Sie sind auf Seite 1von 219

DEPOSITIONAL FACIES AND PETROPHYSICAL ANALYSIS

OF THE BAKKEN FORMATION, PARSHALL FIELD,


MOUNTRAIL COUNTY, NORTH DAKOTA

by
Andrea Simenson
A thesis submitted to the faculty and 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: _______________________
Andrea Simenson

Signed: _______________________
Dr. Stephen A. Sonnenberg
Thesis Advisor

Golden, Colorado

Date ______________

Signed: _______________________
Dr. John D. Humphrey
Professor and Head
Department of Geology
and Geological Engineering

ii
ABSTRACT

The Parshall Field of the Williston Basin was discovered in 2006 by EOG
Resources and is located in Mountrail County, North Dakota. This Devonian-
Mississippian middle Bakken Formation resource play covers some 40 townships in
Mountrail County, North Dakota and is still expanding. The development of horizontal
drilling and modern fracturing techniques has made this field possible.
The Bakken Formation in the field consists of three members: (1) upper shale,
(2) middle dolomitic sandstone, (3) lower shale. The total Bakken interval ranges in
thickness from 70 to 100 feet (21.3 to 30.5 meters). The upper shale is dark-brown to
black organic rich mudstone and averages 18 feet in thickness (5.5 meters). The middle
member ranges in thickness from 30 to 50 feet (9.1 to 15.2 meters). The middle member
is composed of a range in lithologies from bioturbated, argillaceous, calcareous very fine
grained sandstones to low angle planar laminated fine to very fine-grained sandstone to
very fine-grained sandstone and siltstones with centimeter and millimeter scale shale
laminations. The lower shale member is also a dark-brown to black organic rich
mudstone and is approximately 30 feet (9.1 meters) thick in Parshall Field.
The main hydrocarbon productive reservoir in the Parshall Field is the middle
member which has low matrix porosity and permeability and is found at depths of 9,000
to 10,500 feet (2,743 to 3,200 meters). The middle Bakken porosities range from 2 to
12% and permeabilities average 0.02 mD. Some key factors that contribute to the success
of this play include good stratigraphic trapping conditions, generation of hydrocarbons
from the organic rich mudstones, and fractures.
A network of nine stratigraphic and seven structural cross sections was
constructed. Structure cross sections illustrate that the Bakken members dip to the west.
The stratigraphic cross sections illustrate that the Bakken members onlap and thin
towards the east onto the underlying Three Forks Formation.
A petrophysical model was developed to characterize the reservoir rocks at
Parshall Field. Water saturation calculations were done using an Archie saturation
equation and permeability was computed using a generalized Timur equation. These were

iii
then used to determine net pay in Parshall Field which averages 25 feet (7.6 meters) in
thickness.
The upper and lower Bakken were also analyzed for its organic content. Total
organic carbon (TOC) contents were calculated using formation density logs using the
Passey Log R method calibrated to available source rock data. These values were then
used to map the lateral extent of TOC. The TOC values for the upper and lower Bakken
are fairly constant in Parshall Field and decrease towards the west.
Hydrocarbon generation in organic rich facies is interpreted as the cause of
overpressuring in the Bakken which is slightly overpressured at Parshall Field with an
average pressure gradient of 0.60 psi/ft. This was calculated using mud weight data. The
onset of hydrocarbon generation took place in the Cenozoic, and the present-day Bakken
is in the early-mature oil window.
The Parshall Field appears to have a mature-immature boundary forming the
eastern boundary of the oil accumulation. This boundary is likely caused by lack of
migration conduits and pressure from hydrocarbon generation.
The productivity of Parshall Field has increased dramatically since its discovery
in 2006. This study looks at the reservoir characteristics, source rock maturity, and
develops petrophysical model to help better understand Parshall Field.

iv
TABLE OF CONTENTS
ABSTRACT....................................................................................................................... iii
LIST OF FIGURES ......................................................................................................... viii
LIST OF TABLES .............................................................................................................xx
ACKNOWLEDGEMENTS ............................................................................................. xxi
CHAPTER 1 INTRODUCTION .......................................................................................1
1.1 Location of Study Area ................................................................................4
1.2 Research Objectives and Methods ...............................................................4
1.3 Research Contribution .................................................................................6
1.4 Previous Work .............................................................................................6
CHAPTER 2 GEOLOGIC SETTING .............................................................................11
2.1 Regional Stratigraphy and Sedimentology ................................................15
2.1.1 Devonian Rocks .............................................................................19
2.1.2 Upper Devonian Mississippian Rocks ........................................23
2.1.3 Mississippian Rocks.......................................................................27
2.2 Regional Structure .....................................................................................29
2.3 Geologic Overview of Parshall Field .........................................................31
2.4 History of Oil Production in the Williston Basin.......................................32
2.5 Horizontal Drilling Plays in the Bakken Formation ..................................34
2.6 Petroleum System Overview of the Williston Basin .................................37
2.7 Migration....................................................................................................39
CHAPTER 3 FACIES INTERPRETATION AND CORE DESCRIPTIONS ................43
3.1 Methods......................................................................................................43
3.2 Facies Descriptions ....................................................................................46
3.2.1 Facies G. Organic Rich Pyritic Brown/Black Mudstone ...............47
3.2.2 Facies A. Muddy Lime Wackestone ..............................................52
3.2.3 Facies B. Bioturbated, Argillaceous, Calcareous, Very Fine-
Grained Siltstone/Sandstone ..........................................................52
3.2.4 Facies C1. Planar to Undulose Laminated, Shaly, Very Fine-
Grained Siltstone/Sandstone ..........................................................56

v
3.2.5 Facies C2. Symmetrically Ripple to Undulose Laminated, Very
Fine-Grained Siltstone/Sandstone..................................................56
3.2.6 Facies D1. Contorted to Massive, Fine-Grained Sandstone ..........56
3.2.7 Facies D2. Low Angle, Planar to Slightly Undulose, Cross-
Laminated Sandstone with Thin Discontinuous Shale
Laminations....................................................................................59
3.2.8 Facies E1. Finely Inter-Laminated, Bioturbated, Dolomitic-
Mudstone and Dolomitic Siltstone/Sandstone ...............................59
3.2.9 Facies E2. Calcitic, Whole Fossil, Dolomitic-to Lime
Wackestone ....................................................................................62
3.2.10 Facies F. Bioturbated, Shaly, Dolomitic Siltstone .........................62
3.3 Core Descriptions Interpretations ...........................................................62
3.3.1 Van Hook #1-13H (Sec. 13-152N-91W) .......................................66
3.3.2 N & D #1-05H (Sec. 5-152N-90W)...............................................68
3.3.3 Parshall #2-36H (Sec. 36-153N-90W) ...........................................70
3.3.4 Hoff #1-10H (Sec. 10-152N-90W) ................................................70
3.3.5 Long #1-01H (Sec. 1-152N-90W) .................................................73
3.3.6 Patten #1-27H (Sec. 27-153N-89W)..............................................73
3.3.7 Fertile #1-12H (Sec. 12-151N-90W) .............................................76
3.3.8 Jensen #12-44 (Sec. 12-154N-90W) ..............................................78
3.3.9 Deadwood Canyon Ranch #43-28H (Sec. 28-154N-92W)............78
3.3.10 Dobrinski #18-44 (Sec. 18-151N-87W) ........................................82
3.4 Mineralogy and Diagenesis........................................................................84
3.4.1 Diagenetic Features ........................................................................87
3.4.2 Reservoir Quality ...........................................................................87
CHAPTER 4 SUBSURFACE WELL LOG INTERPRETATION AND
DEPOSITIONAL MODEL .......................................................................89
4.1 Methods......................................................................................................89
4.2 Construction of the Cross Sections ............................................................92
4.2.1 Structural Cross Sections ...............................................................92
4.2.2 Stratigraphic Cross Sections ..........................................................92

vi
4.3 Structure Maps .........................................................................................111
4.4 Isopach Maps ...........................................................................................111
4.5 Depositional Model ..................................................................................128
CHAPTER 5 PETROPHYSICAL AND RESERVOIR PROPERTIES
ANALYSIS ..............................................................................................133
5.1 Core Properties Analysis..........................................................................134
5.2 Conventional Well Log Analysis .............................................................152
5.2.1 Data Preperation...........................................................................158
5.2.2 Petrophysical Model ....................................................................159
5.3 Source Rock Characterization..................................................................164
5.3.1 Source Rock Richness Estimation from Formation
Density Log ..................................................................................164
5.3.2 Source Rock Richness Estimation Using the log R
Method .........................................................................................172
5.3.3 Pyrolysis Methods ........................................................................177
5.3.4 Time Temperature Index Method ................................................188
CHAPTER 6 CONCLUSIONS AND RECOMMENDATIONS ..................................191
6.1 Conclusions ..............................................................................................191
6.2 Recommendations ....................................................................................192
REFERENCES CITED ....................................................................................................193

vii
LIST OF FIGURES
Figure 1.1 Index map showing the location of the Williston Basin and the study area
in northwestern North Dakota. Modified from Pitman et al. (2001). ....................2

Figure 1.2 Type log for the Parshall Field showing the overlying Lodgepole
Formation, the Bakken Formation, and the underlying Three Forks Formation. ...3

Figure 1.3 Location of the study area in North Dakota. The Parshall Field is located in
the southern part of Mountrail County in Townships 151-155N and Ranges
88-91W. Modified from LeFever (2008). ...............................................................5

Figure 1.4 Lithologic column describing the lithofacies of the middle member of the
Bakken Formation (modified from LeFever et al., 1991). From Pitman et al.
(2001). .....................................................................................................................8

Figure 2.1 Paleogeographic map of the Late Devonian (360 Ma). Subduction zones
and associated orogenic belts (dashed red lines) surround and separated the
North American craton from the Devonian ocean. The North American craton
was open to the southwest indicated by the yellow arrow. The Williston Basin
and Alberta Basin were connected at that time (W-A) (modified from
Blakey, 2005). .......................................................................................................12

Figure 2.2 Paleogeographic map of the Early Mississippian (345 Ma). The Alberta
Basin (AB) and the Williston Basin (WB) were separated by the Sweetgrass
Arch (SA). The Williston Basin was shielded from the Devonian ocean by the
orogenic belts (dashed red lines) that surrounded the North American craton
and the Transcontinental Arch (TA) to the southeast (modified from
Blakey, 2005). .......................................................................................................13

Figure 2.3 Major structures of the Williston Basin (Gerhard et al., 1990). ......................14

Figure 2.4 Basement tectonic provinces in the Williston Basin (red dashed line)
(Fisher et al., 2005). ..............................................................................................16

Figure 2.5 Generalized stratigraphic column for the Williston Basin, showing units
that produce oil and gas (LeFever, 1992). ............................................................17

Figure 2.6 Generalized southwest northeast structural and stratigraphic cross


section A A from the Bighorn Mountains in Wyoming to the northeastern
side of North Dakota. The Bakken Formation is highlighted in red. The index
map shows the cross section line in red (from Peterson and
MacCary, 1987). ...................................................................................................18

viii
Figure 2.7 Diagram showing basin communication and sedimentation patterns through
time. Areas with arrows indicate open-marine communication: (A) Tippecanoe
sequence, (B) lower Kaskaskia sequence, (C) upper Kaskaskia sequence, (D)
Absaroka sequence (Gerhard et al., 1987). ...........................................................20

Figure 2.8 Thickness distribution map showing the Devonian rocks of the Williston
Basin. Approximated limits of the Prairie salt and source rocks of the Bakken
Formation (Late Devonian Early Mississippian) are shown (Peterson and
MacCary, 1987). ...................................................................................................21

Figure 2.9 Well logs and interpreted lithology of the Tipperary Oil and Gas
Corporation Olsen No. 1 well, SE SW Sec. 26, T160N, R97W, Divide County
(Webster, 1984). ....................................................................................................24

Figure 2.10 Approximate subsurfact limits of the three members of the Bakken
Formation. The northwest to southeast stratigraphic cross section A B shows
the regional extent of the Bakken Formation in North Dakota
(Webster, 1984). ....................................................................................................26

Figure 2.11 Thickness distribution map and generalized rock facies of the Madison
Group (Peterson and MacCary, 1987). .................................................................28

Figure 2.12 Thickness distribution map and generalized rock facies of the Big Snowy
Group (Peterson and MacCary, 1987). .................................................................28

Figure 2.13 Predominant structures of the Williston Basin. Major structural features
include: Antelope Anticline (AT), Billings Anticline (B), Birdtail-Waskada
axis (BW), Cedar Creek Anticline (CC), Hummingbird Synclinorium (HS),
Little Knife Anticline (LK), Nesson Anticline (NS), Poplar Dome (PD),
Roncott High (RH), Sheep Mountain Synclinorium (SM), Weldon-Brockton-
Froid Fault Zone (WBF), Western Nesson Anticline (WN)
(LeFever, 1992). ....................................................................................................30

Figure 2.14 Map showing the expansion of Bakken exploration to the north and west
of Parshall Field. This was the current activity at the end of 2008 (Modified
from Johnson, 2009). ............................................................................................33

Figure 2.15 Map showing the distribution of oil and gas fields in the Williston Basin.
Oil fields are shown in green and gas fields are shown in red
(Longman, 1987). ..................................................................................................35

Figure 2.16 Burial history for four different wells (A= California Oil No. 1 Rough
Creek, B= Texaco No. 1 Clarence Pederson, C= Marathon Oil No. 18-44
Dobrinski, D= Placid Oil No. 36-5 Rosendahl). The upward deflections
represent possible loss of section from major unconformities
(Webster, 1984). ....................................................................................................38
ix
Figure 2.17 Map showing the major tectonic left-lateral shear sets in the Williston
Basin (Gerhard et al., 1990). .................................................................................40

Figure 2.18 Interpretation of the left-lateral shear sets origin shown by an elliptical
strain diagram in the central Rocky Mountains and Williston Basin block
(Gerhard et al., 1990). ...........................................................................................40

Figure 2.19 Map showing the theoretical migration pathways from the major source
rocks to the producing reservoirs in the Williston Basin. The areas of low
migration potential are shown in gray (Gerhard et al., 1990). ..............................41

Figure 3.1 Map showing the location for the ten cores described in this study. Eight
cores: Van Hook #1-13H, N & D #1-05H, Parshall #2-36H, Hoff #1-10H,
Long #1-01H, Patten #1-27H, Fertile #1-12H, and Jensen #12-44 are located
in the Parshall Field, Mountrail County, North Dakota. The Deadwood Canyon
Ranch #43-28H core is from the Sanish Field and the Dobrinski #18-44 core is
from a wildcat well east of Parshall. .....................................................................45

Figure 3.2 Diagram illustrating representative marine ichnofacies. Typical trace fossils
include: 1) Caulostrepsis; 2) Entobia; 3)equinoid borings; 4) Trypanites; 5)
Teredolites; 6) Thalassinoides; 7,8) Gastrochaenolites; 9) Diplocraterion; 10)
Skolithos; 11,12) Psilonichnus; 13) Macanopsis; 15) Diplocraterion; 16)
Arenicolites; 17) Ophiomorpha; 18) Phycodes; 19) Rhizocorallium; 20)
Teichichnus; 21) Planolites; 22) Asteriacites; 23) Zoophycos; 24) Lorenzinia;
25) Zoophycos; 26) Paleodyction; 27) Taphrhelminthopsis; 28) Helminthoida;
29) Cosmorhaphe; 30) Spirorhaphe (Sutter, 2006). .............................................48

Figure 3.3 Core pictures: G, A, B, C, D, E, and F, represent the main facies identified
from Bakken cores available in and near the Parshall Field. G. Organic rich
pyritic brown/black mudstone. A. Muddy lime wackestone. B. Bioturbated,
argillaceous, calcareous, siltstone/sandstone. C. Planar to undulose laminated,
argillaceous siltstone/sandstone. D. Low angle planar to undulose sandstone.
E. Finely inter-laminated, bioturbated, dolomitic siltstone/sandstone. F.
Bioturbated, shaly, dolomitic siltstone. .................................................................49

Figure 3.4 Facies of the Bakken Formation observed from core samples. Facies E is
subdivided into E1 and E2, based on sedimentary features, and is combined
with facies F for log correlations. Facies D is subdivided into D1 and D2 based
on sedimentary structures. Facies C is subdivided into C1 and C2 based on
sedimentary structure differences. The log and core pictures were taken from
the Deadwood Canyon Ranch #43-28H well (core available by permission
from Fidelity Exploration and Production Company). .........................................50

x
Figure 3.5 Core photos of facies G organic rich pyritic brown/black mudstone.
Photos A and B are from the Upper Bakken Shale and photos C and D are
from the Lower Bakken Shale. Small vertical fractures cemented by calcite
and pyrite are visible in photo C. ..........................................................................51

Figure 3.6 Core photos of facies A muddy lime wackestone. Crinoid and brachiopod
shell fragments are more concentrated at the base of the facies. This lowermost
facies of the middle Bakken lies directly above the contact with the lower
Bakken shale as seen in photos A and C. .............................................................52

Figure 3.7 Core photos of facies B bioturbated, argillaceous, calcareous,


siltstone/sandstone. This is the thickest interval in the middle Bakken and is
present in all of the cores described. .....................................................................54

Figure 3.8 Core photo showing Helminthopsis/Sclarituba burrow traces found in


facies B. .................................................................................................................55

Figure 3.9 Core photos of facies C1 planar to undulose laminated, shaly, very fine
grained siltstone/sandstone. This interval is dominated by continuous planar
bedding but there also are sections of wavy laminations. .....................................56

Figure 3.10 Core photos of facies C2 symmetrically ripple to undulose laminated,


very fine-grained siltstone/sandstone. This facies is only present in two cores
described with an average thickness of 3 feet (0.9 meters). .................................58

Figure 3.11 Core photos of facies D1 contorted to massive fine-grained sandstone.


This facies has common micro-faults, microfractures, and slumps representing
soft sediment deformation. ....................................................................................58

Figure 3.12 Core photos of facies D2 low angle, planar to slightly undulose, cross-
laminated sandstone with thin discontinuous shale laminations. Photo C shows
the more massive limestone unit found in the Dobrinski #18-44 core east of
Parshall Field. .......................................................................................................60

Figure 3.13 Core photos of facies E1 finely inter-laminated, bioturbated, dolomitic


mudstone and dolomitic siltstone/sandstone. This facies is present in all of the
cores described. .....................................................................................................61

Figure 3.14 Core photos of facies E2 calcitic, whole fossil, dolomitic-to lime
wackestone. These fossil-rich beds contain brachiopod shells and crinoid
fragments and may represent storm deposits. .......................................................62

Figure 3.15 Core photos of facies F bioturbated, shaly, dolomitic siltstone. This
uppermost facies of the middle Bakken lies directly below the contact with
the upper Bakken shale as seen in photos C and D. ..............................................64

xi
Figure 3.16 Core description legend for Figures 3.17 3.26. ..........................................65

Figure 3.17 Core description of the Van Hook #1-13H. ...................................................67

Figure 3.18 Core description of the N & D #1-05H. ........................................................69

Figure 3.19 Core description of the Parshall #2-36H. ......................................................71

Figure 3.20 Core description of the Hoff #1-10H. This horizontal well was described
from left to right. The red box indicates the location of the core with respect to
the well path taken. The 40 feet (12.2 meters) of core described represents
approximately 5 feet (1.5 meters) of vertical succession. .....................................72

Figure 3.21 Core description of the Long #1-01H. ...........................................................74

Figure 3.22 Core description of the Patten #1-27H. .........................................................75

Figure 3.23 Core description of the Fertile #1-12H. .........................................................77

Figure 3.24 Core description of the Jensen #12-44. .........................................................79

Figure 3.25a Description of the uppermost part of the Deadwood Canyon


Ranch #43-28H. ....................................................................................................80

Figure 3.25b Description of the lowermost part of the Deadwood Canyon


Ranch #43-28H. ....................................................................................................81

Figure 3.26 Core description of the Dobrinski #18-44. ....................................................83

Figure 3.27 Mineralogical analysis measured from QEMSCAN for the Deadwood
Canyon Ranch #43-28H core samples, Mountrail County, North Dakota. ..........85

Figure 3.28 QEMSCAN image of the core plug taken from the Deadwood Canyon
Ranch #43-28H at 10,134.02 feet. The pyrite-filled fractures are shown in
yellow. ...................................................................................................................86

Figure 3.29 Chart showing the postdepositional events observed in sandstones and
siltstones of the middle member of the Bakken Formation (from Pitman
et al., 2001). ..........................................................................................................88

Figure 3.30 Burial and thermal history curve of the Bakken Formation showing
relative timing of major diagenetic events (Pitman et al., 2001). The term
ogs represents overgrowths. ..............................................................................88

xii
Figure 4.1 All wells in the study area. Rasters used are highlighted in pink. Parshall
Field is the outline on the right and Sanish Field is just to the east. The field
outlines were taken from the North Dakota Industrial Commission,
Department of Mineral Resources, Oil and Gas Division website. ......................90

Figure 4.2 The Deadwood Canyon Ranch #43-28H well from Sanish Field is the
type well used for correlation. Facies E and F are combined into one interval,
facies D includes facies D1 and D2, and facies includes facies C1 and C2. ........91

Figure 4.3 Structure section lines shown in red. Base map is the structure on top of the
upper Bakken shale. Parshall Field (to the east) and Sanish Field (to the west)
are outlined in black. .............................................................................................93

Figure 4.4 West-east oriented structural cross section 1-1 showing structural dip
toward the west. This section is north of Parshall Field. Location of the cross
section is shown in Figure 4.3. ..............................................................................94

Figure 4.5 West-east oriented structural cross section 2-2 showing structural dip
toward the west. This section crosses the northern part of Parshall Field and
northern part of Sanish Field. Location of the cross section is shown in
Figure 4.3. .............................................................................................................95

Figure 4.6 West-east oriented structural cross section 3-3 showing structural dip
toward the west. This section crosses the middle of Parshall Field and the
southern part of Sanish Field. Location of the cross section is shown in
Figure 4.3. .............................................................................................................96

Figure 4.7 West-east oriented structural cross section 4-4 showing structural dip
toward the west. This section crosses the southern part of Parshall Field.
Location of the cross section is shown in Figure 4.3. ...........................................97

Figure 4.8 North-south oriented structural cross section 5-5 drawn generally along
structural strike (the section has a high vertical exaggeration). This section is
west of Parshall Field and crosses through the western side of Sanish Field.
Location of the cross section is shown in Figure 4.3. ...........................................98

Figure 4.9 North-south oriented structural cross section 6-6 drawn generally along
structural strike (this section has a high vertical exaggeration). This section
runs through the center of Parshall Field and is almost parallel to the
structural contour lines of the upper Bakken. Location of the cross section is
shown in Figure 4.3. ..............................................................................................99

Figure 4.10 North-south oriented structural cross section 7-7 drawn generally along
structural strike (the section has a high vertical exaggeration). This section is
east of Parshall Field. Location of the cross section is shown in
Figure 4.3. ...........................................................................................................100
xiii
Figure 4.11 Map showing the location of nine stratigraphic sections. Base map is the
Middle Bakken thickness map (in feet). Parshall Field (to the east) and
Sanish Field (to the west) are outlined in black. .................................................101

Figure 4.12 West-east oriented stratigraphic cross section A-A' showing the variations
in the middle Bakken thickness as it thins to the east. Facies C is completely
absent in the east and facies D is not present in some of the wells. This section
is to the north of Parshall Field. Location of the cross section is shown in
Figure 4.11. .........................................................................................................102

Figure 4.13 West-east oriented stratigraphic cross section B-B' showing the variations
in the middle Bakken thickness as it thins to the east. Facies C is completely
absent and facies D is not present in some of the wells in Parshall Field.
Location of the cross section is shown in Figure 4.11. .......................................103

Figure 4.14 West-east oriented stratigraphic cross section C-C' showing the variations
in the middle Bakken thickness as it thins to the east. Facies C and facies D is
only in one of the Parshall Field wells shown here. Location of the cross
section is shown in Figure 4.11. ..........................................................................104

Figure 4.15 West-east oriented stratigraphic cross section D-D' showing the variations
in the middle Bakken thickness as it thins to the east. Facies C and facies D are
present west of the southern part of Parshall Field. Facies D is present in the
Dobrinski #18-44 well east of Parshall Field. Location of the cross section is
shown in Figure 4.11. ..........................................................................................105

