Download as pdf or txt
Download as pdf or txt
You are on page 1of 244

THE USE OF CORE AND BOREHOLE IMAGE LOGS IN A 3D MODEL OF

THE CEPO/POWDER MOUNTAIN AREA, WASHAKIE BASIN, WYOMING.

by

Priya Maraj
A thesis submitted to the faculty and the Board of Trustees of the Colorado School of

Mines in partial fulfillment of the requirements for the degree of Master of Science

(Geology)

Golden, Colorado

Date ______________

Signed: ______________________
Priya Maraj

Approved: ____________________
Dr. Neil F. Hurley
Thesis Advisor

Golden, Colorado

Date ______________

______________________
Dr. Murray Hitzman
Professor and Head
Department of Geology
and Geological Engineering

ii
ABSTRACT

The Washakie basin is in the eastern part of the Green River basin, an area in

southwest Wyoming that has large gas reserves in tight sandstones. The Lewis Shale is a

2,000-2,600 ft (609-792 m) thick interval that is composed of deep-water sandstones and

shales in the study area. Some wells in the study area are excellent gas producers,

whereas others are poor. For example, the CEPO 21-18 well is an excellent well that

produces about 1,700 MCF per day of gas from low-permeability deep-water sandstones

of the Lewis Shale.

The purpose of this study is to build a 3D geological model of the CEPO/Powder

Mountain field area, using borehole image logs and cores as input data for fracture

orientation and intensity. The data set includes 105 mi2 (168 km2) of 3D seismic data, 8

wells with core, and 4 wells with borehole images. Core studies show that partially open

fractures are present exclusively in the sandstone intervals. Borehole images show two

main fracture sets, predominantly in the sandstones, with dip direction/dip values of

213°/ 65° (Fracture Set 1) and 12°/ 54° (Fracture Set 2). Fracture Set 1 is 5 times as

abundant as Fracture Set 2. These appear to be conjugate fracture sets that may have

formed at a time when the maximum stress direction was vertical. Borehole breakouts

iii
interpreted from borehole images suggest that the present-day maximum horizontal stress

orientation is N34°W. This is consistent with other studies in the area.

Production data show wide variations in performance, as indicated by

comparisons of the first 12 months of production in the study area. Completion practices

were not different enough to account for these variations. Gas shows, lost circulation

while drilling, pressure-dependent leakoff from the G-function, and the use of 100-mesh

sand suggest that natural fractures affect production in the area.

Seismic mapping shows a series of N70°W trending faults in the southern part of

the study area. Otherwise, structural relief is minor, and horizons generally dip uniformly

to the west. Vp/Vs and acoustic anisotropy maps suggest that there are some N15°W

trends that may be faults. Nearby studies suggest that oblique-slip faulting is present in

the area.

The geological model was constructed using 3D Move software. Time-to-depth

converted seismic horizons and faults, provided by Stone Energy, were imported to

constrain the model. Two approaches were taken to fracture modeling: (1) attribute-

based fracture modeling, and (2) discrete fracture modeling. The attribute-based model

related curvature to fracture occurrence. This model shows minor variations in curvature.

A subtle N15°W anomaly lies in the vicinity of the Cepo Lewis 21-18 well, the best

producer in the area. This anomaly roughly corresponds to a linear feature seen on the

Vp/Vs and acoustic anisotropy seismic maps. The discrete fracture model used fracture

orientations and intensities from the core and borehole image work. Because core and

iv
borehole image studies showed that fractures were confined to sandstones, the

distribution of fractures in the 3D model was constrained by net-to-gross ratios in 3 wells.

Results show that fractures are common in all 3 wells. The well with the lowest amount

of sandstone, and therefore the fewest fractures, is the best producer. This suggests that

stratigraphic traps are important in this area.

v
TABLE OF CONTENTS

Page
ABSTRACT ...................................................................................................................... iii

TABLE OF CONTENTS................................................................................................... vi

LIST OF FIGURES ......................................................................................................... viii

LIST OF TABLES........................................................................................................... xiv

ACKNOWLEDGMENTS .................................................................................................xv

CHAPTER 1 INTRODUCTION .....................................................................................1

1.1 Research Objectives.................................................................................................1


1.2 Previous Work .........................................................................................................2
1.3 Research Contributions............................................................................................6

CHAPTER 2 GEOLOGIC BACKGROUND..................................................................9

2.1 Location of the Study Area ......................................................................................9


2.2 Regional Stratigraphy ............................................................................................11
2.3 Tectonic Framework ..............................................................................................16
2.4 Petroleum System ..................................................................................................19

CHAPTER 3 BOREHOLE IMAGE LOGS AND CORE .............................................32

3.1 Background ............................................................................................................32


3.2 Methods..................................................................................................................42
3.2.1 Core……………………………………..…………………………….42
3.2.2 Borehole Image Logs…………………………………..……………..44
3.2.3. In situ Stress……………………………………………….…………52
3.3 Results....................................................................................................................56
3.3.1 Core…………………………………...………………………….……56

vi
3.3.2. Borehole Image Logs……………………………………………..…..63
3.3.3. Data comparison……………………….…………………………….82
3.3.4. Vertical fracture spacing………...……………..…...………………..89
3.3.5. Breakouts, in situ stress……………………………….…………......93
3.4 Discussion ................................................................................................................96

CHAPTER 4 WELL COMPLETION, PRODUCTION DATA..................................106

4.1 Background ..........................................................................................................106


4.2 Methods……………………………………………………….……………..…109
4.3 Results…………………………………………………………………..………111
4.3 Discussion ............................................................................................................119

CHAPTER 5 SUPPORTING SEISMIC DATA..........................................................124

5.1 Background ..........................................................................................................124


5.2 Results…………………………………………………………………………..131
5.3 Discussion ............................................................................................................138

CHAPTER 6 3D FRACTURE MODELING ..............................................................146

6.1 Background ..........................................................................................................146


6.2 Methods…………………………………………………………………………146
Discrete Fractures ........................................................................................147
Attribute-Based Fractures ............................................................................159
6.3 Results..................................................................................................................167
6.4 Discussion ............................................................................................................180

CHAPTER 7 CONCLUSIONS ...................................................................................187

REFERENCES ................................................................................................................190

APPENDICES………………………………………………………………………….198

A Core descriptions……………………………………………….…….……..198
B Fractures in borehole image logs…………………………..………………..202
C Production and well data…………………………………………...………..208

vii
LIST OF FIGURES

Figure 1.1 Location map of the study area. .......................................................................3

Figure 2.1 Regional structure map of the Greater Green River basin. ............................10

Figure 2.2 Location map of the field area.........................................................................12

Figure 2.3 Location of the Cretaceous Western Interior Seaway. ...................................13

Figure 2.4. Chronostratigraphic chart of the Upper Cretaceous. ......................................15

Figure 2.5. Regional tectonics of south-central Wyoming. ..............................................17

Figure 2.6. Landsat image of the Cherokee Arch. ............................................................20

Figure 2.7. Left-lateral wrench model of Harding (1974). ...............................................21

Figure 2.8. Hull (2001) interpreted flower structures in his study area............................22

Figure 2.9. Experiments performed by McClay and Bonora (2001). ...............................23

Figure 2.10. Diagram of gas generation, storage and expulsion from coal. ......................26

Figure 2.11. Cross section across the Washakie basin…………………………………...27

Figure 2.12. Plot of cumulative production vs. distance…………………………………29

Figure 2.13. Illustration showing pressure zones...............................................................30

Figure 3.1. Map of the study area showing the location...................................................34

Figure 3.2. Location of cores and image logs in terms of producing intervals.................35

Figure 3.3. Diagram of the Schlumberger FMI tool. ........................................................37

Figure 3.4. Static and Dynamic modes of EMI tool .........................................................39

Figure 3.5. The Electrical Micro Imaging (EMI) tool configuration................................41

viii
Figure 3.6. FMI image showing out-of-phase primary and auxiliary pads. .....................46

Figure 3.7. Dark irregular near-vertical features are open fractures.................................48

Figure 3.8. Example of a healed fracture..........................................................................49

Figure 3.9. Example of an induced fracture......................................................................50

Figure 3.10. Schematic diagram of a wellbore. ................................................................51

Figure 3.11. Diagram illustrates borehole breakouts and induced fractures.....................54

Figure 3.12. Core figure – quart fracture. .........................................................................58

Figure 3.13. Core figure – calcite lined fracture................................................................59

Figure 3.14. Fractures normalized to fractures per 100 ft of core. ....................................62

Figure 3.15. Type of mineralization found in each well....................................................64

Figure 3.16. Stereonet of Powder Mountain 1-13E well ...................................................66

Figure 3.17. Stereonet of all open fractures for the Powder Mountain 23-36 well. ..........67

Figure 3.18. Open fractures, with separated dips -Powder Mountain 23-36 well .............68

Figure 3.19. All open and healed fractures for the Powder Mountain 32-26 well ............69

Figure 3.20. All open fractures for Set 1 for the Powder Mountain 32-26 well................70

Figure 3.21. Fracture Set 1, dips under 80º for the Powder Mountain 32-26 well. ...........71

Figure 3.22. Fracture Set 1, dips over 80º for the Powder Mountain 32-26 well. .............72

Figure 3.23. Stereonet of Fracture Set 2 for the Powder Mountain 32-26 well.................73

Figure 3.24. All open and healed fractures for the Powder Mountain 34-11 well ...........74

Figure 3.25. All open fractures for Set 1 for the Powder Mountain 34-11 well...............75

Figure 3.26. Fracture Set 1, dips under 80º for the Powder Mountain 34-11 well ...........76

ix
Figure 3.27. All open fractures over 80º for the Powder Mountain 34-11 well ...............77

Figure 3.28. Stereonet of all Fracture Set 2 for the Powder Mountain 32-26 well............78

Figure 3.29. Fracture Set 2, dips under 80º for the Powder Mountain 32-26 well ............79

Figure 3.30. Frequencies of dip magnitude for all open fractures ....................................81

Figure 3.31. Dip azimuth and the depth of all open fractures PM 1-13E well. ................83

Figure 3.32. Dip azimuth and the depth of all open fractures PM 23-36 well..................84

Figure 3.33. Dip azimuth and the depth of all open fractures PM 32-26 well..................85

Figure 3.34. Dip azimuth and the depth of all open fractures PM 34-11 well..................86

Figure 3.35. Dip azimuth obtained from borehole image analysis. ..................................87

Figure 3.36. Comparison of data collected from cores and borehole image logs.............88

Figure 3.37. Schematic diagram of core or FMI image.....................................................91

Figure 3.38. Estimation of scatter of a fracture set in orientation and spacing..................92

Figure 3.39. Borehole breakout analysis for the Powder Mountain 1-13E well...............94

Figure 3.40. Graph showing the frequency of the strike azimuth of Shmax ....................95

Figure 3.41. Principle stress directions …………….…………………………....…...….98

Figure 3.42. Principal stress direction for study area.........................................................99

Figure 3.43. Stress directions for study area....................................................................101

Figure 3.44. Idealized right-slip fault regime ..................................................................103

Figure 3.45. Sketch of dip-slip, oblique-slip, and strike-slip faulting………….…...….104

Figure 3.46. Present day stress orientation. .....................................................................105

Figure 4.1. Location of wells with cumulative production rates ....................................108

x
Figure 4.2. G-function analysis for pressure dependent leakoff behavior......................110

Figure 4.3. Cepo-Lewis 21-18 drilling data...................................................................112

Figure 4.4. Powder Mountain 1-13E drilling data .........................................................113

Figure 4.5. Bogey Draw 1-14 drilling data ....................................................................114

Figure 4.6. Mud weight comparison for fluid pressure .................................................117

Figure 5.1. Shear wave splitting .....................................................................................127

Figure 5.2. Various ray paths and wave fields................................................................129

Figure 5.3. Processing products for 3D conventional p-wave seismic data. ..................133

Figure 5.4. Processing products for 3D converted-wave seismic data ...........................134

Figure 5.5. Line 5980. Layer stripping anisotropy ........................................................135

Figure 5.6. Color structure map with contours on the Lewis Shale................................136

Figure 5.7. Fox Hills to Almond isochron maps in fast and slow wave directions. .......137

Figure 5.8. Percent anisotropy map of the surface to Ft Union interval.........................139

Figure 5.9. Fast and slow Vp/Vs ratios on the Fox Hills to the Almond........................140

Figure 5.10. Surface to Ft. Union dominant fracture direction in strike azimuth...........141

Figure 5.11. Lance to Lewis Shale dominant fracture direction in strike azimuth.........142

Figure 5.12. High Vp/Vs areas as potential drilling targets............................................144

Figure 5.13. Color structure map with contours on the Lewis Shale..............................145

Figure 6.1. Structure map on top of Lewis Shale ...........................................................148

Figure 6.2. Fracture parameters used for the grid-based fracture model………………149

Figure 6.3. Fracture length and spatial distribution at Muddy Gap ................................151

xi
Figure 6.4. Method for averaging sand thickness ...........................................................153

Figure 6.5. Plot to calibrate fracture intensity as a function of net-to-gross ratio ..........156

Figure 6.6. Map view of model showing well locations.................................................158

Figure 6.7. Relative connectivity of fractures.................................................................160

Figure 6.8. Fracture parameters used for the attribute-based fracture model .................161

Figure 6.9. Forbidden zone around fractures ..................................................................163

Figure 6.10. Triangulated surface used to calculate curvature .......................................165

Figure 6.11. Cylindrical analysis ....................................................................................166

Figure 6.12. Map view of fractures in Cepo Lewis 21-18 ..............................................168

Figure 6.13. Map view of fractures in PM 1-13E and Bogey Draw 1-14.......................169

Figure 6.14. Relative connectivity of fractures in Cepo Lewis 21-18 ............................170

Figure 6.15. Relative connectivity of fractures in PM 1-13E .........................................171

Figure 6.16. Relative connectivity of fractures in Bogey Draw 1-14.............................172

Figure 6.17. Fracture-wellbore intersections in Cepo Lewis 21-18................................173

Figure 6.18. Fracture-wellbore intersections in PM 1-13E.............................................174

Figure 6.19. Fracture-wellbore intersections in Bogey Draw 1-14 ................................175

Figure 6.20. Cross-section orientation ............................................................................176

Figure 6.21. Cross-section of the Cepo Lewis 21-18 model...........................................177

Figure 6.22. Cross-section of the PM 1-13E model........................................................178

Figure 6.23. Cross-section of the Bogey Draw 1-14 model ...........................................179

Figure 6.24. Curvature map on Lewis Shale...................................................................181

xii
Figure 6.25. Curvature map on Almond .........................................................................182

Figure 6.26. Attribute-based cylindrical analysis ...........................................................183

Figure 6.27. Spatial analysis of fracture intensity...........................................................184

xiii
LIST OF TABLES

Table 2.1. Direct and indirect attributes............................................................................24

Table 3.1. Data for the wells used in the study.................................................................33

Table 3.2. EMI hardware characteristics. .........................................................................42

Table 3.3. Intervals in the FMI log for well PM 1-13E ....................................................45

Table 3.4. Average vertical fracture spacing from core data............................................61

Table 3.5. Summary of the average observed fracture orientations .................................80

Table 4.1. Cumulative production for the first 12 months..............................................107

Table 4.2. Completion practices and status of wells in the study area ...........................115

Table 4.3. Completion practices, Polar Bar Unit #1. ......................................................120

Table 6.1. Number of fractures in 100 ft of core. ...........................................................154

Table 6.2. Net-to-gross ratio for cored intervals.............................................................154

Table 6.3. Net-to-gross ratio for the entire Lewis Shale interval....................................155

Table 6.4. Fracture intensity values used for each fracture set.......................................157

Table 6.5. Results from models of fracture-borehole intersections ................................167

xiv
ACKNOWLEDGMENTS

In every aspect of this study there were a number of people who have willingly

donated their expertise and have been willing to share their data with me. I would like to

thank these institutions for their contributions. David Richards at Midland Valley (MVE)

Inc. helped with 3D Move, Butch Oliver at Triple O Slabbing laid out the cores for me,

Ken Boedeker from EOG provided access and permission to use the well completions data,

Chris Besler and Keith Shanley from Stone Energy supplied the seismic and core data,

Halliburton donated the borehole image data, Janine Carlson helped with everything

involving Unix, Ira Pasternack provided not only his time with Petra, but always offered

good advice, and finally Jennifer Miskimins helped with understanding the completions

data.

Much appreciated financial contributions were made by the Lewis Shale

consortium, the American Association of Petroleum Geologists and the Society of

Professional Well Log Analysts Foundation. BP Trinidad provided a stipend. Neil Hurley

provided various fellowships, including ConocoPhillips and BP, every semester. Lastly I

would like to thank my advisor and committee for their time: Neil Hurley, Tom Davis,

Chuck Kluth and Keith Shanley. Mostly I want to thank Charlie Rourke for all her help,

and all my family and friends that have been there for me all through the years. Most

importantly, this thesis is dedicated to my mother, whose love and support was constantly

felt. Thank you mum for always being there for me, I love you.

xv
1

CHAPTER 1

INTRODUCTION

Estimates for the Greater Green River basin quote a presently recoverable value

of 17-100 TCF (Shanley et al., 2003). Currently, there is much exploratory interest in

this basin within petroleum companies. This study investigates the structural evolution of

a 105 mi2 (272 km2) area in the Washakie basin, south-central Wyoming, commonly

called the CEPO/Powder Mountain area. The data set includes core in 8 wells, borehole

image logs in 4 wells, conventional openhole logs in 10 wells, and 3D seismic data.

1.1 Research Objectives

The objectives of this study are to:

Help resolve the effects of structure as a hydrocarbon resource

mechanism by building a 3D model.

Provide insight into new target areas for exploration, by getting a sense of

the fracture network.

Explain why the production rates of some wells are good, whereas others

are poor, in the area. Specifically look at fracture styles and their role in

production.
2

1.2 Previous Work

This study was conducted as part of the Lewis Shale Consortium at the Colorado

School of Mines. The study incorporates previous work from other students. Figure 1.1

shows the Lewis Shale study area in Wyoming and the location of some of the previous

work in relation to this thesis. Through the consortium, a number of students have

completed work on the Lewis Shale project, relevant to this study. The most relevant of

these, to my study, are Rahmat (2000), Hull (2001), Minton (2002), and Ysaccis (2003).

See Figure 1.1 for the location of their study areas. The figure outlines this study area in

red, Minton’s (2002) cross-section crosses my study area. He described 2 cores in my area

and determined a maximum horizontal stress direction of N26°W. Rahmat (2000) studied

borehole image logs along the southeastern end of this study area. She used orientations of

open fractures in borehole image logs by plotting the strike directions, and in this way she

identified open factures in the Pilgrim Federal and Iverson wells (Great Divide basin, CO.)

oriented in a N40°W direction. Hull (2001) identified stress directions consistent with a

NW orientation for maximum horizontal stress. Hull (2001) completed an interpretation of

a 3D seismic survey, which is outlined in Figure 1.1. He also demonstrated the structural

complexity of the Cherokee Arch, at the southern extent of the study area. He interpreted

the Cherokee Arch structure as being active during deposition of the lower Lewis Shale and

becoming less active during the middle and upper Lewis Shale deposition. Ysaccis (2003)

studied a large 3D seismic survey that overlaps this study area, in which he interpreted the
3

MO

WY
ID
BUSH
LAKE
95W
HAY RESERVOIR
SINK HOLE UT CO
GREAT DIVIDE

STRIKE
UNIT
SIBERIA
RIDGE
TEN MILE DESERT
DRAW SPRINGS CG ROAD 21N
WAMSUTTER
20N PLAYA
ECHO
SPRINGS
FILLMORE
TABLE ROCK/ STANDARD
DELANEY RIM DRAW
STAGE CRESTON
STOP TABLE
ROCK SW
17N

ALKALINE 95W
CREEK
15N
MARAJ
CEPO
POWDER
MOUNTAIN
102W TRITON HULL (2002)
88W
SMITH
12N WYOMING RANCH BAGGS
HIAWATHA COLORADO WEST
WEST SIDE
CANAL
YSACCIS
WINDSOCK/
TEARDROP
BIG
10N HOLE
BLACK
Lewis Shale GREAT MOUTAIN
DIVIDE
MINTON (2002) BLUE
Location Map RAHMAT (2000) GRAVEL
AT T RIBUT E MAP
Lewis cores available 95W CRAIG
NORTH
REMARKS
90W
Figure 1.1. Location map of the study area, with a look at the work of the Lewis Shale
Consortium. Other relevant work includes Minton (2002) Rahmat (2000), Hull (2002) and
Ysaccis (2003).
4

regional structure of the Cherokee Arch.

Other notable work on the Lewis Shale includes Weimer (1960), Blackstone

(1963), Maughan and Perry (1986), Krystinik and DeJarnett (1995), and others.

Geological investigations in the Greater Green River basin began in the 1800’s and have

continued since (Blackstone, 1963). Weimer (1960) presented a concise summary of

Upper Cretaceous stratigraphy in the Rocky Mountain region. The area is characterized

by intertonguing marine and non-marine sediments. As they extend to the east, these

sediments change facies into marine shales. The dominant sedimentary pattern during

the late Cretaceous was deposition during regression of the strandline from west to east.

This was intermittently broken by sharp transgressions of the strandline, causing the

intertonguing pattern that we observe today. Weimer (1960) presented his work using

regional correlation, thus he was able to identify previous miscorrelations that were

carried out on a local scale. He showed that the Lewis Shale at its type locality in the San

Juan basin, New Mexico, is correlative with the upper part of the Mancos Shale of

northwest Colorado and the upper part of the Steele Shale of Wyoming. Thus the “Lewis

Shale” of northwest Colorado and Wyoming is younger than the Lewis Shale at the type

locality, and is correlative with the Bearpaw Shale of Montana. Blackstone (1963)

reported on the importance of structural history on the origin and entrapment of

hydrocarbons. He identified two main structural elements in the area; large northwest

trending crustal folds and low-angle thrusts. It is apparent that the present structural
5

features, formed during the Laramide Orogeny, have been superimposed across an earlier

pattern of deposition.

Krystinik and DeJarnett (1995) integrated biostratigraphic and lithostratigraphic

data in order to create a chronostratigraphic framework for the Campanian and lower

Maastrichtian of the Western Interior Seaway. This chronostratigraphic framework

allows comparison of coeval stratal stacking patterns and key surfaces along the western

margin of the basin and documents some of the complex sequence stratigraphic

relationships that can occur within foreland basins. They found inconsistencies with

published eustatic curves, compared with most of the lithostratigraphic and

chronostratigraphic profiles from their study. This suggested that tectonic activity during

the Campanian and early Maastrichtian probably played a major role in creating the

complex relationships observed in the Western Interior Seaway. Local uplifts and basinal

down warps were a dominant control on relative sea level histories for most of the

western margin of the basin. Major sequence-bounding unconformities, with hundreds of

meters of erosion, are the direct time equivalents of maximum flooding events in other

parts of the basin, reflecting the tectonic complexity typical of foreland basins.

For this study, Stone Energy Corporation in Denver donated a number of

products. Chris Besler of Stone Energy Corporation interpreted 9 horizons and 4 faults in

a 105 mi2 (272 km2) area in a high-quality conventional p-wave 3D seismic area. Both

Western-GECO and AXIS processed the full volume and migrated the post-stack time

volume. Western-GECO then took a smaller area of 15 mi2 (39 km2) to calculate the pre-
6

stack time migration. The azimuthal AVO and velocity were calculated over this smaller

area of the CEPO field.

Stone Energy carried out an interpretation of borehole image logs. I reviewed

these logs for the accuracy of their sinusoidal picks of the fractures. Minton (2002)

described core in the area, however, he did not make note of significant fractures nor

slickensides, something that is investigated in this study.

1.3 Research Contributions

Fractures in 8 cores were described within the study area and 4 borehole image

logs were interpreted for fractures and borehole breakouts. Well completions and

production reports were used to determine if this reservoir responded as a naturally

fractured one and if completion practices had any effect on production variability.

Converted wave seismic data was used to determine if fracture directions were consistent

with those seen in the cores and borehole image logs. Once all this data was collected, it

was used as constraining parameters for a 3D model.

The major conclusions for the study are:

Partially open fractures are present only in sandy units of the cores.

Mineral growths (quartz and calcite) on the walls of the fractures indicate that the

fractures are probably natural.


7

Two conjugate fracture sets were identified from the borehole image logs.

Fracture Set 1 has a dip direction/dip of 213°/65°and Fracture Set 2 is 12°/54°.

Fracture Set 1 has 5 times more fractures than Set 2.

Borehole breakouts show a present-day maximum stress direction of N34°W.

This is consistent with other work in the area.

Wide variations in production, as identified by the first 12 months of production,

cannot be explained as simply the result of completion practices. The highest

cumulative production rate is from the Cepo Lewis 21-18 well (1.8 bcf), while the

lowest cumulative production comes from the Powder Mountain Fed 23-36 (0.07

bcf).

Gas shows, lost circulation while drilling, pressure dependent leakoff, and the use

of 100-mesh sand show that natural fractures affect production.

There is a N70°W trend of faults in southern part of the study area from seismic

data.

There is a N15°W fault/fracture trend identified from Vp/Vs and percent

anisotropy maps.

Two approaches were taken to fracture modeling: (1) attribute-based fracture

modeling, and (2) discrete fracture modeling.

Discrete 3D fracture models show that fractures are common throughout the area.
8

Distribution of fractures in the 3D model was constrained by the amount of

sandstone. Wells with lower fracture densities and lower amounts of sandstone

are the best producers in the field.

