclosing the original data points. If a given
data point lies outside of the boundary region, the prediction involves an extrapolation point. For this test, we
used the diagonal elements h ii of the hat matrix H =
X(X0X)
1X0, where the matrix X is made up of the
attribute values at well locations (Montgomery and
Peck, 1992). The largest value of h ii (e.g., h max) defines
the external boundary of the ellipsoid. Any point that
satisfies

Figure 7. Crossplot of the predicted thickness vs. the actual


thickness values. Note the tight fit and the presence of a single
low thickness value.

not a negative, correlation between integrated trace


and coal thickness.
 The total energy is the sum of all squared trace amplitudes in the window interval. Like the maximum
absolute amplitude, there is a strong positive correlation between this attribute and coal thickness
(Figure 8c). As was the case for the integrated trace
attribute, the inverse relationship between total energy and bed thickness implied by equation 2 seems
counterintuitive.
We suggest that equation 2 integrates the three
attributes in a way that captures subtle variations in
waveform shape. Examination of that equation suggests that the inverse transform attributes (integrated
trace and total energy) only have a significant influence
on the result when the linear attribute (maximum absolute amplitude) is small. They fine-tune the correlation at low-coal thickness values. This is consistent with
the only slight improvement in correlation coefficient
that was obtained using all three attributes (0.87) vs.
the maximum absolute amplitude (0.84). The details of
this interaction are investigated below through forward
modeling.
We obtained some negative values, because our
model (equation 2) is statistical in nature, and these
values were set to zero in our map. These values, as well
as some exceptionally high thickness values close to the
margins of the seismic area, are thought to be caused by
1610

h00 x0 X 0 X
1 x
h00 hmax

is considered to be an interpolation point, and x is a


vector of attribute values at any location in the seismic area. The largest value of h ii, 0.934, is much larger
than the second next larger value, 0.249, and therefore
could be an outlier (Montgomery and Peck, 1992).
However, this observation corresponds to the lowest
measured coal thickness (e.g., 11.85 ft [3.6 m]), and we
demonstrate the importance of this point to the regression analysis in the paragraph below. We therefore
decided to use this value as h max. Table 1 shows performance test results for six points across the seismic
area (locations shown in Figure 9). We note that the
three first h 00 values exceed h max, and consequently
are extrapolation points. These results confirm our previous observations concerning the poor data quality on
the margins of the seismic area. The rest of the h 00
values lie in the boundary and therefore represent interpolation points.
Composite amplitude is commonly used to estimate bed thickness below tuning (e.g., Gochiocco, 1991;
Brown, 1999), although this attribute was not selected
for our analysis using stepwise linear regression analysis. Composite amplitude is the sum of the absolute
amplitudes of reflections identified at the top and base
of the reservoir (Brown, 1999). We measured the Spearmans rank correlation coefficient (r s) between various
amplitude attributes and coal thickness to determine
how composite amplitude ranked in comparison to
other attributes we had generated. The Spearmans
rank correlation coefficient is a measure of the monotonic association between two variables (Lehmann and
DAbrera, 1975). Unlike the correlation coefficient,

Seismic Attribute-Based Characterization of Coalbed Methane Reservoirs

Figure 8. Crossplots of the maximum absolute amplitude (a, d), integrated trace (b, e), and total energy (c, f ) against coal seam
thickness or wedge model thickness. Attributes extracted from the seismic data are in the left column, and attributes extracted from
the wedge model are in the right column. Note the similarity in trends between model results and data.
Marroqun and Hart

1611

Table 1. Hidden Extrapolation Test Results Showing the


Performance of the Regression Model to Predict Coal Thickness
Data Point
1
2
3
4
5
6

h 00

Category

132.780
2.660
223.825
0.254
0.697
0.053

hidden extrapolation
hidden extrapolation
hidden extrapolation
interpolation
interpolation
interpolation

