diagenesis and reservoir characterization of the

112
University of Arkansas, Fayeeville ScholarWorks@UARK eses and Dissertations 12-2012 Diagenesis and Reservoir Characterization of the Pennsylvanian Middle Atoka Formation, Sebastian and Logan Counties, West-Central Arkansas Elvis Chekwube Bello University of Arkansas, Fayeeville Follow this and additional works at: hp://scholarworks.uark.edu/etd Part of the Geology Commons , Sedimentology Commons , and the Stratigraphy Commons is esis is brought to you for free and open access by ScholarWorks@UARK. It has been accepted for inclusion in eses and Dissertations by an authorized administrator of ScholarWorks@UARK. For more information, please contact [email protected], [email protected]. Recommended Citation Bello, Elvis Chekwube, "Diagenesis and Reservoir Characterization of the Pennsylvanian Middle Atoka Formation, Sebastian and Logan Counties, West-Central Arkansas" (2012). eses and Dissertations. 583. hp://scholarworks.uark.edu/etd/583

Upload: others

Post on 28-Jan-2022

3 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Diagenesis and Reservoir Characterization of the

University of Arkansas, FayettevilleScholarWorks@UARK

Theses and Dissertations

12-2012

Diagenesis and Reservoir Characterization of thePennsylvanian Middle Atoka Formation, Sebastianand Logan Counties, West-Central ArkansasElvis Chekwube BelloUniversity of Arkansas, Fayetteville

Follow this and additional works at: http://scholarworks.uark.edu/etd

Part of the Geology Commons, Sedimentology Commons, and the Stratigraphy Commons

This Thesis is brought to you for free and open access by ScholarWorks@UARK. It has been accepted for inclusion in Theses and Dissertations by anauthorized administrator of ScholarWorks@UARK. For more information, please contact [email protected], [email protected].

Recommended CitationBello, Elvis Chekwube, "Diagenesis and Reservoir Characterization of the Pennsylvanian Middle Atoka Formation, Sebastian andLogan Counties, West-Central Arkansas" (2012). Theses and Dissertations. 583.http://scholarworks.uark.edu/etd/583

Page 2: Diagenesis and Reservoir Characterization of the

DIAGENESIS AND RESERVOIR CHARACTERIZATION OF THE PENNSYLVANIAN

MIDDLE ATOKA FORMATION, SEBASTIAN AND LOGAN COUNTIES,

WEST-CENTRAL ARKANSAS

Page 3: Diagenesis and Reservoir Characterization of the

DIAGENESIS AND RESERVOIR CHARACTERIZATION OF THE PENNSYLVANIAN

MIDDLE ATOKA FORMATION, SEBASTIAN AND LOGAN COUNTIES,

WEST-CENTRAL ARKANSAS

A thesis submitted in partial fulfillment

of the requirements for the degree of

Master of Science in Geology

By

Elvis Chekwube Bello

Delta State University

Bachelor of Science in Geology, 2001

December 2012

University of Arkansas

Page 4: Diagenesis and Reservoir Characterization of the

ABSTRACT

The Middle Atoka Formation evolved from a stable passive margin during the early

Pennsylvanian time to a rapidly subsiding basin, with the sedimentary fill thickening greatly

southward during the middle Pennsylvanian time. The basin dips in north-south direction. An

east-west anticlines and synclines were observed.

The purpose of this study is to establish the stratigraphic units, reservoir geometry and

distribution, and infer the depositional environments and reservoir quality. Well correlation and

petrographic studies are used to achieve the goals.

Sandstone Point Count method was used. The Point Counts was divided into five

categories. They include the framework grains (quartz, feldspar, and lithic fragments), accessory

minerals (muscovite and biotite), cementing materials (quartz overgrowths, feldspar

overgrowths, dolomite, and calcite cements), pore-spaces (primary and secondary), and the

“Other” (minerals that cannot be identified under the microscope, and matrix and pyrite).

Vertical sequences of sand bodies that are closely spaced and separated by thick marine

shale intervals, and sandstones that blocky signatures and abrupt bases and tops, were correlated

as genetically related sand bodies. Four stratigraphic units; Borum, Turner, Nichol, and Basham

Sandstone Units were identified.