Cross-sections show the fracture network across the field area.

3D structural models visualize the relative connectivity of fractures.

Lewis Shale production may be affected by the presence of stratigraphic trapping,

more so than structural traps. Laminated sandstones have pinch-outs and shaly

interbeds that act as stratigraphic traps.

The attribute-based fracture model related curvature to fracture occurrence in the

Lewis Shale. Minor variations in curvature could be seen, with a subtle N15°W

trend.
9

CHAPTER 2

GEOLOGICAL SETTING

2.1 Location of the Study Area

The study area is located in the Washakie sub-basin in the easternmost part of the

Greater Green River basin in Sweetwater County, south-central Wyoming. The Greater

Green River basin is an irregularly shaped intermontane basin that comprises part of the

central Rocky Mountain region. It is located between latitude 40°30’ to 43°30’ N and

longitude 107° to 111° W. The basin is divided into four structural sub-basins; the Green

River basin, the Great Divide basin, the Washakie basin and the Sand Wash basin

(Figure 2.1).

The Washakie sub-basin is a structural and topographic basin bounded on the east

by the Sierra Madre Uplift, on the north by the Wamsutter Arch, on the west by the Rock

Springs Uplift, and on the south by the Cherokee Arch (Figure 2.1). The basin is 42 mi

(68 km) in the north-south direction and 54 mi (87 km) in the east-west direction, with an

area of roughly 2,200 mi2 (3,550 km2). The surface elevation in the basin ranges from

6,100 to 8,700 ft (1,860 – 2,650 m) and averages 7,000 ft (2,130 m). The dip of the

Cretaceous strata into the basin is approximately 8° along the eastern flank and 15° along

the western flank (Love, 1970). Precambrian rocks lie at depths of >32,000 ft (9,750 m)

at the center of the Washakie basin (Love, 1970). Approximately 32,000 ft (9,750 m) of
Montana

M
Wyoming

W OU
Idaho

IN NT
D A
R
RI IN
VE S
SWEET
WATE Utah Colorado
ARC H R

UPL
GREAT
I
WL

OVERTHRUST BELT
GREEN DIVIDE R A FT
INS

UPLIFT
WAMSUTTER ARCH
BASIN
RIVER
BASIN WASHAKIE

ROCK SPRINGS
BASIN SI
M ER
AD R A
RE
WYOMING CHEROKEE ARCH
UTAH COLORAD
UINTA O
AX SAND WASH
UP IAL
MOUNTAINS LIF BA
BASIN
T S IN

Figure 2.1. Regional structural map of the Greater Green River basin showing the study area in the box.
The gray shading corresponds to uplifts. Cross-section B-B‘ is shown in Figure 2.12. From Law, (2002).
10
11

Cambrian through Tertiary sedimentary rocks are present in the deeper part of the

Washakie basin (Hale, 1961). The deepest part of the Upper Cretaceous section, in the

basin’s center, is at ~14,000 ft (4,270 m) depth (Roehler, 1990).

The study area, the CEPO/Powder Mountain area, consists of a 105 mi2 (272 km2)

area, which corresponds to the outline of a 3D seismic data set. Figure 2.2 shows the study

area, the outline of the seismic surveys, and locations of borehole image logs and core. The

area covers the main well of interest with one of the highest rates of production in the

basin: the CEPO 21-18 well (Section 18, T14N, R95W). Production rates are around 1,700

MCF per day in this well.

2.2 Regional Stratigraphy

The Lewis Shale in the Washakie basin is a marine siliciclastic deposit in the

Western Cretaceous Interior Seaway. This shallow epeiric seaway covered a large part of

North America from the Gulf of Mexico through Canada (Figure 2.3). The Lewis Shale

was deposited during the Maastrichtian Bearpaw transgression and regression of this

seaway (Winn et al., 1987). The siliciclastic sediments resulted from erosion of the

adjacent Cordilleran highland (Molenaar and Rice, 1988) and from early Laramide

structures such as the Lost Soldier anticline (Winn et al., 1987). This occurred over a 2.4

million year period. The length of time of the transgression and regression suggests that

the Lewis Shale was a 3rd order sequence (Pyles, 2000).

In the Washakie basin, the Lewis Shale overlies the Mesaverde Group and

underlies the Fort Union Formation, Lance Formation and Fox Hills Sandstone. The
12

14N/96W 14N/95W

PM 1-13E

Cepo 21-18

Polar Bar 1
PM Fed 32-26
PM Fed 34-26
PM Fed 23-36

Triton #10
PM Fed 34-11

N Bogey Draw 1-14


6 mi Black Bar

13N/96W 13N/95W
Core 3C-3D survey

Image log 3D survey

Figure 2.2. Location map of the study area showing the seismic survey outline and wells
where core and borehole image logs are available.
13

Figure 2.3. Location of the Cretaceous Western Interior Seaway, in blue, with the outline
of North America, in red. Modified from Gill and Cobban (1970). From Minton (2002).
14

Lewis Shale is about 2,000 to 2,600 ft (600-800 m) thick over most of the study area, but

thins to the west and southwest to a thickness of 800 ft (250 m). Workers in the Lewis

Shale Consortium have divided the formation into three members: the lower shale, the

middle Dad Sandstone, and an upper shale member. The Dad Sandstone is the main

reservoir target within the Lewis Shale. The Lewis Shale represents a major marine

transgression. The shale of the lowermost Lewis is black, carbonaceous, and bioturbated

in places, with abundant shell debris. The maximum extent of the Lewis transgression

was to areas of the Rock Springs uplift in the west and the Wind River uplift in the north

(Winn et al., 1985). The Lewis Shale was primarily deposited by three depositional

mechanisms. They are, in order of increasing depositional energy, pelagic and hemi-

pelagic settling, fluid flows or turbidity currents, and cohesive mass gravity flows

(Minton, 2002).

At the western end of the Washakie basin, along the Rock Springs uplift, the

Mesaverde Group is comprised of four members. The three intervals of interest to this

study are, from the bottom, the Rock Springs Formation, the Ericson Formation, and the

Almond Formation at the top (Weimer, 1960). Figure 2.4 shows the stratigraphic column

of the area. The Mesaverde Group is mainly non-marine to marginal marine. In terms of

petroleum, historically the most commercially significant formation is the Almond. The

Almond is the uppermost sandstone of the Mesaverde Group; it is a fluvial, inter-deltaic,

and marginal marine deposit that lies conformably beneath the Lewis Shale (Miller,

1977). The Almond was deposited in marginal-marine and marine environments during

the final transgression of the Cretaceous seaway into southwestern Wyoming. The
15

Rawlins Uplift
Age Great Divide and
Washakie Basins
Paleogene Fort Union Fm.
Upper Cretaceous
Lance Fm.

Fox Hills SS

Dad Mbr.
Lewis Shale

Almond Fm.
Mesaverde

Ericson Fm.
Gp.

Rock Springs Fm.

Hiatus

Figure 2.4. Chronostratigraphic chart of the Upper Cretaceous, southern Wyoming.


(Modified after Schell, 1973). Stone Energy Company has picked nine seismic horizons,
these are shown with an arrow. The formations of interest to this study include the Fort
Union Formation, Lance A and B Formation, Fox Hills Sandstone, Lewis Shale, Dad
Sandstone Member, Almond Formation, Ericson Formation and the Rock Springs
Formation.
16

Almond intertongues with the Lewis Shale as a result of the second marine transgression

during the Campanian. The Almond ranges in thickness from 300 to 800 ft (90–240 m)

(Roehler, 1990), and is composed of sandstone, siltstone, shale, and coal beds, making

this a potential source rock.

Above the Lewis Shale, the Fox Hills Sandstone is a regressive shallow-marine unit

that contains marine shale and isolated coals locally (Cronoble, 1969). The Fox Hills is

stratigraphically equivalent to part of the Lewis Shale (Pyles, 2000). These formations

commonly interfinger and the contact between the two is gradational in nature. The Lance

Formation conformably overlies the Fox Hills Sandstone. The Lance consists of inter-

bedded carbonaceous shale, coal, siltstone, and sandstones (Weimer, 1960). There is some

interfingering between the marginal- marine facies of the Fox Hills and the nonmarine

sediments of the Lance.

The upper part of the Lewis Shale and the Fox Hills Sandstone represents the

withdrawal of the Cretaceous seaway from the area (Law et al., 1986). The overlying

Lance Formation was deposited primarily in an alluvial plain environment. In response

to continued Laramide deformation and the emergence of adjacent foreland uplifts, an

internal drainage system developed within the Lance Formation (Fouch et al., 1983).

2.3 Tectonic Framework

The study area is located in the southeastern portion of the Greater Green River basin

within two sub-basins, one of which is the Washakie basin. Figure 2.5 shows a
17

Figure 2.5. Regional tectonics of south-central Wyoming,. Structural contours are on


Precambrian basement in km below sea-level. The Washakie and Sand Wash basins are
drawn in a purple box. After Baars et al. (1988). From Minton (2002).
18

tectonic map of the study area and the surrounding region. The Washakie basin is bounded

by the Rock Springs uplift and the Uinta uplift on the west, the Wamsutter arch to the

north, and the Sierra Madre uplift to the east (Minton 2001), as shown in Figure 1.1. The

Greater Green River basin is a part of the Western Interior foreland basin that flanked the

cratonic side of a foreland thrust belt (Jordan, 1981). The foreland basin was created as a

result of the Sevier and Laramide tectonic activity that occurred during the Jurassic through

the earliest Eocene (Baars et al., 1988). Changes in the style of subduction of the Farallon

plate under the North American plate formed a complex variation of thrusting and

subsidence in the region (Cross, 1986). In south-central Wyoming, the Laramide uplift

began during the early part of the Maastrichtian. Evidence of early Laramide compression

in the area is seen at the Lost Soldier anticline (Reynolds, 1976).

The Cherokee arch lies to the south of the Washakie basin. Oblique-slip faulting on

the Cherokee arch, less than 15 mi (25 km) from the study area, has a surface expression of

a right-lateral wrench offset (Snoke, 1997). The Cheyenne belt crosses the study area in an

east-west direction, sub-parallel to and north of the Cherokee arch. The Cheyenne belt is

the suture zone that separates the Wyoming province from the Colorado province. This

accretionary boundary was formed between 1.79 and 1.75 Ga (Snoke, 1997). The

basement rocks in the study area are therefore Archean in age north of the Cheyenne Belt

and Proterozoic in age south of the Cheyenne Belt. The suture zone may have created a

weakness in the crust that has been reactivated at various times during the geologic past.

The Cheyenne belt geosuture is composed of a system of mylonite zones that are

from 0.4 to 4.4 mi (0.25 to 7.1 km) wide where exposed (Snoke, 1997). Surface
19

lineaments at the Cherokee arch are south of the projected line of the Cheyenne belt

(Krugh, 1997). From a Landsat image (Figure 2.6), surface expressions along the

Cherokee arch of this fault system can be mapped, as interpreted by Bader (1987). Here,

trends are interpreted to relate to a right-lateral wrench system. In contrast, Hull (2001)

believed that Harding’s (1974) model of a left-lateral system (Figure 2.7) could be used

to model the system. Direct linkage of movement or reactivation of the Cheyenne belt

and the Cherokee arch has not been established in the literature (Hull, 2001).

Structure is an important component of hydrocarbon production from the Lewis

Shale. Fields along the Cherokee arch produce from structural and structural-stratigraphic

traps. Hull’s (2001) study at the eastern edge of the Cherokee arch in south-central

Wyoming identified a left-lateral wrench system at the Lewis Shale level. He interpreted

complex positive flower structures (Figure 2.8), indicative of lateral movement along a

wrench fault with compression at a bend in the fault. The structure formed at a bend in the

projected path of the Cheyenne belt. A flower-type structure is therefore consistent with

reactivation of the suture zone. This flower structure is one of the criteria used by Harding

(1990) to identify wrench faults and structural inversion due to a change in stress

orientation. Sandbox experiments by McClay and Bonora (2001) illustrated deformation

that can be expected in such cases of strike-slip movement. Figure 2.9 shows the sandbox

model after 10 cm of left-lateral movement, and the interpretations that were made.
20

Figure 2.6. Landsat image of the Cherokee Arch. Interpreted in white are Bader’s (1987)
interpretations with a strain ellipse for a right-wrench fault. Outlined in yellow is the
approximate location of the CEPO/Powder Mountain study area. The Wyoming-
Colorado border is in the blue dotted line. After Bader (1987), from Hull (2001).
21

Figure 2.7. Left-lateral wrench model of Harding (1974). Hull (2001) identified this fault
system in the subsurface of the Sand Wash basin, through the use of 3D seismic data.
From Hull (2001).
22

Figure 2.8. Hull (2001) interpreted flower structures in his study area in the Sand Wash
basin. This is one of the criteria identified by Harding (1990) for wrench fault systems.
The insert shows the location and orientation of the seismic line shown. The master fault
is on the left, with smaller faults that steeped and join at depth. From Hull (2001).
23

Figure 2.9. Experiments performed by McClay and Bonora (2001) illustrate a 10 cm left-
lateral displacement of a sandbox model. The top figure shows the top of the sandbox
experiment and the lower figure shows the interpretations made based on this experiment.
24

2.4 Petroleum System

A petroleum system, as defined by Magoon and Dow (1994), “includes all the

elements and processes needed for an oil and gas accumulation to exist.” These include

made. Hull (2001) believed that this was analogous to the deformation in the eastern

Cherokee Arch. Due to the proximity of my study area, we may expect some of these

same wrench structures. the source, reservoir, seal, and the trap. Much of the gas from

the Greater Green River basin has been referred to as basin-centered gas by workers such

as Law (2002) and Surdam (1997). In general, basin-centered gas accumulations are

regionally pervasive accumulations that are gas saturated, abnormally pressured,

commonly lack a downdip water contact, and have low-permeability reservoirs. In the

context of a petroleum system, there are two types of basin-centered gas systems: a direct

type and an indirect type (Law, 2000). The attributes of these two types of systems are

provided in Table 2.1. Direct and indirect types are distinguished on the basis of source

rock quality; a direct type has a gas-prone source rock, and an indirect type has an oil-

prone source rock. This fundamental difference of source rocks leads to significantly

different characteristics, as shown in Table 2.1.

Table 2.1. Direct and indirect attributes. From Law (2000).

Hydrocarbon Nature of
Source Pressure Thermal
Type Migration Seal Upper
Rocks Mechanism Maturity
Distance Boundary
Gas-
Hydrocarbon Cuts across
Direct prone III Short Capillary > 0.7% Ro
generation statigraphy
kerogen
Oil – Thermal
Lithologic/ Bedding Highly
Indirect prone I/II Short/long cracking of
capillary parallel variable
kerogen oil to gas
25

Estimates of in-place gas resources contained in basin-centered gas accumulations

for the Greater Green River basin, within Cretaceous and Tertiary rocks, are as large as

5,063 TCF (Law et al., 1989), and the mean estimate of recoverable gas is 119.3 TCF

(Law, 1996). The source rock in the Greater Green River basin has been identified as the

Upper Cretaceous and lower Tertiary coal beds and carbonaceous shales in the Fort

Union, Lance, Lewis, Almond, and Rock Springs Formations. Organic matter is largely

gas-prone type III kerogen with additional contributions from thermally cracked oils

sourced from sapropelic coal beds (Garcia-Gonzales et al., 1997). Generation of the gas

occurred by both expulsion and migration during the late Eocene - late Oligocene,

between 40–25 Ma (Law, 2002). Studies of coal from the Almond Formation in the

Greater Green River basin by Garcia-Gonzalez et al. (1997) demonstrated that at the

basin center, most of the oil generated in the coal has been thermally cracked to gas,

whereas at the basin flank, the oil-to-gas reaction has barely begun. These concepts show

that the oil in the coal was generated during the alteration of liptinite macerals to a waxy

oil and a solid residue. The waxy oil was initially stored in porous structures and

subsequently in vesicles as the coal matured under increasing temperature. The primary

migration of the oil occurred as the generation of a sufficient volume of the waxy oil was

able to micro-fracture the coal vesicles. Pore pressure increased and allowed migration

of hydrocarbons out of the coal through interconnecting vesicles, and pores. Figure 2.10

illustrates this process.

The reservoir rocks range from Cretaceous to lower Tertiary sandstones.

Multiple, stacked reservoirs occur in 14,000 ft (4,267 m) intervals (Figure 2.11).


26

A)

E)
B)

C)

D)

Figure 2.10. Diagram of gas generation, storage and expulsion from coal. (A) Immature
coal (Ro < 0.5%) with laminated texture. (B) Mature coal showing development of
porous and vesicular texture due to oil generated and stored in pores (0.8% < Ro < 1.7%).
(C) Cracking of oil to gas increases the pore pressure causing microfractures and the start
of hydrocarbon expulsion (Ro > 1.2%). (D) Expulsion of gas and closure of pores (Ro >
2%). (E) Faulting further fractures the coal and enhances the permeability. From García-
González et al. (1997).
Figure 2.11. Cross section across the Washakie basin. Location of cross section is shown in Figure
2.1. Shaded area is the overpressured gas accumulation. From Law (2002).
27
28

Individual reservoirs range in thickness from 15 to 125 ft (4.6–38 m). The gas-bearing

interval does not commonly contain interbedded, water-bearing reservoirs. The Dad

Sandstone within the Lewis Shale is mainly the result of gravity flows and turbidity

currents. Average porosity across the area is about 13% and in-situ permeability

averages around 0.1 md. (Law, 2002). Numerous workers, including Surdam (1997),

concluded that the area consists of a fluvial-dominated depositional environment, with

marginal marine, deltaic, and barrier bar deposits.

The reservoir is overpressured, with gradients ranging from 0.5 to 0.9 psi/ft (Law,

2002). These anomalously pressured zones can be related to production zones, as shown

in Figure 2.12. Most of the production in this basin comes from these overpressured

zones. Regional seals are formed by capillary pressure and also by extensive bentonite

beds (Zainal, 2001). Thus, local structural and stratigraphic seals are important. Figure

2.13 shows the top of the overpressured zone in a cross-section across the Washakie

basin. Overpressuring is a distinguishing characteristic of these basin-centered gas

accumulations. According to Law (1986), another important feature of these gas

accumulations is that they form downdip from normally pressured, water-bearing

reservoirs (Figure 2.13), and they lack a downdip water contact. However there is a

possibility, that the sands are not in stratigraphic continuity.

Commercial gas has been found to exist in conventional structural and

stratigraphic traps (Shanley, 2003, per comm.) The level of thermal maturity at the top of

accumulation ranges from 0.7 to 0.9%, and is commonly0.8% Ro (Law, 1984). The

depth to accumulations ranges from 8,000 to 11,500 ft (2,438–3,505 m) (Figure 2.13).


29

Figure 2.12. Plot of cumulative production vs. distance to the top of the overpressure
zone for major gas reservoirs (>5 billion cubic feet (BCF)). This plot shows that >80%
of the gas production comes from these overpressured zones. Cumulative gas production
is calculated from total gas production in the history of the field. From Surdam (1997).
30

Figure 2.13. Illustration showing pressure zones for (A) Direct and (B) Indirect basin-
centered gas accumulations. From Law (2002).
31

The gas is of a thermal origin and is generally composed of ~90% methane, ~5% ethane

and longer chain hydrocarbons, ~5% carbon dioxide, and negligible nitrogen.

Condensate ranges from 5 to 70 bbl/mmcf gas.

Other workers in the Greater Green River basin such as Shanley et al. (2003) feel

that the basin-centered gas model is flawed. These workers believe that most of the

reservoirs found in this area are low-permeability conventional gas traps, and once this is

understood, risk can be lowered. An understanding of this leads to a significant

reductions in the available resource estimation, such as that given by Law (2002).
32

CHAPTER 3

BOREHOLE IMAGE LOGS AND CORES

Borehole image logs and core can be used to get an estimation of subsurface fracture

density for input into 3D modeling software. Knowledge of the spacing, orientation and

distribution of fractures in reservoir rocks can lead to a better understanding of the

production characteristics in a fractured reservoir.

3.1 Background

For the study, 4 image logs and 8 cores were examined in order to acquire data on

fractures. Table 3.1 gives the names and locations of the wells used in this study. These

wells are spread across a 4-township area as shown in Figure 3.1. The outer edges of the

survey area lack well data and the south and eastern edge of the survey lack image log data.

Two of the wells, PM 1-13E and PM Fed 34-11, have both types of data available,

however, only in the first well are both types of data at the same depth. Thus, a comparison

of data acquired from core and image logs is only possible in this one well. Figure 3.2

shows the depths of cores and borehole image logs in the ten study wells, as well as the log

tops for three horizons; the Lance, Lewis, and the Almond.
33

Lease name API Number Twnshp Range Sec Operator

Powder Mountain
49-037-24482 13N 96W 11 Stone Energy LLC
Unit 34-11

Westport Oil & Gas


Triton Unit 10 49-037-21922 13N 95W 8
CO. LP

Bogey Draw 1-14 49-037-22991 13N 95W 14 Celsius Energy

Black Bar #1 49-037-23169 13N 95W 30 Celsius Energy

Powder Mountain
49-037-24354 14N 96W 36 Stone Energy LLC
Unit 23-36

Powder Mountain
49-037-24178 14N 96W 26 Stone Energy LLC
34-26X

Powder Mountain Basin Exploration


49-037-24126 Y 14N 96W 26
FED # 32-26 INC

Polar Bar #1 49-037-23037 14N 96W 22 Stone Energy LLC

Powder Mountain
49-037-24237 14N 96W 13 Stone Energy LLC
Unit 1-13E

Cepo Lewis 21-18 49-037-24185 14N 95W 19 EOG Resources INC

Table 3.1. Table giving locations for the wells used in the study. From WOGC website,

(accessed 3-09-03).
34

14N/96W 14N/95W

PM 1-13E

Cepo 21-18

Polar Bar 1
PM Fed 32-26
PM Fed 34-26
PM Fed 23-36

PM Fed 34-11 Triton #10

Bogey Draw 1-14

Black Bar 1

13N/96W 13N/95W
6 mi

Core Borehole Image Log

Figure 3.1. Map of the study area showing the location of the wells with cores and
borehole image logs. Two wells have both core and image logs available; PM 1-13E and
PM Fed 34-11.
35

Bogey
PM CEPO Polar PM PM PM PM Triton Draw Black
1-13E 21-18 Bar 1 32-26 34-26 23-36 34-11 #10 1-14 Bar 1

Lance
11,000

12,000

13,000
Lewis

Almond

14,000

Core

15,000 Perforations

FMI/EMI Log

Well picks

16,000 Inferred picks

Figure 3.2. Location of cores and image logs in terms of producing intervals. Three
producing horizons are shown; the Lance, Lewis and the Almond formations. This figure is
a quick reference to show that there are 3 cores and 2 image logs available in the Lewis
36

A borehole image tool provides a 3D image of the borehole. Of the 4 image logs

available, 1 (PM 1-13E) was run using Schlumberger’s Fullbore Formation MicroImager

(FMI) tool and 3 (PM Fed 32-26, PM Fed 23-36 and PM Fed 34-11) were run using

Halliburton’s Electrical Micro Imaging (EMI) tool. Both tools work in a similar fashion, in

that they take an image of the hole by measuring the electrical resistivity of the rocks and

fluids. These tools allow for visualization of small-scale features such as fractures and

faults. Sedimentary features are also easy to identify, such as dip, paleocurrent direction,

beds, slumps and bed boundaries.

The FMI tool is the more widely recognized imaging tool in industry compared to the

EMI tool; it has 80% coverage of the borehole. Figure 3.3 gives an illustration of the tool,

which consists of 4 pads with each pad having 24 measuring electrodes. Attached to each

pad is a flap with a similar electrode array, allowing the tool to have 192 measuring

electrodes. Electrical currents go 30 in. (76 cm) into the formation to produce an electrical

image. Special focusing circuitry ensures that the measuring currents are forced into the

formation.

The high vertical and azimuthal resolution means that features as small as 0.2 in

(5.08 mm) are visible. However any features smaller than 0.2 in (5.08 mm), depending on

the resistivity contrast with the background rock, can be estimated by quantifying the

current flow to the electrode. In conformance with sampling theory, the FMI image is

sampled at half the resolution (0.1 in, 2.54 mm) vertically and horizontally so that the

theoretical resolution is not compromised. Sampling horizontally for the button diameter
37

Figure 3.3. Diagram of the Schlumberger FMI tool. On the left the tool is shown in an
open position with the pads fully extended. On the right is a close-up view of the pad and
flap. (After Schlumberger FMI brochure website).
38

of 0.2 in (5.08 mm) does this; the buttons are arranged on the pad and flap in two

horizontally offset rows. Thus, fine-scale details, such as 50-micron fractures filled with

conductive fluids, are seen as dark near-vertical lines.

The tool has 2 electrical current components. The high-resolution component

detects micro-resistivity variations in the formation that directly face each button and

provides the data used for imaging and dip interpretation. The variation of resistivity

values in front of each individual button creates the high-resolution image. The low-

resolution component is modulated by the resistivity interval between the upper and lower

pads (Schlumberger web site, 2003). Both signals are rich with petrophysical and

lithological information. The image is normalized through calibration with low frequency,

deeper resistivity measurements from the tool signal or from another resistivity

measurement tool. The information collected is then processed into an image of the

borehole wall. The FMI resistivity image is a 360º image of the borehole wall, which is

“unrolled” and presented as a 2D cylindrical image to be viewed. The data provides a basis

for interpreting bed boundaries, faults, fractures, stacking patterns, paleocurrent directions,

and horizontal compressional forces.