Spearmans rank correlation coefficient works on ranked


data and measures both linear and nonlinear relationships between variables. Table 2 shows that maximum
absolute amplitude and total energy attributes have a
high-ranking correlation with coal thickness (e.g., r s =
0.74 and 0.72, respectively). However, the integrated
trace has a moderate correlation of
0.62. We also note
that composite amplitude has a strong correlation with
coal thickness (0.68). According to Hampson et al.
(2001), the stepwise linear regression methodology

will not select attributes that provide essentially the


same information to the predicted model. As such, the
composite amplitude attribute must provide the same
information provided by the maximum absolute amplitude and/or total energy, both of which scored higher
in Table 2.
We undertook the modeling of various bed thickness and lithology profiles (i.e., do the beds above and
below the coal have the same acoustic impedance?) in
an attempt to determine under what circumstances
composite amplitude might provide a better result than
maximum absolute amplitude. Our results indicated
that different combinations of bed thickness and lithology profiles determine which of the two attributes correlates best to bed thickness. However, the results are
complex, in ways that we do not yet fully understand.
We generated a simple wedge model to examine
the physical basis of the observed empirical relationships between attributes and coal thickness (Figure 10).
This model was generated using the Butterworth zerophase wavelet employed in our original seismic models (Figure 5). From the wedge model results, we extracted the same attributes identified in equation 2,

Figure 9. Attribute-based
thickness map of the lower
Fruitland coal seam. The thickest
coal is predicted for the southeastern region of the survey area,
in accordance with the well
data (Figures 2, 3). Northwestsoutheast striking thickness
trends, delineated by yellow,
red, and green colors, are evident in the southeast part of
the survey area. The numbered
points (1 6) show the location
of data points used to test the
performance of the regression
model to predict coal thickness
(Table 2). Points 13 are identified
as extrapolation points caused
by bad data quality around the
survey margin. Our porosity prediction is not expected to be valid
in these areas. Points 46 are
identified as interpolation points
that are within the expected range
of attribute variation, so the porosity values are considered to
be valid in these areas.
1612

Seismic Attribute-Based Characterization of Coalbed Methane Reservoirs

Table 2. Spearmans Rank Correlation Coefficient between


Seismic Amplitude Attributes and Coal Thickness
Amplitude Attribute
Maximum absolute amplitude
Total energy
Variance in amplitude
Kurtosis amplitude
Average energy
Root mean square amplitude
Composite amplitude
Average reflection strength
Average absolute amplitude
Total absolute amplitude
Integrated trace
Trace length
Amplitude envelope
Total amplitude
Reflection strength slope
Integrate absolute amplitude
Mean amplitude
Amplitude
Skew amplitude

Spearmans Rank
Correlation Coefficient
0.74
0.72
0.72
0.70
0.69
0.69
0.68
0.66
0.63
0.62

0.62
0.58
0.58

0.52

0.49
0.48

0.47

0.41
0.23

then crossplotted them against coal thickness. The


results (Figure 8d, f) show the same trends as those
observed in the data (Figure 8a, c), leading us to conclude that they are truly responding to variations in coal

seam thickness (i.e., the observed correlations are not


spurious; cf. Kalkomey, 1997). On a 3-D crossplot of
attributes against coal thickness (Figure 11a), the lowest coal thickness point appears to be an outlier that
may be associated with bad data. However, an identical
plot of the model results (Figure 11b) shows a trend
that neatly reproduces the trend observed in the data.
This observation leads us to conclude that this was not a
bad data point, therefore justifying its inclusion in the
extrapolation point analyses described previously. Our
previous efforts with seismic modeling to validate
attribute studies (e.g., Hart and Balch, 2000; Hart and
Chen, in press; Tebo and Hart, in press) have likewise
provided important information.
Our attribute-based map of coal seam thickness
(Figure 9) suggests that coal is generally thicker in the
southeast portion of the seismic area, and thickness
patterns show a northwest-southeast trend. This belt of
thicker coal coincides with the local depocenter observed
from the log-based isopach map (Figure 3). We believe
that the thickness trends are geologically plausible because they are consistent with depositional models and
regional mapping of northwest-southeast striking
belts of thicker coal associated with paleoshoreline positions (e.g., Ayers et al., 1994). Furthermore, the area
of thicker coal in the southeast part of the study area
overlies a similar relative thickening at the level of the
underlying Dakota Formation (Hart and Chen, in press),
suggesting that this may have been a localized area of
enhanced subsidence during the Upper Cretaceous.