The sandstones of the Middle Atoka Formation are composed of very fine silt to coarse

quartzarenites, subarkoses, sublitharenites, and litharenites. The reservoirs are heterogeneous and

were divided into two: the amalgamated reservoirs (proximal and medial submarine fans), and

the overbank reservoirs (levee-overbank deposits, crevasse splays, and distal lobes). The

geometries of the reservoirs are elongate and radial depending on the stratigraphic units. Quartz

overgrowths and clay cements are intense and advanced and variable within the two reservoirs.

Page 5: Diagenesis and Reservoir Characterization of the

Porosity loss was significantly caused by compaction and quartz cementation. Secondary

porosity, where it occurs, resulted from the dissolution of the labile grains, probably from the

interaction with migrating organic acids or as a result of the increasing geothermal gradient.

Diagenetic processes either enhance the porosity by dissolving mineral grains, or reducing the

porosity by stimulating growth of clay and quartz minerals.

Amalgamated reservoirs contain higher amount of dissolution, clay cements, and lower

quartz overgrowths. Dissolutions are filled with clay cements. Dissolution when present in

overbank reservoirs are better preserved than is in amalgamated portions. Dissolutions and clay

cements are also higher in the south than is in the northern part. Higher clay content in the south

ensures that the dissolution is almost effectively occluded by clay cements. Consequently there

seem to be no net-gain in porosity despite in the area.

Page 6: Diagenesis and Reservoir Characterization of the

This thesis is approved for recommendation to the

Graduate Council:

Thesis Director:

Dr. Xiangyang Xie

Thesis committee:

Dr. Doy L. Zachry

Dr. Walter Manger

Page 7: Diagenesis and Reservoir Characterization of the

THESIS DUPLICATION RELEASE

I hereby authorize the University of Arkansas Libraries to duplicate this thesis when

needed for research and/ or scholarship.

Agreed

Elvis Chekwube Bello

Refused

Elvis Chekwube Bello

Page 8: Diagenesis and Reservoir Characterization of the

ACKNOWLEDGEMENTS

I would like to give special thanks to Dr. Xiangyang Xie, my thesis advisor, who

suggested the study area, procured the data used for the study from Stephens Production

Company, and arranged the meeting between me and the company personnel. Furthermore, his

ideas and thorough timely review of my work several times have immensely improved the

quality of this thesis.

I also appreciate Dr. Doy L. Zachry for helping me develop my knowledge of diagenesis,

and his invaluable guidance and support. He had been tremendous, explanatory, helping me

whenever and wherever he could. I would also like to thank Dr. Walter Manger for reviewing

and providing technical suggestions that added color to this thesis.

I would like to thank Stephens Production Company, Fort Smith, for the opportunity and

resources provided for this project. I wish to express special thanks to two people at Stephens

Production Company, Robert Liner and Luke Martin, who traveled between Fort Smith and

Fayetteville in order to train me on the use of the Petra-IHS WorkStation. The training was

critical to the development of this thesis and I enjoyed and valued what I learned so much.

Thanks to all my friends in the Geology Department who made my stay at the University

of Arkansas an enjoyable one.

Finally, I extend appreciation to my family who believed, encouraged and supported me

in my efforts to complete this project. Special thanks to my mother Mrs. Comfort John Bello and

Dr. Benjamin John for their consistent support of my graduate studies.

Thank you all.