The image can be displayed in both a “static” and “dynamic” mode (Figure 3.4). The

static mode is used to compare resistivity values over the entire length of the borehole.

This is normalized over the entire logged interval and assigned a color for each resistivity

measurement. The darker colors represent low resistivity rocks and fluids, and the lighter

colors indicate higher resistivity materials. Generally, the light colors correspond to
39

Static Dynamic

Figure 3.4. Static and dynamic modes of the EMI tool. Light color corresponds to

sandstone and darker lower resistivity material is the shalier units. This example is from

the Powder Mountain 23-36 well.


40

sandstone and the dark colors to shale. The dynamic mode is used to enhance small

contrasts in resistivity over a short interval. The dynamic mode normalizes resistivity

contrasts over a certain viewing interval. This creates a full range of available colors over

only that interval. This enhanced contrast allows subtle features to be seen and thus aids in

interpretation.

The maximum temperatures and pressures for the tool are 20,000 psi and 350°F

(175°C). Borehole diameters range from 5 7/8 – 21 in (15 – 53 cm) and the maximum hole

deviation can be 90°. There are 3 logging modes possible: full-bore, four-pad and dipmeter

modes. The logging speeds are 1,800, 3,600 and 5,400 ft/hr, respectively, with a real-time

processed image. The resolution of the FMI is 0.2 in (0.5 cm) (Schlumberger web site,

2003).

The other imaging tool used in my study area is Halliburton’s EMI (Electrical

Micro Imaging) tool (Figure 3.5). The main difference between these tools is that the EMI

has scanning electrode arrays on each of six independent arms and the pads have consistent

contact with the borehole wall, which provides an accurate, sharp image. Electrical images

are made from a multi-pad tool that maps out the resistivity properties. The pads are

capable of rapid sampling and have excellent vertical resolution (0.1 in, 2.54 mm), which is

about the same as the FMI tool. The hardware capabilities for the EMI tool are shown in

Table 3.2.
41

Fiberglass
sleeve

Instrument
section

6–arm
sonde

Imaging
pad

Figure 3.5. The Electrical Micro Imaging (EMI) tool configuration. From Halliburton
EMI brochure.
42

Table 3.2. EMI hardware characteristics

Source Type Induced current

150 microresistivity sensor (25 on each of 6 independent


Sensor Type
swivel pads)

2 rows containing 12 and 13 sensors, respectively 0.3 in (7.6

mm) between rows; 0.2 in (5.08 mm) between sensors on each


Sensor Spacing
row; 0.1 in (2.54 mm) between sensors when both rows are

superimposed

Sampling rate 120 samples/ft at 20 ft/min (6 m/min)

Sampling Size 0.1 in (2.54 mm) (resistivity)

3.2 Methods

3.2.1 Core

Eight cores in the area were examined for fractures. Figure 3.1 shows the location of

these cores and Figure 3.2 shows the depths of the cored intervals. These cores included

the CEPO 21-18, Powder Mountain Unit 1-13 E, Black Bar 1, Bogey Draw 1-14, Powder

Mountain Fed 34-11, Triton Unit 10, Powder Mountain Fed 32-26 and Polar Bar 1. Full

core descriptions, including lithology, have previously been carried out by Minton (2002)

for the CEPO 21-18 and Powder Mountain Unit 1-13 E cores. However, his descriptions

did not include details concerning fractures in these wells. Thus, all 8 cores have been
43

examined solely for fractures. Core descriptions are included in Appendix A. Based on

Kluth (pers. comm.), described items should include:

Orientation and dip magnitude of fractures.

As the core was not oriented, true attitude could not be determined, however, dip

magnitude relative to bedding could be measured.

Height (especially in relation to bed thickness)

Aperture of the fracture

Shear displacement

Mineralization (type and description)

Fracture surface features:

- Slickensides

- Fibrous mineralization

- Pressure solution features

- Plumose structures

Relative age

- to other joint sets

- to other structures

Density / Spacing

- Relationship to lithology

- Relationship to structure

- Relationship to other geological variables


44

3.2.2. Borehole Image Logs

The software used for the FMI log interpretation was Baker Atlas Review/Recall.

Janine Carlson from Colorado School of Mines processed this log data. Minton (2002)

interpreted bed boundaries in this log and identified areas where the tool was stuck and

where washouts occurred (Table 3.3). The overall quality of the log run was poor. The

borehole diameter for the well was 6.125 in (15.56 cm), close to the minimum borehole

diameter requirement for the FMI tool of 5.875 in (14.92 cm). This may have been a

contributing factor to the poor FMI quality. Figure 3.6 shows unsynchronized pads and

tool sticking. Streaked images indicate where the tool got stuck and was then released.

Streaking occurs when the tool travels too fast for it to properly record the image or when

mud or debris builds up on the pads. The rapid rate of movement results when the tool

releases after sticking in response to increased tension on the wireline. Out-of-phase pads

occur due to sticking and jostling of the flaps and possible disruptions to the accelerometer.

A number of criteria were used when investigating FMI logs. Types of fractures

were interpreted. Open fractures appear as dark, irregular, steeply dipping lines on the

borehole image log. Open fractures are more conductive than the surrounding matrix.

Selecting the exact sine wave is difficult because the surfaces are irregular and sensitive

to changes in lithology. There is therefore a margin of error of a few degrees in


45

Table 3.3. This table shows the intervals in the FMI log PM 1-13E where no data was
recorded by the tool due to sticking or other problems.

Bottom depth of interval (ft) Top depth of interval (ft) Number of feet missing

13,434.0 13,425.5 8.5

13,238.0 13,237.0 1.0

13,214.5 13,203.5 11.0

13,193.5 13,180.0 13.5

13,148.3 13,145.0 3.3

13,091.0 13,084.0 7.0

12,996.6 12,980.0 16.6

12,975.0 12,972.0 3.0

12,816.0 12,815.0 1.0

12,806.2 12,802.5 3.7

12,780.0 12,777.6 2.4

12,765.2 12,763.0 2.2

12,749.1 12,748.8 0.3

12,745.2 12,733.5 11.7

12,721.4 12,720.8 0.6


46

Static 12996 ft Dynamic

Vertical scale: 1 ft 13000 ft

Figure 3.6. FMI image showing out-of-phase primary and auxiliary pads. At the top of the
figure we see streaking from where the tool was stuck, then released and slid up the hole
too quickly for an image to be recorded. Out-of-phase pads are common below intervals
where no data are recorded, as we see here. The bed boundaries are shown in green, this
example is from Powder Mountain Unit 1-13E.
47

selecting the exact dip and dip direction of the surface. Examples of open fractures are

shown in Figure 3.7. A second type of fracture identified in the FMI is healed fractures.

These appear light in color, mostly yellow, due to the growth of electrically less resistive

minerals within that fracture. These are illustrated in Figure 3.8. The third types of

fractures are actually micro-faults. This is perhaps the easiest to identify due to the

presence of bed offset. The last type of fractures identified in borehole image logs are

induced fractures (Figure 3.9). Induced fractures may appear dark in color like open

fractures, as they too get filled with low-resistivity drilling mud. There are a few

identifying criteria that can be used to distinguish differences between the two. Induced

fractures tend to cut across beds regardless of lithology, the fractures are sub-vertical, with

almost a 90º dip, and they tend to be much longer than an open fracture, in many cases

exceeding 10 ft (3 m).

An induced fracture is any rock fracture produced by human activities, such as

drilling, accidental or intentional hydrofracturing, core handling, etc. (Kulander et al.,

1990). Induced fractures are produced when stress from drilling pressure exceeds the rock

strength (Figure 3.10). The orientation of these fractures is controlled by present-day

horizontal stress (Kulander, 1988). Induced fractures can be used to interpret the direction

of current horizontal compression. Thus, induced fractures strike parallel to the maximum

horizontal stress direction. Grace and Newberry (1998) identified a number of


48

0° 90°
10,274 ft

10,280
10,280 ft
ft

Vertical scale: 2 ft

Figure 3.7. Dark irregular near-vertical features are open fractures. The dark material is
the low-resistivity drilling mud. Red sine waves are fitted to the fractures and the dip
symbols on the left give the dip track and depth track. On the right are the dips and dip
direction values for each fracture, interpreted by Halliburton. Portion of the EMI log from
the Powder Mountain 32-26 well.
49

0° 90°

12,010

223
53
208
59

47
65
12,016

Vertical scale: 2 ft

Figure 3.8. An example of healed fractures from the Powder Mountain 23-36 well. Two
healed fractures are shown with the red sine waves fitted to them. A fill with a lighter
color, high resistivity material, characterizes them. Depth is shown in feet.
50

11,300 ft

11,300

209

86

205
66

11,306 ft

Vertical scale: 2 ft

Figure 3.9. An example of an induced fracture from the Powder Mountain 34-11 well. The
relatively straight vertical black line is an interpreted induced fracture. The blue sine wave
is fitted to an open fracture, the main distinguishing factor between the two is the steeper
dip of the induced fracture. Both tend to be filled with higher resistivity drilling mud,
giving a dark color. Inclined fracture segments on opposite sides of the wellbore are
actually part of a single induced fracture. The forth pad from the left was not functioning
properly. Depth is shown in feet.
51

Figure 3.10. Schematic diagram of a wellbore. As the core bit is lowered while rotating,
stress fields build up due to the drilling process. These cause the formation of induced
fractures, also known as petal fractures, which appear as the dark red lines cascading down
from the top of the bit (arrow). From Narr (personal communication 2003).
52

criteria that can be used to identify induced fractures. These include:

Fracture never crosses the borehole, that is, it does not make a sine wave,

Often has curvature at termination,

Always open – no mineral fill,

Cannot be micro-faulted,

Oriented parallel to maximum and intermediate principle stresses, usually vertical,

and

Oriented along the least principle stress direction.

Angle from horizontal is nearly vertical (> 75°)

3.2.3. In situ Stress

In situ stress directions can be found by analysis of borehole breakouts and induced

fractures. Breakouts form during or shortly after the drilling process and progress with

time. They are elongations that form as a result of stress concentrations in a non-uniform

stress field (Springer, 1987). Breakout elongation is defined in a plane orthogonal to the

borehole axis, and is perpendicular to the maximum horizontal stress direction (sHmax).

Breakouts are formed by failure along two pairs of conjugate shear fractures, which form

tangential to the borehole circumference (Zheng et al., 1989). Induced hydro-fractures and

open fractures tend to form parallel to the maximum horizontal stress field. Older healed
53

fractures, if oriented parallel to the maximum horizontal stress field, will tend to re-open as

an adjustment to the new stress regime (Knight, 1999).

During the logging process, borehole orientation parameters, including hole

azimuth (HAZI) and vertical hole deviation (DEVI), are continuously measured. One pad

of the four-arm caliper tool (pad 1), is oriented with respect to north, and its azimuth

(P1AZ), is continuously measured. While the logging tool moves up the hole, it rotates

due to cable torque. The tool stops rotating when it encounters a borehole elongation, as

the pads become ‘locked in’ (Knight, 1999). These intervals are potential borehole

breakouts. Numerical data are identified for possible breakouts by screening the data for

four criteria that may also give anomalous numerical data in the log; everything left is

a possible borehole breakout. Figure 3.11 illustrates a borehole elongation or breakout.

Here, a plan view of a borehole with the horizontal maximum and minimum stress is

shown in relation to the orientation of drilling induced fractures.

1. Not all borehole elongations are stress-induced breakouts. Springer (1987)

proposed criteria for recognizing breakouts from borehole image logs, such as the

FMI. The first step is to eliminate elongations caused by logging/drilling artifacts,

such as mudcake formation, washouts and keyseats. A keyseat is


54

Figure 3.11. Diagram illustrates borehole breakouts and induced fractures. Drilling
induced fractures are oriented parallel to the SH max direction, or the horizontal
maximum stress direction. However, borehole breakouts occur in the direction of least
horizontal stress or SH min, perpendicular to the SH max direction. After Minton
(2002) and Rahmat (2000).
55

formed when drill pipe rubs against the high side of the borehole, this also causes

an elongation in the direction of the hole azimuth. These four steps were performed

and elongations remaining after this were identified as possible breakouts.

2. Intervals where the tool was rotating were eliminated. This was determined from

plots of the P1AZ curve, where the curve was changing and not stabilized.

3. Washout zones were eliminated. This is a zone where the smallest caliper is 1 inch

or more larger than the bit size.

4. Intervals where the elongation was less than or equal to 0.25 inches were viewed as

zones of ‘no elongation.’ This can be found byC13-C24(absolute values of

calipers 1 and 3 minus calipers 2 and 4). Elongations caused by mechanical

keyseats were eliminated. These were detected by comparing P1AZ (pad 1

azimuth) (and 90° increments of P1AZ) to HAZI (hole azimuth). If any value of

P1AZ was within ±10° of HAZI or a 90° increment of HAZI, the zone was

eliminated as a breakout. This is because elongation is in the direction of the hole

azimuth.

Barton et al. (1988) proposed a method for using borehole width to estimate stress

magnitudes. However, there is much disagreement, due to the difficulty in obtaining

reliable measurements, and the need exists to better understand the mechanics of rock

failure.
56

3.3 Results

3.3.1. Core

Fractures in all 8 cores in the area were closely studied (Appendix A). One of the

first things noted was the prevalence of bed parallel slickensides. Bed parallel slickensides

can be seen in all cores, although they tended to be more common in the shalier cores. The

slickensides tended to occur near the top and base of the sandier units, close to the shale

beds. They ranged in size from 2 mm to fully horizontal features cutting straight across

the core at almost 6 cm. The slickensides were created after lithification of shale beds, and

are not considered to be sedimentary features. Horizontal slickensides were not observed

along fracture faces. Bed-parallel slickensides in organic rich shales in both cores were

interpreted to be the result of movement due to stress. The amount of movement appears to

be small, as there is no offset seen in the cores.

Evidence for pressure solution is observed in minor stylotization and calcite-filled

fractures. ‘Stylolites’ are defined by Groshong (1975) as “zones of relatively insoluble

residue against which the original fabric elements of the rock are truncated by removal

rather than offset.” The majority of the stylolites were observed along bedding surfaces;

these are large jagged surfaces likely to be associated with diagenetic processes (Hennier,

1984). Displacement along bedding surfaces was never measurable due to the absence of

offset markers across bedding planes. However, evidence for movement along bedding

surfaces can be observed by slickensides.


57

All fractures only occurred in sandstone units; they did not cut lithological

boundaries, and tended to die out in shalier areas. They all tended to be very narrow in

aperture (only a few mm), whereas they were a few cm in length. In all cores, there were

two types of fractures present. The first type contains quartz mineralization in the fracture

aperture. These partially healed fractures have quartz crystals precipitating along the

fracture walls. Fully grown crystal faces can be seen, indicating that the quartz crystals

grew into an open space. The width of these quartz fractures varies, but they ranged from

2-3 mm. An example of a partially filled quartz fracture can be seen in the core

photographs of the Polar Bar #1 well (Figure 3.12). In this example, the growth of quartz

kept the fracture open for the flow of hydrocarbons. The darker areas in the figure are

where no minerals have been able to grow. Thus, we note that the fracture is only partially

healed. A tortuous path is still possible for the movement of hydrocarbons. These types of

fractures are important. Pasternack (Ph.D, dissertation in progress) noted that some of his

wells contain fractures similar to those found in the CEPO area and those shown in Figure

3.12.

The second type of fracture is a calcite-lined fracture. The fracture first formed, but

has since closed due to the precipitation of calcite. However, at some point in the

future, the fracture was then re-opened parallel to the pre-existing fracture, using this as a

plane of weakness. We now have calcite cement on one surface of the fracture and clean

rock on the other. Figure 3.13 gives an example of this second style of fracturing
Figure 3.12. Polar Bar # 1.
Core Depth:
15,229 ft. This partially
healed fracture is 13 cm (7
in) long and up to 2 mm in
aperture. Quartz crystals can
be seen growing in the open
space. The crystal growth
kept the fracture open for
flow of hydrocarbon. Darker
areas are where no mineral
growth has occurred. The
10 mm 10 mm fracture gets wider near the
base and stops at a lithology
change at the base (shale).
58
Figure 3.13. Polar Bar #
1. Core Depth: 15,270
ft. Calcite-lined
fracture, re-opened
along a plane of
weakness. 1 ft (30.4
cm) long fracture, the
white areas inside the
dotted lines are a
calcite-cement filling
and the dark crack on
the left of this is an
open space (0.25 mm).
The fractures dies out
near the base, as the
lithology gets shalier
(shown by the green
arrow).
10 mm
59
60

seen in the cores. A question remains whether re-opened fractures are drilling induced

fractures or not. In Figure 3.13, there is some evidence against this idea,

but the interpretation is equivocal. I suspect that it is a re-opened fracture and not induced

because the fracture dies out slowly as the lithology gets shalier, at the base of the figure.

It does not cut through lithology as a drilling induced fracture may. Secondly, while these

fractures tend to be longer than the quartz-type fractures, they do not get longer than 1 ft

(30.4 cm) in any of the cores seen. However, it is difficult to distinguish, as they both have

a high dip angle.

The number and types of these fractures were counted in each of the 8 cores in the

study area. Figure 3.14 shows the number of fractures found in each core. However, this

diagram may be a little misleading because the lengths of the cores vary. For instance, the

Powder Mountain 34-11 has 145 ft (44.2 m) of core, and 7 fractures were identified.

We can compare this with the Cepo Lewis 21-18 core of 30 ft (9 m) that has 2

fractures identified. Thus, having the average vertical spacing of the fractures may be of

more value. Table 3.4 presents the average vertical fracture spacing in each well. The

table shows that the wells with the highest fracture densities are actually the ones

highlighted in red. They are not the wells with the highest fracture number as shown in

Figure 3.14. The total average spacing of all the wells together in the field is about 1

fracture/22 ft (1 fracture/6.7 m).


61

Table 3.4. Table showing the average vertical fracture spacing from core data, in this
study, this is defined as the length of core over the number of fractures observed. In red are
low fracture spacing values.

Well Average vertical fracture spacing (ft)

PM 34-11 20.7

PM 34-26 12

Polar Bar #1 30

Black Bar #1 12.5

Bogey Draw #1-14 15

PM 1-13E 13.3

Cepo Lewis 21-18 15

Triton #10 60
62

(A) Fractures from core data


8
7
6
# of fractures

5
4
3
2
1
0
PM 34-11 PM 34-26 polar black bogey PM1-13E Cepo triton#10
bar#1 bar#1 draw#1 Lewis 21-
Wells 18

(B) Fractures from core data


9
8
# of fracs per 100 ft

7
6
5
4
3
2
1
0
PM 34-11 PM 34-26 polar black bogey PM1-13E Cepo triton#10
bar#1 bar#1 draw#1 Lewis 21-
Wells 18

Figure 3.14. Graph (A) shows the number of fractures found in each of the 8 cores of the
study area. Due to core length variation fracture values in (B) have been normalized to
fractures per 100 ft of core.
63

Most fractures had some mineralization associated with them. Figure 3.15 shows

the type of mineralization that occurred in each well. In only 1 case was there no

mineralization found in or around a fracture. The quartz-style fractures were found in a

number of wells, but were much less prevalent than the calcite-lined fractures. In one of

the wells, there were a few cases where both types of mineralization occurred. In this case,

they appeared as quartz crystals with some drusy calcite crystals on top.

3.3.2. Borehole Image Logs

There were four wells available in the study area with borehole image logs

available. These are the Powder Mountain 1-13E, Powder Mountain 23-36, Powder

Mountain 32-26, and the Powder Mountain 34-11. For each of the 4 borehole image logs

available, the sine waves fitted to the fractures gave dip and dip direction data (Appendix

B). As Appendix B shows, there were very few healed fractures picked in the image logs.

The Powder Mountain 32-26 had 146 fractures picked, of which only 3 were interpreted as

healed fractures. The Powder Mountain 1-13E had 46 fractures picked and 2 of these were

identified as healed fractures. No healed fractures were identified in any of the other 2

wells. These data were then plotted on stereonets. The data were screened for dips greater

than 80° and these points were removed. The remaining points were interpreted as open

fractures. The removal of the points was done in order to remove induced fractures.
64

Fracture Mineralization Types


7

5
# of mineral

4
3

1
none
0
both PM 34-11 PM 34-26 Polar Bar Black Bar Bogey PM1-13E Cepo Triton #10
#1 #1 Draw #1 Lewis 21-
calcite 18
quartz Wells

Figure 3.15. Graph showing the type of mineralization found in each well. Fractures
tended to be mineralized and calcite was the most common mineral present.
65

Figure 3.16 shows the Powder Mountain 1-13E data. One fracture set was present,

and no fractures were identified with dips over 80°. Figures 3.17 and 3.18 show the data

for the Powder Mountain 23-36 well. The first figure shows all open fractures and Figure

3.18 shows open fractures for Fracture Set 1, divided into dips over and under 80. This

well had one fracture set that could be identified, but in this case there were a number of

fracture dips picked that were greater than 80°, thus 2 stereonets were made to get 2 vector

means for use in the model building. The last 2 wells studied; Powder Mountain 32-26 and

34-11 both had 2 Fracture Sets identified. Figures 3.19 first shows all fractures present in

the EMI log, 2 fracture sets were identified. Figure 3.20 shows all Fracture Set 1 present in

the well, these are of open and healed fractures. Figures 3.21 and 3.22 are Fracture Set 1

broken into dips over and under 80°. Figure 3.23 is also from Powder Mountain 32-26, but

it is of Fracture Set 2, no dips over 80° were identified in this well. Figures 3.24 shows all

fractures present in the Powder Mountain 34-11 well for both open and healed fractures, 2

fracture sets were identified. Figure 3.25 shows all Fracture Set 1 present in the well, these

are of open and healed fractures. Figure 3.26 is Fracture Set 1 broken into dips under 80°

and Figure 3.27 is of fractures picked over 80° for both fracture sets. Figure 3.28 show all

fracture picked for Fracture Set 2, Figure 3.29 is of Fracture Set 2 with dips under 80°.