Figure 4.16 North-south oriented stratigraphic cross section E-E' showing the
variations in the middle Bakken thickness as it thins to the south. Facies C
and facies D are present in all of these wells west of Parshall Field. Location
of the cross section is shown in Figure 4.11. ......................................................106

Figure 4.17 North-south oriented stratigraphic cross section F-F' showing the
variations in the middle Bakken thickness as it thins slightly to the south.
Facies D is only present in the Laredo #26-1 well north of Parshall Field.
This cross section runs through the center of Parshall Field. Location of the
cross section is shown in Figure 4.11. ................................................................107

Figure 4.18 North-south oriented stratigraphic cross section G-G' showing the
variations in the middle Bakken thickness. The middle Bakken thickens
slightly and then thins to the south. This cross section is east of Parshall
Field and shows an increase in Facies B to the south. Location of the cross
section is shown in Figure 4.11. ..........................................................................108

xiv
Figure 4.19 Northwest-southeast oriented stratigraphic cross section H-H' showing
the variations in the middle Bakken thickness as it thins to the southeast.
This section crosses the eastern part of Sanish Field and the center of Parshall
Field. Facies C pinches out into Parshall Field and is only present in the
Parshall SD #1 well which also has facies D. The lower Bakken is also
thinning to the southeast. Location of the cross section is shown in
Figure 4.11. .........................................................................................................109

Figure 4.20 Southwest-northeast oriented stratigraphic cross section I-I' showing the
variations in the middle Bakken thickness as it thins to the northeast. This
section crosses the center of Parshall Field and just south of Sanish Field.
Facies C pinches out into Parshall Field and is only present in the Van Hook
#1-13H well and the Parshall SD #1 well which also has facies D. The lower
Bakken is also thinning to the northeast. Location of the cross section is
shown in Figure 4.11. ..........................................................................................110

Figure 4.21 Structure map on top of the upper Bakken shale. The contour interval
is 250 feet (80 meters). ........................................................................................112

Figure 4.22 Structure map on top of the middle member of the Bakken Formation.
The contour interval is 250 feet (80 meters). ......................................................113

Figure 4.23 Structure map on top of the lower Bakken shale. The contour interval
is 250 feet (80 meters). ........................................................................................114

Figure 4.24 Structure map on top of the Three Forks Formation. The contour interval
is 250 feet (80 meters). ........................................................................................115

Figure 4.25 Structure map on top of the Nisku (Birdbear) Formation. The contour
interval is 250 feet (80 meters). ..........................................................................116

Figure 4.26 Structure map on top of the Prairie Formation. The contour interval is
250 feet (80 meters). ...........................................................................................117

Figure 4.27 Thickness map of the total Bakken Formation. Contour interval is 10 feet
(3 meters). The thickness of the Bakken Formation ranges from 70 feet 100
feet (21.3 meters 30.5 meters) in Parshall Field. .............................................118

Figure 4.28 Thickness map of the upper Bakken shale (facies G). Contour interval
is 3 feet (0.9 meters). The thickness of the upper Bakken member averages
18 feet (5.5 meters) in Parshall Field. .................................................................120

Figure 4.29 Thickness map of the middle member (facies F, E, D, C, B, and A) of the
Bakken Formation. Contour interval is 10 feet (3 meters). The thickness of
the upper Bakken member ranges from 30 feet 50 feet (9.1 meters 15.2
meters) in Parshall Field. ....................................................................................121
xv
Figure 4.30 Thickness map of facies E and F of the middle member of the Bakken
Formation. Contour interval is 5 feet (1.5 meters). The thickness of facies E
and F averages 12 feet (3.7 meters) in Parshall Field. ........................................122

Figure 4.31 Thickness map of facies D of the middle member of the Bakken
Formation. Contour interval is 2 feet (0.6 meters). Facies D is mainly absent
in Parshall Field. .................................................................................................123

Figure 4.32 Thickness map of facies C of the middle member of the Bakken
Formation. Contour interval is 2 feet (0.6 meters). Facies C pinches out in
Parshall Field. .....................................................................................................124

Figure 4.33 Thickness map of facies A and B of the middle member of the Bakken
Formation. Contour interval is 5 feet (1.5 meters). The combined thickness
Of Facies A and B ranges from 17 feet 25 feet (5.2 meters 7.6 meters) in
Parshall Field. .....................................................................................................125

Figure 4.34 Thickness map of the lower Bakken shale (facies G). Contour interval is
5 feet (1.5 meters). The thickness of the lower Bakken member ranges from
20 feet 37 feet (6.1 meters 11.3 meters) in Parshall Field. ...........................126

Figure 4.35 Thickness map of the Three Forks Formation. Contour interval is 10 feet
(3 meters). The thickness of the Three Forks Formation ranges from 222 feet
242 feet (67.7 meters 73.8 meters) in Parshall Field. ...................................127

Figure 4.36 Thickness map from the top of the middle member of the Bakken
Formation to the top of the Prairie Formation. Contour interval is 50 feet
(15.2 meters). ......................................................................................................129

Figure 4.37 Environment of deposition interpretation of the middle Bakken. Facies


are label and colored according to their position on the model. HST (highstand
systems tract), LST (lowstand systems tract), and TST (transgressive systems
tract) are labeled with their corresponding facies associations (modified from
Smith and Bustin, 1996). ....................................................................................131

Figure 5.1 All wells with rasters in the study area. Wells with digital logs used are
highlighted in pink. Parshall Field is the outline on the right and Sanish Field
is just to the east. The field outlines were taken from the North Dakota
Industrial Commission, Department of Mineral Resources, Oil and Gas
Division website. .................................................................................................135

Figure 5.2 Fluid saturation profile of the Patten #1-27H taken from core analysis. .......138

Figure 5.3 Porosity versus permeability plot for the Patten #1-27H taken from
core analysis. .......................................................................................................138

xvi
Figure 5.4 Porosity versus permeability plot for the Hoff #1-10H taken from
core analysis. .......................................................................................................140

Figure 5.5 Fluid saturation profile of the Hoff #1-10H taken from core analysis. .........140

Figure 5.6 Porosity versus permeability plot for the Parshall #2-36H taken from
core analysis. .......................................................................................................141

Figure 5.7a Porosity versus permeability plot for the N & D #1-05H taken from
core analysis. .......................................................................................................142

Figure 5.7b Porosity versus permeability plot for the N & D #1-05H taken from
core analysis. .......................................................................................................142

Figure 5.8 Fluid saturation profile for the N & D #1-05H taken from core analysis. ....143

Figure 5.9a Porosity versus permeability plot for the Long #1-01H taken from
core analysis. .......................................................................................................144

Figure 5.9b Porosity versus permeability plot for the Long #1-01H taken from
core analysis. .......................................................................................................144

Figure 5.10 Fluid saturation profile for the Long #1-01H taken from core analysis. .....146

Figure 5.11 Porosity versus permeability plot for the Van Hook #1-13H taken from
core analysis. .......................................................................................................147

Figure 5.12 Fluid saturation profile for the Van Hook #1-13H taken from
core analysis. .......................................................................................................147

Figure 5.13 Porosity versus permeability plot for the Jensen #12-44 taken from
core analysis. .......................................................................................................148

Figure 5.14 Fluid saturation profile for the Jensen #12-44 taken from core analysis. ...148

Figure 5.15 Porosity versus permeability plot for the Deadwood Canyon Ranch
#43-28H taken from core analysis. .....................................................................149

Figure 5.16 Fluid saturation profile for the Deadwood Canyon Ranch #43-28H
taken from core analysis. ....................................................................................150

Figure 5.17 Porosity versus permeability plot for the Dobrinski #18-44 taken from
core analysis. .......................................................................................................151

Figure 5.18 Fluid saturation profile for the Dobrinski #18-44 taken from core
analysis. ...............................................................................................................153
xvii
Figure 5.19 Porosity versus permeability plot for facies E1 taken from core
analysis. ...............................................................................................................154

Figure 5.20a Porosity versus permeability plot for facies D2 taken from core
analysis. ...............................................................................................................155

Figure 5.20b Porosity versus permeability plot for facies D1 taken from core
analysis. ...............................................................................................................155

Figure 5.21a Porosity versus permeability plot for facies C2 taken from core
analysis. ...............................................................................................................156

Figure 5.21b Porosity versus permeability plot for facies C1 taken from core
analysis. ...............................................................................................................156

Figure 5.22 Porosity versus permeability plot for facies B taken from core
analysis. ...............................................................................................................157

Figure 5.23 Porosity versus permeability plot for facies A taken from core
analysis. ...............................................................................................................157

Figure 5.24 Calculated log for the Deadwood Canyon Ranch #43-28H well. ...............163

Figure 5.25 Average Sw map for the middle Bakken. ....................................................165

Figure 5.26 Oil to water ratio calculated from cumulative oil production
data (bbls/bbls). ...................................................................................................166

Figure 5.27 Picket plot showing the facies D interval from the Deadwood Canyon
Ranch #43-28H well. The average porosity of 4.2 % and 8 ohm-m resistivity
value as a cut-off yield for producible hydrocarbons with water saturation
(Sw) of 40%. .......................................................................................................167

Figure 5.28 Net pay map(in feet) for the middle Bakken member determined from
water saturation values equal to or less than 40%. .............................................168

Figure 5.29 Hydrocarbon pore volume map of the middle Bakken. ..............................169

Figure 5.30 Plot between calculated TOC values from formation density logs and
TOC determined from core analysis (modified from Schmoker and
Hester, 1983). ......................................................................................................173

Figure 5.31 Upper Bakken TOC values calculated using the Schmoker and Hester
method. ................................................................................................................174

xviii
Figure 5.32 Lower Bakken TOC values calculated using the Schmoker and Hester
method. ................................................................................................................175

Figure 5.33 Vitrinite reflectance relationship to level of maturation (modified from


Hood et al., 1975). ...............................................................................................178

Figure 5.34 Comparison of the Schmoker and Hester method and the Passey method
for calculating TOC. ...........................................................................................178

Figure 5.35 Diagram showing the cycle of Rock-Eval pyrolysis analysis and the
corresponding recording (modified from Tissot and Welte, 1984). ...................180

Figure 5.36 Van Krevelen diagram showing types of organic matter. Type I is the
highest kerogen quality due to the highest hydrogen content, type III is the
lowest (from Tissot and Welte, 1984). ................................................................181

Figure 5.37 Modified van Krevelen diagram showing the distribution of the
Bakken Formation, Williston Basin. Published data from Webster (1984)
and Price et al. (1984). Majority of samples indicate a Type-I and II oil
prone kerogen (algal origin). Legend shows source rock data by depth
interval; three Parshall Field wells are indicated in black. .................................184

Figure 5.38 Hydrogen index versus Tmax values for the Bakken Formation,
Williston Basin. Majority of the samples indicate a Type I and II oil prone
kerogen. Kerogen-type lines from Finn and Johnson, 2005. Legend: Bakken
data from Webster (1984) and Price et al. (1984); black data points from
Parshall Field. .....................................................................................................185

Figure 5.39 S2/S3 versus hydrogen index values for the Bakken Formation,
Williston Basin. The majority of the samples indicate a Type-I and II oil
Prone kerogen. Legend: Bakken data from Webster (1984) and Price et al.
(1984); black data points from Parshall Field. ....................................................186

Figure 5.40 Plot comparing Rock-Eval production index and Tmax data for Parshall
Field and North Dakota data. Legend shows source rock data by depth interval
(data from Price et al., 1984; Webster, 1984); black data points from Parshall
Field. ...................................................................................................................187

Figure 5.41 Maturation rate index versus temperature indicating the exponential
relationship between the data. .............................................................................189

Figure 5.42 Deadwood Canyon Ranch #43-28H TTI and burial history (from
Nordeng, 2010). ..................................................................................................190

Figure 5.43 Figure 5.42 Parshall #2-36H TTI and burial history (from Nordeng and
LeFever, 2008). ...................................................................................................190
xix
LIST OF TABLES
Table 1.1 U.S Geological Survey assessment results for the Upper Devonian Lower
Mississippian Bakken Formation, Williston Basin (Pollastro et al., 2008). ...........3

Table 2.1 Table showing the different types of oil classifications found in the Williston
Basin and their corresponding source rock (Burrus et al., 1996). .........................38

Table 3.1 List of cores that were described in this study. Core from the Parshall
Field are highlighted in yellow. ............................................................................44

Table 5.1 Core porosity and permeability measurements and averages listed by well.
The correlation equation and R2 values are from cross plots of porosity versus
permeability. Values marked with (*) indicate permeability samples that
were taken at ambient conditions. .......................................................................137

Table 5.2 Core porosity and permeability measurements and averages listed by facies.
The correlation equation and R2 values are from cross plots of porosity versus
permeability. N/A indicates that not enough data was available. .......................154

Table 5.3 Parameters used to calculate water saturation using the Archie equation. .....161

Table 5.4 Values used in determining water saturation cut-offs. ...................................167

Table 5.5 Summary of log responses to organic matter (modified from Law, 1999). ....170

Table 5.6 Rock-Eval data provided by the NDGS and the USGS for Parshall Field. ....183

Table 5.7 Time-temperature index maturity values. .......................................................189

xx
ACKNOWLEDGEMENTS

I would like to begin by thanking my advisor Dr. Stephen Sonnenberg, for his
guidance and support throughout this research project. My other committee members, Dr.
Rick Sarg, Dr. Manika Prasad, and Robert Cluff for their support throughout the project. I
owe a big thanks to everyone at Discovery Group for their help in editing and general
guidance and assistance for the last two years.
Thanks to the students, professors, and members of the Colorado School of
Mines Bakken Consortium for the valuable discussions and funding throughout the years.
I would also like to thank my parents for their continued love and support
throughout my lifetime of education, as well as my brothers and sisters for helping me
keep things in perspective. Finally, I would like to thank my husband Cole for
everything. Your love, encouragement and support allowed me to start and complete this
program. I could not have done it without you.

xxi
CHAPTER 1
INTRODUCTION

The Bakken Formation, a Devonian-Mississippian black shale and a mixed


sandstone/carbonate unit, is an important source rock for oils produced in the Williston
Basin. Recent advances in horizontal drilling, fracture stimulation, and completion
technology have turned this once uneconomical play into reality. The Bakken horizontal
play began in Montana with the completion of the first well in March 2004 and has now
moved into North Dakota. The focus of this study, Parshall Field discovered in 2006 by
EOG Resources, covers some 40 townships and is still expanding. The field is located in
Mountrail County, North Dakota (Figure 1.1).
The U. S. Geological Survey (USGS) has published an assessment of
undiscovered oil and associated gas resources for the Upper Devonian Lower
Mississippian Bakken Formation consisting of five unconventional Assessment Units
(AU) and one conventional AU (Pollastro et al., 2008). Table 1.1 shows the assessment
results from the USGS. The Parshall Field is included in the Eastern Expulsion Threshold
AU as an unconventional continuous resource by the USGS and has mean undiscovered
reserves of 973 MMBO and 493 BCFG.
The Bakken Formation consists of three members: upper, middle, and lower
(Figure 1.2). The field produces from the middle dolomitic sandstone member and is
sourced from the organic-rich shale of the upper and lower Bakken members. The main
keys to the development of this field are horizontal drilling and multi-stage fracturing of
the middle Bakken dolomitic sandstones. The thickness of the middle Bakken ranges
from 30 to 70 feet (9.1 21.3 meters) and has low matrix porosity and permeability. The
middle Bakken porosities range from 2 to 12% and permeabilities average 0.02 mD. The
vertical depth of the middle Bakken in the Parshall Field area ranges from 9,000 to
10,500 feet (2,743 3,200 meters).
High oil prices combined with new technology (drilling and completion) are keys
to the development of the Parshall Field. The first initial horizontal producer was drilled

1
Figure 1.1 Index map showing the location of the Williston Basin and the study area in
northwestern North Dakota. Modified from Pitman et al. (2001).

2
Table 1.1 U.S Geological Survey assessment results for the Upper Devonian Lower
Mississippian Bakken Formation, Williston Basin (Pollastro et al., 2008).

Figure 1.2 Type log for the Parshall Field showing the overlying Lodgepole Formation,
the Bakken Formation, and the underlying Three Forks Formation.

3
by EOG Resources in 2006 in their Parshall #1-36 well. Currently there are over 400
horizontal development wells that have been drilled in Parshall Field and the neighboring
field to the west, the Sanish Field.

1.1 Location of Study Area


The study area is located in northwestern North Dakota between Townships 147-
164 N and Ranges 84-99 W and covers an area of approximately 5,700 square miles
(14,700 square kilometers) (Figure 1.3). The primary focus of this study is centered on
the Parshall Field area located in southern Mountrail County.

1.2 Research Objectives and Methods


The main objectives of this research were to:
Understand the depositional environment of the Bakken Formation in the
Parshall Field area by determining the Bakken members facies associations,
geometry and continuity. The following are detailed steps which were
performed to understand the depositional setting:
1. Described ten cores in the Parshall Field area: Deadwood Canyon Ranch
#43-28H, Parshall #2-36H, Van Hook #1-13H, Fertile #1-12H, Long #1-
01H, Hoff #1-10H, Patten #1-27H, N & D #1-05H, Jensen #12-44, and
Dobrinski #18-44. These descriptions were used to identify vertical and
lateral facies variations.
2. Correlated the Bakken members and their distinctive units (well log
facies) from vertical wells that were available in the Parshall Field area
(approximately 20 wells in the immediate field area, 913 included in the
entire study area). Structural maps, isopach maps, stratigraphic and
structural cross sections were constructed to support the creation of the
depositional model.
Determine the reservoir quality by applying an integrated petrophysical
analysis approach. The petrophysical analysis and facies interpretation were
combined to determine the relationship between these two factors and what
effect they may have on the Bakken reservoir performance in the Parshall

4
Figure 1.3 Location of study area in North Dakota. The Parshall Field is located in the
southern part of Mountrail County in Townships 151-155 N and Ranges 88-91 W.
Modified from LeFever (2008).

5
Field. Integrated petrophysical analysis in this study covered the following
details:
1. Determined reservoir quality from core analysis reports and the relationships
between rock properties from core measurements.
2. Performed conventional well-log analysis, determined porous interval and net
pay interval within the middle Bakken member reservoir, then correlated these
intervals and mapped the lateral thickness variation in the field.
3. Analyzed facies and rock properties relationships to determine controlling
factors on the reservoir quality.
4. Performed total organic carbon (TOC) calculation from formation density
logs.
5. Mapped the lateral extent of the organic carbon content in the Parshall Field
area. TOC values were plotted to determine the distribution of the source rock
quality.

1.3 Research Contributions


The major contributions of this study are:
Establish depositional facies and paleoenvironmental interpretation of the
Bakken Formation in the Parshall Field area.
Source rock quality analysis from cores and logs and through mapping of its
distribution.
Analysis of rock properties including porosity, permeability, and water
saturation of the Bakken Formation in the Parshall Field area.
Establish the relationship between facies and reservoir rock properties.

1.4 Previous Work


The Mississippian-Devonian Bakken Formation consists of three members: upper
and lower organic-rich black shales, and a middle member that varies from a silty
dolostone or limestone to sandstone lithology (LeFever et al., 1991). Meissners (1978)
Petroleum Geology of the Bakken Formation Williston Basin, North Dakota and
Montana discussed the overall petroleum geochemistry of the Bakken Formation.

6
Meissner described the stratigraphy, reservoir properties, geochemical (source rock)
properties, petrophysical properties, and the theory and controls for localizing reservoir
fracturing in the Bakken. Pitman et al. (2001) published a paper that focused on the
diagenesis and fracture development in the middle member of the Bakken Formation
from samples obtained from wells in North Dakota. Petrographic, geochemical and
reservoir quality analysis were performed on these samples. Carlisle et al. (1992) looked
at horizontal drilling and how fracture susceptibility in the Bakken is lithologically
controlled.
Kerr (1988) and Gerhard et al. (1990) wrote overviews of the geology of the
Williston Basin. Both papers gave a brief description of the Bakken Formation discussing
the basic stratigraphy and tectonics of the formation. Smith and Bustin (2000) studied the
stratigraphy of the Bakken Formation and interpreted the base to be a sequence boundary
(marine flooding surface).
Schmoker and Hester (1990) evaluated the formation resistivity of the shale
members of the Bakken to determine whether or not it is a good indicator for oil
generation and found that the cutoff for mature or immature appears to be around 35
ohm-m. They had previously looked at the organic carbon in the Bakken Formation and
discussed how to calculate the TOC (total organic carbon) values from the density log
(Schmoker and Hester, 1983).
Numerous studies have been published on the geochemistry of the Bakken shales.
Price et al. (1984) looked at the organic metamorphism of the shale in the North Dakota
portion of the Williston Basin. Price and LeFever (1994) compared the oils from the
Bakken and Madison reservoirs in the basin and concluded that the Bakken and Madison
oils are from different source rocks. Their results disagree with those from Williams
(1974) and subsequently with models developed by Dow (1974), Meissner (1978) and
other investigators (including Price et al., 1984). This suggests that further studies need to
be done (this was not addressed in this thesis).
More detailed lithofacies descriptions of the middle Bakken member have been
done by LeFever et al. (1991) as well as from Canter et al. (2009). Seven different
lithofacies were identified by LeFever et al. (1991) as shown in Figure 1.4. The siltstones

7
Figure 1.4 Lithologic column describing the lithofacies of the middle member of the
Bakken Formation (modified from LeFever et al., 1991). From Pitman et al. (2001).

8
and sandstones typically are massive or coarse bedded with rare trough or planar cross
beds. Much of the unit is well-sorted, although bioturbation commonly disturbs the
bedding, especially in the more argillaceous strata (Pitman et al., 2001). Canter et al.
(2009) examined the facies of the middle Bakken in Mountrail County, North Dakota.
They recognized five different facies with some additional subfacies from cores located
in the Sanish and Parshall fields. These facies are similar to those described by LeFever
et al. (1991) but the interpretation of the depositional environment varies for one facies in
particular; the rhythmic, varve-like sandstones. LeFever interprets this interval to be from
a shallow marine tidal environment while Canter et al. (2009) interprets the sandstones to
be from distal storm deposits and distal prodelta hyperpycnal gravity flows.

9
CHAPTER 2
GEOLOGICAL SETTING

The western North American Paleozoic craton was made up of a stable core, the
Canadian shield of older Precambrian rocks and its southwestward extension, the
Transcontinental arch (Peterson and MacCary, 1987). The Transcontinental Arch
separated the North American craton into western and eastern marine shelf areas during
the Early to Middle Paleozoic. To the west of the Transcontinental arch and the Canadian
shield, the west flank of the Paleozoic craton made up the Cordilleran shelf which was
the site of shallow water marine cyclic sedimentation during most of the Paleozoic and
Mesozoic time. This shelf slowly subsided allowing marine Paleozoic carbonates,
sandstones, and shales to accumulate (Figure 2.1 and 2.2). To the west of the shelf lay the
Antler orogenic belt which began active growth in Middle Devonian time. West of this
structure, thick deposits of deep water shale, fine-grained limestone, coarse clastics, and
submarine volcanic deposits accumulated during most of Paleozoic time (Peterson and
MacCary, 1987). The northern Great Plains region was located in what was then in the
eastern portion of the Cordilleran shelf adjacent to the Transcontinental Arch. This region
consisted of several paleostructural elements which influenced sedimentary processes;
the major one being the Williston Basin which began subsiding during Ordovician or
Late Cambrian time (Peterson and MacCary, 1987). The basin is interpreted to have
originated as a craton-margin or continental shelf basin and became an intracratonic basin
during the deformation of the Cordilleran orogen and subsequent crustal additions to the
western continental margin (Gerhard et al., 1990).
The Williston Basin is a near-circular, intracratonic sag that underlies most of
western North Dakota, eastern Montana, northwestern South Dakota, and portions of
Saskatchewan and Manitoba, Canada (Kerr, 1988; Pitman et al., 2001). The boundaries
of the basin are defined by several major tectonic features (Figure 2.3). The Cedar Creek
Anticline flanks the southwestern edge of the basin and to the northeast the sediments
thin onto the Meadow Lake escarpment. The northwestern flank of the basin is the
Sweetgrass-Battle River Arch which separates the Williston Basin from the Alberta
11
Figure 2.1 Paleogeographic map of the Late Devonian (360Ma). Subduction zones and
associated orogenic belts (dashed red lines) surround and separated the North American
craton from the Devonian ocean. The North American craton was open to the southwest
indicated by the yellow arrow. The Williston Basin and Alberta Basin were connected at
that time (W-A) (modified from Blakey, 2005).