Page 9: Diagenesis and Reservoir Characterization of the

TABLE OF CONTENTS

ABSTRACT

ACKNOWLEDGEMENTS

LIST OF FIGURES

LIST OF PLATES

LIST OF TABLES

CHAPTER 1. INTRODUCTION

1.0 Study Area 1

1.1 Purpose of Study 1

1.2 Previous studies 1

1.3 Methodology 4

1.3.1 Well log correlation 5

1.3.2 Petrographic microscopic study 6

1.3.3 Diagenesis 7

1.3.4 Rock texture 10

CHAPTER 2. The ATOKA FORMATION 12

2.1 Geological setting 12

2. 2 Regional stratigraphy of the Atoka Formation 20

2.2.1 Lower Atoka Formation 20

2.2.2 Middle Atoka Formation 21

2.2.3 Upper Atoka Formation 22

CHAPTER 3. WELL LOG CORRELATION 23

3.1 Well log correlation 23

Page 10: Diagenesis and Reservoir Characterization of the

3.2 The depositional model 25

3.3 The depositional environments 27

3.3.1 Amalgamated deposits 28

3.3.1.1 Proximal deposits 28

3.3.1.2 Medial deposits 29

3.3.2 Overbank deposits 29

3.3.2.1 Levee-overbank deposits 29

3.3.2.2 Crevasse splays 30

3.3.2.3 Distal lobes 30

3.4 The Stratigraphic Units 31

3.4.1 Borum Sandstone Unit 35

3.4.2 Turner Sandstone Unit 37

3.4.3 Nichol Sandstone Unit 38

3.4.4 Basham Sandstone Unit 39

3.5 Evaluation of the sandstone reservoir units 40

3.5.1 Borum sandstone reservoir 40

3.5.2 Turner sandstone reservoir 41

3.5.3 Nichol sandstone reservoir 42

3.5.4 Basham sandstone reservoir 43

CHAPTER 4. PETROGRAPHIC STUDY 45

4.1 Composition 45

4.2 Frame work grains 46

4.2.1 Quartz 46

Page 11: Diagenesis and Reservoir Characterization of the

4.2.2 Feldspar 49

4.2.3 Lithic Fragments 49

4.3 Accessory minerals 50

4.4 Cementing materials 51

4.4.1 Quartz and feldspar overgrowths 51

4.4.2 Clay minerals 52

4.5 Grain sizes 53

4.6 Texture 54

4.7 Diagenesis 55

4.8 Tectonic provenance 57

4.9 The reservoir quality 60

CHAPTER 5. CONCLUSIONS 65

REFERENCES 67

APPENDICES 78

Page 12: Diagenesis and Reservoir Characterization of the

LIST OF FIGURES

Figures Page

1. Regional geological setting of Arkoma Basin 3

2. Study area map showing townships and ranges 4

3. Type of well logs used for the study 9

4. Top of the Middle Atoka marker 10

5. Diagrammatic cross sections showing the tectonic development 13

6. Foreland structural features and tectonic elements showing Mississippi Valley 15

7. Major foreland structural features and tectonic elements 16

8. Fan models proposed by Shanmugam and Moiola (1988) 19

9. Types of well log motifs 24

10. Conceptual submarine fan depositional model 26

11. Simplified depiction of a slope-basin environment 27

12. Well log signatures depicting the reservoir types 32

13. Showing clean sands for the four stratigraphic units 33

14. Contour map of the study area 34

15. North-South cross section A-A’ 34

16. East-West cross-section A-A’ 35

17. Index map showing the two cross sections 35

18. Interval isopach map of the Borum Sandstone Unit 36

19. Isopach map of Turner Sandstone Unit 37

Page 13: Diagenesis and Reservoir Characterization of the

20. Interval isopach map of Nichol Sandstone Unit 38

21. Interval thickness of Basham Sandstone Unit 39

22. The isopach map of Borum Gross Sands 40

23. The isopach map of Turner Gross Sands 41

24. The isopach map of Nichol Gross Sands 42

25. The isopach map of Basham Gross Sands 43

26. The isopach map showing sediments dispersal patterns 44

27. Folk’s sandstone classification (Folk, 1980) 45

28. QFL diagram of the Middle Atoka Formation 46

29. Paragenetic Sequence 55

30. Clay minerals distribution in the reservoir rock 56

31. Triangular QmFLt plot showing mean framework grains 58

32. QFL distribution of sedimentary rocks in various tectonic regimes 58

33. Triangular QFL plot of the middle Atoka Formation 59

34. Triangular QmFLt plot middle Atoka Formation 60

35. The percentage abundance of the primary and secondary porosities 78

36. The composition of the framework grains 78

37 Percentage abundance of the component of cementing materials 80

38 showing the distribution of the accessory minerals 80

39. Sanderson 3-9H well modal thin section composition 81