Table 3.4 gives a summary of the observed open fracture orientations from the 4

image logs. Figure 3.30 shows the frequencies of fracture dip for these 4 borehole image
66

healed fracture a)

open fracture

Vector Mean
Dip direction Dip
246° 66°

b)

c) 1 Sigma

2 Sigma

3 Sigma

4 Sigma

5 Sigma

6 Sigma

7 Sigma

8 Sigma

Figure 3.16. (a) Stereonet of all open and healed fractures for the Powder Mountain 1-13E
well. (b & c) Dip magnitudes less than 80°, with contour map of this. This is a Schmidt
lower hemisphere projection. Kamb contours are shown. The plane perpendicular to the
vector mean has a dip direction/dip of 246°/66°. Sigma values represent standard
deviations from a random population.
67

Vector Mean
Dip direction Dip
205° 81°

2 Sigma

4 Sigma

6 Sigma

8 Sigma

10 Sigma

12 Sigma

14 Sigma

16 Sigma

18 Sigma

Figure 3.17. Stereonet of all open fractures for the Powder Mountain 23-36 well. This is a
Schmidt lower hemisphere projection. Kamb contours are shown. The plane perpendicular
to the vector mean has a dip direction/dip of 205°/81°. Sigma values represent standard
deviations from a random population.
68

a) b)

1 Sigma

Vector mean:
Dip direction Dip
197° 67°
10 Sigma

c)
d)

1 Sigma

Vector mean:
50 Sigma Dip direction Dip
209° 86°

Figure 3.18. Stereonets for the Powder Mountain 23-36 well. This is a Schmidt lower
hemisphere projection. a) dip magnitudes > 80° removed, open fractures, b) corresponding
Kamb contours, c) only dip magnitudes > 80°, induced fractures, d) corresponding Kamb
contours. Sigma values represent standard deviations from a random population.
69

healed fracture

open fracture

Figure 3.19. Stereonet of all open and healed fractures for the Powder Mountain 32-26
well. Dip magnitudes are less than 80°. This is a Schmidt lower hemisphere projection.
Kamb contours are shown. Two sets of points can be seen, the main set is circled in the
dotted line, while the second set is circled in the solid line.
70

Vector Mean
Dip direction Dip
192° 72°

5 Sigma

15 Sigma

25 Sigma

35 Sigma

45 Sigma

55 Sigma

65 Sigma

Figure 3.20. Stereonet of all open fractures for Fracture Set 1 for the Powder Mountain 32-
26 well. This is a Schmidt lower hemisphere projection. Kamb contours are shown. The
plane perpendicular to the vector mean has a dip direction/dip of 168°/72°. Sigma values
represent standard deviations from a random population.
71

Vector Mean
Dip direction Dip
199° 67°

2 Sigma
4 Sigma
6 Sigma
8 Sigma
10 Sigma
12 Sigma
14 Sigma
16 Sigma
18 Sigma

Figure 3.21. Stereonet of open fractures for the Powder Mountain 32-26 well. Main
fracture set, with dip magnitudes only < 80°. This is a Schmidt lower hemisphere
projection. Kamb contours are shown. The plane perpendicular to the vector mean has a
dip direction/dip of 199°/67°. Sigma values represent standard deviations from a random
population.
72

Vector Mean
Dip direction Dip
203° 83°

5 Sigma

15 Sigma

25 Sigma

35 Sigma

45 Sigma

55 Sigma

65 Sigma

Figure 3.22. Stereonet of induced fractures for the Powder Mountain 32-26 well. Main
fracture set, with induced fractures alone present, those with dip magnitudes >80°. This is
a Schmidt lower hemisphere projection. Kamb contours are shown. The plane
perpendicular to the vector mean has a dip direction/dip of 203°/83°. Sigma values
represent standard deviations from a random population.
73

Vector Mean
Dip direction Dip
7° 54.7°

1 Sigma

2 Sigma

3 Sigma

4 Sigma

5 Sigma

6 Sigma

7 Sigma

8 Sigma

9 Sigma

Figure 3.23. Stereonet of open fractures for the Powder Mountain 32-26 well. Second
fracture set, with dip magnitudes only < 80°. This is a Schmidt lower hemisphere
projection. Kamb contours are shown. The plane perpendicular to the vector mean has a
dip direction/dip of 7°/55°. Sigma values represent standard deviations from a random
population. No values over 80° were present in this second fracture set.
74

Figure 3.24. Stereonet of all open fractures for the Powder Mountain 34-11 well. No
healed fractures were identified in the fracture interpretation. Dip magnitudes are less that
80°. This is a Schmidt lower hemisphere projection. Kamb contours are shown. Two sets
of points can be seen, the main set is circled in the dotted line, while the second set is
circled in the solid line.
75

Vector Mean
Dip direction Dip
223° 71°

5 Sigma

15 Sigma

25 Sigma

35 Sigma

45 Sigma

55 Sigma

Figure 3.25. Stereonet of all open fractures for Fracture Set 1 for the Powder Mountain 34-
11 well . This is a Schmidt lower hemisphere projection. Kamb contours are shown. The
plane perpendicular to the vector mean has a dip direction/dip of 223°/71°. Sigma values
represent standard deviations from a random population.
76

Vector Mean
Dip direction Dip
211° 63°

2 Sigma
4 Sigma
6 Sigma
8 Sigma
10 Sigma
12 Sigma
14 Sigma
16 Sigma
18 Sigma
20 Sigma
22 Sigma

Figure 3.26. Stereonet of open fractures for the Powder Mountain 34-11 well . Main
fracture set, with dip magnitudes over >80° removed. This is a Schmidt lower hemisphere
projection. Kamb contours are shown. The plane perpendicular to the vector mean has a
dip direction/dip of 211°/63°. Sigma values represent standard deviations from a random
population.
77

Vector Mean:
Dip direction Dip
214° 85°

4 Sigma
8 Sigma
12 Sigma
16 Sigma
20 Sigma
24 Sigma
28 Sigma
32 Sigma
36 Sigma

Figure 3.27. Stereonet of induced fractures for the Powder Mountain 34-11 well. Only
induced fractures, those with dip magnitudes >80 are shown. This is a Schmidt lower
hemisphere projection. Kamb contours are shown. The plane perpendicular to the vector
mean has a dip direction/dip of 214°/85°. Sigma values represent standard deviations from
a random population.
78

Vector Mean:
Dip direction Dip
14° 56°

1 Sigma

5 Sigma

10 Sigma

15 Sigma

20 Sigma

25 Sigma

30 Sigma

35 Sigma

40 Sigma

Figure 3.28. Stereonet of all open fractures for Fracture Set 2 in the Powder Mountain 34-
11 well. This is a Schmidt lower hemisphere projection. Kamb contours are shown. The
plane perpendicular to the vector mean has a dip direction/dip of 14°/56°. Sigma values
represent standard deviations from a random population.
79

Vector Mean:
Dip direction Dip
18° 54°

1 Sigma
2 Sigma
3 Sigma
4 Sigma
5 Sigma
6 Sigma
7 Sigma
8 Sigma
9 Sigma
10 Sigma

Figure 3.29. Stereonet of open fractures for the Powder Mountain 34-11 well. Second
fracture set, with induced fractures removed, those with dip magnitudes >80. This is a
Schmidt lower hemisphere projection. Kamb contours are shown. The plane perpendicular
to the vector mean has a dip direction/dip of 18°/54°. Sigma values represent standard
deviations from a random population.
80

Table 3.5. Summary of the average observed fracture orientations from


borehole image logs.

PM 23-36 PM 1-13E PM 32-26 PM 34-11

Dip Dip Dip Dip

direction Dip direction Dip direction Dip direction Dip

Set 1 197° 67° 246° 66° 199° 67° 211° 63°

Set 2 - - - - 7° 55° 18° 54°

Induced

fracs. 209° 86° - - 203° 83° 214° 85°

Average fracture orientations for Fracture Set 1 and 2.

Dip direction Dip

Set 1 213° 65°

Set 2 12° 54°


81

P M 3 4 -1 1
60

50

40
Fr e q 3 0
20

10

0
10 20 30 40 50 60 70 80 90
d ip

PM 32-26
60

50

40

Freq 30

20

10

0
10 20 30 40 50 60 70 80 90
dip

P M 2 3 -3 6
30
25
20
Fr e q 1 5
10
5
0
10 20 30 40 50 60 70 80 90
d ip

PM 1-13E
5

3
Freq
2

0
10 20 30 40 50 60 70 80 90
dip

Figure 3.30. Frequencies of dip magnitude for all open fractures interpreted from borehole
image logs. Most dips have a high angle, >70°.
82

logs. For all wells, there are a large number of high dip angles, over 70°. As stated above,

all fractures over 80° were removed from the stereonet plot, as they were potential induced

fractures. However, this technique did not discriminate between induced and open or

healed fractures, and we need to keep in mind that the drilling tool may not have caused

some of those fractures that were removed.

As stereonets give no depth information, Figures 3.31 – 3.34 were constructed to

show the relationship between the depths and the fracture dip azimuth for each well. For

Figures 3.31 and 3.34, the wells with one fracture set, we see there is not much spread of

the points, i.e., most of the points are clustered. The next two figures are of more interest.

We can see a main clustering of points at around the same dip azimuth found in the last 2

wells. The second point of note is the larger spread of the data, especially to steeper dip

values near the edge of the graph. Figure 3.35 gives a summary of these last 4 graphs; it

compares the dip azimuth for all fractures, by well. All fractures are plotted here. This

figure is important because it shows that while there are 2 fracture sets, one is dominant

over the other.

3.3.3. Data Comparison

Fracture data was acquired from both core and borehole image logs. A comparison

of these two approaches was possible for one of the wells: PM 1-13E. While other wells

contained both borehole image logs and core, only in this one well were both types of data

available for a 40 ft interval. The results can be seen in Figure 3.36. The
83

PM 1-13E
Dip Azm uith
0 45 90 135 180 225 270 315 360
12800

12850

12900

12950

13000

13050
Depth

13100

13150

13200

13250

13300

13350

13400

Figure 3.31. Graph on the right shows the dip azimuth and the depth of all open fractures
in the Powder Mountain 1-13E well. On the left is a gamma ray log for lithology
comparison. The blue line is the top of the CEPO-Lewis pay sand.
84

PM 23-36
Dip Azm uith
0 45 90 135 180 225 270 315 360
12800

12900

13000

13100

13200
Depth

13300

13400

13500

13600

13700

13800

Figure 3.32. Graph on the right shows the dip azimuth and the depth of all open fractures
in the Powder Mountain 23-36 well. On the left is a gamma ray log for lithology
comparison. The blue line is the top of the CEPO-Lewis pay sand.
85

PM 32-26
Dip Azm uith
0 45 90 135 180 225 270 315 360
10000

10250

10500

10750

11000
Depth

11250

11500

11750

12000

12250

12500

Figure 3.33. Graph on the right gives the dip azimuth and the depth of all open fractures in
the Powder Mountain 32-26 well. On the left is a gamma ray log for lithology comparison.
86

PM 34-11
Dip Azim uth
0 45 90 135 180 225 270 315 360
11200

11400

11600

11800

12000
Depth

12200

12400

12600

12800

13000

Figure 3.34. Chart on the right shows the dip azimuth and the depth of all open fractures in
the Powder Mountain 34-11 well. On the left is a gamma ray log for lithology comparison.
The blue line is the top of the Lance pay sand.
87

Dip Azimuth
45

40

35

30
Frequency

25

20

15

10

0
20 40 60 80 100 120 140 160 180 200 220 240 260 280 300 320 340 360
Dip Azimuth
PM 34-11 PM 23-36 PM 32-26 PM 1-13E

Figure 3.35. Chart showing the dip azimuth obtained from borehole image analysis for all
open fractures in all wells.
88

Core- PM1-13E

FMI - PM1-13
13330
13355

13315
13340
13275

13235 13350

13195

Log Depth
13155 13360
Log Depth

13115
13370
13075

13035 13380
12995

12955 13390

12915
0 1 2
12875 vertical fractures (#per 10 feet)
12835

12795
0 1 2

Vertical fractures(# per 10 feet)

Figure 3.36. Comparison of data collected from cores and borehole image logs. In the PM

1-13E there was a 40 ft interval where both core and FMI image data was available.
89

core has a much higher resolution, like that of a hand sample, and showed us that 3

fractures were present. However, from the FMI log, only 1 fracture could be picked. This

is an important point to keep in mind in constructing the 3D model. Also, the probability

of intercepting vertical fractures with vertical wells is exceedingly small, according to

Lorenz and Hill (1992).

3.3.4. Vertical Fracture Spacing

Borehole image logs may have the advantage over cores. The subsurface is the best

possible place to collect data to understand the effect of processes that affect development

of natural fracture systems, because once brought to the surface, core and rock samples

have lower stress fields, mainly due to the removal of overburden (Narr, 1991; Wu and

Pollard, 1995). This is another reason for the development of unloading fractures.

Knowledge of the spacing, orientation and distribution of fractures in reservoir

rocks can lead to a better understanding of production characteristics in a fractured

reservoir. The high resolution of the FMI log allows it to serve this purpose. Narr (1996)

provided a mathematical solution to estimate average fracture spacing based on observed

borehole-fracture intersections and observed fracture porosity. The only data required are

the dimensions of the imaged borehole and the total height of all sampled fractures. This

method uses the average height of fractures plus their frequency of incidence, as
90

measured in an FMI log. The equation he arrived at is given in Figure 3.37. Sav is the

average fracture spacing, where Ai = aperture, Hi = height, Li = length of fractures

contained in the core, n is the total number of fractures in the core, Aav = average value of

apertures, Wc = width, Hc = height, and Lc = length of the core. Core width Wc is the

diameter perpendicular to the fracture planes. Height Hc is the long dimension, measured

parallel to the core axis, which is the dimension conventionally referred to as length. Lc is

the core diameter parallel to the fractures.

However, as pointed out by Wu and Pollard (2002), FMI logs get little use for

important information such as: (1) surrounding fracture networks, (2) fracture orientations

in 3-D, and (3) intersections of different fracture sets in 3-D. FMI logs ideally can be used

to provide the following data for each fracture: location, individual fractures within a given

section of the borehole (the window). By moving the window along the borehole axis, one

can generate continuous curves for these parameters (Figure 3.38) (Wu and Pollard, 2002).

Statistical methods for estimating subsurface vertical fracture density have been

proposed by both Narr and Lerche (1984) and Aguilera (1988). The first method is a

probabilistic one where the probability of core fracture intersections is determined for a

range of possible densities, after accounting for core diameter variations, bed thickness,

geometric corrections necessitated by bedding dip, and variations in fracture pitch. The

spacing, dip, and dip azimuth. From this information, mean spacing, spacing deviation,

spatial density, connectivity, and uncertainty are then calculated for all
91

Figure 3.37. Schematic diagram of core or borehole image log. The drawing explains the
variables used in Narr’s (1996) equation, to the right, where he estimated average fracture
spacing.
92

Figure 3.38. Moving a window of length, h, along the borehole and taking average values
of fracture parameters in the window provides a continuous estimation of how scattered a
fracture set is in orientation and spacing. dmean is the average fracture spacing. (a) Poorly
developed fracture set, dmean = 0.61 m.; (b) well-developed fracture set, dmean = 0.45 m.; (c)
data collection from a window includes spacing, S ‘; azimuth, θazi; and dip, θdip. (Wu and
Pollard, 2002).
93

true density is then found by determining the vertical fracture density value that yields a

calculated number of fractured beds equal to the observed number of fractured beds.

Aguilera (1988) got similar results by using the simpler method of a binomial theorem

equation.

Some reservoirs with low porosity are known to be productive largely because natural

fractures enhance hydrocarbon delivery to wellbores. (Nelson, 1985). However,

exploration and development decisions must often be made in the face of great uncertainty

about the contribution of fractures to production. This uncertainty stems, in part, from a

scarcity of data on fracture attributes. The role of fractures is commonly deciphered from

well tests and from differences between observed and expected production, instead of from

site-specific observations of fracture porosity, size, and connectivity. The repercussions of

severely limited fracture data include foregone exploration and development opportunities

and the risk of surprises in production responses.

3.3.5. Breakouts and In situ Stress

The borehole breakout analysis conducted for Powder Mountain 1-13E is shown in

Figures 3.39 and 3.40. In the first figure, the strike azimuth of the maximum horizontal

stress direction (Shmax) is shown according to depth. Thus, we are able to see distinctive

Shmax azimuth occurrences versus depth. This is plotted alongside a gamma-


94

Strike Azimuith of Shmax


0 45 90 135 180
12800

12900

13000
Depth

13100

13200

13300

13400

Figure 3.39. Borehole breakout analysis for the Powder Mountain 1-13E well. The graph
on the right shows the strike azimuth of Shmax with depth. On the left is a gamma ray log
for lithology comparison, in yellow are the sandstones and in gray are the shales.
95

25

20
Frequency

15

10

0
0 -15 15-30 30-45 45-60 60-75 75-90 90-105 105- 120- 135- 150- 165-
120 135 150 165 180

Strike Azimuuth, Shmax

Figure 3.40. Graph showing the frequency of the strike azimuth of Shmax. The highest

rate of frequency occurs for the set 150°-165°, whereas the average value for Shmax is

146°.
96

ray (GR) log, in order for us to distinguish in which lithology most of the breakouts occur.

The second figure is a graph giving the frequency of the Shmax values for this well.

3.4 Discussion

Of the types of fractures identified in the core, the quartz and calcite-lined fractures are

below the resolution of both the FMI and EMI tools, so they may not have been detected on

the images. These quartz and calcite-lined fracture types should have appeared as light

colored lines or intervals on FMI and EMI images, as they are partially healed and healed

fractures, respectively. This is because these two minerals are more highly resistive than

the surrounding sandstone rock. These types of features were found in a few cases, but not

at the high rate that would have been expected. From the cores we get information that

would be unattainable from the borehole image logs. Secondary mineral growths on the

walls of the fractures clearly indicate that the fractures were natural.

Looking at the borehole image logs virtually 100% of the fractures picked occurred

in sandstone lithology. Many of the fractures started off in sandstone and continued into

shalier area before dying out, these were categorized as occurring in sandstone. The

borehole image logs give us the best information about the orientation and density of

fractures. Two of the wells appear to have a conjugate fracture set, thus it is probably

natural. The fracture sets are steeply dipping, with both sets having similar strikes, but set
97

one (the predominant set) has steeper, southwest dips. Fracture Set 1 has a dip direction/

dip of 213°/65° and Fracture Set 2 has a dip direction/dip of 12°/54°.

There is evidence from the core studies that induced fractures are common in the

study area. Thus, to remove these erroneous data points, any fracture with a dip angle over

80° was removed. However, there is a possibility that steep natural fractures that were

removed from the data set. There is now a bias so that the dips are shallower than they

actually are. However, the strike would have been unaffected by this assumptions. On

average, almost 60% of the fracture data points were removed as possible induced

fractures.

An illustration of this fracture system can be found in Figure 3.41. However, it

seems that the maximum stress direction is vertical. This has some implications for strike-

slip vs. dip-slip offset. The vertical maximum stress directions, as well as the lack

of horizontal slickensides in the core, both combine to suggest that dip-slip may have been

more common in this area. Figure 3.42 shows what a map view and cross-sectional view

would look like if strike-slip offset were dominant. The dips would also be steep, but the 2

fracture sets would only have a 30° difference in strike azimuth. However, my findings

show a similar strike azimuth. This is only possible if the vertical stress was σ1, maximum,

at the time these fractures formed. In Figure 3.41 we are assuming that this is a conjugate

fracture set where both facture sets formed at the same time. We do know that the tectonic

system is more complex than is represented here, but due to the data set available for use,

this was the interpretation made.


Map view SW Cross-sectional View NE

σ1 is vertical
σH1 max σ1 σ1

Predominant
fracture set
N
100 m

Secondary
σ1= principal stress direction fracture
set

Figure 3.41. Diagram showing the principal stress direction for the study area, as obtained from the
borehole image logs. There is a conjugate fracture set and the maximum stress direction is near
vertical. Secondary fracture set terminates against the dominant set.
98
Mapview SW Cross-sectional View NE
σ
1

Î
σ σ
1 1

Î
Predominant
N fracture set

100 m
Secondary
σ1= principal stress direction fracture set

Figure 3.42. Diagram showing what the present-day principal stress direction would be for a
strike-slip regime.
99
100

Figure 3.43 shows the fracture orientation identified in this study of both fracture sets, in

relation to the maximum present-day horizontal stress direction. For this study, we used

the borehole breakout interpretation to get a σH1 max direction as shown in Figure 3.43.

This, together with the fracture geometry, is what leads us to the interpretation made where

we arrive at the vertical σ1 shown in Figure 3.41.

Christie-Blick and Biddle (1985) have identified a number of criteria for

recognizing a strike-slip fault. These include:

Basement involved faulting, resulting in contrasting types,

The principal displacement zone is sub-vertical at depth,

Upward diverging and rejoining splays, commonly known as flower

structures,

Abrupt variations in thickness and facies in a single stratigraphic unit, both

normal and reverse separation faults in the same profile,

In one profile there are variable magnitude and sense of separation for

different horizons offset by the same fault,

For successive profiles, there are inconsistent dip directions on a single

fault,

For successive profiles, there are variable magnitude and sense of separation

for a given horizon on a single fault,

For successive profiles, there are variable proportions of normal and reverse

separation faults,
101

Map view

N
σH1 max (N34W)

Fracture set 1: N57W

Fracture set 2: N78W

Figure 3.43. Position of fracture set identified in this study in relation to the maximum

present-day horizontal stress direction.


102

Figures 3.44 illustrations this, an idealized map view and cross section of a strike-slip fault

regime. While some of these criteria are being met for this area, most of them are not.

Thus, dip-slip appears to be dominant. Dip-slip is the component of

slip measured parallel to the dip of a fault, as opposed to strike-slip, where the slip is

measured parallel to the strike of a fault. Figure 3.45 illustrates the direction of slip on dip,

oblique and strike-slip faults.

Present-day stress appears to be consistent with those observed in this study. The

fractures observed in image logs are approximately parallel to the Shmax, but conjugate to

the vertical stress. The present-day stress direction obtained from borehole breakouts is

146° (N34W). Figure 3.46 is a present-day stress map of Wyoming, with results from

other nearby studies. Rahmat (2000) interpreted a maximum horizontal stress orientation

to the northwest (N40°W) in the Pilgrim Federal 1-14-28 (T10N-R93W-Sec. 28) and

Iverson 32-5 (T10N-R93W-Sec. 5) wells. Minton (2002) interpreted an orientation of

N26°W from his interpreted drilling induced fractures in the Powder Mountain 1-13E well.

Sunnetcioglu (2001) interpreted a similar result in the Barrel Springs field to the east of my

study area.
103

(A)

(B)

Time-stratigraphic unit
with variable
sedimentary facies.

Figure 3.44. (A) Map view of the spatial arrangement of structures associated with an
idealized right-slip fault. (B) The cross sectional view. From Christie-Blick and Biddle
(1985).
104

Figure 3.45. Generalized sketch of experimental modeling. Configuration of rigid blocks


and master faults before and after deformation for dip-slip, oblique-slip, and strike-slip
faulting. From Schlische et al. (2002).
105

110° 105°
11 10
0 N 5
MT SD

Zoback and Zoback (1989) WY


ID 44°
44°

Maraj study NEB

42°
42° Sunnetcioglu (2001)
UT

Figure 3.46. Present day stress orientation. Shown on the map are stress directions from
studies by Zoback and Zoback (1989) in red, to the west of my study area, and by two
Lewis Shale consortium members, Sunnetcioglu (2001) and Rahmat (2000).
106

CHAPTER 4

WELL COMPLETIONS, PRODUCTION DATA

4.1 Background

Differences in the natural gas production of all the study wells are significant. In

order to compare the wells, I had to derive a method to normalize the production. One

approach is to calculate the estimated ultimate recovery (EUR). For this project, this

method was not used because the projected recovery is propriety information. Thus, a

second method was chosen. Adding the first 12 months of production for each well made

relative comparisons. Most of the wells did not produce for the full 12 months, and this

caused some discrepancies in the calculations. Also, total cumulative production

histories after the first 12 months suggest that Powder Mountain 34-26x may be the best

well in the area, even though this well ranks second in Table 4.1. However, for a general

comparison, this method of calculating cumulative production is useful. Table 4.1 gives

cumulative production data for wells in the study area, using this method. The data are

also shown in Figure 4.1.


107

Table 4.1. Cumulative production for the first 12 months for wells in the study area.

Cum Prod for 1st year


Well Producing Interval
(mcf)

Polar Bar #1 54,484 Lewis

PM Fed 23-36 77,303 Lewis

Triton #10 161,198 Lewis

PM 1-13E 398,568 Lewis

PM 34-26x 1,077,791 Fox Hills

Cepo Lewis 21-18 1,805,427 Lewis


108

14N/96W 14N/95W

PM 1-13E (0.4)

Cepo 21-18 (1.8)

Polar Bar 1 (0.05)


PM Fed 32-26

PM Fed 34-26 (1.1)


PM Fed 23-36 (0.07)

Triton #10 (0.2)

PM Fed 34-11 Bogey Draw 1-14

Black Bar 1

13N/96W 13N/95W
6 mi

Figure 4.1. Map of the study area showing the location of wells with cumulative
production rates in red in bcf (billion cubic feet) for the first 12 months of production.
109

4.2 Methods

Typically, for a fractured reservoir there are many wells with relatively low initial

production and cumulative production, but a few wells with very large values. Thus,

production graphs and reports of the study wells can be used to show production trends.

Drilling and completion reports can be studied in order to see changes in mud weight,

which indicate a change in pressure. This can be the result of encountering a fracture.

Another indicator of fractures from the drilling/ completions data is the sudden high gas

increase. Things of note include trip gas volumes and volumes of mud lost to the

formation.

A method to allow pressure-dependent leakoff to be quantified and the leakoff

mechanism to be identified has been found. This method uses G-function analysis to

identify the leakoff mechanism. The G-function describes fracture pressure decline

behavior, this was analytically shown to be linear for constant leakoff with a wall-

building type of fluid. In 1990, Mukherjee at al. proposed a method of fracture pressure

decline analysis for cases of pressure dependant leakoff, assuming the G-function to be

piecewise linear during pressure decline in naturally fractured reservoirs. Methods were

presented to derive a simple exponential relationship between leakoff coefficient and the

rate of pressure decline (Mukherjee at al., 1990). Figure 4.2 shows results from a case

study by Barree and Mukherjee (1996). The pressure (P) and derivative behavior

(dP/dG) during closure are plotted against G-function (GdP/dG). In this case, a set of
110

Pressure
derivatives

G-function (α=1.0)

Figure 4.2. G-function analysis for pressure dependent leakoff behavior. Bottomhole
pressure (psi) is on the right axis. From Barree and Mukherjee (1996).
111

orthogonal (high angle) fractures is assumed to open when the fracture fluid pressure

exceeds the maximum horizontal stress.

4.3 Results

Production data and well completion practices for the study wells were found on

the Wyoming Oil and Gas Conservation Commission website. These results are provided

in Appendix C. Completion practices and drilling reports for the wells Cepo-Lewis 21-

18, Powder Mountain 1-13E and Bogey Draw Unit #1 can be found in Tables 4.3 to 4.5.

Thus, some comparison is possible. The basic status of the wells can be found in Table

4.2. Completion practices for the wells vary across the area. Figure 4.6 shows the mud

weight with depth, the mud weight increases when the fluid pressure changes in the

reservoir. This occurs at different depths for each of the 3 wells, this shows that pressure

changes across the study area, this could have an affect on gas production. Completion

practices and drilling data for all the wells in the study area was either downloaded from

the Wyoming Oil and Gas Commission website or donated by Ken Boedeker of EOG

Resources.