12
Figure 2.2 Paleogeographic map of the Early Mississippian (345Ma). The Alberta Basin
(AB) and the Williston Basin (WB) were separated by the Sweetgrass Arch (SA). The
Williston Basin was shielded from the Devonian ocean by the orogenic belts (dashed red
lines) that surrounded the North American craton and the Transcontinental Arch (TA) to
the southeast (modified from Blakey, 2005).

13
Figure 2.3 Major structures of the Williston Basin (Gerhard et al., 1990).

14
Foreland Basin. The eastern margin is a gentle ramp up onto the Transcontinental Arch,
while the southern margin is defined today by the Black Hills uplift. Underlying the basin
are three Precambrian tectonic provinces: the Wyoming craton; the Trans-Hudson
orogenic belt; and the Superior craton (Gerhard et al., 1990; LeFever, 1992) (Figure 2.4).
These regional structures along with the eustatic sea-level changes have influenced the
cyclic nature of the Paleozoic rocks.

2.1 Regional Stratigraphy and Sedimentology


At the center of the basin, nearly 16,000 feet (4,877 meters) of sedimentary rocks
are present which range from Cambrian to Tertiary. This accumulation is divided into six
major sequences that represent relative sea-level rise, subsequent sedimentation, and then
relative sea-level drop with accompanying disconformities (Sloss, 1963). These
sequences are the Sauk, Tippecanoe, Kaskaskia, Absaroka, Zuni and Tejas (Gerhard et
al., 1990) (Figures 2.5 and 2.6).
Initial sedimentation in the basin occurred over an irregular Precambrian surface.
These sediments were mostly quartzite and conglomerate with some glauconitic
limestone which entered into the basin from an eastward indentation of the early
Cordilleran shelf during the Cambrian through Lower Ordovician time. It was not until
the end of the Sauk sequences that the Williston Basin became a well defined structural
unit. The Tippecanoe sequence rocks of Middle Ordovician through Silurian age
consisted of carbonate and clastic sediment sequences interrupted by major erosional
events. By the end of this deposition all of the major modern structures in the basin were
present. Some major oil-producing structures present at this time include the Billings,
Little Knife, Antelope, and Nesson anticlines (Gerhard et al., 1990). It is also suggested
that the Meadow Lake Escarpment in Saskatchewan may have been uplifted at the end of
the Tippecanoe.
The Kaskaskia sequence records two regional sea-level rises and an unconformity
that separates the younger Devonian rocks from the older Devonian so the sequence has
been divided into an upper and lower sequence. Prior to deposition of the lower
Kaskaskia sequence, the Transcontinental Arch on the south margin of the basin was
uplifted closing the marine communication with the Cordilleran shelf. This caused a

15
Figure 2.4 Basement tectonic provinces in the Williston Basin (red dashed line) (Fischer
et al., 2005).

16
Figure 2.5 Generalized stratigraphic column for the Williston Basin, showing units that
produce oil and gas (LeFever, 1992).

17
Figure 2.6 Generalized southwest to northeast structural and stratigraphic cross section A
- A' from the Bighorn Mountains in Wyoming to the northeastern side of North Dakota.
The Bakken Formation is highlighted in red. The index map shows the cross section line
in red (from Peterson and MacCary, 1987).

18
reorientation of the seaway to the north into the Elk Point Basin of northwest
Saskatchewan and eastern Alberta. The deposits of the lower Kaskaskia show repeated
transgressive-regressive cycles of mostly carbonates, with few evaporite and clastic
sediments. Reorientation of the basin occurred again during the upper Kaskaskia that
reestablished a connection to the Cordilleran shelf through the Central Montana Trough
(Figure 2.7). The contact between the Bakken Formation and the underlying Three Forks
Formation is the boundary between the lower Kaskaskia and upper Kaskaskia sequence.
These upper Kaskaskia rocks were deposited in a cycle of rapid, trangressive, carbonate
sedimentation and slow, prograding, episodic evaporite sedimentation as circulation in
the basin became restricted (Gerhard et al., 1990).
Devonian and Mississippian strata are separated by an unconformity which
represents uplift and erosion or sea level change. During that time interval, Devonian
strata were uplifted and exposed along the basin margin, while deposition continued in
the deeper portion of the basin. Mississippian sediments were later deposited on the
eroded Devonian surface (Gerhard et al., 1987; LeFever et al., 1991). This thesis focuses
on the Late Devonian Early Mississippian Bakken Formation and the adjacent
formations of the Williston Basin.
Termination by erosion of the upper Kaskaskia sequence took place
approximately during the Mississippian and Pennsylvanian boundary. The rocks of the
Absaroka sequence (Pennsylvanian through Triassic) consist mostly of siliciclastic
sediments with some evaporites and carbonates present. These siliciclastic sediments are
thought to be derived from the uplifts of the Ancestral Rocky Mountains to the south, the
emergent Canadian shield, and possibly from the Sioux Ridge of South Dakota (Gerhard
et al., 1990). The last major marine unit in the basin was the Cretaceous Pierre Shale
which is overlain by the Fox Hills Sandstone, the final regressive clastic unit of the
Cretaceous seaway.

2.1.1 Devonian Rocks


The total thickness of the Lower, Middle, and Upper Devonian beds is more than
2,000 feet (610 meters) in the central part of the basin (Figure 2.8). The Devonian thins

19
Figure 2.7 Diagram showing basin communication and sedimentation patterns through
time. Areas with arrows indicate open-marine communication: (A) Tippecanoe sequence,
(B) lower Kaskaskia sequence, (C) upper Kaskaskia sequence, (D) Absaroka sequence
(Gerhard et al., 1987).

20
Figure 2.8 Thickness distribution map showing the Devonian rocks of the Williston
Basin. Approximated limits of the Prairie salt and source rocks of the Bakken Formation
(Late Devonian Early Mississippian) are shown (Peterson and MacCary, 1987).

21
toward the Central Montana uplift and along the crest of the Cedar Creek Anticline. Both
of these paleostructural features underwent structural growth during the Devonian.
The Devonian of the Williston Basin and central Montana consists of cyclic
sequence of shallow water fossiliferous carbonate, shaly carbonate or shale, and
evaporates, including the Middle Devonian halite found in the Prairie Formation
(Peterson and MacCary, 1987).
The initial deposition consists of red dolomite, siltstone, and shale beds of the
Lower or Middle Devonian Ashern Formation. The Ashern Formation grades upward
into mounds and reef-bearing carbonates of the Middle Devonian Winnipegosis
Formation (100 to 135 feet thick) (30.5 to 41.1 meters). The Winnipegosis Formation
consists of carbonate shelf environments: patch reefs, lagoon, and tidal flat facies, and
also consists of lithofacies of which represent the deeper environments characterized by
laminated limestones, stromatoporoids, and tabulate coral pinnacle reefs (Peterson and
MacCary, 1987).
The Winnipegosis is overlain by the Middle Devonian Prairie Formation (300 to
500 feet thick) (91.4 to 152.4 meters) which consists of a lower unit of anhydrite,
dolomite, thin-shale, and halite beds with minor interbeds of red shale occurring in the
upper part of the halite interval. The Prairie Formation is overlain conformably by the
Dawson Bay Formation, approximately 100 to 175 feet thick (30.5 to 53.3 meters). The
Dawson Bay consists of a single carbonate-evaporite cycle and represents the final phase
of the Middle Devonian deposition (Peterson and MacCary, 1987).
The Upper Devonian units in the Williston Basin (in ascending order, the Souris
River, Duperow, Birdbear (Nisku), and Three Forks formations) consist of several
cycles of carbonate-evaporite and fine clastic beds. These were deposited during the
maximum transgression of the Devonian seaway which had spread across most of the
Rocky Mountain shelf.
The Souris River which conformably overlies the Dawson Bay in the basin center
but is unconformable around the basin borders, consists of argillaceous and fossiliferous
dolomite or limestone capped by anhydritic dolomite or anhydrite (Sandberg and
Hammond, 1958; Wilson, 1967; Peterson and MacCary, 1987). The Duperow Formation
conformably overlies the Souris River Formation, and consists of a cyclical carbonate -

22
evaporite sequence with the presence of abundant skeletal deposits (banks or mounds) of
stromatoporoids, tabulate corals, and other marine organisms (Wilson 1967; Peterson and
MacCary, 1987).
The Birdbear (Nisku) Formation conformably overlies the Duperow Formation
and represents the final carbonate-evaporite cycle of the Devonian. There are four main
depositional environments recognizable in the Birdbear Formation: subtidal, intertidal,
lagoonal, and supratidal (Peterson and MacCary, 1987).
The Three Forks Formation averages about 150 feet (45.7 meters) thick in North
Dakota and underlies the Bakken Formation everywhere in the Williston Basin
(Meissner, 1978). The thickness of the Three Forks Formation reaches up to 250 feet
(76.2 meters) in the central part of the basin. The Three Forks Formation contact with the
Bakken Formation appears conformable in the central part of the basin but near the
margins the two formations are separated by an unconformity (Webster, 1984). The
Three Forks Formation is composed of red and green shale, siltstone, sandstone,
carbonate rock, evaporite sequences, and represents the final regression of the Devonian
sedimentation which was related in part to westward transport of an increasing supply of
clastic materials (Peterson and MacCary, 1987).
The Sanish sandstone name was given to the uppermost bed of the Three Forks
Formation and consists of white, slightly calcareous siltstone or very fine- to fine-
grained sandstone with thickness ranging from 5 to 15 feet (1.5 to 4.6 meters). This bed
locally produces oil from the Antelope Field at the south end of the Nesson Anticline and
is not laterally continuous over a great distance (Kume, 1963). The Sanish has been
interpreted to be from a beach or nearshore marine depositional environment (Webster,
1984).

2.1.2 Upper Devonian Mississippian Rocks


The Upper Devonian to Lower Mississippian Bakken Formation consists of three
members: upper and lower organic-rich black shales and a middle member that varies in
composition from a silty dolostone or limestone to a sandstone lithology (LeFever et al.,
1991) (Figure 2.9). The total Bakken thickness ranges from upwards of 140 feet (42.7
meters) thick in the center of the basin to being completely gone at the edges of the basin.

23
Figure 2.9 Well logs and interpreted lithology of the Tipperary Oil and Gas Corporation
Olsen No. 1 well, SE SW Sec. 26, T160N, R97W, Divide County (Webster, 1984).

24
The limits of the three Bakken members have been mapped showing the pinch out of
each the Bakken members toward the basin margins. Webster (1984) did this for the
North Dakota portion of the Williston Basin. The lower shale is the least extensive, the
middle member more extensive, and the upper shale the most as shown in the Figure
2.10. The lower member consists of dark gray to brownish black to black, massive to
fissile, slightly to highly organic-rich shale (Pitman et al., 2001). The upper member is
lithologically similar to the lower member and consists of dark gray to brownish black to
black, fissile, slightly calcareous, organic-rich shale. The upper member differs from the
lower member in that it lacks crystallized limestone and greenish gray shale beds and has
a higher organic matter content (Pitman et al., 2001). These upper and lower shales have
been interpreted to have been deposited in an offshore marine environment during
periods of sea-level rise (Pitman et al., 2001; Webster 1984). Webster (1984) and
Lineback and Davidson (1982) provided evidence that both of these shales were
deposited in a stratified hydrologic regime indicated by the lack of benthic fauna,
presence of pyrite, and high organic content.
The upper and lower shales of the Bakken Formation are world-class petroleum
source rocks. These shales are easily distinguished on logs and are excellent marker beds
used to create a subsurface correlation within the basin.
The middle member is highly variable and consists of a light gray to medium-dark
gray, interbedded sequence of siltstones and sandstones with lesser amounts of shale,
dolostones, and limestones rich in silt, sand and oolites (Pitman et al., 2001). This middle
member has been interpreted to have been deposited in a shallow water regime following
a rapid sea level drop (Smith and Bustin, 1996). Within the middle member several facies
have been described and LeFever et al. (1991) suggested that the upper part of the
member was deposited in a marine environment with strong current action.
The Bakken Formation received its name from the strata occurring between
depths of 9,615 feet to 9,720 feet (2,930.7 to 2,962.7 meters) in the Amerada Petroleum
Corporation #1 H.O. Bakken deep test, C SW NW Sec. 12, T 157 N, R 95 W, Williams
County, North Dakota (Sandberg and Hammond, 1958). The Bakken is conformably
overlain by the Mississippian Lodgepole Formation.

25
Figure 2.10 Approximate subsurface limits of the three members of the Bakken
Formation. The northwest to southeast stratigraphic cross section A-B shows the regional
extent of the Bakken Formation in North Dakota (Webster, 1984).

26
2.1.3 Mississippian Rocks
The Mississippian section in the Williston Basin consists of two major groups: the
Madison and Big Snowy. The Madison Group is more than 2,000 feet (609.6 meters)
thick in the central Williston Basin (Figure 2.11). The Madison Group consists of three
units. The first or the lower unit is the Lodgepole Limestone, a gray to dark gray
argillaceous to shaly or silty, in places cherty, thin- to medium-bedded limestone. Some
minor amounts of chert and anhydrite are also present in the limestone. The Lodgepole
represents a complex system involving at least three slope-wedge shale units of slope and
basin environments followed by shelf and slope carbonate units. The oldest of the slope-
wedge shales, the Carrington, may be an upslope clinoform equivalent to the Bakken
Shale as there is no evidence for any break between Bakken and Madison sedimentation
(Kerr, 1988). This shale is commonly known as the false Bakken.
The middle unit of the Madison Group, the Mission Canyon Formation, consists
of thick massive fossiliferous to oolitic carbonate rocks. The third unit is the Charles
Formation which consists of anhydrite and halite with interbedded carbonate rock and
shale and is, at least in part, time equivalent to the Mission Canyon. The Lodgepole and
Mission Canyon formations contain higher percentages of argillaceous or finely
crystalline limestone in the central part of the basin and a higher percentage of dolomite
toward the basin margins, particularly toward the south and southwest. There are up to
eight depositional facies recognized in the Lodgepole Formation ranging from a central
basin and basin flank facies to shelf facies toward the basin margin. Widespread
occurrence of cyclic carbonate deposits in the Madison reflects sea level fluctuations and
the stability of the tectonic condition at that time. The absence of medium to coarse
clastics and the widespread nature of individual stratigraphic units suggest that the main
causes of cyclic deposition are eustatic sea level changes coupled with gentle epeirogenic
subsidence of the basin region (Peterson and MacCary, 1987).
The Big Snowy Group consists of the Kibbey, Otter, and Heath formations. These
formations are more than 1,000 feet (304.8 meters) thick in the Central Montana Trough
and more than 600 feet (182.9 meters) thick in the central part of the Williston Basin in
eastern Montana and the western part of North Dakota (Figure 2.12). The Kibbey
Formation consists of 200 to 300 feet (61.0 to 91.4 meters) of red shale, siltstone, and

27
Figure 2.11 Thickness distribution map and generalized rock facies of the Madison
Group (Peterson and MacCary, 1987).

Figure 2.12 Thickness distribution map and generalized rock facies of the Big Snowy
Group (Peterson and MacCary, 1987).

28
sandstone, and is primarily nearshore marine in origin. The middle unit, the Otter
Formation, consists of 200 to 300 feet (61.0 to 91.4 meters) of marine, nearshore, and
tidal flat beds of green shale and minor finely crystalline, commonly stromatolitic, non-
porous limestone or dolomite. The uppermost Heath Formation, consists of 200 to 400
feet (61.0 to 121.9 meters) of dark gray to black, highly organic marine shale, limestone,
minor siltstone or fine-grained sandstone. These beds were deposited under restricted
marine conditions (Peterson and MacCary, 1987).

2.2 Regional Structure


Structural trends in the United States portion of the Williston Basin are
predominantly north- and northwest-trending anticlines (Gerhard et al., 1987; LeFever,
1992) with most prominent structures being the Nesson and Cedar Creek anticlines.
Smaller structural features include the Antelope, Billings, Little Knife, and Poplar
anticlines. The Cedar Creek and Antelope anticlines have northwest trends. The Poplar
Dome is slightly divergent from these two, but also has a general northwest trend. The
structures that have north trends are the Nesson, Billings, and Little Knife anticlines. All
of these structures are oil productive (Figure 2.13).
Clement (1976) reported that the Cedar Creek anticline underwent four major
periods of growth: Early Devonian, Late Devonian, Late Mississippian Triassic, and
post Pliocene. Several upper Paleozoic sedimentary changes also occur in the vicinity
of the anticline due to the thinning or absence of Devonian and Silurian rocks along the
crest of the anticline.
Several periods of uplift and erosion occurred during Late Mississippian to Early
Jurassic. Paleozoic strata in the northern part of the basin were uplifted and differentially
eroded, while strata in the southern part were untouched (LeFever et al., 1991). The
periods of uplift and erosion also occurred during Late Devonian to Early Mississippian
time. During that interval, Devonian strata were uplifted and exposed along the basin
margins while sedimentary deposition continued in the deeper part of the basin.
Mississippian strata were later deposited and eroded the Devonian strata (LeFever et al.,
1991).

29
Figure 2.13 Predominant structures of the Williston Basin. Major structural features
include: Antelope Anticline (AT), Billings Anticline (B), Birdtail-Waskada axis (BW),
Cedar Creek Anticline (CC), Hummingbird Synclinorium (HS), Little Knife Anticline
(LK), Nesson Anticline (NS), Poplar Dome (PD), Roncott High (RH), Sheep Mountain
Synclinorium (SM), Weldon-Brockton-Froid Fault Zone (WBF), Western Nesson
Anticline (WN) (LeFever, 1992).

30
Faulting is associated with two main structural features: the Cedar Creek and
Nesson anticlines (Figure 2.13). These faults were periodically reactivated through time
and are responsible for traps associated with prolific production in the U.S. part of the
Williston Basin (LeFever, 1992).
Structural features present in the Canadian portion of the Williston Basin are
generally related to the salt dissolution of the Middle Devonian evaporites and collapse of
the overlying beds. The present-day edge (dissolution) of the Prairie is illustrated in
Figure 2.13. Several major oil fields in southern Saskatchewan and Manitoba are
associated with salt dissolution of Devonian evaporites. Salt dissolution is also listed as
causing the fracturing in Mondak Fields Mississippian carbonates along the North
Dakota-Montana line and for producing the paleo topography of Nisku deposition which
formed thick pods of reservoir in the Montana portion of the basin.
Structures present in southeastern Saskatchewan are the ElbowHummingbird
Homoclinal flexure, the Hummingbird Synclinorium, the Rocott Anticlinorium, and the
Torquay Rocanville Trend (LeFever, 1992) (Figure 2.13). An additional structural trend
in the southeastern Saskatchewan is the Bismarck Williston lateral fault zone. Several
major oil fields are found in southern Saskatchewan: Rocanville (Bakken), Wapella
(Jurassic), Midale, Benson, Torquay, and Oungre (Mississippian) (LeFever, 1992). These
structural features and trends are what controlled these fields.
A prominent structure in Manitoba is a north-trending zone of the Birdtail-
Waskada Axis, which is an important control on several oil fields in Manitoba. The
accumulations in Daly, Virden, and Waskada fields are controlled by this feature. Several
thickness anomalies in the Devonian and younger strata, caused by dissolution and
collapse features are related to fluid movements along the Precambrian province
boundary strata, and are characteristic of this structural trend (LeFever, 1992). Occasional
asteroidal impacts also created structures that were locally productive such as the Red
Wing Creek structure in North Dakota and the Viewfield Field in Saskatchewan.

2.3 Geologic Overview of Parshall Field


Parshall Field is located in Mountrail County, North Dakota in the southeastern
part of the Williston Basin (Figure 1.3). Parshall Field was discovered in 2006 with the

31
drilling and completion of EOG Resources #1-36 Parshall well (SESE Sec. 36, T153N,
R90W) (LeFever, 2008). The field is presently still expanding towards the west into
Sanish Field and further north in Mountrail County (Figure 2.14). As of September 30,
2010, 214 wells have been drilled in the field. The total thickness of the Bakken
Formation in the Parshall Field ranges from 80 to 130 feet (24.4 to 39.6 meters). The
upper Bakken reaches up to 18 feet (5.5 meters) thick, the middle member ranges from
30 to 70 feet (9.1 to 21.3 meters) thick, and the lower Bakken member ranges from 25 to
45 feet (7.6 to 13.7 meters) thick (Figure 2.14).
The primary oil production of the Parshall Field comes from the middle dolomitic
sandstone member, which has porosities ranging from 2 to 12%. The field development
was driven by advanced drilling and completion technologies. Horizontal drilling and
multi-stage fracture stimulation are keys for successful development in the middle
member reservoir of the Bakken Formation in the Parshall Field.
The contact between the middle member and the upper shale member is sharp
with the upper shale member conformably overlying the middle dolomitic sandstone
member. The upper and lower shales have a high radioactive reading on well logs
associated with the high organic content of the shales. The contact between the middle
member and the lower shale also appears sharp on well logs.
The contact between the lower shale and the Three Forks Formation is an
irregular and sharp contact and in many places of the field the contact is an erosional
contact. A basal lag deposit is found at the bottom of the lower member in contact with
the Three Forks Formation indicating erosion of the Three Forks Formation. This reflects
a hiatus or time interval with no deposition of the Three Forks Formation (LeFever et al.,
1991).

2.4 History of Oil Production in the Williston Basin


Production in the Williston Basin encompasses three states (Montana, North
Dakota, and South Dakota) and two Canadian provinces (Saskatchewan and Manitoba).
The Saskatchewan production is the largest, North Dakota and Montana follow closely,
and lesser amounts come from Manitoba and South Dakota. Petroleum production started
in Montana and North Dakota in the 1920s with gas production from the Cedar Creek

32
Figure 2.14 Map showing the expansion of Bakken exploration to the north and west of
Parshall Field. This was the current activity at the end of 2008 (Modified from Johnson,
2009).

33
anticline (Gerhard et al., 1990; LeFever, 1992). The first oil production in the Williston
Basin was discovered at the Nesson Anticline in North Dakota by the Amerada No. 1
Clarence Iverson in 1951 (Gerhard et al., 1990). The activity in the basin increased
because of this discovery and was followed by discoveries in other portions of the basin.
The oil embargo in 1973 which caused a rapid rise in oil price had a significant
impact on the activities in the basin, and was followed by the discovery of the Red Wing
Creek, Mondak, Little Knife, and the fields along the Billings Nose (Figure 2.15).
Drilling activity in the Williston Basin was then influenced by the fluctuation of the oil
price in the 1980s. Drilling activity in the basin collapsed in 1980 and 1981 along with
the oil price in 1982 (LeFever, 1992).
The earliest discovery of the Bakken Formation occurred in the Antelope Field of
North Dakota in 1953, and development continued into the 1960s. The Antelope Field
averaged only 209 barrels of oil per day (BOPD) (33.2 m3/day) (Sonnenberg et al., 2009).
The first Bakken horizontal well was completed in late 1987 by Meridian Oil Company
with their MOI No. 33-11 well in North Dakota (Sec. 11, T143N, R102W). This well was
completed in the upper Bakken shale for only 258 BOPD (41 m3/day) (LeFever, 1992).
This play continued into the 1990s with more than 20 operators. The Elm Coulee Field of
Richland County Montana was discovered with horizontal drilling in 2000. The concept
of drilling within the middle Bakken was then applied to the Parshall Field area.

2.5 Horizontal Drilling Plays in the Bakken Formation


The Bakken is a significant oil source rock, but the recent success of the Bakken
Formation as a major oil-producing unit comes from the advancement of horizontal
drilling and fracture stimulation technologies. The Bakken Formation has low net pay
thickness intervals, low porosity, low permeability, and only minor natural fracturing.
Now that fracture systems can be artificially generated to open the Bakken up in
horizontal wells, the play has become a reality. Meridian Oil was the first company to
drill and complete a horizontal well in the Bakken. The #33-11 MOI well (Sec. 11,
T143N, R102W, Elkhorn Ranch Field) was completed in September 1987 and produced
258 (BOPD) (41 m3/day) and 299 thousand cubic feet of gas per day (MCFD) (8,466.7

34
Figure 2.15 Map showing the distribution of oil and gas fields in the Williston Basin. Oil
fields are shown in green and gas fields are shown in red (Longman, 1987).