Figure 10. Wedge


model and corresponding
synthetic traces. Attributes
extracted from this model
were used to investigate
the relation between attributes composing the regression model and coal
thickness (Figure 8d, f; 11b).

Marroqun and Hart

1613

Figure 11. Three-dimensional crossplots of coal


thickness vs. the maximum
absolute amplitude and
a combination of inverse
transform attributes (i.e.,
integrated trace and total
energy). (a) Points showing
attributes from seismic data
and well-based coal thickness. (b) Line showing continuous variation in attributes and thickness along
the wedge model (Figure 10).
Note the similarity in trends
between the two results. A
single low-coal thickness
data point appears to be an
outlier, but its presence is
predicted by the model results. The incorporation of
the integrated trace and
total energy attributes into
our regression analysis
(equation 3) helps the
model to fit this point.

Curvature Attribute Analysis


Important opening-mode fractures contribute to
permeability pathways for gas in coalbed methane
reservoirs. Strata that have been faulted or folded may
also influence gas production (Pashin, 1998). Kaiser
and Ayers (1994) pointed out that minor folds and
faults in the San Juan basin enhance fracture and cleat
permeability in the Fruitland.
Horizon attributes, such as various types of curvature, have been successfully used to detect subtle
faults in reservoirs (e.g., Steen et al., 1998, Roberts,
2001). Hart et al. (2002) showed that horizon attributes might also be used to detect highly productive
fracture swarms in tight-gas reservoirs. Accordingly,
we sought to investigate whether curvature maps could
be used to predict the location and orientation of subtle
structural features too small to be detected on vertical
seismic transects, time slices, or horizon slices. We derived 11 curvature attributes (Roberts, 2001) and computed these attributes from the top thick coal seismic
horizon using aperture values of one, three, five, and
seven bins. The aperture value defines the size of a 3  3
grid of points used in the computation of curvature
attributes (Stewart and Wynn, 2000; Roberts, 2001).
1614

With an aperture of one bin, our maps were excessively


noisy. As larger aperture values were used, the resulting
curvature horizon shrank inward from the margins
of the survey and curvature lineaments identified only
regional trends. We thus defined the best results in
terms of delineation and distribution of curvature lineaments. We found that the maximum-, strike-, and
dip-curvature attributes defined below (all derived using
an aperture of three bins) provided the most geologically reasonable results. A depth-converted structure
map, generated by integrating log picks and seismic
horizons (methodology described by Hart, 2000), and
shaded relief maps of the top thick coal seismic horizon helped us to identify and interpret other structural features.
Figure 12 shows the depth-converted structure
map of the top of the thick coal (TopTC seismic pick).
The map shows that the study area is crossed by a
northwest-southeaststriking structural low that approximately coincides with the trend of the thick coal
inferred from both log-based and attribute-based thickness maps (Figures 4, 9, respectively). Shaded relief
maps of this surface (Figure 13a, b) show subtle structural features that are not readily apparent on the
depth-converted structure map, including a variety of

Seismic Attribute-Based Characterization of Coalbed Methane Reservoirs

Figure 12. Depth-converted


structure map (elevation above
sea level in feet) of the top of
the thick coal seam. Note the
northwest-southeast structural
low access cutting the seismic
area and contours, indicating
poor data quality around the
western and northern margins
of the survey area.