The two wells, the Cepo-Lewis 21-18 and the Powder Mountain 1-13E, both have

detailed fracture treatment data available. These data show that natural fractures exist in

the reservoir rocks. The Cepo-Lewis 21-18 had a large fracture treatment that placed

250,000 lbs of sintered bauxite (Carbo-HSP) using a 28#/Mgal Vistar fluid in the
Log Driller's Mud weight Shows Comments core 112
Depth Depth (ppg) fractures
(feet) (feet) shows
9.2 16.2
casing
6500
surface 2349
casing

7000

7500

8000

9146' lost 567 bbls mud

8500 9459' lost 200 bbls mud


9500
9600
9700
9800 9000

9900
10000
10100
10200 9500
10300 10432' - trip gas - 5460 units
10400
10500 10546' - trip gas - 5960 units
10600 10000
10760' lost 40 bbls mud
10700
10800 10860' lost 200 bbls mud
10900 10900' - trip gas - 5280 units
11000 11065' - trip gas - 4280 units
10500
11100
11200
11300 11340' - trip gas - 4160 units
11400 fair-moderate show in Lance
11000
11500 fair-moderate show in Lance
11600
11700
11800
11900 11500

12000
12100
12200
12300 12000

12400
12500 12580' lost 240 bbls mud
12600 12655' - trip gas - 640 units
12700 12500

12800 12634' inter Good show in Lewis. High gas increase --> fracturing. ss is tight. 6'-8' flare 12856' fractured ss 4980 units
12900 casing 12984' trip gas - 2000 units with 8-10' flare
13000 13026' trip gas - 2640 units with 20' flare
13250 13100 13000

13200 30' flare. Excellent show for the Lewis in this area. 13285' trip gas - 7200 units with 20-25' flare. Lost 300 bbls mud
13300 13397' - trip gas - 4800 units with 12-15' flare. Figure 4.3 Cepo-Lewis 21-18 drilling data
13283 13400
Log Driller's Mud weight Shows Comments FMI log 113
Depth Depth (ppg) open FMI fracture
(feet) (feet) core
8.9 15.2
5600
fractures
surface casing 2355 shows
casing
perfs: 6 holes/ft
6100

6600

7100

9000
9100 9112' lost 480 bbls mud
9200 7600

9300
9400
9500 8100
9600
9700
9800 9827' fair-mod show in Ft Union
9900 8600 9924' good show in Ft Union, appears tight
10000
10100 10175' fair-good show in Ft Union
10200 10231' fair-good show in Ft Union
9100
10300
10400
10500 10545' lost 300 bbls mud
10600 9600 10659 lost 750 bbls mud. Trip gas - 6320 units - 8-10' flare
10700
10800 10846' trip gas - 4320 units with 5-6' flare
10900
11000 10100 11071' trip gas - 5920 units with 3-4' flare
11100
11200
11300 10600
11400 11409' lost 450 bbls mud. Trip gas - 3280 units - 3-4' flare
11500
11600 11655' good show in Lance
11700 11100

11800
11900
12000 12056' trip gas - 3680 units with 4-6' flare
11600
12100
12200
12300
12400 12100
12500
12600
12710' 12700
12800 12600 12835'-12836' gas from a fracture 1-3' flare
12900 12704 inter 12982' trip gas 2160 units with 8-10' flare
13000 casing
13100 13100
13115' trip gas 5040 units with 6-8' flare
13200
13330-13370 13300 13330' trip gas 4400 units with 10-12' flare Figure 4.4 Powder Mountain 1-13E drilling data
13409' 13400 13484' trip gas 1440 units with 8-10' flare
13500
Log Driller's Mud weight Shows Comments FMI log 114
Depth Depth (ppg) open FMI fracture
(feet) (feet) core
9.1 14.5
9200
fractures
surface casing 2520 shows
casing

9000
9100 9700

9200
9300
9400
9500
9600
10200
9700
9800
9900
10000
10100
10200 10700
10300
10400
10500
10600
10700
10800 11200
11800 gas kick 2500 units 10-15' flare. 60 bbls mud lost
10900
11000
11100
11200 11204' poor-fair show for the area
11300 11700
11400 11409' poor-fair show for the area
11500 11446' poor-fair show for the area
11600
11700
11800 11800' fair-mod show for the area
11900 12200

12000
12100
12200
12300
12400 12478' poor show, low pressure, tight for the area 12477 1110 units with 15' flare
12700
12500
12600
12700 12791 trip gas 470 units with 25-30' flare
12800
12900
13000 13200 13211' inter 12981 lost 230 bbls mud
13100 casing 13142 lost 182 bbls mud. Trip gas 600 units with 25' flare
13200 13200 trip gas 3500 units with 4-5' flare
13300
13400
13500
13700
13600
13700
13873-13926 13800
13900
14000
14100 14200
Figure 4.5 Bogey Draw #1 drilling data
14200
115

Table 4.2. Completion practices and status of wells in the study area. All depths are in
ft. Spud dates influence completion practices and thus, production of the wells.

Powder Powder Powder


Cepo- Polar Bar
Lease Name Mountain Mountain Mountain
Lewis 21-18 #1
1-13E 32-26 34-26X
49-037- 49-037- 49-037- 49-037- 49-037-
API #
24237 24185 23037 24126 24178
Spud Date 08/21/99 04/15/99 09/30/92 02/27/99
Elevation
6,435 6,416 6,544 6,618
GR
Total Depth 13,503 13,500 15,525 C 13,183
Plug Back 13,459 MI 14,970 O 13,147
Top Perf 13,333 13,258 14,064 N 13,096
Bottom F
13,367 13,294 14,164 13,116
Perf I
Holes/ft 6 MI MI D 2
250,000 lbs E
sintered N
269 bbls T
bauxite
gelled water, I
using a
16440# 100 A
28#/Mgal
Treatment mesh sand MI L MI
Vistar fluid.
& 300400#
100-mesh
20/40
sand.
Carbo-Prop W
Proppant at
8 ppa. E
Producing L
Lewis Shale Lewis Shale Lance L Fox Hills
Interval
Status FL FL FL FL
Abandoned - - 04/19/02 -

PA = plugged and abandoned


FL = flowing MI = missing data
116

Table 4.2. Con’t.

Bogey
Lease Name PM 23-36 PM34-11 Triton #10 Black Bar 1
Draw 1-14
49-037- 49-037- 49-037- 49-037- 49-037-
API #
24354 24482 21922 22991 23169
Spud Date 08/18/00 03/17/01 02/08/82 09/29/93 07/03/93
Elevation
6,575 6,753 6,600 6,740 6,941
GR (ft)
Total Depth
14,300 15,105 14,975 14,204 14,373
(ft)
Plug Back
13,854 12,561 13,420 14,300
(ft)
Top Perf
13,788 12,499 13,259 - 14,056
(ft)
Bottom
13,837 12,660 13,279 - 14,150
Perf (ft)
Holes/ft 4 6 MI - -
Pumped 131
bbl pad. Pad
W/19488
gals &
8000# 100
mesh sand.
Treatment - MI - -
Treated
W/33558
gals &
87000#
20/40 Carbo
HSP.
Producing
Lewis Lance Lewis - Almond
Interval
Status FL PA FL PA PA
Abandoned - 07/04/02 - 06/08/93 11/12/96
117

Cepo Lewis 21-18 (1.8 bcf) PMU 1-13E (0.4 bcf) Bogey Draw 1-14 (dry hole)
9 11 13 15 17 9 11 13 15 17 9 11 13 15 17
10000 10000 10000

11000 11000 11000

12000 12000 12000

13000 13000 13000

Figure 4.6. Comparisons of the variations in fluid pressure as seen in the differences in
mud weight used in 3 wells. Overpressure typically occurs over 12.5 ppg (0.65 psi/ft),
where the red line is shown. The best producer has the highest mud weight. Production
for the first 12 months, in bcf, is shown next to the well names.
118

hydraulic fractures. It was felt that a high-strength ceramic material was needed to prop

the hydraulic fractures open due to the tightness of the formation. The Powder Mountain

1-13E job included a pre-fracture shut in and brief fallout, but similar material was used.

Some 100-mesh sand was used for fluid loss control in the pad stages for both wells. The

use of this material is an indirect indication that natural fractures exist in the reservoir

(Miskimins, pers. comm., 2003). This material is often used to stop leak-off during a

fracturing treatment, it is thought to plug the natural fractures. Results from the

stimulation treatment may have indicated that natural fractures were present, justifying

the use of this 100-mesh sand. Perhaps more importantly, for the Powder Mountain 1-

13E, the shut-in and falloff was analyzed using the G-function method of Barree and

Mukherjee (1996). The falloff shows indications of pressure dependent leakoff in the

curvature of the first derivation and the slight “hump” in the semi-derivative, similar to

Figure 4.2. This justifies the use of the 100-mesh sand in the pad for leakoff control.

The rest of the falloff shows a linear semi-log derivative that indicates normal leakoff and

no sign of fracture closure at the end of the falloff. The lowest pressure measured was

11,420 psi BHP (bottom hole pressure) at 13,350 ft, giving a gradient of 0.85 psi/ft. The

expected pore pressure in the zone is 10,690 psi or 0.80 psi/ft. Thus the range on possible

closure pressure is between these values.

Some of the older wells in the area had different completion practices, which

directly affects how good the production rates were. Drilling and operations experiences

in the area have improved the production in the area, by making better practices more
119

widespread. For instance, the Polar Bar Unit #1 had a spud date of September 1992, and

as Table 4.3 demonstrates, there were things learned from this well that were taken to

other, later wells in the area.

4.4. Discussion

Hydraulic fracturing is critical to connecting sufficient permeability and porosities

for economical production from these vertical wells. Predicting fluid-flow responses

from fractured reservoirs is very difficult because of the complex spatial and geometric

variability of the three-dimensional fracture network. Fractures enhance reservoir quality

and provide pathways for hydrocarbon migration from the underlying Lewis Shale to the

Dad Sandstone member.

The highest cumulative production rate for the first 12 months of production is from the

Cepo Lewis 21-18 well, followed closely by the Powder Mountain 34-26. There is a

large difference between these and the rest of the wells. The lowest cumulative

production comes from the Powder Mountain Fed 23-36 and Polar Bar #1 wells.

The Powder Mountain 1-13E and the Cepo-Lewis 21-18 both had spud dates in

1999. They had similar completion practices and both are quite good producers in the

area. The Cepo-Lewis well has produced 3.5 bcf since August 1999 to June 2003. The

Powder Mountain 1-13E has produced 0.8 bcf from November 1999 to May 2003. The

Polar Bar #1 well was spudded in 1992, it had different completion practices carried out

on it, and it is a poor producer in the area (0.1 bcf from November 1993 to 1999).
120

Table 4.3. Completion practices on the Polar Bar Unit #1. From Flack Petroleum
Consultants (1994).

Date Completions Conducted Possible Problems


Upper Almond
perforated (12,214’-
Acid and chemicals such as Freflo F is not
12,220’) and broken
a recommended breakdown fluid in tight
2/23 - 2/24/93 down with 150 bbls of
gas sands. Causes possible formation
Freflo F. Lower Almond
damage, preferentially use nitrogen.
perforated the next day
(12,324’-12,352’)
* Too much fluid and not enough proppant
used (1)
* 10% methanol used in frac tanks as gel
270 bbls 3% KCl water
3/6 - 3/16/93 stabilizer (2)
loaded in hole.
* 50 lb gel run (3)
* Forced closure used (4)

Packer removes the ability to unload the


Tried to set plug in
well, recommended to snub tubing (with a
4/16 - 4/17/93 packer tail pipe, not able
pump through plug installed in the end of
to.
the tubing)
Perforated tubing and
circulated with 320 bbls Mud caused damage to the Almond.
4/18 - 4/23/93 mud. Drilling mud lost Snubbing unit would be correct application
to formation. Packer and to remove tubing.
plug set above Almond.
*
(1) This yields low conductivity fractures; at these depths there is no conductivity once the
bottom hole pressure is reduced during the production test, this leads to excessive settling
of proppant. In the Almond 133,000 lbs of ceramic used and 79,800 gallons of water.
This is a 1.67 lbs/gal of total ceramics to water. Recommended in this area is foam
instead of water, at a 4.3 lbs/gal ratio.
(2) The Almond contains smectite, illite, kaolinite and chlorite, which, when in contact with
methanol reduces the relative permeability.
(3) Gel residue would be generated which causes a poor cleanup, resulting in a partially
plugged low conductivity proppant pack.
(4) The immediate flow carried back too much proppant to surface and potentially reduced
the concentrations across the perforated interval.
121

There is a general pattern from Tables 4.3 and 4.4 that the older wells in the field

are no longer producing hydrocarbons. However, the Powder Mountain 34-11 is an

anomaly as it was spudded in 2001, but was abandoned a year later. From the

information in Table 4.3, it can be suggested that the Polar Bar #1 could have been a

good producer, if the completion practices were better, as they are today. However, there

is the case of the Triton #10 well that was spudded in 1982 and is still a reasonably good

producer (1.2 bcf from August 1982 to June 2003). Thus, the differences in completion

practices can account for some variations in production, as in the case of the Polar Bar

#1. However, it does not seem to account for some of the other wells, such as the Powder

Mountain 34-11 and Triton #10.

There are a number of other causes of pressure dependent leakoff. Middlebrook

et al. (1995) described some of these:

Fracture extension. A small amount of crack growth takes place after shut-in due

to the equilibration of the fluid pressure within the fracture.

Single and multiple (possibly misaligned) fracture closure(s) in the near-wellbore

region. Misaligned fractures, possibly resulting from stress concentrations

around the perforations and/or wellbore, may have a tendency to close

preferentially in the near-wellbore region. The resulting pressure transient

response may be interpreted as far-field fracture closure if an ideal response is

assumed. Similarly, a pressure transient response resulting from multiple


122

fracture closures may prematurely be interpreted as the minimum in situ stress

when assuming an ideal response.

Closure in an adjacent interval. Due to the highly varying lithology in some

reservoirs, stress tests are often performed within short distances of lithology

interfaces. Even with small injected volumes (2-4 bbls), adjacent layers may be

penetrated resulting in the observance of an average closure and/or multiple

closure pressures.

Tortuosity effects. Large pressure drops (1,000-3,000 psi) are often observed at

shut-in, which indicatives a good cement job and a poor connection between the

created fracture and the wellbore. A continued pressure drop through the tortuous

pathway may inadvertently be interpreted as a fracture flow response (i.e.

leakoff).

The mud weight varies with depth for all 3 wells, the mud weight increases when

the fluid pressure is high. This occurs at different depths for each of the 3 wells, this

shows that pressure changes across the study area, this could have an affect on gas

production. The best producing well, the Cepo Lewis 21-18, had the highest mud weight

used, 16.2 ppg. Whereas the dry hole, had the lowest mud weight of 14.5 ppg.

Evidence for natural fracturing in the Lewis Shale reservoir from production and

drilling data include the high production rates in certain wells and low rates in other

wells, the high gas increases encountered when drilling, the pressure-dependent leakoff
123

from the G-function, as well as the need for some 100-mesh sand to prevent leakoff in the

hydraulic fracturing of the wells.


124

CHAPTER 5

SUPPORTING SEISMIC DATA

P-wave and s-wave seismic data are available in the study area, this chapter

attempts to determine whether they can be used for fracture detection.

5.1 Background

Many petroleum reservoirs are naturally fractured. Furthermore, open fractures

often dominate fluid flow (De Vault, 1997). The volume of oil or gas in place and the

reservoir’s ability to produce are dependent on the fractured state of the reservoir

(Stewart et al., 2003). It has been known for many years that a fractured rock will have a

higher permeability and thus be able to flow hydrocarbons more easily than the same

unfractured rock. It is especially common in older fields to hydraulically fracture the

rock surrounding the well bore to improve production (Swift and Mladenka, 1997).

However, the areal extent of these induced fractures is often limited (Zhu et al., 1996).
125

Understanding the nature of fractures in a reservoir is important to help target

drilling locations, build reservoir models to forecast production history, estimate drainage

areas, and construct secondary and tertiary recovery plans.

Stresses within the earth often cause fractures to open predominantly in one

direction (parallel to Sh max) (Crampin, 2000). Where stresses change, for instance near

a fault, fracture intensity and orientations in which fractures are open can also change.

Thus, analyses of subsurface fracture density and orientation have become particularly

important as petroleum reservoir characterization tools (Jenner, 2001).

The determination of fracture density and orientation from seismic data has been a

subject of considerable research. In surface seismic exploration, two basic body waves

are generated, compressional (P) waves, and shear (S) waves. P-waves create particle

motions in the direction that they travel, and their velocity (Vp) is governed by the bulk

modulus, shear modulus, and density. However, S-waves in an isotropic medium induce

particle motion perpendicular to the propagation direction and only the shear modulus

and density govern their velocity (Vs) (Reasnor, 2001). The shear wave propagating in

fracture planes splits into a mode that is polarized parallel to the fractures (called S1, or

the fast shear wave), and a mode that is polarized perpendicular to the fractures (called

S2, or the slow shear wave). Because the different modes travel with different vertical

velocities, time differences between the S1 and S2 modes can be used to estimate the

shear wave splitting parameter, γ (Thomsen, 1986), which can be related to fracture

density.
126

When shear waves are propagated through an azimuthally anisotropic media, such

as a reservoir with vertical, open fractures aligned in one direction, their particle

displacement will polarize into 2 orthogonal directions (Figure 5.1). One direction is

parallel to the open fracture orientation (S1–wave) and the other is perpendicular (S2–

wave). This assumes that ray incidence is near vertical or perpendicular to the horizontal

layer or reservoir. Waves polarized perpendicular to the cracks can deform the rock

easily because of their orientation with respect to the zones of weakness or fractures.

Therefore, this wave experiences low effective rigidity and travels at a slow velocity.

The wave polarized parallel to the fractures will experience higher rigidity and travel at a

higher velocity. These waves travel independently at different velocities and will,

therefore, result in 2 reflections for any impedance contrast. Recording of these 2

different waves with geophones will result in 2 separate time series. Changes in the time

difference between the 2 series results where azimuthal anisotropy occurs or changes

from one layer to the next (Thomsen, 1988).

Azimuthal anisotropy is most likely due to aligned fractures, which are the result

of preferentially aligned regional tectonic stress. This is not expected to vary greatly with

depth (Crampin and Atkinson, 1985). Time differences between the 2 time series for

corresponding events provide a robust measurement, poorly resolved in time, of average

anisotropy over extended depth intervals. Comparison of reflection amplitudes of

corresponding events on the 2 time series (S1 and S2) provides a measure, highly resolved

in time, of local anisotropy differences (Thomsen, 1988). This essentially means that
127

Figure 5.1. Shear waves can be used to indicate dominant fracture directions by
identifying areas high in porosity and permeability. The shear wave splits into a fast and
slow component aligned parallel or perpendicular to the fracture direction, respectively.
After Benson and Davis (2001).
128

differences in reflection amplitudes in the split S-wave time series allow identification

and characterization of fractures of the corresponding unit or formation. Figure 5.2

shows the converted wave survey geometry used in this survey, the 2 orthogonally

oriented upcoming mode-converted S-waves (in yellow) are recorded primarily on the

horizontal geophones (X & Y) and are of primary importance when investigating S-wave

birefringence for fracture detection (Van Dok et al., 1997).

Winterstein (1992) pointed out that anisotropy means a variation in physical

properties with direction. Therefore, when considering velocities, ‘azimuthal anisotropy’

means the variation in velocity with direction (azimuth). Although often described

as‘azimuthal anisotropy,’ the phenomenon is not a variation of amplitude or velocity with

direction of propagation, but with direction of polarization. Although there are other

factors that can also cause azimuthal anisotropy, such as regional stress or depositional

fabric and diagenesis, there has been a strong correlation observed between shear wave

birefringence (shear wave splitting) and fractures (Crampin, 1994). The technique used

in this study is an anisotropy approach where overburden effects can be removed by

layer-stripping analyses. A technique known as a 2Cx2C Alford rotation (Alford, 1986)

is carried out where both sources and receivers are rotated to a new orientation. The

resulting traces show the characteristic minimization of energy in the off-diagonal

components. If we apply this rotation to all orthogonal pairs, they can be summed to
129

Figure 5.2. Diagram illustrating the various ray paths and wave fields associated with
3D/3C surface seismic using a conventional P-wave source. From Van Dok et al. (1997).
130

create a single set of four components oriented in the principal directions (Van Dok et al.,

1997).

An example in the Barinas Basin, Venezuela, by Perez et al. (1999) shows the

high potential of P-waves to detect fracture effects on seismic wave propagation. The

distribution of fractures in this reservoir was obtained previously using measurements of

shear-wave splitting from P-S converted waves from the same data set. The P-wave data

were then used to see if the data could yield the same information using azimuthal

variation of P-wave AVO responses. The final results obtained from the

azimuthal P-wave AVO analysis corroborate the fracture orientation obtained previously

using P-S converted waves (Perez et al., 1999).

Shear-wave seismic surveys are more expensive to acquire and process than P-

wave surveys. Thus, although shear wave splitting is a well-known technique for fracture

detection, it is not widely used over large areas. Studies have shown that PS (converted

or multicomponent) waves are cheaper and can also be used to detect fractures. Workers

such as Gaiser (2000) and Van Dok et al. (1997) have shown that converted-wave data

can be used for anisotropy analysis. Converted waves are generated with a P-wave

source and converted into shear waves at a reflection boundary. Converted-wave

seismology requires recording of seismic data with three-component (multicomponent)

receivers. The three components measure displacement of the ground, usually in two

horizontal and one vertical direction (Figure 5.1). Measuring three components of
131

ground displacement enables the recording of compressional (P) and shear (S) waves,

which represent the full complement of “body” waves in seismology.

Although converted wave data are more complex to process (Thomsen, 1999),

shear wave splitting can be observed on the upward traveling shear wave portion of the

wave propagation (Bale et al., 2000). This can be used as a fracture detection tool.

Multicomponent recording provides significantly more information about rock and fluid

properties of reservoirs and their changes than can be achieved from conventional P-wave

seismic surveys alone. Multicomponent seismology enables us to define weak or

mechanically less-competent zones in the subsurface that are often associated with

fractured rock (Davis, 2001). The main use of multicomponent seismic methods has

been in the characterization of reservoir rock properties, e.g. lithology, porosity and

fractures. Sometimes pore structure changes alone can be distinguished, which is akin to

distinguishing fractured from nonfractured rock.

5.2 Results

For this study, Stone Energy Corporation in Denver provided a number of

products. Chris Besler of Stone Energy Corporation interpreted nine horizons and four

faults in a 105 mi2 (272 km2) area from high-quality conventional p-wave 3D seismic

data. Both Western-Geco and Axis processed the full volume and migrated the post-

stack time volume. Figure 5.3 shows the different processing products from each group.
132

Western-Geco then took a smaller area of 15 mi2 (39 km2) to calculate the pre-stack time

migration. The azimuthal AVO (amplitude versus offset) and velocity were

calculated over this smaller area of the CEPO field. The flow chart in Figure 5.4 shows

this available geophysical data.

The main analysis conducted of interest to this study is the layer-stripping

anisotropy analysis performed by Western-Geco over the 15 mi2 (39 km2) 3D converted-

wave area. Figure 5.5 shows line 5980 of the seismic survey with the horizons that have

been layer-stripped using Western-Geco’s anisotropy techniques. Next is a structure map

(Figure 5.6) on the Lewis Shale. The horizon dips about 3° to the west, otherwise there is

not much structure on this surface. The structure maps in the area do not change

significantly with depth.

The interpretation of the converted wave volumes produced a series of maps.

Isochron maps are first shown in the Fox Hills to Almond intervals for the fast and slow

shear-wave data from the converted wave data set (Figure 5.7). The isochron maps

illustrate possible fault and fracture patterns. Figure 5.7 shows a possible interpretation

of faults, in which there are at least 2 major fault geometries interpreted. Next, is a

percent anisotropy map generated using the surface to Ft. Union interval (Figure 5.8).

Anisotropy maps show different velocities that occur in different directions, which allow

us to infer the orientation of fractures in a field. However, anisotropy maps can also

illustrate a number of things including lithology/facies changes and stratigraphic trends.


133

Figure 5.3. Chart showing the processing products for 3D conventional p-wave seismic
data provided by Stone Energy Company of Denver. The survey area was 105 mi2 (272
km2).
134

Figure 5.4. Table showing processing products for 3D converted-wave seismic data, by
Western-Geco and Explortech. The coverage of the 3C survey was 15 mi2 (39 km2).
135

Layer-Stripping Horizons
All-azimuth fast s-wave

Ft Union

Lance

Lewis

Almond

Rock Sp

Figure 5.5. Line 5980. Layer stripping anisotropy has been conducted by Western-Geco
on 5 horizons shown. Time (msec) is on the vertical axis.
136

6 mi

Figure 5.6. Colour structure map with contours on the Lewis Shale. The survey is the
entire 105 mi2. The wells are labeled and faults are shown.
137

(PS1) Fast

(PS2) Slow

6 mi

Figure 5.7. Fox Hills to Almond isochrons in fast and slow wave directions. Wells are
labeled. The dotted lines are interpretations of possible fault geometries based on PS
differences. From Stone Energy Corporation.
138

percent anisotropy map generated using the surface to Ft. Union interval (Figure 5.8).

Anisotropy maps show different velocities that occur in different directions, which allow us

to infer the orientation of fractures in a field. However, anisotropy maps can also illustrate

a number of things including lithology/facies changes and stratigraphic trends. It is

important to note that the fault mapped to the south is not parallel to the trends seen in the

black dotted lines. Instead, it is at an oblique angle. Figure 5.8 has been interpreted with

black dotted lines to show trends seen in the seismic. There appears a strong ‘Z” pattern,

typical of strike-slip regions. Lastly, the fast and slow Vp/Vs ratios are shown in the Fox

Hills to Almond interval (Figure 5.9). Vp/Vs1 ratios typically show stratigraphic trends in

the data whereas Vp/Vs2 data show structural discontinuities, such as faults and fractures.

Figures 5.10 and 5.11 show the dominant fracture direction for the intervals: (1) surface to

the Ft. Union, and (2) Lance to Lewis Shale. Using an algorithm to try to identify patterns

from anisotropy data creates the dominant fracture direction maps.