35
m3/day) from the upper Bakken shale (LeFever, 2005). The success of this well set off
the horizontal drilling phase of the upper Bakken shale play that continued into the early
1990s. Product prices declined significantly in the 1990s and, along with somewhat
unpredictable production of the upper Bakken shale, brought this phase to a close
(Sonnenberg et al., 2009).
The Elm Coulee Field of Richland County, Montana renewed interest in
horizontal drilling in the Bakken with its discovery in 2000. The key well for this play
was the Kelly/Prospector #2-33 Albin FLB well (Sec. 33, T24N, R57E, Richland
County). The middle Bakken was chosen to be perforated in this vertical well based on
shows seen on the mud log and flowed 157 BOPD (25.0 m3/day) for the first 20 days of
production in March of 1996 and was still making 80 BOPD (12.7 m3/day) after three
months. The results of this well were encouraging and thus created the concept that a
large field existed here (Sonnenberg et al., 2009). This area was termed by Dick Findley
(a Billings independent) the sleeping giant because it had been drilled through many
times without the Bakken Formation being looked at as a potential reservoir. Several re-
entries or recompletions were done in the late 1990s to pursue the play but it really took
off in 2000 when horizontal drilling began in the middle member.
The Elm Coulee discovery and development prompted operators to also target the
middle Bakken in North Dakota. Michael Johnson noticed that wireline logs of the
middle Bakken in Mountrail County resembled those from the Elm Coulee Field in
Montana. In 2005 EOG Resources drilled the Nelson Farms #1-24H well (Sec. 24,
T156N, R92W) and demonstrated that this horizontal well with its large hydraulic
fracture stimulation of the middle Bakken could recover significant oil along the eastern
flank of the Williston Basin in Mountrail County (Nordeng, 2010). The following year
EOG Resources drilled the Parshall #1-36 well, the discovery well for Parshall Field, and
the Parshall #2-36 well that had production rates in excess of 500 BOPD (79.5 m3/day).
Most of the laterals drilled in Mountrail County are on the order of 5,000 feet (1,524
meters) and are fracture stimulated with at least a single stage or, in many cases, 10 to 12
individual stages (Nordeng, 2010). This discovery has not only focused attention on the
Bakken Formation in Mountrail County but also throughout the Williston Basin.

36
2.6 Petroleum System Overview of the Williston Basin
The Williston Basin hosts three main types of source rocks: 1) Type I lacustrine
lipid-rich algae, bacteria and sporopollenin (oil-prone): found in the Ordovician
Winnipeg Formation shales (average of 0.42 % total organic content); 2) Type II
marine phytoplankton, zooplankton, and bacteria (oil and gas- prone): found in the Late
Devonian Early Mississippian Bakken Formation (average 11.33 % TOC, with
maximum value of 20.00 %); and 3) Type III terrestrial vascular plants (gas-prone):
found in the Pennsylvanian Tyler Formation (average 2.68 % TOC with maximum value
of 9.07 %) (Williams, 1974; Webster, 1984).
Several studies have published oil family classification in the Williston Basin as
shown in Table 2.1. Recent oil family classification in the Williston Basin is divided into
five families: A, D, B, C, and G (Osadetz et al., 1994; Burrus et al., 1996). Table 2.1
shows oil type classification from Williams (1974) and Dow (1974) with type I, II, and
III oil classification. According to a burial history diagram from Webster (1984), the
Bakken enters the oil window in Late Cretaceous time (70 Ma) (Figure 2.16). The
Bakken Formation is considered to be a major source in the Williston Basin. Webster
(1984) interpreted onset of oil generation to occur at a depth of 9,000 feet (2,743.2
meters) for the Bakken Formation. Expulsion of this oil occurred when burial depths
reached between 9,500 and 10,000 feet (2,895.6 and 3,048 meters) (Webster, 1984).
Meissner (1978) recognized that the overpressuring of the Bakken was caused by
the generation of petroleum which in turn caused fracturing of adjacent rocks. These
fractures can be conduits for the petroleum migration from the Bakken Formation
(Gerhard et al., 1990). Due to its high original organic carbon content, immature Bakken
is in-part kerogen supported. As kerogen matures in this zone it becomes pliable, loses
strength and begins to convert to liquid hydrocarbons. A volumetric increase associated
with this conversion causes locally elevated pressure cells to develop. The increased
pressure induces the formation of microfractures. Oil is expelled from the thermally
weakened kerogen through these microfractures (Coskey et al., 2009). Webster (1984)
determined that the center area of generation is located in McKenzie County, North
Dakota and the second center area is located in Sheridan County, Montana (Gerhard et
al., 1990).

37
Table 2.1 Table showing the different types of oil classifications found in the Williston
Basin and their corresponding source rock (Burrus et al., 1996).

Figure 2.16 Burial history for four different wells (A= California Oil No. 1 Rough Creek,
B= Texaco No. 1 Clarence Pederson, C= Marathon Oil No. 18-44 Dobrinski, D= Placid
Oil No. 36-5 Rosendahl). The upward deflections represent possible loss of section from
major unconformities (Webster, 1984).

38
2.7 Migration
Gerhard et al. (1990) suggested that migration from the major source rocks to the
reservoir rocks was controlled by tectonic fractures and fractures caused by hydrocarbon
generation. One hypothesis suggests that petroleum migrated along the large fracture
systems in the basin related to the original left-lateral shear sets (Brockton-Froid and
Colorado-Wyoming) to the north. This is also supported by the presence of linear features
having the same orientation that are consistently reactivating along the Nesson Anticline
(Gerhard et al., 1990) (Figure 2.17 and 2.18). An apparent geometric relationship
between petroleum production and major structural trends is easily traced into the
generation area of the basin for the Bakken and deeper rocks (Gerhard et al., 1990). For
example petroleum production in the northeastern part of the basin appears to be a direct
result of migration from the Bakken along the fracture trends into traps. The Bakken oil
has migrated preferentially along the fractures in overlying Mississippian rocks, sealed
from the lower rocks by Prairie Formation evaporates and by younger Devonian rocks
that lack effective permeability (Gerhard et al., 1990).
The Weldon-Brockton-Froid fault system has very effective fracture permeability
shown by drill-stem tests (Gerhard et al., 1990; LeFever et al., 1991). These fractures link
petroleum migration from the basin center to many places, including the oil accumulation
in the Cedar Creek Anticline (Figure 2.19). The Weldon-Brockton-Froid fault system is
also connected to the petroleum production from southern Saskatchewan that may have
aided petroleum migration to the trap area. Most traps are in the Mississippian carbonates
capped by extensive calcrete surfaces in southern Saskatchewan (Gerhard et al., 1990).

39
Figure 2.17 Map showing the major tectonic left-lateral shear sets in the Williston Basin
(Gerhard et al., 1990).

Figure 2.18 Interpretation of the left-lateral shear sets origin shown by an elliptical strain
diagram in the central Rocky Mountains and Williston Basin block (Gerhard et al., 1990).

40
Figure 2.19 Map showing the theoretical migration pathways from the major source rocks
to the producing reservoirs in the Williston Basin. The areas of low migration potential
are shown in gray (Gerhard et al., 1990).

41
CHAPTER 3
FACIES INTERPRETATION AND CORE DESCRIPTIONS

There are 29 cores reported for the Bakken interval in Mountrail County and
Ward County, North Dakota but only ten cores were described in this study. Synthesis of
these data helps to delineate facies variation, determine the sediment source, and to create
a depositional model for the Bakken Formation in the Parshall Field and nearby
surrounding areas.
The core dataset includes eight cores available in Mountrail County (Van Hook
#1-13H, N & D #1-05H, Parshall #2-36, Hoff #1-10H, Long #1-01H, Patten #1-27H,
Fertile #1-12H, and Jensen #12-44) at the North Dakota Geological Survey (NDGS) Core
Library in Grand Forks, North Dakota, and one other core available via permission by
Fidelity Exploration and Production Company (Deadwood Canyon #43-28H). One
additional core (Dobrinski #18-44) in Ward County which is also located at the NDGS
Core Library was described to support the paleoenvironmental and facies interpretation.
Table 3.1 shows the cores that were described in this study. The locations of the ten cores
described are shown in Figure 3.1.
Other supporting data such as thin sections, well logs, and isopach maps were
used to conduct an integrated analysis and to support the depositional environment
interpretation. The idea of analyzing different scales of data is to produce a more precise
interpretation based on the sedimentologic and stratigraphic record.
Thin sections from one core (Deadwood Canyon #43-28H) were examined and
interpreted for general framework mineralogy and pore structure. Mineralogical analysis
was also obtained from core samples using QEMSCAN technology to give a more
detailed look at mineralogy identification, porosity, and mineralogical differences with
changes in depth of the core.

3.1 Methods
Cores were described for their lithologic variations, grain-size distribution,
mineralogical composition, fossil and trace fossil associations, rock textures, and
43
Table 3.1 List of cores that were described in this study. Cores from the Parshall Field are highlighted in yellow.

44
Figure 3.1 Map showing the location for the ten cores described in this study. Eight
cores: Van Hook #1-13H, N & D #1-05H, Parshall #2-36H, Hoff #1-10H, Long #1-01H,
Patten #1-27H, Fertile #1-12H, and Jensen #12-44 are located in the Parshall Field,
Mountrail County, North Dakota. The Deadwood Canyon Ranch #43-28H core is from
the Sanish Field and the Dobrinski #18-44 core is from a wildcat well east of Parshall.

45
sedimentary structures. Grain size was visually measured using a hand lens. Thin section
analysis and QEMSCAN analysis was conducted to determine a more precise grain size
range and lithological composition. QEMSCAN is an automated electron beam
technique providing quantitative mineralogical data. It combines features found in other
analytical instruments such as a Scanning Electron Microscope (SEM), Electron Probe
Micro Analyzer (EPMA), and X-ray diffraction (XRD) spectrometers
(http://www.fei.com/applications/industry/automated-mineralogy/qemscan.aspx).
The presence of calcite and dolomite minerals in core sample were identified
using 10% HCl. The difference of the HCl response between the calcite and dolomite is
the most distinct characteristic to identify them with calcite being more reactive to HCl
compared to dolomite. In addition, thin section analysis over various Bakken intervals
was also performed to help in facies determination.
To support the facies descriptions and paleoenvironmental interpretation, well-log
cross sections and a series of isopach maps from various Bakken members were
constructed. Images of individual facies were taken from the North Dakota Geological
Surveys website that can be referenced with a small subscription fee from the following
link: https://www.dmr.nd.gov-oilgas/.

3.2 Facies Descriptions


A facies is a body of rock characterized by a particular combination of lithology,
physical and biological structures that exhibit an aspect different from the bodies of rock
above, below, and laterally adjacent (Walker, 2006). Determination of facies is one of
the main objectives of this study. Different data such as cores, thin sections, and
QEMSCAN data were examined to help characterize and interpret these facies.
Trace fossil identification is an important part of this study since the cores
described are burrowed to highly bioturbated. Trace fossils represent a unique blend of
sedimentological and paleontological information that lead to the interpretation of
depositional environments. Ichnofacies (trace fossils associations) record the behavior of
active, in situ organisms to a much greater extent than body fossils (Sutter, 2006). Trace
fossil study involves the grouping of trace fossil association and characteristics into a
habitual ichnofacies. There are nine general ichnofacies that have been recognized (Suter,

46
2006), each named after a representative ichnogenus: Trypanites, Teredolites,
Glossifungites, Psilonichnus, Skolithos, Cruziana, Zoophycos, and Nereites (Figure 3.2).
These trace fossil associations reflect numerous environmental factors, such as substrate
consistency, food supply, hydrodynamic energy, salinity, and organic levels.
Ten conventional cores from the Parshall Field area were analyzed in this study:
Van Hook #1-13H, Fertile #1-12H, Long #1-01H, Hoff #1-10H, Patten #1-27H, Parshall
#2-36, N & D #1-05H, Deadwood Canyon #43-28H, Jensen #12-44 and Dobrinski #18-
44 (Figure 3.1). The Hoff #1-10H core is a horizontal core. All of the cores show the
same characteristics and facies with seven facies identified based on the presence of
sedimentary structures, grain size variation, trace fossil association and intensity, rock
textures, and fossil content (Figure 3.3). These facies, designated A-G (modified from
Canter et al., 2009), show a good match with the characteristic changes observed from
thin section study.
Facies E is subdivided into facies E1 and facies E2, based on bioturbation
intensity and sedimentary features. Facies D is subdivided into facies D1 and facies D2,
based on sedimentary structures. Facies C is subdivided into facies C1 and facies C2,
based on sedimentary structure differences. Facies associations from cored intervals in
the Parshall Field along with other cores from outside the Parshall Field were used to
determine the paleoenvironmental setting of the Bakken members in the field. Seven
different facies identified from core data in the Parshall Field have specific stratigraphic
position shown in the well log data (Figure 3.4).

3.2.1 Facies G. Organic Rich Pyritic Brown/Black Mudstone


Facies G is characterized by dark gray, brownish-black to black, fissile, non-
calcareous, carbonaceous, and bituminous shale (Figure 3.5). This mudstone interval
contains silt-size dolomites, calcites, and quartz. Pyrite is generally disseminated
throughout the shale, but also occurs in laminations. This facies also contains calcitic
fossil fragments which are difficult to identify. Fractures are generally cemented by
pyrite and calcite. This facies composes both the upper and lower Bakken shales.

47
Figure 3.2 Diagram illustrating representative marine ichnofacies. Typical trace fossils
include: 1) Caulostrepsis; 2) Entobia; 3)equinoid borings; 4) Trypanites; 5) Teredolites;
6) Thalassinoides; 7,8) Gastrochaenolites; 9) Diplocraterion; 10) Skolithos; 11,12)
Psilonichnus; 13) Macanopsis; 15) Diplocraterion; 16) Arenicolites; 17) Ophiomorpha;
18) Phycodes; 19) Rhizocorallium; 20) Teichichnus; 21) Planolites; 22) Asteriacites; 23)
Zoophycos; 24) Lorenzinia; 25) Zoophycos; 26) Paleodyction; 27) Taphrhelminthopsis;
28) Helminthoida; 29) Cosmorhaphe; 30) Spirorhaphe (Suter, 2006).

48
Figure 3.3 Core pictures: G, A, B, C, D, E, and F, represent the main facies identified
from Bakken cores available in and near the Parshall Field. G. Organic rich pyritic
brown/black mudstone. A. Muddy lime wackestone. B. Bioturbated, argillaceous,
calcareous, siltstone/sandstone. C. Planar to undulose laminated, argillaceous
siltstone/sandstone. D. Low angle planar to undulose sandstone. E. Finely inter-
laminated, bioturbated, dolomitic siltstone/sandstone. F. Bioturbated, shaly, dolomitic
siltstone.

49
Figure 3.4 Facies of the Bakken Formation observed from core samples. Facies E is
subdivided into E1 and E2, based on sedimentary features, and is combined with facies F
for log correlations. Facies D is subdivided into D1 and D2 based on sedimentary
structures. Facies C is subdivided into C1 and C2 based on sedimentary structure
differences. The log and core pictures were taken from the Deadwood Canyon Ranch
#43-28H well (core available by permission from Fidelity Exploration and Production
Company).

50
Figure 3.5 Core photos of facies G organic rich pyritic brown/black mudstone. Photos
A and B are from the Upper Bakken Shale and photos C and D are from the Lower
Bakken Shale. Small vertical fractures cemented by calcite and pyrite are visible in photo
C.

51
3.2.2 Facies A. Muddy Lime Wackestone
Facies A consists of a muddy lime wackestone that contains crinoids and
brachiopod shell fragments (Figure 3.6). This thin bed is immediately above the lower
Bakken shale and could represent the first influx of carbonates into the system. Porosity
is intergranular with calcite and pyrite as the predominant cements. The amount of pyrite
increases as the contact with the lower Bakken is approached. This facies ranges in
thickness from 1 to 5 feet (0.3 to 1.5 meters), averaging 2.9 feet (0.9 meters), which
makes it harder to be picked up by wireline logging tools. When it is thick enough, the
gamma ray reading will show a cleaner signal than the overlying facies B.
Core porosity values for this interval range from 2.7% - 6.1% and average 4.5%.
The permeability values range from 0.0001 mD 0.0057 mD and average 0.0012 mD.

3.2.3 Facies B. Bioturbated, Argillaceous, Calcareous, Very Fine-Grained


Siltstone/Sandstone
Facies B consists of bioturbated, argillaceous, calcareous sandstones and
siltstones (Figure 3.7). Strong bioturbation makes identification of many trace fossils
difficult but some common Helminthopsis/Sclarituba burrow traces have been identified
(Figure 3.8). Some calcite filled vertical fractures are present and local calcareous
concretions are found within the facies. Intergranular and microfracture porosity are
present with calcite cement and some pyrite cement. Thickness of this interval ranges
from 3 to 34 feet (0.9 to 10.4 meters) and averages 20 feet (6.1 meters). This is the
thickest interval in the middle Bakken and is found in all of the cores described. The
facies shows up on wireline logs as a thick section of shaly sand.
Core porosity values for this interval range from 2.2% - 9.8% and average 5.7%.
Permeability values range from 0.0001 mD 0.03 mD and average 0.0015 mD. The
porosity values are favorable for a target zone but because of the homogeneous nature of
this highly bioturbated section permeability measurements are extremely low.

52
Figure 3.6 Core photos of facies A muddy lime wackestone. Crinoid and brachiopod
shell fragments are more concentrated at the base of the facies. This lowermost facies of
the middle Bakken lies directly above the contact with the lower Bakken shale as seen in
photos A and C.

53
Figure 3.7 Core photos of facies B bioturbated, argillaceous, calcareous,
siltstone/sandstone. This is the thickest interval in the middle Bakken and is present in all
of the cores described.

54
Figure 3.8 Core photos showing Helminthopsis/Sclarituba burrow traces found in facies
B.

55
3.2.4 Facies C1. Planar to Undulose Laminated, Shaly, Very Fine-Grained
Siltstone/Sandstone
Facies C1 mainly consists of finely laminated shaly sandstone and siltstone. These
laminations are on the millimeter and centimeter scale (Figure 3.9). There are some wavy
laminated sections which could possibly be microbial influence (or colonization) of the
sediment; however the section is dominated by continuous planar laminations.
Intergranular and minimal amounts of intercrystaline porosity are present. On well logs
this interval is identified by convergence of the neutron porosity and density porosity
curves and by clean (low) gamma ray readings. Thickness of this interval ranges from 2
to 14 feet (0.6 to 4.3 meters) with an average of 7.5 feet (2.3 meters).
Core porosity values for this interval range from 2.5% - 10.3% and average 6.3%.
Permeability values range from 0.0001 mD 0.01 mD and average 0.0026 mD.

3.2.5 Facies C2. Symmetrically Ripple to Undulose Laminated, Very Fine-Grained


Siltstone/Sandstone
Facies C2 consists of finely laminated to rippled, shaly, very fine-grained
sandstone (Figure 3.10). Some hummocky cross stratification is present as well as some
localized bioturbation. The laminations can look crenulated to irregular from possible
microbial influence (or colonization) of the laminations giving it a stromatolitic character.
There is also considerable calcite cement present. Thickness for this interval averages 3
feet (0.9 meters). The thinness of this facies makes it hard to be picked out using wireline
logs.
Core porosity values for this interval range from 2.3% - 7.5% and average 5.4%.
Permeability ranges from 0.0005 mD 0.027 mD and averages 0.0079 mD.

3.2.6 Facies D1. Contorted to Massive, Fine-Grained Sandstone


Facies D1 consists of a muddy, contorted to massive, fine-grained sandstone that
has common micro-faults, microfractures, and slumps representing soft sediment
deformation (Figure 3.11). Some localized microbial influenced (or colonized) sections
have been found in this facies. This facies is only seen in two of the cores described.

56
Figure 3.9 Core photos of facies C1 planar to undulose laminated, shaly, very fine-
grained siltstone/sandstone. This interval is dominated by continuous planar bedding but
there also are sections of wavy laminations. The irregular bedding surfaces may be due to
microbial influence (or colonization) of the sediment.

57
Figure 3.10 Core photos of facies C2 symmetrically ripple to undulose laminated, very
fine-grained siltstone/sandstone. This facies is only present in two cores described with
an average thickness of 3 feet (0.9 meters).

Figure 3.11 Core photos of facies D1 contorted to massive fine-grained sandstone. This
facies has common micro-faults, microfractures, and slumps representing soft sediment
deformation.

58
On well logs, this facies appears just below the cleanest gamma ray readings of facies D2
but are still cleaner than the underlying facies C2. Thickness of this interval ranges from
2 to 5 feet (0.6 to 1.5 meters) observed in two cores.
Core porosity values range from 2.0% - 2.6%, averaging 2.3%. Permeability in
this facies ranges from 0.0003 mD 0.0012 mD and average 0.0008 mD.

3.2.7 Facies D2. Low Angle, Planar to Slightly Undulose, Cross-Laminated


Sandstone with Thin Discontinuous Shale Laminations
Facies D2 consists of a light brown to light grey, parallel to undulating laminated,
low angle cross-laminated sandstone (Figure 3.12). This facies lacks bioturbation and can
be highly cemented by calcite. Some calcite filled fractures are also present. This facies is
present in only four of the cores described. Intergranular porosity is abundant with calcite
as the dominant cement. Thickness of the interval ranges from 0 to 22 feet (0.0 to 6.7
meters) and average 8 feet (2.4 meters) where present. This facies has the cleanest
signature on the gamma ray reading from wireline logs (it has been called the clean
bench) and the neutron and density come together.
Core porosity values range from 2.5% - 12.8%, averaging 4.3% and permeability
ranges from 0.0001 mD 0.055 mD, averaging 0.0042 mD. This facies is considered to
be included as a target interval for horizontal drilling in the Bakken but is mostly absent
in the Parshall Field area.

3.2.8 Facies E1. Finely Inter-Laminated, Bioturbated, Dolomitic-Mudstone and


Dolomitic Siltstone/Sandstone
Facies E consists of an interbedded dark grey, highly bioturbated siltstone, and
light gray, very fine grained, thin parallel laminated sandstone (Figure 3.13). Locally
strong and moderate bioturbation and microbial influence of the sediment is present.
Thickness of the interval ranges from 5 to 11 feet (1.5 to 3.4 meters), averaging 7.5 feet
(2.3 meters). At the base of this facies there is a thin organic-rich mudstone which can be
clearly seen on the gamma ray log as a high gamma ray value and a low bulk density.
Core porosity ranges from 0.5% - 11.3% with an average of 5.7% for this facies.
Permeability values range from 0.0001 mD 0.083 mD and has an average permeability

59
Figure 3.12 Core photos of facies D2 low angle, planar to slightly undulose, cross-
laminated sandstone with thin discontinuous shale laminations. Photo C shows the more
massive limestone unit found in the Dobrinski #18-44 core east of Parshall Field.

60
Figure 3.13 Core photos of facies E1 finely inter-laminated, bioturbated, dolomitic
mudstone and dolomitic siltstone/sandstone. This facies is present in all of the cores
described.

61
of 0.0062 mD. This interval is the most dolomitic zone of the Bakken and is one of the
main targets for drilling.

3.2.9 Facies E2. Calcitic, Whole Fossil, Dolomitic-to Lime Wackestone


Facies E2 consists of a calcitic, whole fossil, dolomitic-to lime wackestone with
fossil-rich beds (Figure 3.14). These fossil-rich beds contain crinoid fragments and
brachiopods found as articulated shells, single valves, or shell fragments. This facies can
be interspersed with facies E1 and is present in all but two of the cores. Thickness of the
interval averages much less than one foot (< 0.1 meters). Only one sample in this facies
had core data showing a porosity of 3.3% and permeability of 0.0308 mD. This facies is
not well detected on wireline logs because the thin beds are below the vertical resolution
of the logs.

3.2.10 Facies F. Bioturbated, Shaly, Dolomitic Siltstone


Facies F consists of a bioturbated, shaly, dolomitic siltstone that lies immediately
below the contact with the upper Bakken shale (Figure 3.15). Little structure is seen
within the facies and the pyrite gives it a lighter yellow, mottled look. This facies is not
present in all of the cores. Thickness of the interval ranges from 0.2 to 3 feet (0.05 to 0.9
meters), averaging 1.5 feet (0.5 meters). This facies is hard to identify on wireline logs
because of its proximity to the highly radioactive overlying shale that overshadow the
signal.
Core porosity and permeability in this interval are low due to closely packed
dolomite crystals. Porosity in this facies ranges from 4.0% - 7.6% with average of 6.0%.
Permeability is very low measured from core, ranging from 0.0005 mD 0.075 mD with
an average of 0.0482 mD.