northwest-southeast and northeast-southwesttrending


structures. The poor data quality along the western margin of the survey is obvious from this display.
Maximum curvature shows the locations where
the horizon forms anticline, syncline, or flat-dipping
surfaces as positive, negative, and zero curvature values, respectively (Roberts, 2001). Lineaments of high
positive maximum curvature values are well defined
across the seismic area (Figure 13c) and generally correspond to features mapped from the shaded relief
maps (Figure 13a, b). Strike curvature describes the
shape of the surface into areas of valley shapes (negative curvature values) and ridge shapes (positive curvature values) (Roberts, 2001). Curvature lineaments
on the strike curvature map (Figure 13d) display are less
pronounced. However, these lineaments mostly correspond to structures delineated on maximum curvature
display (Figure 13c). Dip curvature shows relief variations in the surface and is a measure of the rate of
change of dip in the maximum dip direction, i.e., at
right angles to the strike curvature (Roberts, 2001).
The dip curvature display (Figure 13e) shows curvature
lineaments that are bigger and more continuous in
comparison to lineaments mapped on other displays.
According to Roberts (2001), this attribute will tend to
exaggerate local relief along the surface.

The lineaments mapped from these displays


(Figure 13a, e) indicate the presence of low-amplitude
structural features. A seismic transect that shows the
expression of typical curvature-defined features is shown
in Figure 14. Lineament azimuths were analyzed to determine if a consistent trend existed between mapped
lineaments and cleat strike from the San Juan basin
(Figures 1, 15a). The analysis was performed on binary
images (e.g., black lines on white background) with
processing image software (methodology described by
Hart et al., 2002). The software extracted the orientation of each lineament, and the lineations so defined
were plotted as rose diagrams. Mapped lineaments from
shaded relief maps (Figure 13a, b) show major trends that
strike toward 45 and 145j (Figure 15c, d, respectively).
These lineaments are subparallel to major face-cleat
strikes identified in outcrop (Tremain et al., 1994) and
probably reflect real structural features. In contrast,
lineaments observed on maximum-, strike-, and dipcurvature maps (Figure 13c, e) show multimodal patterns (Figure 15e, g, respectively). To identify clusters
in lineament trends, we combined all mapped lineaments into a single rose diagram (Figure 15h). Although
this composite rose diagram shows considerable scatter,
major trends centered at 45 and 145j are concordant
with face-cleat strikes as measured in outcrop.
Marroqun and Hart

1615

1616

Seismic Attribute-Based Characterization of Coalbed Methane Reservoirs

Figure 14. Seismic transect BB0 through


Fruitland Formation at Rosa field showing the expression in cross section of
curvature-defined lineaments (arrows)
at the thick coal level that might be related to minor structural features. Line
location shown in Figure 13a e.

Acquisition artifacts could conceivably influence


the orientation of structures derived from our horizon
attribute analysis. The orientation of acquisition artifacts was inferred from amplitude striping on the TopTC
amplitude horizon. Amplitude lineaments show a strong
preferred trend of 45j (Figure 15b), the acquisition
orientation. Although we identified some lineaments
that strike in this direction, other northeast-southwest
trends are different and may be differentiated from
acquisition artifacts because of their orientation (cf.
Hart et al., 2002). Analysis of outcrop and core cleat
orientation (Tremain et al., 1994) (Figure 1) and production trends (Kaiser and Ayers, 1994) from the
Fruitland suggest that both northwest-southeast and
northeast-southwest striking structural trends could
be present in our study area. We sought to further reduce the problem of discerning real structure features
from noise by focusing on lineaments that were iden-

tified via several approaches. Lineaments identified


using only one approach were considered to be possibly
the result of map generation (e.g., gridding) or processing artifacts (Hesthammer and Fossen, 1997).
We created a map combining our coal thickness
prediction with curvature analysis results and a bubble
plot of cumulative gas production normalized by the
number of years of production to evaluate stratigraphic
and structural controls on coalbed methane production
(Figure 16). Stratigraphic factors controlled the development of the lower Fruitland seam and, indirectly, the
distribution of coalbed methane resources. As previously discussed, the predicted coal thickness trends are
geologically realistic. The combination of stratigraphic and structural data suggests that the relative decrease in coal thickness at the north-central part of
the seismic area is probably caused by paleotopographic variations on which peat swamps developed;