40. State 2-22 well modal thin section composition 82

41. Carter 2-26 well modal thin section composition 83

42. Hale 5-19 well modal thin section composition 84

Page 14: Diagenesis and Reservoir Characterization of the

43. Holcomb 1-36 well modal thin section composition 85

44. Schultz-Teeter 3-18H well modal thin section composition 86

LIST OF PLATES

Plate Page

1. Photomicrographs of Quartz and Others 47

2. Photomicrographs of Textures and Others 48

3. Photomicrographs of Overgrowths, Sorting and Others 50

4. Photomicrograph of Authigenic Minerals and Others 51

5. Photomicrographs of Plant Fragments, Muscovite and Others 52

6. Photomicrograph of Deformation and Others 54

LIST OF TABLES

Tables Page

1. Stratigraphic column of the Atoka Formation 6

2. Informal sandstone units of middle Atoka Formation 6

3. Petrographic Data of Samples of Schultz-Teeters 3-18H 87

4. Petrographic Analysis of Thin-Sections obtained from Holcomb 1-39 well 88

5. Petrographic Analysis of Thin-Sections obtained from Hale 5-17H Well 89

6. Petrographic Analysis of Thin-sections obtained from Hale 5-17H well cont. 90

7. Petrographic Analysis of Thin-Sections obtained from Carter 2-26 well 91

8. Petrographic Analysis of Thin-Sections obtained from Carter 2-26 well cont. 92

9. Petrographic Data of Samples of State 2-22H Well 93

Page 15: Diagenesis and Reservoir Characterization of the

10. Petrographic Data of Samples of Sanderson 3-9H Well 94

11. Well Names and Gross Sands of the four sandstone units 95

12. Percentages of the framework grain components for all the depths 96

Page 16: Diagenesis and Reservoir Characterization of the

1

CHAPTER 1

INTRODUCTION

1.0 Study area

The study area is located on the southern part of the Arkoma Basin (Figure 1). It is

situated north of the Ouachita Mountains and south of the Backbone and Washburn anticlines.

Geographically, it is limited to the west-central Arkansas, encompassing eight townships area. It

covers a total area of 288 square-miles, encompasses Townships 7N-5N and Ranges 32W-27W

in Sebastian and Logan Counties (Figure 2).

1.1 Purpose of Study

The purpose of this study is to establish the stratigraphic units, determine the reservoir

distribution and geometry, depositional systems, and document the reservoir quality of the

Middle Atoka Formation.

1.2 Previous studies

As one of the main gas producing units in the Arkoma Basin, the Atoka Formation has

been a major research topic and exploration target since 1930s. Croneis (1930) defined the

surface extent of the Atoka Formation within Arkansas, including the distribution, thickness,

lithology, fossils, and the petrologic characters. Caplan (1957) determined the thickness of the

Atoka Formation within northwestern Arkansas using subsurface data. Scull (1961) using well

log correlation, provided three-fold division of the Atoka Formation in a series of cross sections

for the Arkoma basin. Buchanan and Johnson (1968) produced the structural cross section of the

Bonanza Field with lettered correlation markers A-C (descending order) to describe the tectonic

history of the Arkoma basin. They designated marker B and C as the middle and lower Atokan

boundaries respectively.

Page 17: Diagenesis and Reservoir Characterization of the

2

Merewether (1961) discovered the southward thickening trend of the Atoka Formation.

However, Buchanan and Johnson (1968) attributed the thickening to the growth faulting in the

Middle Atoka Formation. Thomas (1983) while working on the southern Arkoma basin in

western Arkansas was able to determine the southward thickening trends of the three members of

the Atoka Formation. Gilbreath and Haley (1972) mapped the thickness of the Cecil series, an

informal Lower Atoka member of the Atoka formation in Arkansas. Merewether (1961) and

Merewether and Haley (1972) were able to identify and describe the formation contacts, folds

and structures of the Atoka Formation during the mapping of the Delaware and Knoxville

quadrangles.