3.3 Core Descriptions Interpretations


Core descriptions were drawn using lithologic and sedimentary structures
symbols (Figure 3.16). Grain size profiles were derived from hand lens and thin section
observations.

62
Figure 3.14 Core photos of facies E2 calcitic, whole fossil, dolomitic-to lime
wackestone. These fossil-rich beds contain brachiopod shells and crinoid fragments and
may represent storm deposits.

63
Figure 3.15 Core photos of facies F bioturbated, shaly, dolomitic siltstone. This
uppermost facies of the middle Bakken lies directly below the contact with the upper
Bakken shale as seen in photos C and D.

64
Figure 3.16 Core description legend for Figures 3.17 3.26.

65
Both horizontal and vertical fractures were observed mainly in the upper and
lower Bakken shale members from cores and in lesser amounts in the middle member of
the Bakken Formation. Some microfractures were also identified from thin section
observations. Open cracks and fractures are best observed if the cores are wet. The
horizontal fractures observed from core may have formed as the result of hydrocarbon
generation from the rock especially in the upper part of the middle Bakken member due
to the close proximity to the upper shale member with its high organic content.

3.3.1 Van Hook #1-13H (Sec. 13-T152N-R91W)


The Van Hook #1-13H well is located in the middle of the Parshall Field,
Mountrail County, North Dakota (Figure 3.1). The cored interval in this well is from
9,460 feet 9,583 feet (2,883 2,921 meters) and consists of the lower and middle
member of the Bakken Formation (Figure 3.17). Seven facies and subfacies are identified
in this core within the Bakken section: facies G, facies A, facies B, facies C1, facies C2,
facies E1, and facies E2.
Facies G is present in this core only in the lower Bakken member, the upper
Bakken was not cored. The contact between the underlying Three Forks Formation and
lower Bakken shale is not sharp showing that it is conformable here. It consists of a
brown to black mudstone with some interbedded limestone units. Some fractures are
present and are filled with calcite and pyrite. Facies G is interpreted to have formed
during anoxic marine conditions indicated by the presence of pyrite, high organic matter,
and rare benthic fauna. Facies A consists of a skeletal lime wackestone with abundant
crinoids and shell fragments that are filled with pyrite. The contact between facies A and
the lower Bakken shale (facies G) is erosional with a shell lag surface present.
Facies B consists of an argillaceous, calcareous, bioturbated, very fine-grained
sandstone to siltstone. Helminthopsis/Sclarituba is the dominate trace fossil in this facies,
with possible local Planolites. Brachiopod and crinoid shell fragments are present as well
as some small calcite-filled vertical fractures. The contact between the underlying facies
A and facies B is transitional here.

66
Figure 3.17 Core description of the Van Hook #1-13H.

67
Facies C1 and C2 is planar to undulatory laminated interval that is highly
calcareous. There is a transitional contact between facies B and facies C1. Local calcite
concretions are present throughout this facies. Further up this section, the dominantly
parallel laminations grade into more symmetrically ripple-laminated. Facies C2 shows
the presence of wave activity and facies C1 shows possible hummocky cross stratification
suggesting shallower water depth than the underlying facies. There is evidence for some
tidal influence in the rhythmically laminated beds.
Facies E1 consists of locally discontinuous, parallel to undulose laminated, very
fine-grained sandstone to siltstone. Soft sediment deformation, including slump features
also occurs in this facies. The contact with facies C2 is non conformable with the base
represented by a tempestite deposit. This deposit is indicative of a major storm event that
occurred in shallow water conditions because these sediments are disturbed and re-
deposited by the energy of waves. The very top of this core consists of facies E2; a fossil
rich storm lag surface and a calcareous wackestone that has shell fragments.

3.3.2 N & D #1-05H (Sec. 5-T152N-R90W)


This core is located in the middle of Parshall Field (Figure 3.1). The cored
interval in this well is from 9,393 feet 9,485 feet (2,863 2,891 meters) and consists of
the upper, middle, and lower members of the Bakken Formation (Figure 3.18). Six facies
and subfacies are identified in this core within the Bakken section: facies G, facies A,
facies B, facies E1, facies E2, and facies F.
Facies A consists of a lime skeletal wackestone with crinoid and brachiopod shell
fragments. The contact with the lower Bakken (facies G) is not sharp with a lag surface
present. Facies B consists of an argillaceous, calcareous, bioturbated very fine-grained
sandstone with some shell fragments. There are sparse horizontal and vertical calcite-
filled fractures present. The contact with facies is A is gradational.
Facies E1 consists of parallel interbeds of bioturbated sandstones and argillaceous
siltstone. The contact with the underlying facies B appears to be sharp but there might be
core missing here. Sparse vertical fractures are present here. There are fossil rich beds
present as well (facies E2) interbedded with facies E1. Facies E1 becomes muddier up

68
Figure 3.18 Core description of the N & D #1-05H.

69
section and less bioturbated. The contact between facies F and the upper Bakken shale
(facies G) is gradational over a foot (0.3 meters).

3.3.3 Parshall #2-36H (Sec. 36-T153N-R90W)


This core is located in the middle of Parshall Field (Figure 3.1). The cored
interval in this well is from 9,248 feet 9,306 feet (2,819 2,836 meters) and consists of
the middle and upper members of the Bakken Formation (Figure 3.19). Six facies and
subfacies are identified in this core within the Bakken section: facies A, facies B, facies
E1, facies E2, facies F, and facies G.
Facies A consists of an intraclastic skeletal lime wackestone and is approximately
one foot (0.3 meters) thick. There was no lower Bakken (facies G) cored so the lower
contact could not be viewed. Facies B conformably overlies facies A and consists of
bioturbated, argillaceous, calcareous sandstone to siltstone with common Helminthopsis/
Sclarituba burrow traces. A couple of horizontal fractures were also identified in the
facies. The contact between facies B and facies E1 is non-conformable. Facies E1
consists of some soft sediment deformation structures including microfaults and sand
injection features. Localized symmetrical ripples have also been identified in sandier
intervals near the top of the facies. Vertical fractures are present but a very sparse in this
facies. Facies E2 is intermixed with facies E1 and consists of fossil lag surfaces. The
contact between facies E and facies F is gradational. Facies F is present here but it is only
0.2 feet (< 0.1 meters) thick. The contact with the upper Bakken shale (facies G) is
gradational.

3.3.4 Hoff #1-10H (Sec. 10-T152N-R90W)


This core was taken in a horizontal well located in the middle of Parshall Field
(Figure 3.1). The facies identified from this core is: facies E1 (Figure 3.20). No contacts
were present in the core of the overlying or underlying facies because this horizontal core
only cored facies E1. Facies E1 consists of finely, planar to undulose laminated,
dolomitic siltstones interlaminated with dolomitic mudstones. Local bioturbation is
present as well as some soft sediment deformation features and contorted bedding.

70
Figure 3.19 Core description of the Parshall #2-36H.

71
Figure 3.20 Core description of the Hoff #1-10H. This horizontal well was described
from left to right. The red box indicates the location of the core with respect to the well
path taken. The 40 feet (12.2 meters) of core described represents approximately 5 feet
(1.5 meters) of vertical succession.

72
Calcite-filled horizontal and vertical fractures are present throughout the core. Laminated
intervals appear to be influenced by microbial activity (or colonization). An algal interval
was preserved at 9,535 feet (2,906 meters).

3.3.5 Long #1-01H (Sec. 1-T152N-R90W)


This core is located to the north of the Fertile #1-12H core and is in the southern
part of Parshall Field (Figure 3.1). The cored interval in this well is from 9,135 feet
9,199 feet (2,784 2,804 meters) and consists of the lower, middle, and upper members
of the Bakken Formation (Figure 3.21). Eight facies and subfacies are identified in this
core within the Bakken section: facies G, facies A, facies B, facies C1, facies D2, facies
E1, facies E2, and facies F.
Facies A is a skeletal mud lime wackestone and shows a fairly basal scour
contact with the lower Bakken shale (facies G). This facies has a gradational contact with
the overlying facies B. Facies B is an argillaceous, bioturbated, very fine-grained
sandstone/siltstone that has some small calcite-filled horizontal fractures. The main type
of burrow traces here are Helminthopsis/Sclarituba and with possible Planolites traces.
The contact between facies B and the overlying, very thin section of facies C1 and
D2 is sharp at a depth of 9,151 feet (2,789 meters). This very thin section (0.08 feet; < 0.1
meters) of planar laminated muddy fine-grained sandstone (facies C1) is followed by
another very thin section (0.2 feet; < 0.1 meters) of low angle, laminated sandstone
(facies D2). The contact between facies D2 and the overlying facies E1 appears to be
gradational as burrowing decreases as you move up the core and becomes more
laminated (possibly microbial influenced). Fossil rich beds (facies E2) occur within facies
E1 and are more frequent at the top of the section. Facies F is present in the top 2 feet
(0.6 meters) of the middle Bakken member and has an erosional contact with the
overlying upper Bakken shale (facies G) and a gradational contact with the underlying
facies E.

3.3.6 Patten #1-27H (Sec. 27-T153N-R89W)


This core is located on the eastern side of Parshall Field (Figure 3.1). The cored
interval in this well is from 9,054 feet 9,086 feet (2,760 2,769 meters) (Figure 3.22).

73
Figure 3.21 Core description of the Long #1-01H.

74
Figure 3.22 Core description of the Patten #1-27H.

75
Only four facies and subfacies are identified in this core within the Bakken section: facies
D2, facies E1, facies F and facies G.
Facies D2 is different here than it is in other cores. Here facies D2 consists of
alternating units of cross-bedded, bioclast-rich, very fine- to fine-grained sandstone and
sandy skeletal lime grainstone. There are some vertical fractures present that are filled
with calcite cement at 9,077 feet (2,767 meters).
The contact between facies D2 and E1 is relatively sharp. Facies E1 and facies F
are similar to other cores showing bioturbation as well as algal interval development. The
contact between facies F and the upper Bakken shale (facies G) appears quite sharp but it
appears that some bits of core might be missing.

3.3.7 Fertile #1-12H (Sec. 12-T151N-R90W)


The Fertile #1-12H well is located to the southeast of the Van Hook #1-13H well
(Figure 3.1). The cored interval in this well is from 9,365 feet 9,421 feet (2,854 2,872
meters) and consists of the lower, middle, and upper members of the Bakken Formation
(Figure 3.23). Six facies and subfacies are identified in this core within the Bakken
section: facies G, facies A, facies B, facies E1, facies E2, and facies F.
Facies G is present in both the upper and lower Bakken. The contact with the
underlying Three Forks Formation is not present in this core. The contact between facies
A and the lower Bakken shale (facies G) is erosional with a shell lag surface present.
Facies B are very similar to the facies described in the Van Hook #1-13H core and has a
transitional contact with the underlying facies A. Unlike the Van Hook #1-13H core,
calcite filled fractures are absent in facies B.
Facies E1 unconformably overlies facies B and consists of finely, planar to
undulose laminations that are microbial influenced. Several fossil rich lags are also
deposited in this core in facies E2 intermixed with facies E1. Facies F consists of a
bioturbated, shaly, dolomitic sandstone and has a sharp basal contact with facies E1. The
contact between facies F and the upper Bakken (facies G) show a scoured surface as the
depositional environment became deeper.

76
Figure 3.23 Core description of the Fertile #1-12H.

77
3.3.8 Jensen #12-44 (Sec. 12-T154N-R90W)
This core is located to the northern part of Parshall Field in Mountrail County,
North Dakota (Figure 3.1). The cored interval in this well is from 9,157 feet 9,216 feet
(2,791 2,809 meters) and consists of the middle member of the Bakken Formation
(Figure 3.24). Six facies and subfacies are identified in this core within the Bakken
section: facies G, facies A, facies B, facies E1, facies E2, and facies F.
These facies are very similar to the other cores described in this study. Facies G is
a pyritic planar laminated brown mudstone. Facies A is a muddy lime wackestone and
has an erosional contact with the underlying facies G. Facies B is a bioturbated,
argillaceous, very fine-grained sandstone/siltstone and the contact with the underlying
facies A is transitional. The contact between facies B and the overlying facies E1 is
erosional.
Facies E1 is an argillaceous, very fine-grained sandstone and has alternating beds
of bioturbated siltstones and planar laminated siltstone. Facies E2 is a calcitic, whole
fossil wackestone and is interbedded with facies E1. Facies F is a bioturbated, shaly,
dolomitic siltstone and has a transitional basal contact with facies E1 and an erosional
contact with the overlying upper Bakken (facies G).

3.3.9 Deadwood Canyon Ranch #43-28H (Sec. 28-T154N-R92W)


This core is located to the west of Parshall Field, in Sanish Field (Figure 3.1). The
cored interval in this well is from 10,077 feet 10,268 feet (3,071 3,130 meters) and
consists of the middle and lower members of the Bakken Formation as well as the Three
Forks Formation (Figure 3.25a and Figure 3.25b). Nine facies and subfacies are identified
in this core within the Bakken section: facies G, facies A, facies B, facies C1, facies C2,
facies D1, facies D2, facies E1, and facies E2.
Facies G is present at the bottom of the core representing the lower Bakken shale
member. It consists of a laminated, black to brown, organic rich mudstone. Facies A is a
muddy lime wackestone that contains crinoids and brachiopod shell fragments. The
contact between facies G and facies A is not present in the core. Facies B consists of
bioturbated, argillaceous sandstones and siltstones. Strong bioturbation makes
identification of many trace fossils very hard but some common Helminthopsis/

78
Figure 3.24 Core description of the Jensen #12-44.

79
Figure 3.25a Description of the uppermost part of the Deadwood Canyon Ranch #43-
28H.

80
Figure 3.25b Description of the lowermost part of the Deadwood Canyon Ranch #43-
28H.

81
Sclarituba burrow traces have been identified. A few calcite-filled horizontal fractures
were found in this facies. Facies B has a transitional contact with the underlying facies A.
Facies C1 mainly consists of finely planar to irregular laminated (millimeter to
centimeter scale) shaly sandstone and siltstone. Facies C2 is a finely laminated to
undulose, shaly, fine-grained sandstone. Local hummocky cross stratification is present.
Facies D1 is a muddy, contorted to massive, fine-grained sandstone that has common
micro-faults, microfractures, and slumps representing soft sediment deformation. Facies
D2 is composed of light brown to light grey, parallel- to undulating-laminated, low angle
cross-stratified, fine-grained sandstone. The contacts between facies B, C1, C2, and D1
are all gradational. The contact between facies D1 and D2 is sharp contact. Facies E1 is
an interbedded dark grey, highly bioturbated siltstone, and light gray, very fine-grained,
thin parallel-laminated sandstone. Facies E2 is a fossil rich bed that is located at the top
of the core. The contact with the overlying upper Bakken is not present in this core.

3.3.10 Dobrinski #18-44 (Sec. 18-T151N-R87W)


This core is located to the east of the Parshall Field in Ward County, North
Dakota (Figure 3.1). The cored interval in this well is from 8,629 feet 8,667 feet (2,630
2,642 meters) and consists of the upper, middle, and lower members of the Bakken
Formation (Figure 3.26). This core has a much thinner middle member section compared
to the other cores. Eight facies and subfacies are identified in this core within the Bakken
section: facies G, facies A, facies B, facies D1, facies D2, facies E1, facies E2, and facies
F.
Facies G is present at the base of the core and consists of brown to black, planar
laminated mudstone. The contact with the overlying facies A is sharp. Facies A is a
skeletal lime wackestone and grades into facies B, a highly bioturbated very fine-grained
sandstone to siltstone with common Helminthopsis/Sclarituba burrow traces. This core is
missing the C facies (both parts) even though both parts of the D facies are present. The
contact between facies B and facies D1 is also sharp.
The D facies is slightly different from the other cores described. Facies D1 is still
a contorted to massive facies that shows soft sediment deformation, but facies D2 appears
to be a much more massive limestone unit with some faint planar laminations. This is

82
Figure 3.26 Core description of the Dobrinski #18-44.

83
overlain by facies E1 with a sharp contact at the base. The irregularly laminated siltstones
and mudstones of facies E1 is intermixed with a few fossil rich beds of facies E2. This
then grades up into facies F, a bioturbated, shaly, dolomitic siltstone. This is overlain by
the upper Bakken (facies G) that is mostly brown with some black and is planar
laminated with very few fractures present. The contact with the underlying facies F is
sharp.

3.4 Mineralogy and Diagenesis


Mineralogy of the Bakken Formation in the Parshall Field area was analyzed with
a focus on how it changes with present depth of burial. Thin sections from the Deadwood
Canyon Ranch #43-28H core were examined and interpreted for general framework
mineralogy and pore structure. Mineralogical analysis was also obtained from twelve
core plugs using QEMSCAN technology (Figure 3.27).
One core plug was analyzed from the lower Bakken member (facies G) just below
the contact with the middle member (10,146.12 feet) (3,092.5 meters). This core plug
consists mainly of quartz and illite/smectite with some smaller amounts of feldspars and
carbonates. The high amount of quartz present in the lower Bakken is thought to come
from radiolarians present in the facies. Four core plugs were taken from facies B and
consist mainly of quartz, calcite, dolomite, and illite/smectite (10,120.07 10,135.99
feet) (3,084.6 3,089.4 meters). One of these core plugs also showed almost five percent
pyrite from the presence of pyrite filled fractures (Figure 3.28).
Two core plugs were taken from facies C1 and have similar mineralogy to that of
facies B (10,109.72 feet and 10,113.57 feet) (3,081.4 meters and 3,082.6 meters). There
are slightly greater amounts of quartz present in this facies. One core plug was taken from
facies D2 and shows the highest amount of quartz in the middle Bakken as well as calcite
(10,095.87 feet) (3,077.2 meters). Four core plugs were taken from facies E (10,079.1
10,083.07 feet) (3,072.1 3,073.3 meters). These samples show a decrease in quartz and
increase in dolomite content towards the top of the middle Bakken. The most dolomitic
sections are found in facies E1 near the top of the core.

84
Figure 3.27 Mineralogical analysis measured from QEMSCAN for the Deadwood
Canyon Ranch #43-28H core samples, Mountrail County, North Dakota.

85
Figure 3.28 QEMSCAN image of the core plug taken from the Deadwood Canyon
Ranch #43-28H at 10,134.02 feet. The pyrite-filled fractures are shown in yellow.

86
3.4.1 Diagenetic Features
A generalized summary of diagenetic features observed from the sandstones and
siltstones of the middle member are shown in Figure 3.29. Pitman et al. (2001)
reconstructed burial and thermal history curves of the Bakken Formation and showed
relative timing of major diagenetic events in the deep part of the Williston Basin (Figure
3.30). Postdepositional alteration in the middle Bakken includes mechanical and chemical
compaction, diagenetic cementation, and mineral dissolution. A detailed look at the
diagenetic features of the Parshall Field is not addressed in this thesis.

3.5 Reservoir Quality


The most common controls on sandstone porosity and permeability in
unconventional reservoir rocks are thermal maturity, framework grain composition, and
sorting (Pitman et al., 2001). The most important influence on the reservoir quality for
Parshall Field appears to be the thermal maturity of the adjacent shales. There are no
significant influences from grain sorting and composition because there are only minor
variations between wells.
Core porosity and permeabilities for the middle Bakken in Parshall Field are
typically very low. Porosity values range from 1.0% 13% (average about 5.0%) and
permeability values range from 0.0001mD 1.9mD (fractured sample) (average
0.03mD). These characteristics suggest that something else is aiding in the reservoir
quality.
The upper middle Bakken facies have a higher dolomite content than the lower
middle Bakken facies and show slightly better average porosity values (facies E and F).
The upper middle Bakken facies also have more fractures and microfractures than the
lower members. These fractures locally enhance the quality of the reservoir by enhancing
the permeability.
The thermal maturity of the Bakken shales adjacent to the sandstones and
siltstones can vary widely and thus can have an important influence on the reservoir
quality. The thermal maturity for Parshall Field is addressed later in chapter 5.

87
Figure 3.29 Chart showing the postdepositional events observed in sandstones and
siltstones of the middle member of the Bakken Formation (from Pitman et al., 2001).

Figure 3.30 Burial and thermal history curve of the Bakken Formation showing relative
timing of major diagenetic events (Pitman et al., 2001). The term ogs represents
overgrowths.

88
CHAPTER 4
SUBSURFACE WELL LOG INTERPRETATION AND DEPOSITIONAL MODEL

Raster logs were acquired from 13 vertical wells in the Parshall Field and 913
wells in the total study area (Figure 4.1) with the goal of correlating and mapping facies
determined from core interpretations. Of the 913 total wells, 682 wells had density-
neutron porosity logs, 913 wells had gamma ray logs, 625 wells had sonic logs, and 895
wells had resistivity logs. All of the raster logs were obtained from MJ Systems and from
the North Dakota Geological Surveys website at: https://www.dmr.nd.gov-oilgas/. Well
logs are not typically run in the newer horizontal wells in Parshall Field but were run in
the vertical pilot holes prior to horizontal drilling. Wells that did not penetrate the
complete Bakken Formation were not used in this study.

4.1 Methods
All available digital raster copies of the geophysical logs were loaded into the
geologic mapping and correlation software, Petra, and key geologic information, such as
operator tops, test intervals, cores, production data, etc., were entered in the appropriate
sections of the database. The Deadwood Canyon Ranch #43-28H was chosen as the main
type well because of its modern log suite and completeness of the middle Bakken
member facies (Figure 4.2). For correlation purposes facies E and F were combined into
one interval due to the intermixing of facies E1 and facies E2 and the relative thinness of
facies F which is commonly overshadowed by the high radioactivity of the overlying
shale. Facies D includes facies D1 and D2 and facies C includes facies C1 and C2. The
upper Bakken member, lower Bakken member, Three Forks Formation, and the
underlying Nisku (Birdbear) Formation tops were also picked.
These formation tops were picked in the 913 wells with rasters and used to create
subsurface structure and isopach maps. Several north-south and west-east cross sections
were constructed to determine subsurface structure and stratigraphic thicknesses from
each unit.

89
Figure 4.1 All wells in the study area. Rasters used are highlighted in pink. Parshall Field
is the outline on the right and Sanish Field is just to the east. The field outlines were
taken from the North Dakota Industrial Commission, Department of Mineral Resources,
Oil and Gas Division website.

90
Figure 4.2 The Deadwood Canyon Ranch #43-28H well from Sanish Field is the type
well used for correlation. Facies E and F are combined into one interval, facies D
includes facies D1 and D2, and facies includes facies C1 and C2.

91
4.2 Construction of the Cross Sections
Seven structural and nine stratigraphic cross sections were built for this study
which were used to determine structural variations and depositional trends. Wells were
chosen based on location and wells that had both resistivity rasters and neutron-density
rasters. The depositional model was created from log facies correlation with the core
facies interpretation (see Chapter 3).

4.2.1 Structural Cross Sections


Seven structural cross sections across the Parshall Field area, oriented north-south
and west-east, were constructed to understand the structural framework of the Bakken
and adjacent formations in the Parshall Field area (Figure 4.3). Cross sections 1-1, 2-2,
3-3, and 4-4 have west-east orientation across the Parshall Field area (Figure 4.4, 4.5,
4.6, and 4.7). Cross sections 5-5, 6-6, and 7-7 are orientated north-south (Figure 4.8,
4.9, and 4.10).
Structural cross sections and maps show that the Bakken members dip toward the
west of Parshall Field at an average rate of approximately 65 feet per mile. No major
faults were indicated during log correlation. Structural cross sections also show that the
Bakken members thin to the east of the field as they get closer to the edge of the basin.

4.2.2 Stratigraphic Cross Sections


Nine stratigraphic cross sections were built for this study (Figure 4.11). These
stratigraphic cross sections were correlated to understand changes in middle Bakken
deposition as well as understand the thickness variations and facies variations within the
Bakken members and their vertical and lateral extents. Those nine stratigraphic cross
sections are named: A-A, B-B, C-C, D-D, E-E, F-F, G-G, H-H, and I-I (Figure
4.12 through Figure 4.20). Cross sections A-A through G-G are the same wells chosen
for the structure cross sections 1-1 through 7-7.
The top of the upper Bakken was selected as the datum of the stratigraphic cross
sections. It is clear that the middle member pinches-out toward the east of the field,
shown by both structural and stratigraphic cross sections and subsurface mapping.