Figure 13. Horizon attributes derived from seismic horizon corresponding to the top of the thick coal, with their interpreted
lineaments. (a) Shaded relief map with illumination from the northwest (subtle structural features show up as reddish lineaments).
(b) Shaded relief map with illumination from the northeast (subtle structural features show up as reddish lineaments). (c) Maximum
curvature map. Positive curvature (concave down) lineaments are shown in red and yellow, and negative curvature (concave up)
lineaments are shown in dark blue and purple. (d) Strike curvature map. Positive curvature (concave down) lineaments are shown in red
and yellow, and negative curvature (concave up) lineaments are shown in dark blue and purple. (e) Dip curvature map. Positive
curvature (concave down) lineaments are shown in yellow and orange, and negative curvature (concave up) lineaments are in green
and blue. Also shown is the location of seismic transect BB0 shown in Figure 14.
Marroqun and Hart

1617

Figure 15. Rose diagrams derived from


mapped lineaments: (a) principal cleat
strike measured in outcrop by Tremain
et al. (1994) (Figure 1); (b) amplitude
horizon, showing strong northeast-southwest trend related to acquisition footprint; (c) shaded relief display with illumination source position northwest (Figure 13a); (d) shaded relief display with
illumination source position northeast
(Figure 13b); (e) maximum curvature
display (Figure 13c); (f) strike curvature
display (Figure 13d); (g) dip curvature
display (Figure 13e); and (h) the combination of all lineaments derived from
horizon attributes. Two dominant trends
are apparent in (h a) northwest-southeast set (apparently corresponding to
domain 2 orientations) and a bimodal
northeast-southwest trend. The latter
probably contains trends associated with
acquisition artifacts (see b) and real
structural trends. Hart et al. (2002)
describe a methodology for distinguishing acquisition artifacts from real structural trends that have similar orientations.

areas of topographic highs may have produced thin


coal accumulations. Despite the lack of well control in
this area, the depth-converted structure map (Figure 12)
shows a subtle structural high in this area that supports this interpretation.
Structural features mapped through horizon attribute analysis also appear to influence coalbed methane production. The best producing wells (e.g., wells
producing gas exceeding 80 mmcf/yr [2.3  106 m3/yr])
are concentrated in areas of thicker coal and in the
vicinity of structural lineaments, especially northwestsoutheast striking features. Another area of aboveaverage gas production is near the west-central margin
of the 3-D survey area, but the poor quality of seismic
data in the margins makes the interpretation uncertain. Four wells that are found in a patch of thicker
1618

coal and in the proximity of lineaments (outlined by


yellow dashed line in Figure 16) have lower coalbed
methane production. These wells have the longest
records of production in Rosa field (e.g., 1988 2001),
so engineering factors, including improvements in completion technologies with time, may be at least partially
responsible for the lack of correlation to thickness and
structural lineaments. As a general rule, most wells with
lower normalized cumulative production (e.g., gas production lower than 30 mmcf/yr [0.8  106 m3/yr]) are
drilled in areas of lower coal accumulation that are away
from structural lineaments. We conclude that the integration of attribute-based coal thickness, structural
trends, and production data yields a consistent interpretation that explains much of the variability in coalbed methane production.

Seismic Attribute-Based Characterization of Coalbed Methane Reservoirs

Figure 16. Bubble plot of


cumulative production normalized by number of years of
production and curvature lineaments superimposed over
predicted thickness map. Lineaments mapped from shaded
relief lineaments are in violet,
maximum-curvature lineaments
are in black, strike-curvature
lineaments are in light blue,
and dip-curvature lineaments
are in yellow. Except for the
wells circled by the yellow
dashed line, the best production seems to be associated
with wells located where the
coal seam is thickest (red and
yellow) and are associated
with structural lineaments.