5.3 Discussion

Even though a number of products were generated from the converted wave data,

it is difficult to interpret many of these maps. Velocity can be related to effective stress,

a higher velocity has a higher effective stress as it has a higher fluid pressure. Thus

pressure can be seen on the Vp/Vs maps. As lithology is highlighted by Vp/Vs1 maps,

then net-to-gross can be mapped. From the previous chapters we saw that the best

producing well had a low net-to-gross ratio and a high fluid pressure. This would show
139

FAULT

6 mi

Figure 5.8. Percent anisotropy map of the surface to Ft Union interval, from the
converted wave dataset. The lineament interpreted is a possible trend that will also be
seen on some of the following maps. Wells are labeled. From Stone Energy
Corporation.
140

Vp/Vs1
Fox Hills - Almond

6 mi

Vp/Vs2
Fox Hills - Almond

Figure 5.9. Fast and slow Vp/Vs ratios on the Fox Hills to the Almond. The lines are
interpretations of possible fault trends, the pink lineament was also seen on a previous
map. A strong orientation of 135° can be interpreted. Wells are labeled. From Stone
Energy Corporation.
141

6 mi

Figure 5.10. Surface to Ft. Union dominant fracture direction in strike azimuth. One
fault has been mapped in the southwest (arrow). North = 0 degrees, south = 180 degrees.
Fractures are oriented at an average of 135 deg angle. From ongoing work from Stone
Energy Corporation.
142

6 mi

Figure 5.11. Lance to Lewis Shale dominant fracture direction in strike azimuth. One
fault has been mapped in the southwest (arrow). North = 0°, south = 180°. Fracture
direction has changed from the surface to be oriented at an average of 225° angle. From
ongoing work from Stone Energy Corporation.
143

up on the Vp/Vs maps as a high value. Figure 5.12 shows these areas. Faults control

deposition by providing an area to be filled. Compartments can be interpreted in about a

165° orientation. However, looking at the dominant fracture direction map of the surface

to Fort Union there is a fracture set, at a 135º angle, similar to that seen in the cores and

borehole image logs. This orientation changes again at the depth of the reservoir

horizons, the Lance to the Lewis Shale, to an angle of 255º. These orientations were not

seen in the core and borehole image data analyzed. The percent anisotropy map shows a

‘Z’-shaped pattern, however, it is quite possible that we are seeing a fluvial system

overprinting a northwest fracture set, giving us a false impression of a ‘Z’-shaped pattern.

A point of interest is the structure contour map on the Lewis Shale (Figure 5.13).

Unevenly spaced contour intervals can be identified, especially around the interpreted (in

pink) N15°W fault trend from Figures 5.8 and 5.9. The contour interval between 6,600 ft

and 6,800 ft is widely spaced, lending credibility to the fact that some sort of structure

exists in that area.


144

Vp/Vs1
Fox Hills - Almond

6 mi

Vp/Vs2
Fox Hills - Almond

Figure 5.12. Fast and slow Vp/Vs ratios on the Fox Hills to the Almond. The circles are
high Vp/Vs values, where low net-to-gross values and high pressure areas may be found.
Wells are labeled. From Stone Energy Corporation.
145

6 mi

Figure 5.13. Colour structure map with contours on the Lewis Shale, from Figure 5.9.
The pink lineament shows the interpreted N15°W fault trend that has been interpreted on
previous maps.
146

CHAPTER 6

3D FRACTURE MODELING

6.1 Background

The 3D model had to be constrained by data obtained through the use of cores and

well logs. Once these were studied and data were obtained, the model could be built

using these parameters. The software used to build the 3D geological model was 3D

Move supplied by Midland Valley. Because some important constraints needed to build

the model could not be found in the core and logs, some assumptions had to be made.

Thus, more than one answer is possible. This exercise was intended to visualize the

spatial and orientation data, in the hope that this will lead to insights in the study area.

6.2 Methods

The first process for the model construction was to check the 3D seismic

interpretation for a good correlation. One must ensure that the interpretation does not

jump horizons, and that the interpretation was also something that was geologically

possible for the area. This did not take much time as there is little complex structural
147

change across the area, the horizons are layered, and the interpreted faults in the area

have little or no offset across them.

Due to the limitations on the computing power of the computers available, the

size of the model had to be decreased to a representative area. The full study area could

not be modeled. Models were constructed in 3D Move using 2 methods.

a. Discrete fractures using the grid-based fracture module.

b. Attribute-based fractures.

The first models were generated using the grid-based fracture module. Three

hypothetical models for excellent, good and poor conditions were modeled. A 0.25 mi2

(0.4 km2) area was chosen, as illustrated in Figure 6.1, in the center of the study area, to

account for edge effects. The vertical depth of the model was 100 ft. This first method

incorporates statistical data from the borehole image logs and cores studied.

6.2.1 Discrete Fractures:

For the calculations to be run, a grid had to be picked. This grid must be small

enough to highlight values of interest in the data, but large enough so that the calculation

could be run quickly. The size of the grid used for fracture generation was 1,600x1,600

m, the cell count was 10x10 m and the cell size was 160x160 m. The 3 cases were

chosen to be modeled using parameters obtained from the Cepo-Lewis 21-18, Powder

Mountain 1-13E, and the Bogey Draw 1-14 wells. The toolbox is shown in Figure 6.2.
148

200000

190000

180000

170000

0.25 mi2

160000

150000

640000 650000 660000 670000 680000 690000 700000

Figure 6.1. Structure contour map on the top of the Lewis Shale, outlined in the box is
the 0.25 mi2 area used for detailed fracture modeling.
149

Figure 6.2. Fracture parameters used for the grid based fracture model. The ‘length’
button is a pull down menu that includes orientation, aspect ratio and intensity.
150

In the toolbox, there is a pull down menu that includes length (which is selected in the

figure), orientation, aspect ratio and intensity. The parameters used for each of these

were:

Fracture length of 984 ft (300 m) was used for Fracture Set 1 and 426 ft (130 m)

was used for Fracture Set 2. Cores and borehole image logs do not provide

information on the spatial distribution and interconnectivity of fractures away

from the wellbore. Outcrop data was needed here. A study by Harstad et al.

(1995) in the Frontier Formation provided this data. Figure 6.3 shows the

location of Muddy Gap and illustrates the spatial distribution of fractures in that

outcrop. We had some reservations about using this outcrop study to constrain

our horizontal fracture length. First we need to keep in mind that uplift creates

stress that increases the number of fractures seen in outcrop, second, Harstad et al.

(1995) carried out this study on the Frontier Fm., a different facies from the one

studied here. Harstad et al. (1995) concluded that the spatial distribution of

regional fractures is controlled by bed thickness, with fewer and longer fractures

per unit area as bed thickness increases. Bed thickness, and therefore mechanical

stratigraphy, is similar in the two studies.

From the borehole image logs, a dip direction of 213º and a dip of 65º was

obtained for Fracture Set 1 and a dip direction of 12º and a dip of 54º was

obtained for Fracture Set 2.


151

N
Idaho
Wyoming
Muddy Gap

Washakie basin

Utah Colorado

100 mi

100 ft
Bed Thickness = 20 ft

Figure 6.3. Diagram showing fracture length and spatial distribution at an outcrop of the
Frontier Formation at Muddy Gap, Wyoming. The top map shows the location of Muddy
Gap with respect to the Washakie Basin. From Harstad et al. (1995).
152

The aspect ratio is the ratio of the longest axis (h) verses the shortest axis (l) for a

modeled fracture. In order to calculate the bed thickness, the average sand

thickness was needed for both the entire modeled interval and for the entire Lewis

Shale interval. The shale cutoff used was arbitrarily 105 API units. It was

assumed that any shale less than 2 ft (0.6 m) thick would not stop a fracture. This

we gathered from the core studies. This led to the bed thickness calculation, as

illustrated in Figure 6.4. Using the calculation shown, the average sand thickness

was found. The average bed thickness was 20 ft (6 m). From Harstad et al.

(1995), the average fracture length was 426 ft (130 m). Thus, the aspect ratio is

h/l = 0.04.

Intensity (N) values varied for the 3 cases. The N value is the number of open

fractures generated per grid cell. The procedure below outlines the steps used to

calculate the open fracture intensity that was needed for modeling.

1. Determine the number of open fractures in a given cored interval for 3

wells. Table 6.1 shows these results. The results were normalized to

number of open fractures per 100 ft of core.


Gamma Ray 150 h1 = 28 ft
0 Caliper 15
h2 = 4 ft
Top of Gross
interval h3 = 16 ft

7200 (hi / i) = 16 ft (4.8 m)

h1
Average sand thickness for
all wells = Σ (hi / i) = 20 ft
(6 m)
7220

h2 Figure 6.4. Powder Mountain


23-36. Method used for
perfs averaging sand thickness in
7240 Lewis Shale. The EMI log
h3 covered this entire interval.
Base of Gross
Interval
153
154

Table 6.1. Number of fractures in 100 ft, from core data.

Thickness of # of open
Well #/100 ft
Core (ft) fractures
Cepo-Lewis 21-
33 2 6.1
18
Powder Mountain
40 3 7.5
Unit 1-13E
Bogey Draw 1-14 56 4 7.1

2. Determine the net-to-gross ratio within the cored intervals using the

gamma ray logs (Table 6.2). The key premise is that the fractures are

present in the sands, but not the shales. This assumption was supported by

the core studies and image data.

Table 6.2. Net-to-gross ratio for cored intervals.


Wells Cored interval (ft) Net-to-Gross Ratio
Cepo-Lewis 21-18 13,250-13,283 0.9
Powder Mountain Unit 1-
13,330-13,370 0.87
13E
Bogey Draw 1-14 13,873-13,929 0.73

3. Plot the net-to-gross ratios for the cored intervals against the number of

fractures seen in the cored intervals, normalized to the number of fractures

per 100 ft of core (Figure 6.5). A line was fit, which allowed us to use the
155

graph to calculate the number of fractures in a 100 ft interval for any net-

to-gross value.

4. Determine the net-to-gross ratio for the entire Lewis Shale interval (Table

6.3). Depth is from log tops. Table 6.3 also shows the predicted fracture

intensity for each well from Figure 6.5.

Table 6.3. Net-to-gross ratio for the entire Lewis Shale interval.
Predicted
Net-to- Fracture
Wells Top Lewis Base Lewis
Gross Intensity of
borehole #/100 ft
Cepo-Lewis 21-
12,333 13,397 0.3 2
18
Powder
Mountain Unit 1- 12,476 13,503 0.4 3.5
13E
Bogey Draw 1-14 12,078 13,868 0.8 6.3
156

PM 1-13E
8
Bogey Draw 1-14
7

6
Cepo-Lewis 21-18
5
# frac/100 ft

0
0 0.2 0.4 0.6 0.8 1
N/G

Figure 6.5. Plot to calibrate fracture intensity as a function of net-to-gross ratio (N/G).

See text for discussion.


157

5. From the PM 32-26 and PM 34-11 wells (Figures 3.15 and 3.19), there

was approximately a 5:1 ratio of the abundance of Fracture Set 1 to

Fracture Set 2. This number was used to assign relative importance to

each fracture set.

6. In 3D Move, fractures were then modeled using arbitrary intensity values

(Table 6.4). Fractures of Set 1 and Set 2 crossed, this was the only option

allowed by the software.

7. In order to determine the number of fractures that intersected boreholes,

the area had to be sampled to ensure that the average fracture intensity

matched the expected value for each of the 3 wells. A grid of 25 evenly

spaced pseudo-wells was used to test the intensity values expected (Figure

6.6). These fracture intensity values are shown in Table 6.3.

Table 6.4. Fracture intensity values used for each fracture set.

1st year of (N) Fracture Set 1 (N) Fracture Set 2


Wells
production (bcf) (# seeds/unit cell) (# seeds/unit cell)
Cepo-Lewis 21-18 1.8 7 2
Powder Mountain 1-
0.4 10 3
13E
Bogey Draw 1-14 Dry Hole 50 10
158

100 m

Figure 6.6. Map view of the model area showing the location of the grid of wells used to
sample fracture intersections with the borehole.
159

For the grid-based fracture set, once the models were populated with fractures

using these parameters, the validation of a fracture network could be achieved by

undertaking an analysis of the fracture connectivity (3D Move online help, 2003). By

decoupling topological characteristics from the fracture geometry, we can depict 3D

fractures and their intersections as vertices and edges in a graphical representation.

(Figure 6.7). The objects themselves are placed as nodes within the graphical

representation and intersections are edges. This allows the relative connectivity tool to

describe relationships between discrete objects. Relative connectivity gives an indication

of the overall connectivity of a fracture with respect to the rest of the fracture network. It

incorporates both a measure of the connectivity to all other fractures in its own connected

component, and the size of the component. This allows identification of the part of the

fracture network that will most likely be productive. The formula below describes this,

where C is the connected component size and B is the total topological distance to other

fractures in the same connected component:

relative connectivity = (C –1) ^3 / B

6.2.2. Attribute-Based Fractures:

The second model was built using the attribute-based fracture generator, where

bed curvature was used to constrain fracture intensity. The toolbox for attribute-based

fractures is shown in Figure 6.8. A seed is a point of stress concentration around a


160

Figure 6.7. Relative connectivity can be found by visualizing objects such as fractures
and their intersections as geometries and as a graph.
161

Figure 6.8. Fracture parameters used in the attribute-based fracture model. The upper
bounding bed is the top of the Lewis Shale. The lower bounding bed is the top of the
Almond Sandstone.
162

structural defect that would trigger failure and initiate a fracture. Thus, each seed is a

potential fracture. Again, due to limitations in the computing power, this exercise was

solely intended to investigate whether or not the curvature present on the top of the Lewis

Shale to the top of the Almond Sandstone was sufficient to delineate areas of higher

fracture intensity. Therefore, a small seed value was used to decrease the computing

time. The seeds per step generated were 656 ft (200 m) for 3 runs. The seed probability

is the probability that a given location would grow a fracture (3D Move online help,

2003). This controls the fracture distribution, and the Lewis Shale curvature map was

used as a probability surface. The growth orientation of the fracture tips was kept

constant at the average strike azimuth of 320º. This is the strike value obtained from the

borehole image logs (140°) added to 180º. The impedance is whether or not a fracture tip

will continue to propagate and grow (3D Move online help, 2003). The spatial

impedance was left as a constant across the bounding beds. A 0 value means that it will

not be impeded. The fracture impedance value was left as the default 1. This implies

that the approaching fracture will meet an open fracture surface and be impeded. The

Forbidden Zone is the volume around each fracture in which other fractures cannot

propagate. This is a zone of stress reduction around the side of a tension fracture that

inhibits the growth and propagation of other fractures (Figure 6.9). However, this is only

active for fractures of the same set, that is, a fracture set is not influenced by another

fracture set.
163

fracture
forbidden zone

projection

width

tip angle

Figure 6.9. Diagram showing the forbidden zone around fractures, where other fracture
seeds are not allowed to be generated.
164

The curvature tool gives a measure of the rate of change of dip across a surface.

This is calculated for each vertex by taking the normals to each of the triangles (the side

opposite the bisector) that surround the vertex, and calculating an average normal for the

vertex. A vertex is the point of intersection of triangles as shown in Figure 6.10. This

calculation is weighted to account for the area of the triangles, that is, larger triangles

have greater weight. A value for curvature is then produced from the divergence of

normals of the triangles surrounding a vertex to the average normal.

The average cylindricality was measured for the Lewis Shale surface, in order to

help identify fold trends and surface deviations. The average cylindrical vector is derived

from the pole to the best-fit plane of the surface triangle normals. This is the pole to the

profile plane of any folding. This analysis compares the orientation of the normal of the

surface triangle with the orientation of the average cylinder vector. A normal to the

surface triangle at 90º to the average cylinder vector has a deviation attribute of 0 (Figure

6.11). An irregular, non-cylindrical folded surface would contain triangle normals that

do not lie on the best-fit plane. Deviations from this best-fit plane will have a deviated

attribute value, and this is viewed on a colour map as values other than 0. Figure 6.11

also shows the curvature analysis toolbox, on which the plunge and azimuth of the

surface analyzed is also found.


165

vertex

Figure 6.10. Triangulated structure of a surface used to calculate curvature. The red
triangles intersect at a point (vertex). The blue arcs are the angles adjacent to the vertex
that is used, together with the area of the red triangles, to calculate curvature.
166

Figure 6.11. Diagram showing the theory behind cylindrical analysis. The surface
triangle normals (green lines) lie in a single plane, the average cylindrical vector (red
arrow) is oriented perpendicular to this plane. The curvature toolbox is shown on the
side.
167

6.3 Results

Three cases were generated. Figure 6.12 shows an example of the fracture

density around the Cepo Lewis 21-18 well. Figures 6.13 shows the fracture density

around the Powder Mountain 1-13E and the Bogey Draw 1-14 wells. Figures 6.14 to

6.16 show the relative connectivity of the fractures. Highly connected areas are shown in

red and less connected areas are in blue. Figures 6.17 to 6.19 show the fracture

intersections in each well and the graph shows the frequency distribution of fracture-

borehole intersections for each model. The yellow discs are intersections between the

boreholes and Fracture Set 1 and the green discs are the intersections between the

boreholes and Fracture Set 2. Models were run and sampled in an attempt to match

fracture intensity values shown in Table 6.3. These results are shown in Table 6.5.

Table 6.5. Results from models of fracture-borehole intersections, with expected results.
Modeled Results of
Predicted Fracture
Wells Fracture-Borehole
Intensity of borehole
intersection
Cepo-Lewis 21-18 2 1.75

Powder Mountain
3.5 2.85
Unit 1-13E

Bogey Draw 1-14 6.3 6.5

Figure 6.20 shows the orientation of a cross section across the model, from southwest to

northeast. Figures 6.21 to 6.23 show these cross sections.


168

100 m

Figure 6.12. Map view of the Cepo-Lewis 21-18 discrete fracture model. Fracture Set 1
is in yellow/ green, and Fracture Set 2 is in blue. Hypothetical wells are in red, laid out in
a grid fashion in the center of the model. The vertical height of the model is 100 ft (30
m). The bottom figure shows the cross-cutting relationship of the fractures, due to the
difference in dip.
169

A
N

100 m

B)

Figure 6.13. Map view of the a) Powder Mountain 1-13E and b) Bogey Draw 1-14
discrete fracture models. Fracture Set 1 is in yellow/green, and Fracture Set 2 is in blue.
Hypothetical wells are in red, laid out in a grid fashion in the center of the model. The
vertical height of the model is 100 ft (30 m). The fracture density increases drastically
for (b).
170

100 m

Color

mi
mi
ma

Figure 6.14. Map view of the Cepo-Lewis 21-18 showing 3D fracture relative
connectivity. Low areas are in blue and red is high connectivity. The red lines show
hypothetical well locations. Most of the area remains blue, unconnected to other
fractures in 3D. The figure on the bottom shows some unconnected fractures, especially
at the top of the figure, in blue and the intersecting fractures (red circles) that appear in
green.
171

Color scale: 100 m

min
mid
max

Figure 6.15. Map view of the Powder Mountain 1-13E showing relative connectivity.
Low areas are in blue and red is high connectivity. Most of the model is moderately
connected to other fractures.
172

100 m

Figure 6.16. Map view of the Bogey Draw 1-14 well showing relative fracture
connectivity. Low areas are in blue and red is high connectivity. The red lines show
hypothetical well locations. There is a large connected fracture network in the center of
the model. The figure on the bottom shows part of this network, and the high density of
connected fractures.
Cepo-Lewis 21-18
9
8 N= 25 wells
7
6
5
4
3

Frequency
2
mean = 1.75
1
0
0 01 21 23 34 4
# of fractures

Figure 6.17. Cepo-Lewis 21-18 well showing


fracture intersections, as disks, with 25 arbitrary
boreholes. Fracture Set 1 is in yellow and
Fracture Set 2 is in green. The graph shows the
frequency of intersections, with the average
being 1.75 intersections per well. The red and
white disks are the bounding units, 100 ft thick.

N
173
PMU 1-13E
8
7
6
N= 25 wells
5
4
3

Frequency
2
1 mean = 2.85
0
0
1 2 3 4 5 6
# of fractures

Figure 6.18. Powder Mountain 1-13E


well showing fracture intersections,
as disks, with 25 arbitrary boreholes.
Fracture Set 1 is in yellow and
Fracture Set 2 is in green. The graph
shows the frequency of intersections,
with the average being 2.85
intersections per well. The red and
white disks are the bounding units,
100 ft thick.
174
Bogey Draw #1
6
5 N= 25 wells
4
3
2

Frequency
1
0
mean = 6.5
0 1 12 2 3 34 4 5 5 6 67 7 8 89 9101011111212
# of fractures

Figure 6.19. Bogey Draw 1-14 well showing fracture


intersections, as disks, with 25 arbitrary boreholes. Fracture Set 1
is in yellow and Fracture Set 2 is in green. The graph shows the
frequency of intersections, with the average being 6.5
intersections per well. The red and white disks are the bounding
units, 100 ft thick. The figure on the right illustrates the
difference in dip of the 2 fracture sets.
N
175
176

800 m

Figure 6.20. Map view of the model showing the orientation of the cross sections in
Figures 6.21, 6.22 and 6.23.
W N

Northeast-Southwest cross section of Cepo-Lewis 21-18 discrete fracture model. Wells are green vertical lines. Fractures can be seen in yellow for Fracture
g to the southwest, and in blue for Fracture Set 2, dipping to the northeast.
W NE

Northeast-Southwest cross section of Powder Mountain 1-13E discrete fracture model. Wells are green vertical lines. Fractures can be seen in yellow for Fracture S
o the southwest, and in blue for Fracture Set 2, dipping to the northeast.
Northeast-Southwest cross section of Bogey Draw #1-14 discrete fracture model. Wells are green vertical lines. Fractures can be seen in yellow for Fracture Set 1, d
and in blue for Fracture Set 2, dipping to the northeast.
180

Figure 6.24 shows the curvature map generated for the Lewis Shale. The

curvature map generated for the Almond is shown in Figure 6.25. The cylindrical

analysis map shows a dip to the west (275º azimuth) and a plunge of 2.8º (Figure 6.26).

Figure 6.27 shows the spatial analysis conducted on the Lewis Shale to show the line

intensity of the fractures, which is a measure of the line length per grid cell. This is a

guide to the fracture density. The map shows lower fracture intensity near the center of

the model and a higher intensity in the south, near the mapped faults.

6.4 Discussion

Perhaps the most difficult parameter needed for the 3D model building is

horizontal fracture density. The distance between fractures within beds cannot generally

be determined from a core (Narr and Lerche, 1984). Subsurface data are limited to either

simply counting fractures in core, or indirectly gaining responses, such as those found in

production records. These parameters can be influenced by a number of other factors,

such as well completion techniques, as shown in Chapter 4. A variety of factors

influence the spacing of fractures. These include lithology and the preference for

fractures to propagate in sands rather than in shales. Also, the proximity to tectonic

structures is important, because these are in highly strained areas. A second assumption

made that will affect the modeling was about the fracture length used from outcrop
181

N 800 m

PM 1-13E

Cepo 21-18

Polar Bar 1 PM Fed 32-26


PM Fed 34-26
PM Fed 23-36

Triton #10
PM Fed 34-11
Bogey Draw 1-14

Black Bar 1

Color scale:
min
mid
max

Figure 6.24. Curvature map on Lewis Shale. The edges should be disregarded because
of edge effects. The wells are labeled and the interpreted 3D seismic fault planes are
shown as the solid colored areas. The pink line is the N15W lineament discussed in
chapter 5.
182

N 800 m

PM 1-13E

Cepo 21-18

Polar Bar 1 PM Fed 32-26


PM Fed 34-26
PM Fed 23-36

Triton #10
PM Fed 34-11
Bogey Draw 1-14

Black Bar 1

Color scale:
min
mid
max

Figure 6.25. Curvature map on the Almond. The edges should be disregarded because of
edge effects. The curvature is relatively homogeneous in the center. The wells are
labeled.
183

N 800 m

PM 1-13E

Cepo 21-18

Polar Bar 1 PM Fed 32-26


PM Fed 34-26
PM Fed 23-36

Triton #10
PM Fed 34-11

Bogey Draw 1-14

Black Bar 1

Color scale:
min
mid
max

Figure 6.26. Cylindrical analysis, dips to the west (275.57º) at an angle of 2.84º. The
deviations seen are more folded areas. The N15W pink lineament is shown.
184

N PM 1-13E (0.4) 800 m

Cepo 21-18 (1.8)

Polar Bar 1 (0.05) PM Fed 32-26


PM Fed 34-26 (1.1)
PM Fed 23-36 (0.07)

Triton #10 (0.2)


PM Fed 34-11

Bogey Draw 1-14

Black Bar 1

Color scale:
min
mid
max

Figure 6.27. Map view of top of the Lewis Shale showing the spatial analysis of fracture
intensity, constrained to the curvature map. Wells are labeled with the 12 month
cumulative production in bcf.
185

studies at Muddy Gap. Unloading and decompaction create stress that causes an increase

in the number of fractures seen in outcrop to those seen in the subsurface. Also, although

the facies are different, the bed thickness is similar in this study and the work done at

Muddy Gap (Harstad et al., 1995).

The assumption made in Chapter 3 of removing data points with a dip magnitude

over 70°, causes some question. By including these steeply dipping natural fractures, the

average dip values of fractures would have been higher than actually used for the

modeling. Due to the increased dip magnitude, each of the 25 sample wells in the

discrete fracture modeling section would have encountered fewer fractures than actually

modeled. Thus, an even higher fracture density perhaps should have been modeled.

However, as the same treatment was applied to all these modeled scenarios, the relative

fracture density is still valid.

The Lewis Shale curvature map (Figure 6.24) used as a probability constraint

shows similar patterns to the Almond curvature map (Figure 6.25). This suggests that

these slight structural features are not just a product of correlation, but are actually

present in the area. Thus, it was a viable idea to use curvature maps to constrain fracture

distribution. The curvature map shown in Figure 6.24 has the interpreted pink fault from

Figures 5.8, 5.9 and 5.12. The curvature map shows that there is a higher curved area

along this possible fault.