The investigation of the Bloyd-Atoka boundary by Zachry and Haley (1975) showed that

Henbest (1953) and other previous workers had misidentified the Middle Bloyd Sandstone as the

basal sandstone of the Atoka Formation. In 1975, Zachry and Haley established the proper

position of the Bloyd-Atoka boundary and published detailed reports on the position of the

Bloyd-Atoka boundary. Other works in the area include studies that established the

sedimentation and plate tectonics of the Ouachita Mountains and Arkoma Basin; the distribution

and the evidence of normal faulting in the East Central portion of the Arkoma Basin; and the

sedimentation transition from the shelf to the basin in the Oklahoma – Arkansas area (Shields,

1961; Briggs and Roeder, 1975; Sutherland and Manger, 1979).

Page 18: Diagenesis and Reservoir Characterization of the

3

Figure 1. Regional maps showing geological setting of Arkoma Basin (modified from

Encyclopedia of Arkansas).

Arkoma Basin

Ozark Plateau

Ouachita Mountains

West Gulf Coastal Plain

Page 19: Diagenesis and Reservoir Characterization of the

4

Figure 2: Study area map showing the townships and ranges.

Additionally, there are many unpublished Master theses at the Department of

Geosciences, University of Arkansas on the middle Atoka Formation. They include Williams,

(1983), Fritsche, (1981), Normand, (1986), Harris, (1983), Mollison, (1983), and Gatson, (1985).

1.3 Methodology

Gamma Ray, Resistivity Logs, and Conductivity Logs are used for the correlation of

wells in the area (Figure 3). The stratigraphic units, distribution, geometry, and the depositional

environments of the sandstone reservoirs are made through the correlation of fifty wells.

Petrographic studies of fifty thin-sections obtained from six-wells are combined with the

geophysical studies to determine the reservoir quality.

T6N R31W T6N R30W T6N R29W T6N R28W

T5N R31W T5N R30W T5N R29W T5N R28W

•West central Arkansas

•Sebastian & Logan

Counties

•Townships 5 N – 6 N

•Ranges 28 W – 31 W

•288 square miles

SebastianLogan

Crawford Franklin

Scott

Polk

Yell

Johnson

MontgomeryGarland

Pope

Perry

Study area

Page 20: Diagenesis and Reservoir Characterization of the

5

1.3.1 Well log correlation

All the well logs used for the study are converted to raster logs (Figure 3). The

determination of the lithologies, well correlations, and the construction of the cross sections are

accomplished by the use of IHS-Petra-Workstation. The top of the middle Atoka Formation is

marked by low resistivity shale. It is recognized on well log by the blocky-type signature of the

Conductivity Log (Figure 4). This blocky type signature is recognized and correlatable across the

study area. The vertical variation of the log signatures was used to infer changing depositional

environments. Wells that have full/partial penetration of the middle Atoka Formation were

selected for the correlation. In total, ninety-eight wells are selected. Fifty well logs are finally

used for the correlation of wells and the construction of isopach and contour maps. A series of

cross sections are made across the area to ascertain the dips, structures, and the stratigraphic

units of the middle Atoka Formation. Gross isopach maps are used to determine the stratigraphic

thickness, geometry, and the distribution of the sandstone reservoirs in the area.

In this study, four sandstone units of the middle Atoka Formation are identified. They are

the Borum, Turner, Nichol, and the Basham Sandstone Units. These units are the parallel

equivalence of the Morris, Areci, Bynum, Freiburg and Casey Sandstone Units used in other

previous studies. In addition, the Basham Sandstone Unit is composed of both Morris and Areci

Sandstone Units of the previous studies.

Page 21: Diagenesis and Reservoir Characterization of the

6

System Series Formation Sandstone Unit

Desmoinessian

Carpenter A

Alma

Carpenter B

Morris

Areci

Bynum

Freiburg

Casey

Sells

Jenkins

Dunn B

Dunn C

Paul Barton

Cecil Spiro

Pennsylvanian

MorrowanBloyd

Hale

Middle

Upper

Lower

Hartshorne

AtokaAtokan

Table 1. Stratigraphic column of the Atoka Formation, Arkoma Basin, Arkansas (after

Zachry 1983).