DISCUSSION
Several studies (e.g., Schultz et al.,1994; Hirsche et al.,
1997; Kalkomey, 1997; Pennington, 1997; Hart, 1999,
Hart and Chen, in press; Hampson et al., 2001) have
examined the methodology and concepts to be used
when conducting a seismic attribute study (i.e., our
attribute-based prediction of coal thickness), and a full
review is beyond the scope of this paper. In this study,
we used seismic modeling to identify seismic horizons
because we lacked velocity logs with which to make
well ties. The seismic models allowed us to identify and
carefully map the horizons of interest. The empirical
relationship we derived between attributes and physical properties had a high degree of statistical significance, and we successfully identified hidden extrapolation points in areas of poor seismic data quality.
The coal thickness map was geologically reasonable,
and the results, when integrated with curvature analysis, helped to explain patterns in coalbed methane production in the study area.
Our horizon curvature attribute analyses identified
two dominant structural trends, one striking northwest-

southeast and the other northeast-southwest. These orientations are consistent with basin-scale mapping of
face-cleat orientations in outcrop (Tremain et al., 1994)
and production trends from the Fruitland (Kaiser and
Ayers, 1994). Multicomponent 3-D seismic data (including a multicomponent 3-D vertical seismic profile)
and AVO analysis from the Cedar Hill field (approximately 50 km [30 mi] west of our study area) also
indicated the presence of northwest-southeast and
northeast-southwest structural trends (Shuck et al.,
1996; Ramos and Davis, 1997). The highest coalbed
methane production from that area is associated with
northwest-southeast trending structures or from areas
of enhanced fracture density where different structures
intersect. These results are in general agreement with
the results presented in this paper; however, we emphasize that our study was conducted with conventional P-wave 3-D seismic data instead of the more
expensive approaches used at Cedar Hill field.
Our simple seismic modeling efforts neglected
elastic attenuation effects in coal seams but were able to
reproduce complex attribute behaviors (e.g., Figure 11).
Hughes and Kennett (1983) used seismic models to
Marroqun and Hart

1619

study attenuation in coal and concluded that loss of


seismic energy is small because coal seams are thin in
comparison to the predominant wavelength. We conclude that simple acoustic models are useful for understanding coal-bearing rocks, at least for coals that
are seismic thin beds.

CONCLUSIONS
The controls on coalbed methane production are many
and include a variety of geologic, hydrodynamic, and
engineering factors. In this paper, we have shown the
potential application of using 3-D seismic and curvature attributes to predict coalbed thickness and the
location of enhanced permeability zones in a coal seam
from the lower part of the Fruitland Formation in our
study area. We draw the following general conclusions:
1. Seismic attribute studies may be used to predict coal
thickness. In our case, a combination of three attributes (maximum absolute amplitude, integrated
trace, and total energy) yielded a multivariate linear
regression expression that has a 0.87 correlation coefficient with the input data. The map generated
from this analysis has geologically reasonable coal
thickness trends. Coal thickness variations that strike
northwest-southeast are related to the paleoshoreline orientation. An area of thicker coal accumulation in the southeast part of our study area appears to
be related to a local depocenter that was present
throughout at least some of the Late Cretaceous.
2. Horizon attributes derived from seismic mapping of
the top of the coal, including shaded relief and various measures of curvature, identified geologically
reasonable structural features in the study area. We
needed to vary the aperture of our curvature calculations to generate these maps. Structures that strike
northwest-southeast and northeast-southwest correspond to the orientation of face cleats measured
in outcrop.
3. Integration of our attribute-derived thickness map,
our map of seismically defined subtle structures, and
coalbed methane production data yield a picture that
is consistent with known or inferred geological controls on coalbed methane production. The best production is associated with thick coal deposits that are
close to seismically defined structures.
4. Seismic modeling is a useful technique for validating
seismic attribute studies. Seismic modeling helped
us to confirm that the attributes selected by the step1620

wise linear regression methodology were indeed


physically related to coal thickness, thus minimizing the possibility of basing our results on spurious
correlations. This same seismic modeling helped us
to identify that an apparent outlying data point
should be kept for the analysis and was not caused
by bad data.
5. Our results were derived from mediocre-quality
conventional (P-wave) 3-D seismic data. Similar results should be obtainable for coalbed methane reservoirs elsewhere.