The well with the lowest amount of sandstone and the least fractures is also the

best producer of the 3 modeled wells. This is because the fractures were constrained to
186

the sandier units. The intensity values used for the three models are a factor of net-to-

gross ratios, which varies considerably between wells (Table 6.3). It seems that the

element of stratigraphic trapping may have more importance than first thought. The

laminated sandstones have pinchouts and shaly interbeds that act as traps. This is similar

to many other Rocky Mountain fields, for example the Rulison field in the Piceance

Basin, Colorado. The hydrocarbon reservoirs found here in the Williams Fork Formation

are found in the lower net-to-gross stacked fluvial sands in the anticlinal nose (Cole and

Cumella, 2003). At Rulison field, the nose of the structure provides a trap for the gas.

The thinner sands are able to provide a stratigraphic trap for migrating gas. Although the

depositional environment is different in the Rulison verses CEPO/Powder Mountain

areas, the similarity in low net-to-gross interval trapping may be valid.


187

CHAPTER 7

CONCLUSIONS

The purpose of this study is to build a 3D geological model of the CEPO/Powder

Mountain field, using borehole image logs and cores as input data for fracture orientation

and intensity. The data set includes 105 mi2 (168 km2) of 3D seismic data, 8 wells with

core, and 4 wells with borehole images.

The major conclusions for the study are:

Partially open fractures are present only in sandy units of the cores.

Mineral growths (quartz and calcite) on the walls of the fractures indicate that the

fractures are probably natural.

Two conjugate fracture sets were identified from the borehole image logs.

Fracture Set 1 has a dip direction/dip of 213°/65°and Fracture Set 2 has a dip

direction/dip of 12°/54°.

Fracture Set 1 has 5 times more fractures than Set 2.

Borehole breakouts show a present-day maximum stress direction of N34°W,

which is consistent with other work in the area.


188

Wide variations in production, as identified by first 12 months of production,

cannot be explained as simply the result of completion practices. The highest

cumulative production rate is from the Cepo Lewis 21-18 well (1.8 bcf), while the

lowest cumulative production comes from the Powder Mountain Fed 23-36 (0.07

bcf). Some dry holes are present in the area.

Gas shows, lost circulation while drilling, pressure dependent leakoff, and the use

of 100-mesh sand show that natural fractures affect production.

Highest mud weight was used in the well with the best production. Higher

amounts of shales in the best producing well may sustain overpressure.

There is a N70°W trend of faults in southern part of the study area from seismic

data.

There is a N15°W fault/fracture trend identified from Vp/Vs and percent

anisotropy maps.

Two approaches were taken to fracture modeling: (1) attribute-based fracture

modeling, and (2) discrete fracture modeling.

Discrete 3D fracture models show that fractures are common throughout the area.

The distribution of fractures in the 3D model was constrained by the amount of

sandstone. The modeled well with the lowest fracture density had the lowest

amount of sandstone and was the best producer in the field.

3D structural models visualize the relative connectivity of fractures. Wells with

more sandstone had a more highly connected fracture network.


189

Lewis Shale production may be affected by the presence of stratigraphic trapping,

more so than structural traps. Laminated sandstones have pinch-outs and shaly

interbeds that act as stratigraphic traps. Therefore, wells with lower net-to-gross

ratios may provide better traps.

The attribute-based fracture model related curvature to fracture occurrence in the

Lewis Shale. Minor variations in curvature could be seen, with a subtle N15°W

trend. The same trend could be seen in the Lewis Shale cylindrical map, and

some of the seismic maps.


190

REFERENCES

Alford, R. M., 1986, 56th SEG: Shear data in the presence of azimuthal anisotropy:
Dilley, Texas.

Baars, D. L., B. L. Bartleson, C. E. Chapin, B. F. Curtis, R. H. De Voto, J. R. Everett,


R.C. Johnson, C. M. Molenaar, F. Peterson, C.J. Schenk, J. D. Love, I. S. Merin, P. R.
Rose, R. T. Ryder, N. B. Waechter, L. A. Woodward, 1988, Basins of the Rocky
Mountain region: in Sloss, L. L., ed., Sedimentary cover-North American craton: U.S.,
The Geology of North America, Vol. D-2, GSA, p. 109-165.

Bader, J. W., 1987, Surface and subsurface structural relations of the Cherokee Ridge
Arch, south-central Wyoming: Unpublished M.S. Thesis, San Jose State University, San
Jose, California, 68 p.

Bale, R., Dumitru, G., and Probert, T., 2000, Analysis and stacking of 3-D converted-
wave data in the presence of azimuthal anisotropy (abs): 70th Annual International
Meeting, Society of Exploration Geophysics, 1189-1192.

Barree, R.D., and Mukherjee, H., 1996, Determination of pressure dependent leakoff and
its effect on fracture geometry, SPE 36424 71st Annual Technical Conference and
Exhibition, 10 p.

Blackstone, D. L. Jr., 1963, Development of geologic structure in central Rocky


Mountains in Memoir #2: Backbone of the Americas: Tectonic History from Pole to
Pole: AAPG Special Publication, p. 160-179.

Biddle, K. T., and Christie-Blick, N., 1985, Deformation and basin formation along
strike-slip faults in Biddle, K. T., and Christie-Blick, N., eds., Strike-slip deformation,
basin formation, and sedimentation, SEPM Special Pub No. 37, p. 1-35.

Cole, R. and Cumella, S., 2003, Stratigraphic architecture and reservoir characteristics of
the Mesaverde group, southern Piceance basin, Colorado: RMAG and AIPG Field Trip
Guidebook, p. 386-449.

Crampin, S., and Atkinson, B.K., 1985, Microcracks in the earth’s crust: First Break, v. 3,
No. 3, p. 16-20.

Crampin, S., 1994, The fracture criticality of fractured rock: Geophysical Journal
International, v. 118, p. 428-434.
191

Crampin, S., 2000, The potential of shear-wave splitting in a stress-sensitive compliant


crust, A new understanding of pre-fracturing deformation from time-lapse studies (abs):
Annual International Meeting, Society of Exploration Geophysics, p. 1520-1523.

Cronoble, J. M., 1969, South Baggs-West Side Canal gas field, Carbon County,
Wyoming and Moffat County, Colorado: in Barlow, J. A., ed., Wyoming Geological
Association Guidebook, 21st Annual Field Conference, p. 121-137.

Cross, T. A., 1986, Tectonic controls of foreland basin subsidence and Laramide style
deformation, western United States: in Allen, P. A., and Homewood, P., ed, Foreland
basins: International Association of Sedimentologists Special Publication 8, p. 15-39.

Davis, T. L., 2001, Multicomponent seismology - The next wave: Geophysics, v. 66, No.
1, p. 49.

DeVault, B., 1997, 3-D Seismic prestack multicomponent amplitude analysis, Vacuum
Field, Lea County, New Mexico: Unpublished Ph.D. Dissertation, Colorado School of
Mines, Golden, CO, 208 p.

Fouch, T.D., Lawton, T.F., Nichols, D.J., Cashion, W.B., and Cobban, W.A., 1983,
Patterns and timing of synorogenic sedimentation in Upper Cretaceous rocks of central
and northeast Utah, in M. Reynolds and E. Dolly, eds., Mesozoic paleogeography of the
west-central United States: Rocky Mountain Section of SEPM, Rocky Mountain
Paleogeography Symposium 2, p. 305–336.

Gaiser, J. E., 2000, Advantages of 3-D PS-wave data to unravel S-wave birefringence for
fracture detection: 70th Ann. Internat. Mtg., Soc. Expl. Geophyics., Expanded Abstracts,
p. 1201–1204.

García-González, M., Surdam, R.C., and Lee, M. L., 1997, Generation and expulsion of
petroleum and gas from Almond Formation coal, Greater Green River basin, Wyoming:
AAPG Bulletin, v. 81, p. 62-81.

Gill, J. R., and W. A. Cobban, 1973, Stratigraphy and geologic history of the Montana
Group and equivalent rocks, Montana, Wyoming, and North and South Dakota: U. S.
Geological Survey Professional Paper 776, 37 p.

Grace, L. M., and Newberry, B. M., 1998, Geological application of dipmeter and
borehole electrical images, Workshop in Denver, CO., July 1998, Schlumberger Oilfield
Services, Version 8.1.
192

Groshong, Jr., R.H., 1975, Strain, fractures and pressure solution in natural single-layer
folds: GSA Bulletin v. 86. p. 1363-1376.

Hale, L. A., 1961, Late Cretaceous (Montana) stratigraphy, eastern Washakie Basin,
Carbon County, Wyoming, in G. J. Wiloth, ed., Symposium on Late Cretaceous rocks,
Wyoming and adjacent areas: Wyoming Geological Association Guidebook, Sixteenth
Annual Field Conference, Green River, Washakie, Wind River, and Powder River Basins, p.
129-137.

Halliburton EMI brochure (2003).

Harding, T. P., 1974, Petroleum traps associated with wrench faults: AAPG Bulletin, v.
58, p.1290-1304.

Harding, T. P., 1990, Identification of wrench faults using subsurface structural data:
criteria and pitfalls: AAPG Bulletin, v. 74, p. 1590-1609.

Harstad, H., Teufel, L. W., and Lorenz, J. C., 1995, Characterization and simulation of
naturally fractured tight gas sandstone reservoirs: SPE 30573 Annual Technical
Conference, p. 437-446.

Haun, J. D., 1961, Stratigraphy of post-Mesaverde Cretaceous rocks, Sand Wash basin
and vicinity, Colorado and Wyoming: in Wiloth, G. J., ed., Wyoming Geological
Association Guidebook, Symposium on late Cretaceous rocks of Wyoming, p. 116-124.

Hennier, J.H., 1984, Structural analysis of the Sheep Mountain anticline, Bighorn Basin,
Wyoming.: Unpublished Master’s Thesis, Texas A&M University, College Station, TX.

Hull, E.A. 2001, 3-D seismic interpretation of tectonics and sedimentation within the
Upper Cretaceous Lewis Shale and Fox Hill Sandstone, Baggs, Wyoming: Unpublished
Master’s Thesis, Colorado School of Mines, Golden, Colorado, 73 p.

Jenner, E., 2001, Azimuthal anisotropy of 3-D compressional wave seismic data,
Weyburn field, Saskatchewan,Canada: Unpublished Ph.D. dissertation, Colorado School
of Mines, Golden, CO, 210 p.

Jordan, T. E., 1981, Thrust loads and foreland basin evolution, Cretaceous, Western
United States: AAPG Bulletin, v. 65, p. 2506-2520.
193

Knight, C.N., 1999, Structural and stratigraphic controls on the Mesaverde reservoir
performance: North La Barge field, Sublette county, Wyoming: Unpublished PhD.
dissertation, Colorado School of Mines, Golden, CO, 217 p.

Krugh, K. A., 1997, U-Pb thermochronologic constraints on the early Proterozoic


tectonic evolution of the Hartville uplift, southeast Wyoming: Unpublished M. S. Thesis,
University of Wyoming, Laramie, Wyoming.

Krystinik L. F. and DeJarnett, B. B., 1995, Lateral variability of sequence stratigraphic


framework in the Campanian and Lower Maastrichtian of the Western Interior Seaway :
in J. C. Van Wagoner and G. T. Bertram, eds., Sequence stratigraphy of foreland basin
deposits, AAPG Memoir 64, p. 11–25.

Kulander, B. R., Dean, S. L., and Ward, B. J., 1990, Fractured core analysis:
Interpretation, logging and use of natural and induced fractures in core, AAPG Methods
in exploration series 8, AAPG, Tulsa, 88 p.

Laubach, S.E., 2003. Practical approaches to identifying sealed and open fractures.:
AAPG Bulletin, v. 87, p. 561–579.

Law, B.E., Pollastro, R.M. and Keighin, C.W. 1986, Geologic characterization of low-
permeability gas reservoirs in selected wells, Greater Green River Basin, Wyoming,
Colorado, and Utah: in C. Spencer and R. Mast, eds., Geology of tight gas reservoirs:
AAPG Studies in Geology, v. 29, p. 253–269.
Law, B. E., and Spencer, C. W. eds., 1989, Geology of tight gas reservoirs in the Pinedale
anticline area, Wyoming and at the Multiwell Experiment site, Colorado: U.S. Geological
Survey Bulletin 1886, p. 39–61.
Law, B. E., 1996, Southwestern Wyoming province (037): in D. L. Gautier, G. L. Dolton,
K. I. Takahashi, and K. L. Varnes, eds., 1995 national assessment of United States oil and
gas resources - results, methodology, and supporting data: U.S. Geological Survey
Digital Data Series DDS-30, Release 2, 1 CDROM.

Law, B. E., 2000, What is a basin-centered gas system?: Basin-centered gas symposium:
Rocky Mountain Association of Geologists, 8 p.

Law, B. E., 2002, Basin-centered gas system: AAPG Bulletin, v. 86, p. 1891–1919.

Love, J. D., 1970, Cenozoic geology of the Granite Mountains area, Central Wyoming:
USGS Professional Paper 495-C, 154 p.
194

Magoon, L. B., and Dow, W. G. eds., 1994, The petroleum system - from source to trap:
AAPG Memoir 60, 655 p.

Maughan, E. K., and Perry, W. J. Jr., 1986, Lineaments and their tectonics implications in
the Rocky Mountains and adjacent plain areas: in J. A. Peterson, ed., Paleotectonics and
Sedimentation in the Rocky Mountain Region, United States, AAPG Memoir 41 , p. 41–
53.

McClay, B., and Bonora, M., 2001, Analog models of restraining stepovers in strike-slip
fault systems: AAPG Bulletin, v. 85, p. 233-260.

Middlebrook, M. L., Aud, W. W., and Harkrider, J. D., 1995, An evolving approach in
the analysis of stress test pressure decline data: SPE 29599, p. 549-566.

Miller, F. X., 1977, Biostratigraphic correlation of the Mesaverde Group in southwestern


Wyoming and northern Colorado, in Exploration frontiers of the central and southern
Rockies: Rocky Mountain Association of Geologists, p. 117–137.

Minton, G.E. 2002, Subsurface study of the Lewis Shale in the southern Washakie and
Sand Wash basins using borehole images, cores, well log and seismic data: Unpublished
Master’s Thesis, Colorado School of Mines, Golden, Colorado, 220 p.

Molenaar, C. M., and D. D. Rice, 1988, Cretaceous rocks of the Western Interior Basin:
in Sloss, L. L., ed., The Geology of North America, Vol. D-2, Sedimentary Cover-North
American Craton: U.S., GSA, p. 77-82.

Mukherjee, H., Larkin, S., and Kordziel, W., 1990, Extension of fractured pressure
decline curve analysis to fissured formations: SPE 21872, Proc., SPE Rocky Mountain
Regional/Low Permeability reservoirs Symposium, Denver (1991), p. 671-685.

Narr, W., and Lerch, I., 1984, A method for estimating subsurface fracture density in
core: AAPG Bulletin, v. 68, p. 637–648.

Narr, W., 1991, Fracture density in the deep subsurface: techniques with application to
Point Arguello oil field: AAPG Bulletin, v. 75, p. 1300–1323.

Narr, W., 1996, Estimate average fracture spacing in subsurface rock: AAPG Bulletin, v.
80, p. 1565–1586.

Nelson, R. A., 2001, Geological analysis of naturally fractured reservoirs, 2nd ed., 336 p.
195

Perez, M. A., Gibson, R.L., and Toks, M.N., Detection of fracture orientation using
azimuthal variation of P-wave AVO responses: Geophysics, v. 64, No. 4, p. 1253–1265.

Pyles, D. R., 2000, A high frequency sequence stratigraphic framework for the Lewis
Shale and Fox Hills Sandstone, Great Divide and Washakie Basins, Wyoming:
Unpublished M.S. Thesis, Colorado School of Mines, Golden, Colorado, 261 p.

Rahmat, N. A. M. N., 2000, Borehole image analysis of the Cretaceous Lewis Shale,
Sand Wash basin, Moffat County, Colorado: Unpublished Master’s Thesis, Colorado
School of Mines, Golden, Colorado, 156 p.

Reasnor, M., 2001, Forward modeling and interpretation of multicomponent seismic data
for fracture characterization, Weyburn field, Saskatchewan: Unpublished Master’s thesis.
Colorado School of Mines, Golden, CO, 171 p.

Reynolds, M. W., 1976, Influence of recurrent Laramide structural growth on


sedimentation and petroleum accumulation, Lost Soldier area, Wyoming:
AAPG Bulletin, v. 60, p. 12-33.

Roehler, H.W., 1990, Stratigraphy of the Mesaverde Group in the central and eastern
Greater Green River Basin, Wyoming, Colorado, and Utah: U.S. Geological Survey
Professional Paper 1508, 52 p.

Schlische, R. W., Withjack, M. O., and Eisenstadt, G., 2002, An experimental study of
the secondary deformation produced by oblique-slip normal faulting: AAPG Bulletin, v.
86, p. 885–906.

Schlumberger website brochure (accessed 3/10/03)


http://www.slb.com/oilfield/index.cfm?id=id3139

Springer, J., 1987, Stress orientations from wellbore breakouts in the Coalinga region:
Tectonics, v. 6, p. 677-676.

Shanley, K.W., Cluff, R.C., Shannon, L.T., and Robinson, J.W. 2003, Controls on
prolific gas production from low-permeability sandstone reservoirs in basin-centered
regions: Implications from the Rocky Mountain region for resource assessment, prospect
appraisal, and risk analysis (abs.): AAPG annual meeting-Utah, p. A156.

Snoke, A. W., 1997, Geologic history of Wyoming within the tectonic framework of the
North American cordillera: Wyoming State Geological Survey Public Information
Circular.
196

Stewart, R.R., Gaiser, J.E., Brown, R.J., and Lawton, D.C., 2003, Converted-wave
seismic exploration: Applications: Geophysics, v. 68, No. 1, p. 40–57.
Surdam, R. C., 1997, A new paradigm for gas exploration in anomalously pressured
“tight gas sands” in the Rocky Mountain Laramide basins, in R. C. Surdam, ed., Seals,
traps, and the petroleum system: AAPG Memoir 67, p. 283–298.

Swift, T. E., and Mladenka, P., 1997, Technology tackles low-permeability sand in south
Texas, Oil and Gas Journal, v. 95, p. 68.

Thomsen, L., 1986, Weak elastic anisotropy: Geophysics, v. 51, p. 1954-1966.

Thomsen, L., 1988, Reflection seismology over azimuthally anisotropic media:


Geophysics, v. 53, p. 304-313.

Thomsen, L., 1999, Converted-wave reflection seismology over inhomogeneous,


anisotropic media: Geophysics, v. 64, p. 678-690.

Van Dok, R. R., Gaiser, J. E., Jackson, A.R., and Lynn, H. B., 1997, 3-D converted-wave
processing: Wind River Basin case history (abs.): 67th Ann. Internat. Mtg., Soc. Expl.
Geophysics, Expanded Abstracts, p.1206–1209.

Weimer, R. J., 1960, Upper Cretaceous stratigraphy, Rocky Mountain area: AAPG
Bulletin, v. 44, p. 1-20.

Winn, R.D. Jr., Bishop, M. G. and Gardner, P.S. 1985, Lewis Shale, south-central
Wyoming: shelf, delta front, and turbidite sedimentation: Wyoming Geological
Association 36th Annual Field Conference Guidebook, p. 113–130.

Winn R. D. Jr., Bishop, M. G. and Gardner, P. S. 1987, Shallow-water and sub-storm-


base deposition of Lewis Shale in Cretaceous Western Interior seaway, south-central
Wyoming: AAPG Bulletin, v. 71, p. 859-881.

Winterstein, D. F., 1992, Discussion on field measurements of azimuthal anisotropy: first


60 meters, San Fransisco bay area, CA, and estimation of the horizontal stresses’ ratio
from Vs1/Vs2 by Lynn, H. B., with reply by the author: Geophysics, v. 57, p. 653-656.

Wu, H. and Pollard., D.D. 2002, Imaging 3-D fracture networks around boreholes:
AAPG Bulletin, v. 86, p. 593–604.

Wyoming Oil and Gas Conservation Commission Website. http://wogcc.state.wy.us/


accessed: 3rd Sept 2003.
197

Ysaccis, J., 2003, Structural interpretation of the Cherokee Arch, south central Wyoming,
using 3-D seismic data and well logs, Unpublished Master’s Thesis, Colorado School of
Mines, Golden, Colorado.

Zainal, Z., 2001, Subsurface geology and petroleum system of the Lower Lewis Shale in
the Sand Wash basin, northwest Colorado. Unpublished Master’s Thesis, Colorado
School of Mines, Golden, Colorado.

Zheng, Z., Keneny, J., and Cook, N.G.W., 1989, Analysis of borehole breakouts: Journal
of Geophysical Research, v. 94, p. 7171-7182.

Zhu, X., Gibson, J., Ravindran, N., Zinno, R., and Sixta, D., 1996, Seismic imaging of
hydraulic fractures in Carthage tight sands: A pilot study: The Leading Edge, v. 15, No.
3, p. 218-224.

3D Move Online help, 2002.


198
APPENDIX A

PM 34-26 PM 34-11
log depth vertical frac length of frac mineral log depth vertical fracs length of frac mineral
(ft) # (cm) (ft) (cm)
12020 0 13360 0
12015 0 13365 0
12010 0 13370 0
12005 0 13375 0
12000 0 13380 2 22 calcite
11995 2 120 open 22 calcite
120 open 13385 2 1
11990 0 3
11985 0 13390 2 2 open
11980 0 3 open
11975 0 13395 0
11970 1 12 open 13400 2 3 open
11965 2 10 quartz 5 open
11 calcite 13405 0
13410 2 3 open
5 open
13415 2 5 open
60 calcite
13420 0
13425 0
13430 0
13435 0
13440 0
13445 0
13450 0
13455 0
13460 0
13465 0
13470 0
13475 0
13480 0
13485 1 120 calcite
13490 0
13495 0
13500 0
199

Polar Bar #1 Triton #10


log depth (ft) vertical fracs length of frac mineral log depth (ft) vertical fracs length of frac mineral
(cm) (cm)
15225 1 13 quartz 13280 0
15230 0 13285 0
15235 0 13290 0
15240 0 13295 1 54 open
15245 0 13300 0
15250 0 13305 0
15255 0 13310 0
15260 0 13315 0
15265 0 13320 0
15270 1 30 calcite 13325 0
15275 0 13330 0
15280 0 13335 0
15285 0
200

Black Bar #1 Bogey Draw 1-14


log depth (ft) vertical fracs length of frac mineral log depth (ft) vertical fracs length of frac mineral
(cm) (cm)
14060 0 13875 0
14065 0 13880 0
14070 3 25 calcite 13885 0
20 calcite 13890 0
7 calcite 13895 0
14075 0 13900 0
14080 0 13905 2 65 calcite
14085 3 1 calcite out of core calcite
7 calcite 13910 2 36 calcite
36 calcite 36 calcite
14090 0 13915 0
14095 0 13920 0
14100 0 13925 0
14105 0 13930 0
14110 0
14115 1 15 calcite
14120 1 16 calcite
14125 0
14130 0
14135 0
200

mineral

calcite
calcite
calcite
calcite
201

PM 1-13E CL21-18
log depth (ft) vertical fracs length of frac mineral log depth (ft) vertical fracs length of frac mineral
(cm) (cm)
13330 0 13250 0
13340 2 11 calcite 13260 0
22 calcite 13270 0
13350 0 13280 2 15 calcite
13360 0 18 calcite
13370 1 19 calcite 13290 0
13380 0
13390 0
APPENDIX B 202

32-26 23-36
Depth Fracture Type Dip Direction Dip Depth Fracture Type Dip Direction Dip