92
Figure 4.3 Structure section lines shown in red. Base map is the structure on the top of
the upper Bakken shale. Parshall Field (to the east) and Sanish Field (to the west) are
outlined in black.

93
94
Figure 4.4 West-east oriented structural cross section 1-1 showing structural dip toward the west. This section is north of
Parshall Field. Location of the cross section is shown in Figure 4.3.
95
Figure 4.5 West-east oriented structural cross section 2-2 showing structural dip toward the west. This section crosses the
northern part of Parshall Field and northern part of Sanish Field. Location of the cross section is shown in Figure 4.3.
96
Figure 4.6 West-east oriented structural cross section 3-3 showing structural dip toward the west. This section crosses the
middle of Parshall Field and the southern part of Sanish Field. Location of the cross section is shown in Figure 4.3.
97
Figure 4.7 West-east oriented structural cross section 4-4 showing structural dip toward the west. This section crosses the
southern part of Parshall Field. Location of the cross section is shown in Figure 4.3.
98
Figure 4.8 North-south oriented structural cross section 5-5 drawn generally along structure strikes (the section has a high
vertical exaggeration). This section is west of Parshall Field and crosses through the western side of Sanish Field. Location of
the cross section is shown in Figure 4.3.
99
Figure 4.9 North-south oriented structural cross section 6-6 drawn generally along structural strike (this section has a high
vertical exaggeration). This section runs through the center of Parshall Field and is almost parallel to the structural contour
lines of the upper Bakken. Location of the cross section is shown in Figure 4.3.
100
Figure 4.10 North-south oriented structural cross section 7-7 drawn generally along structural strikes (the section has a high
vertical exaggeration). This section is east of Parshall Field. Location of the cross section is shown in Figure 4.3.
Figure 4.11 Map showing the location of nine stratigraphic sections. Base map is the
middle Bakken thickness map (in feet). Parshall Field (to the east) and Sanish Field (to
the west) are outlined in black.

101
102
Figure 4.12 West-east oriented stratigraphic cross section A-A' showing the variations in the middle Bakken thickness as it
thins to the east. Facies C is completely absent in the east and facies D is not present in some of the wells. This section is to
the north of Parshall Field. Location of the cross section is shown in Figure 4.11.
103
Figure 4.13 West-east oriented stratigraphic cross section B-B' showing the variations in the middle Bakken thickness as it
thins to the east. Facies C is completely absent and facies D is not present in some of the wells in Parshall Field. Location of
the cross section is shown in Figure 4.11.
104
Figure 4.14 West-east oriented stratigraphic cross section C-C' showing the variations in the middle Bakken thickness as it
thins to the east. Facies C and facies D is only in one of the Parshall Field wells shown here. Location of the cross section is
shown in Figure 4.11.
105
Figure 4.15 West-east oriented stratigraphic cross section D-D' showing the variations in the middle Bakken thickness as it
thins to the east. Facies C and facies D are present west of the southern part of Parshall Field. Facies D is present in the
Dobrinski #18-44 well east of Parshall Field. Location of the cross section is shown in Figure 4.11.
106
Figure 4.16 North-south oriented stratigraphic cross section E-E' showing the variations in the middle Bakken thickness as it
thins to the south. Facies C and facies D are present in all of these wells west of Parshall Field. Location of the cross section
is shown in Figure 4.11.
107
Figure 4.17 North-south oriented stratigraphic cross section F-F' showing the variations in the middle Bakken thickness as it
thins slightly to the south. Facies D is only present in the Laredo #26-1 well north of Parshall Field. This cross section runs
through the center of Parshall Field. Location of the cross section is shown in Figure 4.11.
108
Figure 4.18 North-south oriented stratigraphic cross section G-G' showing the variations in the middle Bakken thickness. The
middle Bakken thickens slightly and then thins to the south. This cross section is east of Parshall Field and shows an increase
in Facies B to the south. Location of the cross section is shown in Figure 4.11.
109
Figure 4.19 Northwest-southeast oriented stratigraphic cross section H-H' showing the variations in the middle Bakken
thickness as it thins to the southeast. This section crosses the eastern part of Sanish Field and the center of Parshall Field.
Facies C pinches out into Parshall Field and is only present in the Parshall SD #1 well which also has facies D. The lower
Bakken is also thinning to the southeast. Location of the cross section is shown in Figure 4.11.
110
Figure 4.20 Southwest-northeast oriented stratigraphic cross section I-I' showing the variations in the middle Bakken
thickness as it thins to the northeast. This section crosses the center of Parshall Field and just south of Sanish Field. Facies C
pinches out into Parshall Field and is only present in the Van Hook #1-13H well and the Parshall SD #1 well which also has
facies D. The lower Bakken is also thinning to the northeast. Location of the cross section is shown in Figure 4.11.
Stratigraphic cross sections also show that the thickness of the middle member is fairly
constant in the center of the Parshall Field and thins toward the east of the field.
Looking more closely, facies E and F are relatively a constant thickness in the
Parshall Field area. The core description in the Deadwood Canyon Ranch #43-28H shows
that this facies occurs at the uppermost part of the middle Bakken member. Facies D and
facies C appear to be largely confined to paleolow areas and is mainly absent in Parshall
Field. Facies B is present in all of the wells and varies in thickness but it does pinchout to
the south and east. Facies A is very thin and is not always present on the well logs.
The lower Bakken, middle Bakken, and upper Bakken appear to onlap and thin
out toward the east onto the Three Forks Formation. The Bakken and the Three Forks
have an unconformable contact relationship. LeFever et al. (1991) interpreted this contact
to mean that the Three Forks was exposed to erosion before the lower Bakken shale was
deposited.

4.3 Structure Maps


Structure maps were constructed for the Bakken, Three Forks, Nisku, and Prairie
Evaporite formations (Figure 4.21 Figure 4.26). The Bakken, Three Forks, and Nisku
formations all dip toward the west southwest with similar structure dip within the
Parshall Field (approximately 65 feet per mile). No obvious faults are present based on
the log interpretations. However, faulting is reported in the literature (Murray, 1968)
principally along the flanks of the Nesson Anticline and Antelope Field. The Nesson
Anticline structure is located to the west of the Parshall Field. This structural feature can
be seen clearly from the Bakken, Three Forks, Nisku, and Prairie structure maps.

4.4 Isopach Maps


The depositional pattern of the Bakken Formation, as related to the overlying and
underlying strata, is by nine interval thickness maps (Figure 4.27 4.36). The Bakken
Formation in Parshall Field ranges in thickness from 70 feet 100 feet (21.3 meters
30.5 meters) with increasing thickness to the west (Figure 4.27). This trend is similar to
the isopach of the middle Bakken member where the thickest middle member interval is
also thickening to the west.

111
Figure 4.21 Structure map on top of the upper Bakken shale. The contour interval is 250
feet (80 meters).

112
Figure 4.22 Structure map on top of the middle member of the Bakken Formation. The
contour interval is 250 feet (80 meters).

113
Figure 4.23 Structure map on top of the lower Bakken shale. The contour interval is 250
feet (80 meters).

114
Figure 4.24 Structure map on top of the Three Forks Formation. The contour interval is
250 feet (80 meters).

115
Figure 4.25 Structure map on top of the Nisku (Birdbear) Formation. The contour interval
is 250 feet (80 meters).

116
Figure 4.26 Structure map on top of the Prairie Formation. The contour interval is 250
feet (80 meters).

117
Figure 4.27 Thickness map of the total Bakken Formation. Contour interval is 10 feet (3
meters). The thickness of the Bakken Formation ranges from 70 feet 100 feet (21.3
meters 30.5 meters) in Parshall Field.

118
The upper Bakken shale has an average thickness of 18 feet (5.5 meters) which
increases towards the west and thins to the east (Figure 4.28). Within Parshall Field this
interval is relatively constant with some slight thickening in an area in the southeast part
of the field.
The middle Bakken member thins to the east of Parshall Field and thickens to the
west (Figure 4.29). The maximum thickness of the middle member of the Bakken
Formation in Parshall Field is 50 feet (15.2 meters) and has a minimum thickness of
approximately 30 feet (9.1 meters). Just to the west of Parshall Field in Sanish Field
(where the Deadwood Canyon Ranch #43-28H core is located) the middle member is
almost 70 feet (21.3 meters) thick. Facies E and F of the middle member averages 12 feet
(3.7 meters) thick and is relatively constant in Parshall Field (Figure 4.30). This interval
does show some thickening to the north of Parshall Field. Facies D is almost entirely
absent from Parshall Field (Figure 4.31). This interval is present in most of Sanish Field
and to the north of Parshall Field. Facies C pinches out in Parshall Field and is thinning to
the east through Sanish Field (Figure 4.32). Facies B is the thickest interval of the middle
Bakken member in Parshall Field with a maximum thickness of 25 feet (7.6 meters)
(Figure 4.33). Facies A is the thinnest interval in the middle Bakken with an average
thickness of 2.9 feet (0.9 meters). The identification of facies A on wireline logs is
dependent on the presence of the lower Bakken shale. The wireline signature for facies A
is a small response on the gamma-ray signature that is sometimes difficult to identify on
older logs where the shale signature is less defined (Figure 4.2). Facies A and B were
combined in thickness mapping because of this issue.
The lower Bakken shale is approximately 30 feet (9.1 meters) thick in the Parshall
Field area but increases sharply in thickness to the west up to approximately 50 feet (15.2
meters) (Figure 4.34). The depocenter for the lower Bakken is elongate north to south and
is located west of Parshall Field.
The thickness of the Three Forks Formation in Parshall Field ranges from 222 feet
242 feet (67.7 meters 73.8 meters) (Figure 4.35). The Three Forks Formation thins
east of the field, is relatively constant immediately to the west of the field, and begins to
thin further west of Sanish Field. The depocenter for the Three Forks Formation is
oriented north-south and lies just west of Parshall Field.

119
Figure 4.28 Thickness map of the upper Bakken shale (facies G). Contour interval is 3
feet (0.9 meters). The thickness of the upper Bakken member averages 18 feet (5.5
meters) in Parshall Field.

120
Figure 4.29 Thickness map of the middle member (facies F, E, D, C, B, and A) of the
Bakken Formation. Contour interval is 10 feet (3 meters). The thickness of the upper
Bakken member ranges from 30 feet 50 feet (9.1 meters 15.2 meters) in Parshall
Field.

121
Figure 4.30 Thickness map of facies E and F of the middle member of the Bakken
Formation. Contour interval is 5 feet (1.5 meters). The thickness of facies E and F
averages 12 feet (3.7 meters) in Parshall Field.

122
Figure 4.31 Thickness map of facies D of the middle member of the Bakken Formation.
Contour interval is 2 feet (0.6 meters). Facies D is mainly absent in Parshall Field.

123
Figure 4.32 Thickness map of facies C of the middle member of the Bakken Formation.
Contour interval is 2 feet (0.6 meters). Facies C pinches out in Parshall Field.

124
Figure 4.33 Thickness map of facies A and B of the middle member of the Bakken
Formation. Contour interval is 5 feet (1.5 meters). The combined thickness of Facies A
and B ranges from 17 feet 25 feet (5.2 meters 7.6 meters) in Parshall Field.

125
Figure 4.34 Thickness map of the lower Bakken shale (facies G). Contour interval is 5
feet (1.5 meters). The thickness of the lower Bakken member ranges from 20 feet 37
feet (6.1 meters 11.3 meters) in Parshall Field.

126
Figure 4.35 Thickness map of the Three Forks Formation. Contour interval is 10 feet (3
meters). The thickness of the Three Forks Formation ranges from 222 feet 242 feet
(67.7 meters 73.8 meters) in Parshall Field.

127
A thickness map was constructed from the top of the middle member of the
Bakken Formation to the top of the Prairie Formation to see if there were any variations
that could have been caused by salt dissolution within the Prairie (Figure 4.36). No
significant thicks or thins were identified in Parshall Field.

4.5 Depositional Model


The lower and upper members of the Bakken Formation are interpreted to have
been deposited in an offshore marine environment during periods of sea-level rise
(Webster, 1984; LeFever et al., 1991; Smith and Bustin, 1996). The Devonian sea had
withdrawn to the limit of the lower Bakken shale caused by major uplift and erosion
around the basin margin. The deposition of the upper and lower shale members are
interpreted to have been deposited in a stratified hydrologic regime under anoxic bottom-
water conditions (Webster, 1984; Lineback and Davidson, 1982). Anoxic conditions are
indicated by the presence of pyrite, absence of bioturbation, and high organic-matter
content.
The depositional model interpreted for the middle Bakken in the Parshall Field is
an offshore to tidally influenced inner shelf environment. Facies A is interpreted to have
been deposited in an offshore marine environment. Wave and current features are not
present and there appears to be suspension deposition as well as deposit-feeder traces in
this facies. The contact between facies A and the lower Bakken shale (facies G) is
erosional with a shell lag surface present. Christopher (1961) proposed an interpretation
of this facies being the initial transgression following a sea-level drop based on the rapid
deposition of the clays and silts, and the preservation of abundant brachiopods. Facies B
is interpreted to have been deposited in a lower shoreface setting. This highly bioturbated
zone is reworked by deposit feeding ichnofauna and lacks evidence for wave activity.
Facies C is interpreted to have been deposited in a lower - middle shoreface environment.
The interlaminated sandstone and muddy siltstone may represent tidal processes with
microbial influence. Some evidence of wave activity is present in symmetrically ripple
laminated sections and possible hummocky cross stratification is also present. The
contacts between facies A and facies B and the contact between facies B and facies C is
gradational suggesting that they are all part of the same system. These facies represent

128
Figure 4.36 Thickness map from the top of the middle member of the Bakken Formation
to the top of the Prairie Formation. Contour interval is 50 feet (15.2 meters).

129
a highstand systems tract where relative sea level was constant, there was low
sedimentation, and there was lots of bioturbation (Figure 4.37).
The contact between the underlying facies C and facies D is an erosional to non-
depositional sequence boundary. This represents the beginning of the fall of sea level in
the basin. Facies D is interpreted to have been deposited in a middle shoreface
environment. This facies is the coarsest grained and only shows up in a two of the cores
described. It is a possible distributary channel that is tidally influenced. Mapping of this
facies suggests that the sediment source came from areas north and east of the field and
were deposited as longshore drift sediments near the coastline (Figure 4.31). As sea level
was falling, sedimentation rates increased resulting in the lack of bioturbation in this
facies. This represents the lowstand systems tract of the middle Bakken (Figure 4.37).
The top of facies D is marked by a flooding surface as relative sea level began to rise.
Facies E1 is interpreted to have been deposited in a lower shoreface environment.
The sandstone beds may be interpreted as tempestites that are strongly reworked by
biogenic action and represent storm deposits. Facies E2, which is interbedded with facies
E1, are fossil rich beds that may also represent storm deposits. The contact between
facies E and F is gradational when present in the cores. Facies F is interpreted to have
been deposited in an offshore environment reflecting the transgression of the upper
Bakken Sea. Both facies E and F of the middle Bakken together with the upper Bakken
shale (facies G) represents a transgressive systems tract with a rapid sea level rise, low
sedimentation rates, and higher amounts of bioturbation present then the underlying
facies (Figure 4.37).
The sediment source (erosional products from older formations) comes from areas
north of the field. The carbonate and silt-sized sediments are believed to be products of
erosion and deposition related to accommodation space in the Parshall Field area.

130
Figure 4.37 Environment of deposition interpretation of the middle Bakken. Facies are
label and colored according to their position on the model. HST (highstand systems
tract), LST (lowstand systems tract), and TST (transgressive systems tract) are labeled
with their corresponding facies associations (modified from Smith and Bustin, 1996).

131
CHAPTER 5
PETROPHYSICAL AND RESERVOIR PROPERITES ANALYSIS

Determining porosity and saturation, the correlation between porosity and


permeability, and the controls of facies on reservoir properties were evaluated for the
Parshall Field. Different types of data were analyzed to get a more precise prediction of
the reservoir properties in the middle Bakken. This includes routine core analysis reports,
digitized raster images, formation water analyses, and source rock analyses.
Routine core analysis reports including porosity, permeability, and residual
saturations (water, oil) were available for eight wells in Parshall Field, one well in Sanish
Field, and one well to the east of Parshall Field. These reports were downloaded from the
North Dakota Industrial Commission, Department of Mineral Resources, Oil and Gas
Division (NDIC) website. This data was used to calibrate the log model and constrain the
range of plausible parameters used to compute porosity, permeability, and saturations
from logs.
There are 913 vertical wells in the study area that contain gamma ray, density-
neutron porosity, sonic, and/or resistivity raster logs. Of these, 135 wells were selected
for well log analysis based on the availability of open-hole logs and the locations of the
wells. Forty-six wells were digitized from raster images and 89 wells were available to
download from the NDIC website in the form of LAS (Log ASCII Standard Format)
files. The curves digitized (depending on log type and curve availability) included two to
three resistivity curves, gamma ray, sonic transit time, bulk density, neutron porosity, and
caliper. The 135 wells selected for petrophysical analysis had the following data:
83 wells with Gamma Ray, Resistivity, Density, Neutron, Sonic
44 wells with Gamma Ray, Resistivity, Density, Neutron
6 wells with Gamma Ray, Resistivity, Sonic
1 well with Gamma Ray, Resistivity, Density, Sonic
1 well with Gamma Ray, Resistivity, Neutron, Sonic

133
The wells included in this study are shown in Figure 5.1. These digital logs were
interpreted using LESA (Log Evaluation System Analysis) software from the company
Digital Formation (Version 8.7.2). Once the calculations are done, the digits are loaded
into Petra for mapping purposes.
A limited number of formation water analyses were available in and near
Parshall Field from the NDIC website. Three water analyses were from 2008 and one was
from 1986. These were used in determining formation water resistivity in the well-log
analysis.
Source-rock analysis data was compiled from published works by Webster
(1984) and Price et al. (1984) as well as from the NDIC website and the United States
Geological Survey. These source-rock analyses were used to determine the maturity and
kerogen type of the upper and lower Bakken shale members.

5.1 Core Properties Analysis


Routine core analysis provides information on porosity, permeability, water
saturation, oil saturation, and grain density. Porosity is defined as the percent of void
space, or that volume within rock that can contain fluids. The core laboratory
measurement of porosity determines the effective porosity and is commonly made by a
Boyles law technique using Helium as the experimental gas (Monicard, 1980).
Permeability is the ability of rocks to transmit fluids and is strongly controlled
by the pore throat size. Measurements are typically made by recording the pressure drop
resulting from a steady state flow of air moving through samples. In this case the sample
is dry and saturated with air, and the applied pressure gradient is such that laminar flow is
maintained. Since only one fluid is present, the result of the measurement is called
absolute permeability. Samples with either natural fractures or drilling induced fractures
should be excluded from the permeability measurement because they do not represent the
matrix permeability.
Saturation refers to the amount of a specific fluid (gas, oil, or water) in a rock.
One hundred percent (100%) water saturation indicates that all the pore space is occupied
by water (Monicard, 1980). The measurement of saturation is expressed as the percentage
of a specific fluid filling pore spaces in the rock. A typical core analysis does not show a

134
Figure 5.1 All wells with rasters in the study area. Wells with digital logs used are
highlighted in pink. Parshall Field is the outline on the right and Sanish Field is just to the
east. The field outlines were taken from the North Dakota Industrial Commission,
Department of Mineral Resources, Oil and Gas Division website.

135
total of 100% fluid saturation since the rock has been removed from its subsurface
environment of higher pressure and temperature (in situ conditions) to a surface
environment that has much lower pressure and temperature. This change in conditions
allows formation fluids to escape, especially gas. Drilling mud may also invade the rock
pores and may displace some of the original fluid.
Core analysis reports from seven cores within the Parshall Field (Patten #1-27H,
Hoff #1-10H, Parshall #2-36, N & D # 1-05H, Long #1-01H, Van Hook #1-13H, and
Jensen #12-44) and two cores from outside the field (Deadwood Canyon Ranch #43-28H,
and Dobrinski #18-44) were analyzed.
Several plots were made to analyze the relationship between porosity and
permeability and to determine relationships within the reservoir. Semi-log plots of linear
porosity versus log permeability places porosity along the X-axis and log permeability
along the Y-axis. Regression analysis can then be performed to test how well porosity
and permeability correlate with each other. The coefficient of determination (R2) reflects
the degree of correlation between porosity and permeability. Higher porosity and
permeability values represent better reservoir quality.
Core porosity and permeability plots vary from core to core and are summarized
in Table 5.1. The Patten #1-27H from Parshall Field is likely the best to depend on for
valid porosity and permeability data because it was taken using the most recent coring
technology and analysis. The average permeability in this core measured 0.0009 mD and
the average porosity is 4.2%. The highest porosity and permeability interval can be found
at depth 9,069 feet 9,074 feet (2,764.2 meters 2,765.8 meters), within facies E from
core description. A fluid saturation profile shows that these facies E and D2 are within
the oil-saturated zone with an average oil saturation of 41.2% (Figure 5.2). The average
water saturation values for facies E and D2 is 34.9%. The correlation between porosity
and permeability has an R2 value of 0.5849 (Figure 5.3).

136
Table 5.1 Core porosity and permeability measurements and averages listed by well. The correlation
equation and R2 values are from cross plots of porosity versus permeability. Values marked with (*)
indicate permeability samples that were taken at ambient conditions.

137
Figure 5.2 Fluid saturation profile of the Patten #1-27H taken from core analysis.

Figure 5.3 Porosity versus permeability plot for the Patten #1-27H taken from core
analysis.

138
The Hoff #1-10H is a horizontal well that had core analysis work done all within
facies E. Porosity values range from 0.5% - 8.9% and permeability values range from
0.0001 mD 0.009 mD. The correlation between porosity and permeability is poor with
an R2 of 0.1385 (Figure 5.4). The scatter of data on this plot demonstrates the complexity
of facies E. This facies varies from more laminated sections to bioturbated sections that
change the permeability and porosity of the rock. The fluid saturation profile in this core
has an average oil saturation of 28.1% and an average water saturation of 23.4% (Figure
5.5).
Core analysis from the Parshall #2-36 core shows that porosity values range from
2.5% - 8.7% and permeability varies from 0.0002 mD 0.019 mD. The best porosity and
permeability zone occurs at depth interval of 9,274 feet 9,285 feet (2,826.7 meters
2,830.1 meters) which includes facies E with an average porosity of 6.8% and an average
permeability of 0.005 mD. Again the correlation is low at an R2 value of 0.2386 caused
by the different characteristics of the facies (Figure 5.6). There is decent porosity below
this interval but the permeability values decrease dramatically. Oil saturation values were
not reported for this core so no fluid saturation profile was created.
Core analysis from the N & D #1-05H core shows porosity values range from
2.7% - 9.0% and permeability varies from 0.0001 mD 0.28 mD. Facies F2, E, B, and A
are present in this core which gives the poor correlation coefficient of 0.1021 (Figure
5.7a). The higher permeability values present in this core are from fractured samples and
when removed a much higher but still mediocre R2 value of 0.5342 occurs (Figure 5.7b).
The fluid saturation profile for the middle Bakken reservoir in this core has an average oil
saturation of 32.0% and an average water saturation of 22.2% (Figure 5.8).
The Long #1-01H core analysis shows porosity values range from 2.2% - 8.1%
and permeability varies from 0.0001 mD 0.738 mD. This highest permeability value
comes from a sample with a horizontal fracture present. The average permeability for this
core not including this specific measurement is 0.016 mD. The plot between porosity and
permeability shows a correlation (R2) of 0.3016 (Figure 5.9a). When the fractured sample
in facies B is removed the correlation rises to 0.5497 (Figure 5.9b). Facies F2, E, B, and
A are present in this core which would explain the low correlation between porosity and

139
Figure 5.4 Porosity versus permeability plot for the Hoff #1-10H taken from core
analysis.

Figure 5.5 Fluid saturation profile of the Hoff #1-10H taken from core analysis.

140
Figure 5.6 Porosity versus permeability plot for the Parshall #2-36H taken from core
analysis.

141
Figure 5.7a Porosity versus permeability plot for the N & D #1-05H taken from core
analysis.

Figure 5.7b Porosity versus permeability plot for the N & D #1-05H taken from core
analysis.