REFERENCES CITED
Ayers, W. B. Jr., W. A. Ambrose, and J. S. Yeh, 1994, Coalbed methane in the Fruitland Formation, San Juan basin Depositional
and structural controls on occurrence and resources, in W. B.
Ayers, Jr., and W. R. Kaiser, ed., Coalbed methane in the Upper
Cretaceous Fruitland Formation, San Juan basin, New Mexico
and Colorado: Bureau of Economic Geology Bulletin, v. 146,
p. 13 40.
Brown, A. R., 1999, Interpretation of three-dimensional seismic
data, 5th ed.: AAPG Memoir 42, 514 p.
Chen, Q., and S. Sidney, 1997, Seismic attribute technology for
reservoir forecasting and monitoring: The Leading Edge, v. 16,
p. 445 456.
Dugan, T., 2002, A history of Fruitland Formation coalbed methane
development, in the San Juan basin, New Mexico and Colorado:
AAPG Rocky Mountain Section Annual Meeting, 2002 Technical Program, p. 23.
Gardner, G. H. F., L. W. Gardner, and A. R. Gregory, 1974, Formation velocity and density The diagnostic basics for stratigraphic traps: Geophysics, v. 36, p. 770 780.
Gochioco, L. M., 1991, Tuning effect and interference reflections
from thin beds and coal seams: Geophysics, v. 56, p. 1288 1295.
Gochioco, L. M., 1992, Modeling studies of interference reflections
in thin-layered media bounded by coal seams: Geophysics,
v. 57, p. 1209 1216.
Hamilton, W., 1987, Plate-tectonic evolution of the western U.S.A:
Episodes, v. 10, p. 271 276.
Hampson, D. P., J. S. Schuelke, S. James, and J. A. Quirein, 2001,
Use of multiattribute transforms to predict log properties from
seismic data: Geophysics, v. 66, p. 220 236.
Hart, B. S., 1999, Geology plays key role in seismic attributes studies:
Oil & Gas Journal, v. 97 (July 12), p. 76 80.
Hart, B. S., 2000, 3-D seismic interpretation: A primer for
geologists: SEPM Short Course Notes 48, 123 p.
Hart, B. S., 2002, Validating seismic attribute studies: Beyond
statistics: The Leading Edge, v. 21, p. 1016 1021.
Hart, B. S., and R. S. Balch, 2000, Approaches to defining reservoir
physical properties from 3-D seismic attributes with limited
well control: An example from the Jurassic Smackover Formation, Alabama: Geophysics, v. 65, p. 368 376.
Hart, B. S., and M. A. Chen, in press, Understanding seismic
attributes through forward modeling: The Leading Edge, v. 23.
Hart, B. S., R. Pearson, and G. C. Rawling, 2002, 3-D seismic
horizon-based approaches to fracture-swarm sweet spot definition in tight-gas reservoirs: The Leading Edge, v. 21, p. 220
236.
Hesthammer, J., and H. Fossen, 1997, The influence of seismic noise