Sandstone unit

(This study)

Morris

Areci

Bynum Nichol

Freiburg Turner

Casey Borum

Morrowan

Formation Member Sandstone unit

Pennsylvanian

System

Basham

Upper Atoka Formation

Middle Atoka

Lower Atoka Formation

Bloyd

Atoka

Table 2. Informal sandstone unit of middle Atoka Formation used by petroleum

companies and this study.

1.3.2 Petrographic microscopic study

Fifty thin-sections were made from the well cuttings obtained from four vertical and two

horizontal wells. The thin-sections were impregnated with blue epoxy in order to reveal the

porosity. Glycerin liquids were applied on the surfaces of the thin sections and covered with the

Page 22: Diagenesis and Reservoir Characterization of the

7

Glass-Microscopic-Cover-Slip. A hand operated mechanical counting stage was attached on the

stage of the Nikon-Transmitted-Light Microscope. This was used to construct the counting grid.

Irregular grids of 91-300 were made on each thin sections based on their quality. It was

necessary to repeat grid lines in some cases. Magnifications of 10X and 20X were used for point

counts, 20X and 40X for the photography, while 40X was used for the grain size measurements.

Depending on the quality thin sections each contained 250-300 point counts. Therefore,

some of the data may not reflect the true composition characters.

The point-counts were divided into five categories. They include the framework grains,

accessory minerals, cementing materials, pore-spaces, and “others”. The framework grains

include quartz (monocrystalline quartz and polycrystalline quartz), feldspar (orthoclase,

plagioclase, and microcline), and the lithic fragments (cherty fragments, metamorphic and the

sedimentary fragments), cementing materials (clay minerals, quartz overgrowth, feldspar

overgrowth, dolomite, and calcite cements), “others” (pyrite, matrix, and other minerals that

cannot be identified under the microscope), accessory mineral (muscovite and biotite), and the

porosity.

The percentages of each category was calculated and tabulated in Tables 4 to 10. The

occurrences of some particular constituents did not coincide with the cross-polar point during

counting and are therefore recorded as zero percent. This is not an indication of the absence of

such constituent minerals, but should rather be considered as occurring in trace amounts.

1.3.3 Diagenesis

Diagenesis is defined as modifications made to rocks prior to burial, as well as their

complete destruction during continued weathering (Burley et al, 1984). The modifications

include changes occurring immediately after deposition, through compaction, lithification,

Page 23: Diagenesis and Reservoir Characterization of the

8

alteration, and total transformation of the mineral grains. Diagenesis persists toward the onset of

metamorphism. The diagenetic processes have negative impact on the reservoir quality by

altering the porosity and permeability.

Page 24: Diagenesis and Reservoir Characterization of the

9

Figure 3. Type log used for the study

Resistivity Log

Conductivity Log

Gamma-Ray

Holcomb 1-36

Page 25: Diagenesis and Reservoir Characterization of the

10

Figure 4 Top of the Middle Atoka marker that is capped by low resistivity shale.

1.3.4 Rock texture

Grains were randomly selected on thin-sections for grain size measurement. The long

axes of such minerals were measured to determine the grain size and distributions. Pore throats

Top of middle

Atoka marker:

Top of low

blocky

conductivity

motif toward the

track

Gregory

Page 26: Diagenesis and Reservoir Characterization of the

11

were also measured to obtain the pore throat diameter. All the measurements were then

converted to millimeter using Nikon-Microscopic-Calibrated-Values of the Microscope

Objective and the Ocular Mike Michrometer. The degree of sorting, roundness and sphericity

were visually estimated. Ternary plots were made for quartz, feldspar and lithic fragments to

determine the composition of the sandstones (Folk, 1980) and the tectonic provenance

(Dickinson, 1985).