10274 open 49 70 12747 open 198 85


10278 open 213 73 12754 open 219 83
10279 open 257 74 12757 induced 202 88
10281 open 237 77 12776 open 210 84
10287 open 203 56 12781 open 223 85
10291 induced 213 82 12792 induced 209 88
10295 induced 192 84 12806 induced 295 87
10307 open 193 79 12820 induced 197 87
10313 open 74 48 12826 induced 217 87
10334 open 339 51 12852 open 234 83
10374 open 77 77 12855 open 236 83
10404 open 351 69 12858 open 226 82
10406 open 5 77 12861 open 225 82
10440 open 50 71 12866 open 208 74
10452 induced 226 84 12874 induced 232 87
10459 induced 210 82 12877 open 207 64
10465 open 322 73 12878 open 227 66
10510 induced 197 83 12886 induced 293 86
10527 open 41 77 12900 open 172 68
10546 open 180 67 12901 open 189 64
10550 open 354 52 12902 induced 155 84
10562 open 214 67 12905 open 179 69
10564 open 220 79 12918 open 196 65
10566 induced 237 81 12919 open 191 73
10576 induced 246 86 12927 induced 190 89
10584 induced 190 88 12946 open 198 65
10591 induced 197 86 12947 induced 215 87
10611 open 165 62 12960 induced 222 85
10613 open 111 44 12972 induced 216 86
10621 induced 207 88 12989 induced 211 84
10633 induced 190 84 13032 open 213 77
10642 open 294 64 13036 induced 6 88
10643 open 294 64 13100 open 185 80
10657 open 186 79 13102 induced 192 86
10669 induced 197 87 13108 open 179 83
10676 open 218 78 13185 open 343 43
10678 open 224 66 13252 open 304 76
10681 open 255 75 13757 induced 205 86
10684 induced 224 88 13764 open 209 36
10691 induced 194 81 13794 induced 223 87
10701 open 174 74 13805 induced 195 87
10702 induced 174 83 13830 induced 213 87
10715 induced 216 81
10728 open 214 77
10734 induced 243 83
10787 open 126 49
10788 open 167 47
10791 open 98 31
10792 open 177 30 203
10803 open 164 72
10806 induced 208 80
10808 open 146 72
10826 open 273 79
10826 open 177 72
10829 open 211 69
10841 induced 213 84
10843 induced 221 80
10848 induced 250 84
10867 induced 187 87
10874 open 24 62
10881 induced 173 80
10886 open 328 41
10891 open 181 69
10895 open 200 75
10898 induced 211 80
10900 induced 194 87
10920 induced 200 88
10928 open 208 72
10932 induced 198 84
10945 induced 195 86
10951 induced 254 84
10955 open 217 64
10961 induced 223 83
10964 induced 208 82
10966 induced 204 85
10971 induced 203 83
10995 induced 193 85
11009 induced 190 83
11010 induced 191 81
11016 open 216 61
11035 induced 180 82
11047 induced 220 83
11060 open 299 63
11068 open 246 75
11071 induced 206 82
11075 open 209 75
11104 induced 223 81
11112 open 152 67
11128 open 177 74
11129 open 180 74
11142 open 218 76
11162 induced 201 87
11169 induced 187 83
11172 induced 197 82
11180 open 147 59
11210 open 1 48
11214 open 247 69
11230 open 166 74
11262 open 203 76
11272 open 221 74
11274 open 206 69
11276 induced 218 81 204
11290 open 1 76
11292 induced 180 80
11352 open 197 72
11354 induced 197 81
11374 induced 223 82
11378 induced 214 81
11434 induced 191 84
11452 induced 205 85
11455 induced 201 84
11458 open 166 70
11460 open 198 75
11466 induced 203 85
11470 open 213 73
11475 open 177 72
11485 open 173 58
11492 induced 214 81
11522 HEALED 225 26
11524 induced 178 81
11549 open 180 76
11558 open 21 76
11714 open 196 68
11727 induced 200 80
11786 open 24 46
11792 open 191 76
11794 open 236 67
11795 open 223 71
11796 open 207 71
11840 induced 214 82
11862 induced 207 81
11866 open 200 77
11912 open 18 58
11952 induced 198 80
11968 induced 195 82
11971 open 195 78
11972 open 210 66
11976 induced 211 82
12014 HEALED 223 53
12014 HEALED 208 59
12016 open 65 47
12040 open 352 61
12042 open 169 52
12052 open 207 71
12055 induced 207 83
12347 open 308 62
12497 open 4 62
205

34-11 1-13E
Depth Fracture Type Dip Direction Dip Depth Fracture Type Dip Direction Dip

11230 open 215 85 12818 induced 254 87


11236 open 36 80 12824 induced 247 85
11253 open 199 85 12829 induced 239 86
11258 open 214 77 12832 induced 263 84
11279 open 211 85 12836 open 225 74
11287 open 244 86 12840 induced 247 88
11296 induced 206 85 12846 induced 253 83
11302 induced 209 86 12851 induced 228 83
11304 open 205 66 12852 induced 231 83
11308 open 325 56 12856 induced 228 84
11316 open 206 70 12858 induced 245 87
11323 open 201 63 12860 induced 256 87
11324 open 207 66 12863 induced 248 85
11326 open 262 81 12866 induced 245 83
11332 open 255 84 12868 induced 246 88
11335 open 243 84 12884 open 235 75
11338 induced 271 84 12887 induced 243 85
11342 open 269 71 12890 induced 229 86
11354 open 250 42 12893 induced 238 84
11357 induced 218 85 12896 open 250 75
11360 open 207 64 12898 induced 261 83
11362 open 232 55 12901 induced 250 86
11371 open 204 73 12903 induced 258 85
11373 open 208 53 12906 open 265 78
11374 open 195 54 12914 open 250 79
11393 open 208 70 12940 induced 248 87
11412 open 238 53 12952 induced 243 84
11419 open 212 84 12963 induced 288 87
11422 induced 205 84 12967 induced 276 86
11440 open 212 86 12968 induced 241 89
11450 open 123 28 12977 induced 221 83
11456 open 197 35 13000 induced 242 85
11462 open 351 47 13006 induced 262 87
11465 induced 180 84 13007 induced 256 84
11469 induced 172 85 13010 induced 245 83
11472 open 197 69 13013 open 244 73
11474 open 203 82 13045 open 249 78
11483 open 197 70 13151 HEALED 6 19
11546 open 201 70 13309 HEALED 228 38
11548 induced 223 87 13336 open 193 38
11564 induced 202 86 13353 induced 250 85
11571 open 182 75 13355 induced 264 89
11576 open 150 77 13356 induced 255 86
11580 open 215 78 13361 induced 59 87
11581 open 182 86 13373 induced 266 86
11600 open 218 58 13375 induced 232 86
11604 open 198 85 13386 induced 236 86
11644 open 252 87
11664 open 23 78 206
11669 induced 226 85
11688 induced 215 86
11699 open 172 70
11725 induced 224 83
11728 induced 248 84
11775 open 200 75
11778 open 211 83
11782 open 231 84
11787 induced 204 83
11808 open 206 74
11811 open 194 80
11819 induced 211 84
11828 open 220 77
11849 open 187 84
11855 open 239 78
11923 open 252 62
11954 open 210 65
11969 open 218 44
11969 open 218 43
11974 open 29 38
12003 open 310 44
12004 open 337 48
12008 open 205 63
12010 open 229 64
12011 open 54 43
12019 open 232 72
12032 open 188 71
12033 open 226 69
12035 open 26 68
12036 open 227 61
12040 open 234 69
12042 open 24 63
12047 open 187 62
12050 open 218 78
12053 open 190 49
12056 open 204 52
12057 open 31 71
12058 open 33 69
12061 open 212 77
12065 open 188 77
12072 induced 216 86
12091 open 226 71
12096 induced 35 85
12096 open 203 74
12099 open 216 81
12109 open 206 70
12114 open 214 64
12116 open 219 78
12133 open 192 42
12141 induced 209 87
12154 open 191 59
12166 open 200 86
12167 open 184 77 207
12215 open 288 57
12217 open 378 45
12226 open 223 82
12233 open 231 72
12240 induced 222 86
12256 induced 212 87
12279 open 249 69
12279 open 249 72
12282 open 249 60
12283 open 73 56
12299 induced 23 85
12341 open 154 69
12357 open 164 71
12358 open 164 72
12365 induced 263 82
12505 open 233 70
12506 open 238 64
12515 induced 206 86
12536 open 228 61
12537 open 39 86
12544 open 216 85
12565 open 75 51
12569 induced 205 85
12573 open 243 59
12610 open 210 23
12611 open 192 57
12628 induced 334 84
12637 open 223 82
12649 induced 32 85
12701 induced 203 85
12707 open 320 62
12730 open 199 77
12747 open 196 76
12752 open 230 72
12778 open 279 59
12802 open 262 68
12910 open 27 49
12924 open 196 51
12926 open 23 59
12926 open 170 67
12932 open 259 54
12960 open 4 79
12963 open 192 77
12967 open 173 83
208

APPENDIX C

Powder Mountain 1-13E:

Cumulative Production
From Nov 1999 To May 2003
Oil 0 BBLS
Gas 805,248 MCF
Water 4,543 BBLS

PM 1-13E
50000

40000
Gas Prod (MCF)

30000

20000

10000

0
may

may

may
july

july
mar

mar

mar
May
jan

Sept

jan

jan
Mar

Jul
sept

sept
Jan
nov

nov

nov

nov

1999 2000 2001 2002 2003


Date

The table shows cumulative oil, gas and water production for the Powder Mountain 1-
13E well. Data was acquired starting from November 1999 till May 2003. The graph
below this shows the monthly production for gas, for this time. Production is out of the
Lewis Shale.
209

Cepo Lewis 21-18:

Cumulative Production
From Aug 1999 To Jun 2003
Oil 257 BBLS
Gas 3,508,896 MCF
Water 670 BBLS

Cepo Lewis 21-18


300000

250000
Gas Prod (MCF)

200000

150000

100000

50000

0
jan

jan

jan

jan
may

may

may

may
Aug

nov

jul

nov

jul

nov

jul

nov
mar

mar

mar

mar
sept

sept

sept
1999 2000 2001 2002 2003

The table shows cumulative oil, gas and water production for the Cepo Lewis 21-18 well.
Data was acquired starting from August 1999 till June 2003. The graph below this shows
the monthly production for gas, for this time. Production is out of the Lewis Shale.
210

Polar Bar #1

Cumulative Production
From Nov 1993 To Nov 1999
Oil 0 BBLS
Gas 142,624 MCF
Water 1,621 BBLS

Polar Bar #1
10000
Cum Gas Prod (MCF)

8000

6000

4000

2000

0
nov

sept

sept

sept

sept

sept

sept
jan

may

jan

may

jan

may

jan

may

jan

may

jan

may
1993 1994 1995 1996 1997 1998 1999
Year

The table shows cumulative oil, gas and water production for the Polar Bar #1 well. Data
was acquired starting from November 1993 till November 1999. The graph below this
shows the monthly production for gas, for this time. Production is out of the Lewis
Shale.
211

Powder Mountain 34-26X

Cumulative Production
From Jun 1999 To May 2003
Oil 444 BBLS
Gas 3,298,309 MCF
Water 7,308 BBLS

PM 34-26x - Fox Hills

140000

120000

100000
Gas Prod (MCF)

80000

60000

40000

20000

0
may

may

may

may
nov

nov

nov

nov
jan

jan

jan

jan
mar

mar

mar

mar
sept

sept

sept

sept
jul

jul

jul

jul
1999 2000 2001 2002 2003

The table shows cumulative oil, gas and water production for the Powder Mountain 34-
26x well. Data was acquired starting from August 1999 till June 2003. The graph below
this shows the monthly production for gas, for this time. Production is out of the Fox
Hills horizon.
212

Powder Mountain 34-26

Cumulative Production
From Jun 1999 To May 2003
Oil 444 BBLS
Gas 3,298,309 MCF
Water 7,308 BBLS

PM 34-26x - Fox Hills

140000

120000

100000
Gas Prod (MCF)

80000

60000

40000

20000

0
may

may

may

may
mar

mar

mar

mar
jul

jul

jul

jul
nov

nov

nov

nov
sept

sept

sept

sept
jan

jan

jan

jan
1999 2000 2001 2002 2003

The table shows cumulative oil, gas and water production for the Powder Mountain 34-
26X well. Data was acquired starting from June 1999 till May 2003. The graph below
this shows the monthly production for gas, for this time. Production is out of the Fox
Hills horizon.
213

Powder Mountain 23-36

Cumulative Production
From Feb 2001 To May 2003
Oil 0 BBLS
Gas 165,145 MCF
Water 849 BBLS

PM 23-36 - Lewis Shale

12000

10000

8000
Gas Prod (MCF)

6000

4000

2000

0
mar

mar

mar
sept

sept
may

may

may
jan

jan
jul

jul
nov

nov

2001 2002 2003

The table shows cumulative oil, gas and water production for the Powder Mountain 23-36
well. Data was acquired starting from February 2001 till May 2003. The graph below
this shows the monthly production for gas, for this time. Production is out of the Lewis
shale.
214

Powder Mountain 34-11

CONFIDENTIAL WELL
NO PRODUCTION DATA AVAILABLE
215

Triton # 10
Cumulative Production
From Aug 1982 To Jun 2003
Oil 19 BBLS
Gas 1,226,056 MCF
Water 125 BBLS
Triton #10 - Lewis Shale

40000

35000

30000
Gas Prod (MCF)

25000

20000

15000

10000

5000

0
1982
1983

1984

1985

1986

1987

1988

1989

1990

1991

1992

1993

1994

1995

1996

1997

1998

1999

2000

2001

2002

2003
Triton #10 - Lew is Shale

8000
7000

6000
Gas Prod (MCF)

5000
4000
3000

2000
1000
0
1999 2000 2001 2002 2003

The table shows cumulative oil, gas and water production for the Triton #10 well. Data was
acquired starting from August 1982 till June 2003. The top graph shows the monthly production
for gas, for this time. The second graph illustrates the monthly production from January 1999 till
June 2003, as a comparison with other wells in the area. Production is out of the Lewis Shale.
216

Well Completions:

API # 49-037-24237
Lease Name: POWDER MOUNTAIN UNIT 1-13E
Location: SE NE 13 TOWNSHIP 14 NORTH RANGE 96 WEST, LONGITUDE
108.14667, LATITUDE 41.18806
Spud Date: 08/21/1999
Operator: STONE ENERGY LLC
Elevation GR: 6,435 ft
Total Depth: 13,503 ft
Plug Back: 13,459 ft
Status: FL
Source of Status: FORM 2, AS OF 03/2003

Top Perfs (ft) Bottom Perfs (ft) Holes/Ft Purpose


13,333 13,367 6 PRODUCTION

Top Interval Bottom


Treatment
(ft) Interval (ft)

269 BBLS GELLED WATER, 16440# 100 MESH


13,333 13,367
SAND & 300400# 20/40 CARBO-PROP

Reservoir Class: G Completion Date: 11/17/1999 Date First Production: 11/17/1999

Producing Formation Depth (ft)


LEWIS CEPO SAND 13,286

Logs:
NEUTRON-C-LITHO-DENSITY
PLATFORM EXPRESS ARRAY INDUCTION W/LINEAR CORRELATION GR
PLATFORM EXPRESS NEUUTRON TRIPLE DETECTOR DENSITY
BOREHOLE COMPENSATED SONIC
DUAL INDUCTION - GR
CEMENT BOND GAMMA RAY & COLLARS
MUD
217

API # 49-037-24185
Lease name: CEPO LEWIS 21-18
Location: L-4 18 TOWNSHIP 14 NORTH RANGE 95 WEST; LONGITUDE 108.14139
LATITUDE 41.18193
Spud Date: 04/15/1999
Operator: EOG RESOURCES INC
Elevation GR: 6,416 ft
Total depth: 13,500 ft
Status: FL
Source of status: FORM 2, AS OF 06/2003
Completion Date 08/22/1999

Geological Markers Depth (ft)


LANCE 10,458
FOX HILLS 12,166
LEWIS 12,353

Logs:
NEUTRON-C-TRIPLE DETECTOR DENSITY
MUD
BHC SONIC
CEMENT BOND COLLAR LOCATOR GAMMA RAY
DIPOLE SONIC GAMMA RAY COLLAR LOCATOR
MICROLOG
DUAL INDUCTION - ARRAY INDUCTION
218

API # 49-037-23037
Lease Name: POLAR BAR 1
Location: NW SE 22 TOWNSHIP 14 NORTH RANGE 96 WEST, LONGITUDE
108.1908, LATITUDE 41.16736
Spud date: 09/30/1992
Completion Date: 10/28/1999
Operator: STONE ENERGY LLC
Elevation GR: 6,544 ft
Total depth: 15,525
Plug back: 14,970
Status: FL
Source of status: FORM 4
Notice of abandoment: 04/19/2002

Casing Size Casing Depth (ft)


7 5/8 13,498
5 1/2 12,983

Top Interval (ft) Bottom Interval (ft) Measured Producing Interval

12,367 12,869 MD LANCE

Completions:
IP Oil Bbls 0 IP Gas Mcf 453 IP Water Bbls 0
Reservoir Class G Date First Production 11/01/1999 TD Formation ALMOND

Logs available:
CEMENT BOND
LITHODENSITY/COMPENSATED NEUTRON/GAMMA RAY
SONIC STC
BHC SONIC GAMMA RAY
DIGITAL SONIC DELTA T AND SPHI
CEMENT BOND

Top Perfs (ft) Bottom Perfs (ft)


14,064 14,164
219

PM 32-26 CONFIDENTIAL WELL

API # 49-037-24126 Y

Permit to Drill Filed By BASIN EXPLORATION INC

Lease Name: POWDER MOUNTAIN FED # 32-26

Location: SW NE Sec 26 T14N R96W

Approved 10/14/1999
220

API # 49-037-24178
Lease name: POWDER MOUNTAIN 34-26X
Location: SW SE 26 TOWNSHIP 14 NORTH RANGE 96 WEST, LONGITUDE
108.17278, LATITUDE 41.15111
Spud date: 02/27/1999
Completion Date: 06/09/1999
Operator: STONE ENERGY LLC
Elevation GR 6,618 ft
Total Depth: 13,183 ft
Plug Back: 13,147 ft
Status: FL
Source of Status: FORM 2, AS OF 03/2003

Hole Size (in) Casing Size (in) Casing Depth (ft) Weight Cement
12 1/4 9 5/8 3,385 36 1185
6 1/8 4 1/2 13,009 13.5 350
8 3/4 7 10,569 23 1050

Top Interval (ft) Bottom Interval (ft) Measured Producing Formation


13,096 13,116 MD FOX HILLS

Completions:
IP Oil Bbls 0 IP Gas Mcf 3,857 IP Water Bbls 2
Reservoir Class G Date First Production 06/09/1999 TD Formation FOX HILLS

Geological Markers Depth (ft)


LANCE SANDSTONE 11,497
MIDDLE LANCE SHALE 12,178
MIDDLE LANCE SAND 12,259
JONAH SHALE 12,372
221

JONAH SAND 12,455


YELLOW POINT SHALE 12,562
YELLOW POINT SAND 12,609
WARDELL SHALE 12,744
WARDELL SAND 12,802
FOX HILLS SAND 13,045

Logs available:
MUD
HOSTILE ENVIRONMENT SPECTRAL DENSITY DUAL SPACED NEUTRON
HOSTILE ENVIRONMENT FULLWAVE SONIC
HOSTILE ENVIRONMENT DUAL INDUCTION
SPECTRAL DENSITY DUAL SPACED NEUTRON
HIGH RESOLUTION INDUCTION
THERMAL MULTIGATE DECAY LITHOLOGY
CEMENT BOND
HOSTILE ENVIRONMENT CEMENT BOND

Perfs:
Top Perfs (ft) Bottom Perfs (ft) Holes/Ft Purpose
13,096 13,116 2 PRODUCTION
222

API # 49-037-24354
Lease name: POWDER MOUNTAIN UNIT 23-36
Location: NE SW 36 TOWNSHIP 14 NORTH RANGE 96 WEST, LONGITUDE
108.15888, LATITUDE 41.14000
Spud date: 08/18/2000
Completion Date 02/14/2001
Operator: STONE ENERGY LLC
Elevation GR 6,575
Status: FL
Source of status: FORM 2, AS OF 03/2003
Total depth: 14,300 ft
Plug back: 13,854 ft

Hole Size (in) Casing Size (in) Casing Depth (ft) Weight Cement
12 1/4 9 5/8 3,487 36 1225
8 3/4 7 12,710 26/29 1770

Top Interval (ft) Bottom Interval (ft) Measured Producing Formation


13,788 13,837 MD LEWIS

Top Interval Bottom


Treatment
(ft) Interval (ft)
Pumped 131 BBL PAD. PAD W/19488 GALS & 8000# 100
13,827 18,837 MESH SAND. TREATED W/33558 GALS & 87000# 20/40
Carbo HSP.

Completions:
IP Oil Bbls 0 IP Gas Mcf 526 IP Water Bbls 0
Reservoir Class G Date First Production 02/26/2001 TD Formation LEWIS

Geological Markers Depth (ft)


FORT UNION 9,821
LANCE UPPER 10,302
LANCE MIDDLE 11,418
223

LANCE LOWER 11,934


FOX HILLS 12,792
LEWIS 12,953
LEWIS SAND 13,750

Logs available:
HIGH RESOLUTION INDUCTION
HIGH RESOLUTION INDUCTION RUN 2
MICRO-ELECTRIC
SPECTRAL DENSITY DUAL SPACED NEUTRON
SPECTRAL DENSITY DUAL SPACED NEUTRON RUN 2
FULL WAVE SONIC (DELTA "T")
FULL WAVE SONIC DELTA "T" RUN 2
FULL WAVE SONIC DTC/DTS SEMBLANCE PLOT
MECHANICAL ROCK PROPERTIES 5"=100'

Top Perfs (ft) Bottom Perfs (ft) Holes/Ft Size Purpose

13,788 13,837 4 .225 PRODUCTION


224

API # 49-037-24482
Lease Name: POWDER MOUNTAIN UNIT 34-11
Location: SW SE 11 TOWNSHIP 13 NORTH RANGE 96 WEST, LONGITUDE
108.17333, LATITUDE 41.10750
Spud date: 03/17/2001
Operator: STONE ENERGY LLC
Elevation GR: 6,753 ft
Status: PA 07/04/2002
Source of status: FORM 4
Total depth: 15,105 ft
Plug back: 12,561 ft
Notice of abandonment: 04/19/2002

Hole Size
Casing Size (in) Casing Depth (ft) Weight Cement
(in)
12 ¼ 9 5/8 3,510 36 1600
8 3/4 7 12,969 29 & 26 2850

Top Interval (ft) Bottom Interval (ft) Measured Producing Formation


12,395 12,625 MD LANCE

Reservoir Class G Completion Date 05/28/2001 Formation LANCE

Geological markers Depth (ft)


FORT UNION 10,258
LANCE 12,250
LEWIS 13,546

HIGH RESOLUTION INDUCTION


SPECTRAL DENSITY DUAL SPACED NEUTRON
HIGH RESOLUTION INDUCTION
SPECTRAL DENSITY DUAL SPACED NEUTRON

Top Perfs (ft) Bottom Perfs (ft) Holes/Ft

12,499 12,660 6
225

API # 49-037-21922
Lease Name: TRITON UNIT 10
Location: NW NE 8 TOWNSHIP 13 NORTH RANGE 95 WEST, LONGITUDE
108.11527, LATITUDE 41.11851
Spud date: 02/08/1982
Operator: WESTPORT OIL & GAS COMPANY LP
Elevation GR: 6,600 ft
Elevation KB 6,562 ft
Status: FL
Source of status: FORM 2 AS OF 06/2003
Total depth: 14,975 ft
Plug back: 13,420 ft

Casing Size (in) Casing Depth (ft)


13 3/8 2,530
9 5/8 12,530

Completions:
IP Oil Bbls 0 IP Gas Mcf 949 IP Water Bbls 0
Reservoir Class G Completion Date 08/16/1982 TD Formation ERICSON

Geological markers Depth (ft)


LANCE 9,738
FOX HILLS 12,168
LEWIS 12,454
LEWIS A 12,715
LEWIS B 13,019
LEWIS C 13,258
ALMOND 14,333
ERICSON 14,806

Available logs:
ACOUSTIC CEMENT BOND GAMMA RAY (06/01/1982)
ACOUSTIC CEMENT BOND GAMMA RAY (07/08/1982)
226

ACOUSTIC CEMENT BOND GAMMA RAY (07/30/1982)


BOREHOLE COMPENSATED SONIC
NEUTRON FORMATION DENSITY
CASED RESERVOIR ANALYSIS
DUAL INDUCTION - SFL
DUAL SPACING THERMAL NEUTRON DECAY TIME

Top Perfs (ft) Bottom Perfs (ft)


13,259 13,279
227

API # 49-037-22991
Lease Name: BOGEY DRAW 1
Location: SW 14 TOWNSHIP 13 NORTH RANGE 95 WEST, LONGITUDE
108.06677, LATITUDE 41.09250
Operator: CELSIUS ENERGY COMPANY
Spud date: 09/29/1993
Elevation GR: 6,740 ft
Status: PA 06/08/1993
Source of status: Well Logs
Total depth: 14,204 ft

Casing Size (in) Casing Depth (ft)


7 5/8 13,200

Logs available:
DUAL INDUCTION - SFL - GAMMA RAY W/LINEAR CORRELATION
BHC SONIC GAMMA RAY
BHC SONIC GAMMA RAY (11/16/1993)
BHC SONIC GAMMA RAY (11/16 & 12/05/1993)
NEUTRON-C-LITHO-DENSITY GAMMA RAY (11/16/1993)
NEUTRON-C-LITHO-DENSITY GAMMA RAY (11/16 & 12/3/1993)
F-LOG GAMMA RAY
FORMATION MICRO IMAGER MONITOR GAMMA RAY
ARRAY SONIC GAMMA RAY
FORMATION ANALYSIS
228

API # 49-037-23169
Lease Name: BLACK BAR 1
Location: NE NW 30 TOWNSHIP 13 NORTH RANGE 95 WEST, LONGITUDE
108.14157, LATITUDE 41.07341
Spud date: 07/03/1993
Operator: CELSIUS ENERGY COMPANY
Elevation GR: 6,941 ft
Elevation KB 6,965 ft
Status: PA 11/12/1996
Source of status: WELL FILES
Total depth: 14,373
Plug back: 14,300
Notice of abandonment: 02/07/1994
Subsequent notice of abandonment: 11/12/1996

Casing Size (in) Casing Depth (ft)


7 5/8 13,170

Completions:
IP Oil Bbls 0 IP Gas Mcf 597 Formation DRY
Reservoir Class G TD Formation ALMOND IP Water Bbls 133

Geological Markers Depth (ft)


WASATCH 4,429
FORT UNION 7,225
LANCE 10,916
FOX HILLS 12,028
LEWIS 12,223
ALMOND 14,048

Logs available:
CEMENT BOND GAMMA RAY W/VARIABLE DENSITY
PHASOR INDUCTION GAMMA RAY W/LINEAR CORRELATION
LITHODENSITY/COMPENSATED NEUTRON/GAMMA RAY
MUD

You might also like