142
Figure 5.8 Fluid saturation profile for the N & D #1-05H taken from core analysis.

143
Figure 5.9a Porosity versus permeability plot for the Long #1-01H taken from core
analysis.

Figure 5.9b Porosity versus permeability plot for the Long #1-01H taken from core
analysis.

144
permeability due to facies heterogeneity. The fluid saturation profile in this core has an
average oil saturation of 33.8% and an average water saturation of 21.8% (Figure 5.10).
Core analysis from the Van Hook #1-13H well shows that porosity values range
from 2.5% - 11.3% and permeability values vary from 0.0004 mD 0.105 mD. The plot
between porosity and permeability shows a correlation (R2) of 0.482 (Figure 5.11). This
is interpreted to be caused by the highly heterogeneous bed characteristics within facies
E, C2, C1, and B, which range from highly bioturbated sandstones to laminated facies.
The fluid saturation profile shows that the middle Bakken reservoir in this core has an
average oil saturation of 30.6 % and an average water saturation of 18.6% (Figure 5.12).
Core analysis from the Jensen #12-44 shows that porosity values range from
2.7% 8.8% and permeability varies from 0.02 1.9 mD (Figure 5.13). The higher
permeability values are from fractured samples. Permeability measurements were taken
at ambient conditions for this well so these values are much high than in other core
analyses but the general trend of porosity versus permeability is similar to those of other
cores. The correlation between porosity and permeability in this core analysis has a
correlation (R2) of 0.598. The fluid saturation profile shows that the middle Bakken
reservoir in this core has an average oil saturation of 46.9% and an average water
saturation of 40.3% (Figure 5.14).
Two core analyses from wells located outside the Parshall Field were also
analyzed. The Deadwood Canyon Ranch #43-28H core analysis shows that porosity
values range from 1.7% - 11.0% and permeability varies from 0.0001 mD - 0.229 mD
(Figure 5.15). The plot between porosity and permeability shows a correlation coefficient
(R2) of 0.4659. All of the facies are present except for facies F2 with each having their
own characteristics which explains the poor correlation. The fluid saturation profile for
the middle Bakken reservoir in this core has an average oil saturation of 31.4% and an
average water saturation of 49.2% (Figure 5.16).
Core analysis from the Dobrinski #18-44 (Ward County, North Dakota) shows
that porosity values range from 1.1% - 6.4% and permeability varies from 0.01 mD
0.05 mD (Figure 5.17). Permeability measurements were taken at ambient conditions for
this well so these values are much high than in other core analyses but the general trend

145
Figure 5.10 Fluid saturation profile for the Long #1-01H taken from core analysis.

146
Figure 5.11 Porosity versus permeability plot for the Van Hook #1-13H taken from core
analysis.

Figure 5.12 Fluid saturation profile for the Van Hook #1-13H taken from core analysis.

147
Figure 5.13 Porosity versus permeability plot for the Jensen #12-44 taken from core
analysis.

Figure 5.14 Fluid saturation profile for the Jensen #12-44 taken from core analysis.

148
Figure 5.15 Porosity versus permeability plot for the Deadwood Canyon Ranch #43-28H
taken from core analysis.

149
Figure 5.16 Fluid saturation profile for the Deadwood Canyon Ranch #43-28H taken
from core analysis.

150
Figure 5.17 Porosity versus permeability plot for the Dobrinski #18-44 taken from core
analysis.

151
of porosity versus permeability is similar to those of other cores. The correlation between
porosity and permeability has a correlation (R2) of 0.613. Five different facies are present
in this core which explains the poor correlation between porosity and permeability. The
fluid saturation profile shows that this well is not a producer and is completely wet with
an average water saturation of 89.2% (Figure 5.18).
The correlation between porosity and permeability for each of the wells studied
depends on which facies and how many are present in the core which are summarized in
Table 5.2. Facies E1 has the most scattered data because of the interbedded laminated and
bioturbated sections and has an R2 value of 0.2135 (Figure 5.19). Facies D2 shows a good
correlation of porosity and permeability values with R2 values of 0.7242 (Figure 5.20).
The slope of the trend for facies D2 and D1 shows that these rocks build permeability
much faster than the other lithotypes. Facies C2 and C1 are only in the Deadwood
Canyon Ranch #43-28H and the Van Hook #1-13 cores with limited data in facies C2
(only four data points) (Figure 5.21). Facies C1 shows a good correlation between
porosity and permeability with a R2 value of 0.6321. Facies B correlation of porosity and
permeability has a wide range of values since this facies is bioturbated and has a R2
values of 0.3943 (Figure 5.22). Facies A shows a good correlation between porosity and
permeability and has a R2 values of 0.6117 (Figure 5.23). Facies F1 only has one data
point and facies F2 only has three data points so the relationships for these were not
plotted.
All of the cores with fluid analysis had some oil saturation except for the
Parshall #2-36H well in which oil saturations were not reported. Water saturations for the
wells in Parshall Field are generally less than 40%. The Dobrinski #18-44 well to the east
of Parshall Field has water saturation values near 90%. The Deadwood Canyon Ranch
#43-28H well in Sanish Field has water saturations that range from 80% down to 20% in
the producing interval.

5.2 Conventional Well Log Analysis


Well logs record the physical properties of the rocks in the subsurface
environment, including their electrical properties, porosity, lithology, mineralogy,
permeability, and water saturation (Asquith et al., 2004). The well logs were calibrated to

152
Figure 5.18 Fluid saturation profile for the Dobrinski #18-44 taken from core analysis.

153
Table 5.2 Core porosity and permeability measurements and averages listed by facies.
The correlation equation and R2 values are from cross plots of porosity versus
permeability. N/A indicates that not enough data was available.

Figure 5.19 Porosity versus permeability plot for facies E1 taken from core analysis.

154
Figure 5.20a Porosity versus permeability plot for facies D2 taken from core analysis.

Figure 5.20b Porosity versus permeability plot for facies D1 taken from core analysis.

155
Figure 5.21a Porosity versus permeability plot for facies C2 taken from core analysis.

Figure 5.21b Porosity versus permeability plot for facies C1 taken from core analysis.

156
Figure 5.22 Porosity versus permeability plot for facies B taken from core analysis.

Figure 5.23 Porosity versus permeability plot for facies A taken from core analysis.

157
the measured core data (porosity, permeability, and fluid saturations) to minimize errors
from log readings. The curves digitized and used (depending on log type and curve
availability) included two to three resistivity curves, gamma ray, sonic transit time, bulk
density, neutron porosity, and caliper.
The gamma-ray log measures the degree of natural radioactivity of the rock. The
gamma-ray tool utilizes a sodium-iodide scintillation detector to measure the natural
gamma ray radiation of the formation and 5-window spectroscopy to resolve the detected
spectrum into the three most common components of the naturally occurring radiation:
potassium, thorium, and uranium. This tool is commonly run with an electrical tool which
measures the resistivity of the formation.
The density log is used to determine formation density and estimate formation
porosity. The density logging tool emits gamma rays from a chemical source at the
bottom of the tool that enter the surrounding rocks where some gamma rays are absorbed.
The remaining gamma rays survive to reach scintillation counters mounted about 18 and
24 inches (0.5 meters and 0.6 meters) above the source. The number of gamma rays
arriving at the far detector is inversely proportional to the electron density of the rock,
which in turn is proportional to the actual rock density.
The neutron log is used to evaluate formation porosity. It measures the change in
neutron flux with distance from a source and converts it to a calibrated apparent porosity
value. The rate of decrease, represented by the ratio of the near to far count rates, is
primarily due to the hydrogen content of the formation.
The sonic or acoustic log employs a sonde that generates a compressional sound
wave that travels to the wall of the borehole, along which it propagates and generates, in
turn, a reflected wave that is received by the tool. The speed of the sound waves can be
used to determine porosity (Asquith et al., 2004).

5.2.1 Data Preparation


A BITSIZE curve was created, based on information in the well log header. The
BITSIZE curve is the size of the borehole drilled. The BITSIZE and DRHO (density
correction curve) curves were used to determine a BADHOLE flag curve. The
BADHOLE flag curve was used as a qualitative quality indicator in the wells with

158
adequate data. The density log is often invalid if the hole is more than three inches out of
gauge. If the data appeared valid, it was used in the calculations. If the density appears
washed out, all of the calculations were nulled. In almost all of the wells, bad hole was
restricted to the upper and lower Bakken shale; the middle Bakken was usually in good
shape. The environmental corrections module was run. This corrected the deep resistivity
log for wellbore environmental conditions, producing an RT curve, and in the process
created a flushed zone resistivity curve, RXO. In general, the correction to the deep
resistivity curve was minimal. Zones for creation of interpretive parameters were created
from previously-picked tops, with the zone interval running from its top to the top below.
The zones used were Bakken E, Bakken D, Bakken C, Bakken B, and Bakken A.

5.2.2 Petrophysical Model


Shale volumes were calculated from the gamma ray after determining GRclean
and GRshale values for each zone in each well from a log display. The clean point was
picked at the lowest GR value in the cleanest section of the lower Lodgepole Formation,
and the shale point was chosen from a shale at the top of the Three Forks Formation.
Calculation of the gamma ray index is the first step in determining the volume of shale
and it is done using:

(1)
Where: GRI = gamma ray index
GR = gamma ray reading of formation
GRcl = minimum gamma ray (clean sand or carbonate)
GRsh = maximum gamma ray (shale)

This is then used in the Vshale calculation using the Steiber method:

(2)

159
Where: VSH = volume of shale
GRI = gamma ray index

Once a Vshale curve was calculated, total porosity was computed from the
density, sonic and neutron logs by taking the matrix and fluid parameters using the
Wyllie time-average equation for both the density porosity and sonic porosity and using
the matrix parameter for the neutron log (using a limestone matrix of 2.71 g/cm3).

(3)
Where: Density = density derived porosity
matrix = matrix density (2.71g/cm3)
bulk = formation bulk density
fluid = fluid density (1.0 g/cm3)

(4)
Where: Sonic = sonic derived porosity
tlog = interval transit time in the formation
tmatrix = interval transit time in the matrix (48 sec/ft)
tfluid = interval transit time in the fluid of the formation (189 sec/ft)

Effective or shale corrected porosity (corrected for clay bound water) was then
calculated for each log by subtracting the product of Vshale times the apparent porosity of
the shale. The shale values were determined by inspection of several cross-plots of
porosity vs. shale content with the average values used being:
Density of shale = 2.65 g/cm3
Neutron porosity of shale = 0.20 on Limestone matrix
DT shale = 95 sec/ft
The effective porosities were then compared to core data to confirm the validity of the
porosity model. Cross plot porosities are then made to help eliminate the effects that

160
different lithologies have on different logging tools. This was done for the neutron,
density, and sonic logs.
Water saturation calculations are now performed using a simple Archie
saturation equation:

(5)
Where: Sw = water saturation of uninvaded zone
Rw = formation water resistivity
Rt = formation resistivity (uninvaded zone)
= porosity
a = tortuosity factor
m = cementation exponent
n = saturation exponent

Porosity values are used from the neutron/density cross plot porosity, Rt is taken from the
deep resistivity curve, Rw is determined by water analyses to be 0.02 ohm-m, and a, m,
and n are determined from core data and log data. The tortuosity factor (a) is assumed to
be 1, the saturation exponent (n) is 1.74, and the cementation exponent (m) varies for
each facies as shown below in Table 5.3. The saturation exponent chosen was based on
communications with Julie LeFever from the North Dakota Geological Survey. The
cementation exponent values were calibrated with core data.

Table 5.3 Parameters used to calculate water saturation using the Archie equation.

161
The product of a formations water saturation (Sw) and its porosity () is the
bulk volume of water (BVW).

BVW = Sw * (6)

Where: BVW = bulk volume water


Sw = water saturation of uninvaded zone
= porosity

If values for bulk volume water, calculated at several depths in a formation, are constant
or very close to constant, they indicate that the zone is of a single rock type and at
irreducible water saturation (Swirr). This is important because if a zone is at irreducible
water saturation, the water in the formation does not move because it is held onto grains
by capillary pressure, therefore hydrocarbon production should be water free (Asquith et
al., 2004). Swirr was estimated by dividing a constant bulk volume water irreducible of
5% by the log determined porosity at each depth:
Swirr = 0.05 / (7)

Permeability was computed using a general form of the Timur equation (from LESA):

(8)
Where: Kcoef = 62500
KKPHIEXP ~ 5.0 9.25
KKSWIEXP = 2.0
PHIX = cross plot porosity
SWI = irreducible water saturation

The permeability exponent KPHIEX was set by zone to best approximate the in situ core
permeabilities. An example of all these calculations is in Figure 5.24 showing the
Deadwood Canyon Ranch #43-28H well.

162
Figure 5.24 Calculated log for the Deadwood Canyon Ranch #43-28H well.

163
Average calculated water saturations for the middle Bakken were plotted to look
at the distribution (Figure 5.25). Parshall Field has water saturation values ranging from
25% to approximately 40%. The area to the northwest of Parshall Field has average water
saturations of 50%. The general southwest-northeast trend of the feature may be related
to lineaments. Oil generated may have been expelled north and up-dip, leaving this area
more water prone (Hill et al., 2007). The higher production of water in this area
compared to Parshall Field is also demonstrated by mapping the percent of oil over water
from cumulative data (Figure 5.26).
Resistivity logs were evaluated to determine a cut-off value for the determination
of net pay. By using average porosity values determined from core porosities and well
log neutron-density effective porosity calculations for each facies, the true resistivity
values needed in order to produce hydrocarbons with 40% water saturation were
determined (Table 5.4) (Figure 5.27). These Rt values determined from Pickett plot
analysis were selected as the cut-off values to determine the net pay interval. The pay
thickness of the middle Bakken reservoir ranges from 14 feet - 40 feet (4.3 meters 12.1
meters) and averages 25 feet (7.6 meters) in Parshall Field (Figure 5.28).
A hydrocarbon pore volume map was created by multiplying several maps. A grid
of the net thickness of the middle Bakken was multiplied by the grid of the average
porosity map by the grid of average oil saturation (1-Sw) (Figure 5.29).

5.3 Source Rock Characterization


Several different parameters were look at involving the Bakken source rock.
These include measured and calculated total organic carbon values, pyrolysis data, and
the time temperature index method. Each is discussed in the sections below.

5.3.1 Source Rock Richness Estimation from Formation Density Log


Total organic carbon (TOC) values were calculated from formation density logs
by using a method from Schmoker and Hester (1983). Organic carbon values were also
calculated from resistivity and porosity logs using the method from Passey et al. (1990).

164
Figure 5.25 Average Sw map for the middle Bakken.

165
Figure 5.26 Oil to water ratio calculated from cumulative oil production data (bbls/bbls).

166
Table 5.4 Values used in determining water saturation cut-offs.

Figure 5.27 Picket plot showing the facies D interval from the Deadwood Canyon Ranch
#43-28H well. The average porosity of 4.2 % and 8 ohm-m resistivity value as a cut-off
yield for producible hydrocarbons with water saturation (Sw) of 40%.

167
Figure 5.28 Net pay map (in feet) for the middle Bakken member determined from water
saturation values equal to or less than 40%.

168
Figure 5.29 Hydrocarbon pore volume map of the middle Bakken.

169
The calculated TOC values were compared to limited measured TOC data provided by
the North Dakota Geological Survey.
Source rock richness can be determined by measuring the total organic carbon
(TOC) present in a rock. Different conventional well logs which can be used to determine
source rock quality are gamma ray, density, neutron, sonic, resistivity, and pulse neutron
(Table 5.5). In this study, organic carbon content was calculated from formation density
logs. This method was chosen because of the availability and widespread distribution of
wells with a formation density log across the area.

Table 5.5 Summary of log responses to organic matter (modified from Law, 1999).

Log/Property Response/Value for Comments


Organic Matter (OM)
Gamma ray (GR) High High GR caused by U; can be linear
or uranium (U) with OM; U not always present
Density Low (approx. 1 g/cm3) Similar to pore fluids
Neutron High Due to hydrogen in OM
Sonic High transit time Estimates vary from 150 to more than
200 sec/ft
Resistivity High May not affect log response unless
generated hydrocarbons occupy pores
Pulsed neutron High carbon-oxygen Most direct measurement of carbon;
ratio needs inorganic correction

The total organic carbon (TOC) values of the upper and lower Bakken shales
were calculated from 128 formation density logs available in the Parshall Field area. The
calculation of TOC values was made using the Schmoker and Hester (1983) method.
They calculated the organic carbon content for 159 locations in North Dakota and 107 in
Montana. They show an empirical correlation between TOC measured from core and
TOC calculated from formation density log (the presence of pyrite is important because
of its high density of 5.0 g/cm3). The upper and lower Bakken shales have disseminated
and laminated pyrite observed from core. The occurrence of pyrite should be considered
as an important part of the upper and lower Bakken shale density and should be taken
into account in the calculation of organic-carbon content from density logs. Using the
same approach from Schmoker (1979) in Appalachian Devonian shale, his previous study

170
has shown that there is a linear increase of pyrite with organic matter. The linear increase
is represented by the equation:
p = 0.135o + 0.0078 , (9)

where p is fractional volume of pyrite and o is the fractional volume of organic matter.
In the Schmoker and Hester (1983) study, the upper and lower members of the
Bakken Formation are treated as a four component system, consisting of rock matrix,
interstitial pore, pyrite, and organic matter (represented by m, i, p, and o). The formation
density is then a function of densities and fractional volumes of these four components:
= o o + p p + i i + (1 ( o + p + i )) m . (10)

The equation above requires a simplification to a two variable expression


relating o and . To reduce the number of unknowns in the above equation, several
assumptions were made. One assumption is that the porosity of the upper and lower shale
members is low and does not vary enough to alter formation density significantly. It is
also assumed that, at low porosities, density differences between pore-fluid types can be
neglected and terms combined to a weighted average matrix density assuming average
constant porosity, mi. Therefore,
= o o + p p + (1 ( o + p )) mi , (11)

By substituting p using equation 9, setting p equals to 5.0 g/cm3 and rearranging terms
equation 11 becomes:
o = ( 0.9922 mi 0.039) /( o 1.135 mi + 0.675) , (12)
Weight-percent organic carbon content (TOC) is frequently used to measure the organic
richness, and is related to the fractional volume of organic matter ( o ) by the equation:

TOC = o (100 o ) /( R ), (13)


where R is the ratio of weight-percent of organic matter to weight-percent organic carbon.
By substituting for o using equation 12, equation 13 becomes:

TOC = [(100 o )( 0.9922 mi 0.039)]/[( R )( o 1.135 mi + 0.675)], (14)


The parameters of o, mi, and R are estimated from available data and equation 10 is then
reduced to the form:
TOC = ( A / ) B. (15)

171
A and B are constants that are specifically calculated for a particular formation,
member, or area. The values of A and B are based on an organic matter density ( o ) of

1.01 g/cm3, a matrix density ( mi ) of 2.68 g/cm3, and a ratio between weight-percent of
organic matter and organic carbon (R) of 1.3. The final equation that can be applied to the
Bakken shales from their study is:
154.4
TOC = 57.261 , (16)

where TOC is organic carbon content (weight percent) and is formation density
(g/cm3). A comparison of calculated organic carbon content and measured TOC values
from Bakken shale cores from 39 wells in North Dakota (Schmoker and Hester data) is
shown in Figure 5.30. The ideal situation is shown by the red dashed line. The average
absolute difference between organic-carbon content determined from core samples and
that calculated from density logs is 1.1% (Schmoker and Hester, 1983).
In this study, the same approach (Schmoker and Hester, 1983) was applied to
the Parshall Field area. Figure 5.31 and Figure 5.32 shows the distribution of the upper
and lower Bakken TOC values in the Parshall Field area. The upper Bakken values
average 12% TOC and the lower Bakken averages 13% TOC.

5.3.2 Source Rock Richness Estimation Using the Log R Technique


The total organic carbon (TOC) values of the upper and lower Bakken shales
were also calculated from the resistivity, density, sonic, and neutron logs available in the
Parshall Field area. These calculations of TOC values were made using the log R
technique from Passey et al. (1990). In their study, the upper and lower members of the
Bakken Formation are treated as a three component system, consisting of the rock matrix,
the solid organic matter, and the fluid(s) filling the pore spaces. This method involves the
overlying of a properly scaled porosity log on a resistivity curve (preferably from a deep-
resistivity log). In either hydrocarbon reservoir rocks or organic-rich non-reservoir rocks,
a separation between the curves occurs. The baseline values are determined when the two
curves overlie each other over a significant depth range in a fine-grained, non-source
rock. The magnitude of the curve separation (designated as log R) in non-reservoirs is

172
Figure 5.30 Plot between calculated TOC values from formation density logs and TOC
determined from core analysis (modified from Schmoker and Hester, 1983).

173
Figure 5.31 Upper Bakken TOC values calculated using the Schmoker and Hester
method.

174
Figure 5.32 Lower Bakken TOC values calculated using the Schmoker and Hester
method.

175
calibrated to total organic carbon and maturity, and allows for depth profiling of organic
richness to be assessed (Passey et al., 1990).
Any type of porosity log can be used to calculate log R. The algebraic
expression for the calculation of log R from the sonic/resistivity overlay is:

log R = log10 (R/Rbaseline) + 0.02 x (t tbaseline) (17)

Where: log R = curve separation measured in logarithmic resistivity cycles


R = resistivity measures in ohm-m by the logging tool
t = measured transit time in sec/ft
Rbaseline = resistivity corresponding to the tbaseline value when the
curves are baselined
0.02 = based on the ratio of -50 sec/ft per one resistivity cycle

Empirically derived equations for calculating curve separation when substituting the
density or neutron log for the transit-time curve are:

log RNeu = log10 (R/Rbaseline) + 4.0 x (N Nbaseline) (18)


and
log RDen = log10 (R/Rbaseline) 2.50 x (b baseline) (19)

Where log RNeu and log RDen are log R separations based on the density and
neutron logs, respectively, N is the neutron porosity reading from the well log in the
interval of interest, Nbaseline is the neutron porosity baseline value, and b and baseline are
the density values for the source rock and baseline intervals, respectively (Passey et al.,
1990).
The log R separation is linearly related to TOC and is also a function of
maturity of the rocks. The log R separation can be transformed directly to TOC if the
maturity (in level of organic metamorphism units, LOM) can be determined or estimated
(Hood et al., 1975). This can be obtained from a variety of sample analyses (e.g. vitrinite

176
reflectance, thermal alteration index, or Tmax). Tmax has been found to be linearly
related to vitrinite reflectance (Ro) by Jarvie et al., (2004) with the equation:

Ro = 0.0180(Tmax) 7.16 (20)

Vitrinite reflectance (%Ro) is a measure of the percentage of incident light reflected from
the surface of vitrinite particles in a sedimentary rock (Law, 1999). Oil and gas zone
boundaries can be established using vitrinite reflectance data. The boundaries are
approximate and vary according to kerogen type. Once vitrinite reflectance has been
calculated, this value is converted to LOM by using Figure 5.33 from Hood et al., (1975).
The empirical equation for calculating TOC from log R is:

TOC = ( log R) x 10 (2.297-0.1688 x LOM) (21)

Where TOC is the total organic carbon content measured in weight percent, and LOM is
the maturity. A LOM of 7 corresponds to the onset of maturity for oil-prone kerogen, and
an LOM of 12 corresponds to the onset of overmaturity for oil-prone kerogen (Passey et
al., 1990). For the Parshall Field area, a LOM of 7.8 is used in the calculations for TOC
in the upper Bakken and a LOM of 7.5 is used for the lower Bakken (Figure 5.33).
Comparison of both the Schmoker and Hester method and the Passey method
are shown in Figure 5.34. From observation the Schmoker and Hester values match the
core data more closely than the Passey method. The Passey method relies on knowing the
level of maturity of the area. No vitrinite reflectance data is available for the Parshall
Field, so this value becomes harder to determine through the Passey method.

5.3.3 Pyrolysis Methods


Pyrolysis is the decomposition of organic matter by heating in the absence of
oxygen. In pyrolysis analysis, the organic matter is pyrolyzed and then combusted. The
amount of hydrocarbons and carbon dioxide that is released is then measured. In Rock-
Eval pyrolysis, which is the most widely used pyrolysis technique, the sample is placed in
a vessel and heated progressively to 550C under an inert atmosphere. The S1 peak is a

177
Figure 5.33 Vitrinite reflectance relationship to level of maturation (modified from Hood
et al., 1975).

Figure 5.34 Comparison of the Schmoker and Hester method and the Passey method for
calculating TOC.