Page 27: Diagenesis and Reservoir Characterization of the

12

CHAPTER 2

ATOKA FORMATION

The Atoka Formation is lower Pennsylvanian in age. It is composed of sequences of

alternating sandstone and shale units. It is overlain by the Hartshorne Sandstone of Desmoinesian

age, and unconformably lies on the top of the Morrowan Series. It is divided informally into

lower, middle, and upper Atoka Formation. The Atoka Formation is one of the main gas

producing units in the Arkoma Basin.

2.1 Geological Setting

The Arkoma Basin is 250 miles long and 20 to 50 miles wide (Branan, 1968). It lies

immediately north of the Ouachita Mountains to which it is closely related. Previous studies

suggest that the development of the Arkoma Basin and the Ouachita Orogenic Belt involves the

collision of continental plates (Viele, 1979; Houseknecht and Kacena, 1983; Lillie et al 1983).

Houseknecht (1986) proposed five different tectonic-stratigraphic stages of the development of

the basin (Figure 5) with the middle Atoka Formation forming between stages three and four;

from early Mississippian/earliest Atokan to the middle Atokan times.

Page 28: Diagenesis and Reservoir Characterization of the

13

Figure 5: Diagrammatic cross sections showing the tectonic development of the southern

margin of North America and the Ouachita orogenic belt much shortened from north to south

(from Houseknecht, 1986). (A) late Precambrian- earliest Paleozoic records the onset of

extension. (B) During the late Cambrian-earliest Mississippian, the Ouachita Ocean is at its

widest and deepest. (C) In early Mississippian-earliest Atokan times, the Ouachita Basin is

segmented by the development of an accretionary wedge, into a forearc basin to the south and a

trench basin to the north. (D) During early to middle Atokan time, thrusting of the accretionary

wedge onto the southern margin of North America occurred. (E) In the late Atokan-

Desmoinesian, thrusting of the Ouachita orogeny onto North America continued.

North Arkoma Basin Ouachita Mountains

Sabine

uplift

South

?

Page 29: Diagenesis and Reservoir Characterization of the

14

The first stage records the onset of the basin extension and the development of a Proto-

Atlantic Oceanic Basin on the southern margin of the North America Continent during late

Precambrian and Cambrian time. The sediments associated with this rifting include carbonates,

shale and sandstone. These sediments were deposited on a stable shelf environment. Walper

(1977), and Houseknecht and Weaverling (1983) maintained that the onset of the Anadarko and

the Reelfoot Rift Basin development also began at this time. The Anadarko Basin evolved into

an aulacogen, because of the intrusive and extrusive igneous activity. The Reelfoot Basin was

inactive before the igneous activity (Walper, 1977).

The southern margin of North American Continent preserves a belt of deformed

Paleozoic strata extending from the eastern margin of the Mississippi Embayment in west-central

Alabama to the Marathon and Solitario Uplifts of west Texas, and from there into the adjacent

northern Mexico. This folded and faulted belt is the Ouachita Orogen and only about 20 percent

of its nearly 2500 kilometers is actually exposed. The bulk of the Ouachita Orogen is in the

subsurface beneath the Mesozoic and Cenozoic successions of the Gulf Coastal Plain and

Mississippi Embayment. Consequently, the only significant outcrops of the deformed strata

comprising the orogen are the Ouachita Mountains of southwestern Arkansas and southeastern

Oklahoma, and the Marathon Uplift, which is actually a fenster through the Cretaceous of the

Edwards Plateau in west Texas. (A fenster is formed when an underlying block of thrust-faulted

strata is exposed in a small area by erosive action - http://www.answers.com/topic/thrust-fault).

The Ouachita Orogen resulted from the collision of northern South America (as part of

Gondwana) with the southern midcontinent of North America (as part of Laurussia). In addition

to the Ouachita-Marathon Fold-Belt Mountains, the collision also formed a series of foreland

basins between the deformed orogenic belt and the stable cratonic region.

Page 30: Diagenesis and Reservoir Characterization of the

15

Most of these basins produce petroleum, particularly the shale resource plays of the

Arkoma Basin, and the Fort Worth Basins.