Seismic Attribute-Based Characterization of Coalbed Methane Reservoirs

in structural interpretation of seismic attribute maps: First Break,


v. 15, p. 209 219.
Hughes, V. J., and B. L. N. Kennett, 1983, The nature of seismic
reflections from coal seams: First Break, v. 1, p. 9 18.
Hirsche, K., J. P. Hirsche, and L. Mewhort, 1997, The use and
abuse of geostatics: The Leading Edge, v. 16, p. 253 258.
Kaiser, W. R., and W. B. Ayers, Jr., 1994, Coalbed methane production, Fruitland Formation, San Juan basin: Geologic and
hydrologic controls, in W. B. Ayers, Jr. and W. R. Kaiser, ed.,
Coalbed methane in the Upper Cretaceous Fruitland Formation, San Juan basin, New Mexico and Colorado: Bureau of
Economic Geology Bulletin, v. 146, p. 187 207.
Kalkomey, C. T., 1997, Potential risks when using seismic attributes
as predictors of reservoir properties: The Leading Edge, v. 16,
p. 247 251.
Laubach, S. E., and C. M. Tremian, 1994, Tectonic setting of the
San Juan basin, in W. B. Ayers, Jr. and W. R. Kaiser, eds.,
Coalbed methane in the Upper Cretaceous Fruitland Formation, San Juan basin, New Mexico and Colorado: Bureau of
Economic Geology Bulletin, v. 146, p. 9 11.
Lehmann, E. L., and H. J. M. DAbrera, 1975, Nonparametrics:
Statistical methods based on ranks: San Francisco, HoldenDay, 457 p.
Lorenz, J. C., and S. P. Cooper, 2003, Tectonic setting and
characteristics of natural fractures in Mesaverde and Dakota
reservoirs of the San Juan basin, New Mexico: Geology, v. 25,
p. 3 14.
Montgomery, D. C., and E. A. Peck, 1992, Introduction to linear
regression analysis, 2d ed.: New York, John Wiley & Sons, Inc.,
527 p.
Pashin, J. C., 1998, Stratigraphy and structure of coalbed methane
reservoirs in the United States: An overview: International
Journal of Coal Geology, v. 35, p. 209 240.
Pennington, W. D., 1997, Seismic petrophysics: An applied science

for reservoir geophysics: The Leading Edge, v. 16, p. 241


244.
Ramos, A. C. B., and T. L. Davis, 1997, 3-D AVO analysis and
modeling applied to fracture detection in coalbed reservoirs:
Geophysics, v. 62, p. 1683 1695.
Roberts, A., 2001, Curvature attributes and their application to 3-D
interpreted horizons: First Break, v. 19, p. 85 99.
Schultz, P. S., S. Ronen, M. Hattori, P. Mantran, and C. Corbett,
1994, Seismic guided estimation of log properties: Part 3. A
controlled study: The Leading Edge, v. 13, p. 770 776.
Shuck, E. L., T. L. Davis, and R. D. Benson, 1996, Multicomponent
3-D characterization of a coalbed methane reservoir: Geophysics, v. 61, p. 315 330.
Steen, ., E. Sverdrup, and T. H. Hansson, 1998, Predicting the
distribution of small faults in a hydrocarbon reservoir by combining outcrop, seismic and well data, in G. Jones, Q. J. Fisher,
and R. J. Knipe, eds., Faulting, fault sealing and fluid flow in
hydrocarbon reservoirs: Geologic Society (London) Special
Publication 147, p. 27 50.
Stewart, S. A., and T. J. Wynn, 2000, Mapping spatial variation in
rock properties in relationship to scale-dependent structure
using spectral curvature: Geology, v. 28, p. 691 694.
Taner, M. T., F. Koehler, and R. E. Sheriff, 1979, Complex seismic
trace analysis: Geophysics, v. 44, p. 1041 1063.
Tebo, J., and B. S. Hart, in press, Use of volume based 3-D seismic
attribute analysis to characterize physical property distribution:
A case study to delineate reservoir heterogeneity at the Appleton field, SW Alabama: Journal of Sedimentary Research.
Tremain, C. M., S. E. Laubach, and N. H. Whitehead, III, 1994,
Fracture (cleat) patterns in Upper Cretaceous Fruitland Formation coal seams, San Juan basin, in W. B. Ayers Jr. and W. R.
Kaiser, ed., Coalbed methane in the Upper Cretaceous Fruitland Formation, San Juan basin, New Mexico and Colorado:
Bureau of Economic Geology Bulletin, v. 146, p. 87 102.

Marroqun and Hart

1621

Вам также может понравиться