178
measurement of free hydrocarbons present in the sample before the analysis (Figure
5.35). The next peak, S2, is the volume of hydrocarbons that formed during thermal
pyrolysis of the sample; this peak can be used to estimate the ability of the source rock to
generate hydrocarbons. The S3 peak is the result of the CO2 generated from the pyrolysis
S4 records the residual carbon from the sample. Tmax is the temperature at which the
maximum rate of hydrocarbon generation occurs in a kerogen sample during pyrolysis
analysis (Law, 1999).
Source rock richness is represented by the total organic carbon (TOC) content and
can be calculated from pyrolysis analysis. The percentage of TOC can be calculated using
the following formula (Law, 1999):
%TOC = [0.082(S1 + S2) + S4]/10 (22)
Units are usually given as weight percent organic carbon per weight of dry rock
(milligrams hydrocarbon per gram of rock).
The amount of hydrogen relative to the amount of organic carbon in the sample
is represented by the hydrogen index (HI) value. The hydrogen index (HI) of a sample
can be calculated by using the following equation (Law, 1999):
HI = S2 (mg/g)/%TOC 100 (23)
The most common technique to determine source rock quality based on Rock-
Eval data is to compare hydrogen index to oxygen index. The oxygen index (OI) is the
amount of oxygen relative to the amount of organic carbon present in a sample. The S3
peak can be used to determine the total amount of oxygen present in a sample according
to the following formula (Law, 1999):
OI = S3 (mg/g)/%TOC 100 (24)
A plot of hydrogen index (HI) and oxygen index (OI), known as a modified van
Krevelen diagram, can be used to determine the kerogen type present in a source rock
(Figure 5.36). The type of kerogen present in a rock determines its quality. Type I
kerogen is the highest quality (the more oil prone the kerogen the higher the quality) and
type III is the lowest. Type I has the highest hydrogen content; type III, the lowest.
Source rock quality and thermal maturity can be determined from the kerogen
types contained in the organic matter. In the Parshall Field, 28 core chips were analyzed

179
Figure 5.35 Diagram showing the cycle of Rock-Eval pyrolysis analysis and the
corresponding recording (modified from Tissot and Welte, 1984).

180
Figure 5.36 Van Krevelen diagram showing types of organic matter. Type I is the highest
kerogen quality due to the highest hydrogen content, type III is the lowest (from Tissot
and Welte, 1984).

181
to determine the kerogen type and maturity index. One sample is from the Parshall #2-
36H well, two samples are from the Parshall SD #1 well, and 25 samples are from the
Jensen #12-44 well (Table 5.6). This data is provided by NDIC website and the USGS
and is publicly accessible. The TOC and Rock-Eval pyrolysis analysis results were
plotted to determine the organic richness, hydrocarbon potential and maturity of the
Bakken. The modified van Krevelen diagram shows that the kerogen type is Type I and II
(Figure 5.37 Figure 5.39). These results were compared to the Rock-Eval pyrolysis
results from the North Dakota data (Webster, 1984 and Price, 1984).
The maturity of a homogeneous source section can be accessed from the
production index (PI). PI is calculated from Rock-Eval data:
PI = S1/(S1 + S2) (25)
Hunt (1979) suggests that PI from 0.08 to 0.4 is characteristic of source rock in
the oil window. Based on the plot of Production Index (PI) and maturity based on Tmax,
the Parshall Field samples are near the boundary of the oil generation zone (Figure 5.40).
These samples were compared to the data from North Dakota from Webster (1984) and
Price (1984) and show a good match at the same depth interval.
Hydrocarbon generation appears to occur at a low thermal maturity (420-430 oC
Tmax range). Extremely high TOC values are present in the upper and lower Bakken
which can lead to early fractionation and mobilization of source-rock bitumen from rocks
with Type I or II organic matter. Hydrocarbon generation in organic rich facies is
interpreted as the cause of overpressuring in the Bakken which causes fractures. Parshall
Field is slightly overpressured with an average pressure gradient of 0.60 psi/ft. This was
calculated using mud weight data. These fractures have allowed the hydrocarbons to
move into the middle member. According to Jarvie and Johnson (2008) these low
thermally mature rocks appear to be the source of the oil rather than migrated oil in
Parshall Field. Just to the east of Parshall Field these source rocks are immature.

182
Table 5.6 Rock-Eval data provided by the NDGS and the USGS for Parshall Field.

183
Figure 5.37 Modified van Krevelen diagram showing the distribution of the Bakken
Formation, Williston Basin. Published data from Webster (1984) and Price et al. (1984).
Majority of samples indicate a Type-I and II oil prone-kerogen (algal origin). Legend
shows source rock data by depth interval; three Parshall Field wells are indicated in
black.

184
Figure 5.38 Hydrogen index versus Tmax values for the Bakken Formation, Williston
Basin. Majority of the samples indicate a Type I and II oil prone kerogen. Kerogen-type
lines from Finn and Johnson, 2005. Legend: Bakken data from Webster (1984) and Price
et al. (1984); black data points from Parshall Field.

185
Figure 5.39 S2/S3 versus hydrogen index values for the Bakken Formation, Williston
Basin. The majority of the samples indicate a Type-I and II oil prone kerogen. Legend:
Bakken data from Webster (1984) and Price et al. (1984); black data points from Parshall
Field.

186
Figure 5.40 Plot comparing Rock-Eval production index and Tmax data for Parshall Field
and North Dakota data. Legend shows source rock data by depth interval (data from Price
et al., 1984; Webster, 1984); black data points from Parshall Field.

187
5.3.4 Time Temperature Index Method
The time and depth of oil generation from petroleum source rocks containing type
II kerogen (like the Bakken) can be determined using time-temperature index (TTI) using
the Lopatin (1971) method.
TTI = t (26)
Where: t = time in millions of years
= Maturation rate index

The maturation rate index is a measure of how fast oil is being generated.
= r (T/10)-10.5 (27)
Where: = Maturation rate index
T = temperature (oC)
r = change in reaction rate with change in temperature (assumed to be 2)
This is based on the idea that chemical rates double for every 10oC increase in
temperature (Figure 5.41).
According to Waples (1980), the onset of oil generation occurs at a time-
temperature index of maturity (TTI) of 15 (Table 5.7). For the Bakken a TTI value of 15
occurs at approximately 9,000 feet (2,743 meters) (Pitman et al., 2001). Burial history
and time-temperature index plots have been made for the Deadwood Canyon Ranch #43-
28H well (Sanish Field) and the Parshall #2-36H well (Parshall Field) by Nordeng (2010)
and Nordeng and LeFever (2008) (Figure 5.42 and Figure 5.43). The burial history curves
show that there is rapid subsidence and deposition from the Late Devonian to the Early
Mississippian. More gentle subsidence occurred until the late Cretaceous and then
another rapid subsidence occurred. The time-temperature index plots indicate that oil
generation for both of these wells occurred during the latest rapid subsidence in the
Cretaceous for the Deadwood Canyon Ranch #43-28H and in the Cenozoic for the
Parshall #2-36H.

188
Figure 5.41 Maturation rate index versus temperature indicating the exponential
relationship between the data.

Table 5.7 Time-temperature index maturity values.

189
Figure 5.42 Deadwood Canyon Ranch #43-28H TTI and burial history (from Nordeng,
2010).

Figure 5.43 Parshall #2-36H TTI and burial history (from Nordeng and LeFever, 2008).

190
CHAPTER 6
CONCLUSIONS AND RECOMMENDATIONS

6.1 Conclusions

The two main objectives of this study were to: 1) understand the depositional
setting and mechanism of the Bakken Formation in the Parshall Field by determining and
correlating depositional facies, 2) determine the reservoir quality by applying integrated
petrophysical analysis approach.
The conclusions made from this research are:
The middle Bakken in the Parshall Field has nine distinctive facies, representing a
offshore to tidally influenced deltaic environment.
Middle Bakken facies were deposited from middle shoreface distal offshore
environment characterized by sedimentary structures and ichnofacies observed
from cores. Thus, water depth, sediment supply, and depositional energy control
the facies variation within the middle Bakken in the Parshall Field.
Middle Bakken pay interval in the Parshall Field averages 25 feet (7.6 meters) in
thickness and include facies E1, C2, C1, and B.
TOC calculation of the upper and lower Bakken shale showing a good correlation
with the core analysis result, with average organic content of 12% in the upper
Bakken and 13% in the lower Bakken in Parshall Field.
Average formation fluid pressure gradient in Parshall Field is 0.60 psi/ft and is
slightly overpressured which is believed to be caused by hydrocarbon generation
within the Bakken Formation.
Hydrocarbon was generated in the Cenozoic and the present-day Bakken source
rock is in the early-mature oil generation zone according the burial history curves
of the Parshall #2-36H and the Deadwood Canyon Ranch #43-28H.

191
6.2 Recommendations

A detailed look at the mineralogy and diagenesis of the middle Bakken at


Parshall Field would be beneficial. This could help determine the rocks ease to fracture
and therefore enhance the permeability of this area.
Faults and fractures are not abundantly present based on observation from the
available data. The availability of seismic data or even time-lapse seismic can be a
valuable input for creating subsurface mapping, identifying fractures zone, and
monitoring the fractures development within the Parshall Field area. Borehole image data
can also be used to characterize and identify the presence of fractures.
More advanced well log analysis tools and more advanced core data would help
evaluate the reservoir in more detail. This could include special core analysis that would
look at relative permeability, wettability, and capillary pressure. Nuclear magnetic
resonance (NMR) logging tools would also be useful for getting information about
formation fluids and porosity and better calibrating the saturation model.

192
REFERENCES CITED

1987, By. Chemical and Mineral Analysis with QEMSCAN: FEI Company. Electron
Microscope | Electron Microscopes | Scanning Electron Microscope | SEM |
Transmission Electron Mircroscope | TEM | DualBeam | Focused Ion Beam | FIB
| FEI Company. Web. 1 Oct. 2010.
<http://www.fei.com/applications/industry/automated-mineralogy/qemscan.aspx>.

NDIC Oil & Gas Division Home Page. Department of Mineral Resources. Web. 1 Oct.
2010. <https://www.dmr.nd.gov/oilgas/>.

Asquith, G., D. Krygowski, S. Henderson, and N. Hurley, 2004, Basic well log analysis,
2d ed.: AAPG Methods in Exploration Series 16, 244 p.

Blakey, R. (2005), North American paleogeographic maps, early Mississippian


(345Ma). Paleogeography and geologic evolution of North America. Web. 1 Oct.
2009. < http://jan.ucc.nau.edu/rcb7/namM345.jpg>.

Blakey, R. (2005), North American paleogeographic maps, late Devonian (360Ma).


Paleogeography and geologic evolution of North America. Web. 1 Oct. 2009.
<http://jan.ucc.nau.edu/rcb7/namD360.jpg>.

Burrus, J., K. Osadetz, S. Wolf, B. Doligez, K. Visser, and D. Dearborn, 1996a, A two-
dimensional regional basin model of Williston basin hydrocarbon systems: AAPG
Bulletin, v. 80, no. 2, p. 265-291.

Burrus, J., K. Osadetz, S. Wolf, and K. Visser, 1996b, Physical and numerical modelling
constraints on oil expulsion and accumulation in the Bakken and Lodgepole
petroleum systems of the Williston Basin (Canada-USA): Bulletin of Canadian
Petroleum Geology, v. 44, no. 3, p. 429-445.

Carlisle, J., L. Druyff, M. Fryt, J. Artindale, and H. Von Der Dick, 1992, The Bakken
Formation - An Integrated Geologic Approach to Horizontal Drilling, in
Schmoker, J., E. Coalson, and C. Brown, eds., Geologic Studies Relevant to
Horizontal Drilling: Examples from Western North America, p. 215-226.

Canter, L., O. Skinner, and M. D. Sonnenfeld, 2009, Facies and Mechanical Stratigraphy
of the Middle Bakken, Mountrail County, North Dakota, Abstract, Rocky
Mountain Section of SEPM Luncheon, January 27, 2009.

Christopher, J. E., 1961, Transitional Devonian - Mississippian Formations of southern


Saskatchewan: Saskatchewan Mineral Resources Report 66, Regina,
Saskatchewan, 103 p.

193
Clement, J. H., 1976, Geologic history key to accumulation of Cedar Creek (abs.): AAPG
Bulletin, v. 60, no. 11, p. 2067-2068.

Coskey, R. J. and J. E. Leonard, Bakken Oil Accumulations - What's the Trap?,


American Association of Petroleum Geologists Annual Convention, Denver,
Colorado, Presentation Session, Search and Discovery Article #90090, June 7-10,
2009, http://www.searchanddiscovery.net.

Dow, W. G., 1974, Application of Oil-Correlation and Source-Rock Data to Exploration


in Williston Basin: AAPG Bulletin, v. 58, no. 7, p. 1253-1262.

Fischer, D. W., J. A. LeFever, J. A. Sorensen, S. A. Smith, L. D. Helms, R. D. LeFever,


S. G. Whittaker, E. N. Steadman, and J. A. Harju, 2005, The influence of
tectonics on the potential leakage of CO2 from deep geological sequestration units
in the Williston Basin: Plains CO2 Reduction (PCOR) Partnership, 17 p.

Flannery, J., and J. Krauss, 2006, Integrated Analysis of the Bakken Petroleum System:
U.S. Williston Basin: American Association of Petroleum Geologists Annual
Convention, Houston , TX., Poster Session Presentation, Search and Discovery
Article #10105, April 10-12, 2006, http://www.searchanddiscovery.net.

Gerhard, L. C., S. B. Anderson, and D. W. Fischer, 1990, Petroleum geology of the


Williston Basin: AAPG Memoir, v. 51, p. 507-559.

Gerhard, L. C., S. B. Anderson, and J. A. LeFever, 1987, Structural history of the Nesson
anticline, North Dakota, in M. W. Longman, ed., Williston Basin: anatomy of a
cratonic oil province, Denver, Colorado, Rocky Mountain Association of
Geologists, p.337-354.

Gerhard, L. C., S. B. Anderson, J. A. LeFever, and C. G. Carlson, 1982, Geological


development, origin, and energy mineral resources of Williston Basin, North
Dakota: AAPG Bulletin, v. 66, no. 8, p. 989-1020.

Hill, R., D. J. Howard, G. Jones, G. Scott, and J. L. Jarrell, 2007, The Play Report
Bakken Shale, Patchwork of Prospectivity. 29 p. www.rseg.com.

Hood, C. C., C. M. Gutjahr, and R. L. Heacock, 1975, Organic metamorphism and the
generation of petroleum: AAPG Bulletin, v. 59, no. 6, p. 986-996.

Hunt, J. M., 1979, Petroleum Geochemistry and Geology, Second Edition, Freeman W.
H. and Co., San Francisco, 617pp.

Jarvie, D. M., and M. S. Johnson, 2008, Maturity and Organofacies Assessment of


Bakken Shale: Implications for New Areas for Exploration and Production,
AAPG Rocky Mountain Section, Search and Discovery Article #90092, July 9-
11, 2008, http://www.searchanddiscovery.net.
194
Johnson, M. S. 2009, Parshall Field, North Dakota - Discovery of the Year for the
Rockies and Beyond, American Association of Petroleum Geologists Annual
Convention, Denver, Colorado, Presentation Session, Search and Discovery
Article #20081, June 7-10, 2009, http://www.searchanddiscovery.net.

Kerr Jr., S. D., 1988, Overview: Williston Basin Carbonate Reservoirs: in Goolsby, S.M.
and M.W. Longman, eds., Occurrence and Petrophysical Properties of Carbonate
Reservoirs in the Rocky Mountain Region: RMAG Guidebook, p. 251-274.

Kume, J., 1963, The Bakken and Englewood Formations of North Dakota and
northwestern South Dakota: North Dakota Geological Society Bulletin 39, 87 p.

Law, C. A., 1999, Chapter 6: Evaluating Source Rocks, in Beumont, E. A. and N. H.


Foster, eds., Treatise Handbook 3: Exploring Oil and Gas Traps, AAPG Special
Publication, p. 6-1 - 6.33.

LeFever, J. A., C. D. Martiniuk, E. F. R. Dancsok, and D. F. Lund, 1991, Petroleum


potential of the middle member, Bakken Formation, Williston basin, in
Christopher, J. E. and F. Haidl, eds., Proceedings of the sixth international
Williston basin symposium: Saskatchewan Geological Society, Special
Publication 11, p. 7494.

LeFever, J. A., 1992, Horizontal drilling in the Williston Basin, United States and Canada
in Schmoker, J. W., E. B. Coalson, C. A. Brown, eds., Geological studies relevant
to horizontal drilling: examples from western north America, Denver, Colorado,
Rocky Mountain Association of Geologists, p. 177-197.

LeFever, J., 2005, Oil production from the Bakken Formation: A short history: North
Dakota Geological Survey Newsletter, v. 32, no. 1, p. 1-6.

LeFever, J., 2008, Whats Happening at Parshall, North Dakota: North Dakota
Geological Survey Newsletter, v. 35, no. 1, p. 1-2.

Lineback, J. A., and M. L. Davidson, 1982, The Williston basin sediment-starved


during the Early Mississippian, in J.E. Christopher and J. Kaldi, eds., Fourth
International Williston Basin Symposium, Saskatchewan Geological Society
Special Publication 6, p. 125-130.

Longman, M. W., ed., 1987, Williston Basin: anatomy of a cratonic oil province: Rocky
Mountain Association of Geologists, p. iii.

Meissner, F. F., 1978, Petroleum geology of the Bakken Formation, Williston basin,
North Dakota and Montana, in D. Rehrig, ed., The economic geology of the
Williston basin: Proceedings of the Montana Geological Society, 24th Annual
Conference, p. 207227.

195
Monicard, R. P., 1980, Properties of reservoir rocks: core analysis: Houston, Gulf
Publishing Company, 168 p.
Murray, G. H., 1968, Quantitative fracture study - Sanish pool, McKenzie County, North
Dakota: AAPG Bulletin, v. 52, no. 1, p. 57-65.

Nordeng, S., 2010, A brief history of oil production from the Bakken Formation in the
Williston Basin: North Dakota Geological Survey Newsletter, v. 37, no. 1, p. 5-9.

Nordeng, S. H., 2010, The Bakken source system: Emphasis on the Three Forks
Formation: 18th Williston Basin Petroleum Conference and Expo, North Dakota
Geological Survey, PowerPoint presentation.

Nordeng, S. H., and J. LeFever, 2008, The Bakken: A question of maturity: 16th Annual
Williston Basin Petroleum Conference, North Dakota Geological Survey,
PowerPoint presentation.

Nordquist, J. W., 1953, Mississippian stratigraphy of northern Montana: Billings


Geological Society, p. 68 - 82.

Osadetz, K. G., L. R. Snowdon, and P. W. Brooks, 1994, Oil families in Canadian


Williston Basin (Southwestern Saskatchewan): Bulletin of Canadian Petroleum
Geology, v. 42, p. 155-177.

Passey, Q. R., S. Creaney, J. B. Kulla, F. J. Moretti, and J. D. Stroud, 1990, A practical


model for organic richness from porosity and resistivity logs: AAPG Bulletin, v.
74, no. 12, p. 1777-1794.

Pemberton, S. G., M. Spilla, A. J. Pulham, T. Saunders, J. A. MacEachern, D. Robbins,


and I. K. Sinclair, 2001, Ichnology and sedimentology of shallow to marginal
marine systems: Geological Association of Canada, Short Course Notes, v. 15,
343 p.

Peterson, J. A., and L. M. MacCary, 1987, Regional stratigraphy and general petroleum
geology of the U.S. portion of the Williston Basin and adjacent areas, in M. W.
Longman, ed., Williston Basin: Anatomy of a Cratonic Oil Province: Denver,
Colorado, Rocky Mountain Association of Geologists, p. 9-43.

Pitman, J. K., Price, L. C., and LeFever, J. A., 2001, Diagenesis and fracture
development in the Bakken Formation, Williston Basin: Implications for reservoir
quality in the middle member: U.S. Geological Survey Professional Paper 1653,
17 p.

196
Pollastro, R. M., T. A. Cook, L. N. R. Roberts, C. J. Schenk, M. D. Lewan, L. O. Anna,
S. B. Gaswirth, P. G. Lillis, T. R. Klett, and R. R. Charpentier, 2008, Assessment
of undiscovered oil resources in the Devonian-Mississippian Bakken Formation,
Williston Basin Province, Montana and North Dakota: USGS National
Assessment of Oil and Gas Fact Sheet, 2 p.

Price, L. C., T. Ging, T.Daws, A. Love, M. Pawlewicz, and D. Anders, 1984, Organic
metamorphism in the Mississippian - Devonian Bakken shale, North Dakota
portion of the Williston Basin in Woodward J., F. F. Meissner, J. L. Clayton, eds.,
Hydrocarbon Source Rocks of the Greater Rocky Mountain Region, Denver,
Colorado, Rocky Mountain Association of Geologists, p. 83-134.

Price, L. C. T., and J. LeFever, 1994, Dysfuntionalism in the Williston Basin: the
Bakken/mid-Madison petroleum system: Bulletin of Canadian Petroleum
Geology, v. 42, p. 187-218.

Sandberg, C. A., and C. R. Hammond, 1958, Devonian system in Williston Basin and
central Montana: AAPG Bulletin, v. 42, no. 10, p. 2293-2334.

Schmoker, J. W., 1979, Determination of organic content of Appalachian Devonian


shales from formation-density logs: AAPG Bulletin, v. 63, p. 1504-1509.

Schmoker, J. W., and T. C. Hester, 1983, Organic carbon in the Bakken Formation,
United States portion of Williston basin: AAPG Bulletin, v. 67, no. 12, p. 2165
2174.

Schmoker, J. W., and T. C. Hester, 1990, Formation Resistivity as an Indicator of Oil


Generation Bakken Formation of North Dakota and Woodford Shale of
Oklahoma: The Log Analyst, January-February, 1990, p. 1-9.

Sloss, L. L., 1963, Sequences in the cratonic interior of the North America: Geological
Society of America Bulletin, v. 74, p. 93-114.

Smith, M. G., and R. M. Bustin, 1996, Lithofacies and paleoenvironments of the Upper
Devonian and Lower Mississippian Bakken Formation, Williston Basin: Bulletin
of Canadian Petroleum Geology, v. 44, p. 495-507.

Smith, M. G., and R. M. Bustin, 2000, Late Devonian and Early Mississippian Bakken
and Exshaw Black Shale Source Rocks, Western Canada Sedimentary Basin: A
Sequence Stratigraphic Interpretation: AAPG Bulletin, v. 84, no. 7 (July 2000) p.
940-960.

Sonnenberg, S. A., and A. Pramudito, 2009, Petroleum geology of the giant Elm Coulee
Field, Williston Basin: AAPG Bulletin, v. 93, no. 9 (September 2009) p. 1127-
1153.

197
Suter, J. R., 2006, Facies Models Revisited: Clastic Shelves in Posamentier, H. W., and
R. G., Walker, Facies Models Revisited, Society for Sedimentary Geology
(SEPM), p. 375.

Tissot, B. P., and D. H. Welte, 1984, Petroleum Formation and Occurrence, 2nd edition:
Springer Verlag, Heidelberg, 699 p.

Walker, R. G., 2006, Facies Models Revisited in Posamentier, H. W., and R. G., Walker,
Facies Models Revisited, Society for Sedimentary Geology (SEPM), p. 375.

Waples, D. W., 1980, Time and temperature in petroleum formation: application of


Lopatin's method to petroleum exploration: AAPG Bulletin, v. 64, no. 6 (June
1980) p. 916-926.

Webster, R. L., 1984, Petroleum source rocks and stratigraphy of the Bakken Formation
in North Dakota in Woodward J., F. F. Meissner, J. L. Clayton, eds., Hydrocarbon
source rocks of the greater Rocky Mountain region, Rocky Mountain Association
of Geologists, p. 57 - 82.

Williams, J. A., 1974, Characterization of oil types in Williston Basin: AAPG Bulletin, v.
58, no. 7, p. 1243-1252.

Wilson, J. L., 1967, Carbonate-evaporite cycles in lower Duperow Formation of


Williston Basin: Canadian Petroleum Geology Bulletin, v. 15, p. 230-312.

198

Das könnte Ihnen auch gefallen