Figure 6: Major foreland structural features and tectonic elements adjacent to the

Ouachita Fold-Belt as interpreted by Denison, 1989. Note that the Arkoma Basin is not shown to

connect with the Black Warrior Basin across the Mississippi Embayment (modified after Miall,

2008).

Ouachita Mountains

Ozark Plateau

Ouachita Salient

?

Page 31: Diagenesis and Reservoir Characterization of the

16

Figure 7: Major foreland structural features and tectonic elements adjacent to the

Ouachita Fold-Belt as interpreted by Viele and Thomas (1989). Note that those authors indicate

that the Arkoma Basin is continuous with the Black Warrior Basin across the Mississippi

Embayment (modified after Miall, 2008).

Nevertheless, it has been a common interpretation in many articles to extend the eastern

Arkoma Basin of Arkansas across the Mississippi Embayment and connect it to the Black

Warrior Basin in western Mississippi (Figure 3). While there is no question that the Ouachita

fold-belt extends through this area as proved by drilling and geophysics (Ginzburg, et al, 1983;

Mickus and Keller, 1992), the Mississippi Embayment is part of an aulacogen that includes the

Anadarko Basin. That aulacogen represents failed Precambrian to Cambrian rifting of the

southern North American craton, when it was part of the Precambrian supercontinent Rodinia,

and later Pannotia. A remnant structure from that rifting is the Mississippi Valley Graben, also

Ozark Plateau

A

Arkoma Basin

Ouachita Mountains

Metamorphic belt

Frontal

belt

Page 32: Diagenesis and Reservoir Characterization of the

17

known as the Reelfoot Graben, which remains tectonically active to this day. The Mississippi

Valley Graben disrupts any structural continuity between the Arkoma and Black Warrior Basins,

which are not connected across the Mississippi Embayment.

The second stage of the basin development occurs from late Cambrian to earliest Atokan

age. This stage is marked by the onset of the ocean basin closure, accommodated by southward

subduction beneath Llanoria (Houseknecht and Kacena 1983). The occurrence of the detrital

sediment in the basin is an indication of orogenic provenance (Morris 1974). The local

abundance of volcanic fragments in the Stanley Formation of the Ouachita Mountain

(Houseknecht and Kacena 1983) is evidence that the subduction was underway in Mississippian

time, although the exact time is difficult to determine. Houseknecht and Kacena (1983) and

Nicholas and Waddell (1982) documented Carboniferous volcanic fragments in the subsurface,

south of Ouachita Mountains. This represents traces of the magmatic arc that developed along

the northern margin of Llanoria, which marks the onset of the accretionary wedges that are

associated with subduction zones. Carbonates, shale and sandstones are deposited on the

continental passive margin. The sediment thickness of the Arkoma Basin increases gradually

southward towards the Ouachita Mountains.

The third stage of the development of Arkoma Basin is composed of rocks of the Lower

and the Middle Atoka Formation. The thickness increases from north (1000/1300 feet) to south

(over 15,000 feet) perpendicular to the axis of the basin (Zachry and Pierre-Orly, 2008). The

passive margin shelf remains the site of deposition at this stage, which culminated with the

deposition of the basal Spiro Sandstones. Houseknecht and Kacena (1983) and Dickson (1974)

suggested that the more than 5 km thick, thin-bedded alternation of sandstones and shales were

deposited in the deep remnant ocean basin in a relatively short time. Houseknecht and Kacena

Page 33: Diagenesis and Reservoir Characterization of the

18

(1983) and Thomas (1985) propose that some of the sediments deposited into the basin were

derived from the east, transported westward in a longitudinal fashion, and ultimately deposited

on the submarine fans settings and the surrounding abyssal plains (Figure 8).

The fourth stage occurred during early to middle Atokan time. At this stage, the closure

of the remnant ocean basin by subduction was complete, and the northward advancing

subduction complex was being thrusted over the rifted continental margin of the North America

(Houseknecht and Kacena 1983). The bending of the southern margin of the North American

continental crust was caused by the weakening subduction of the continental crust and the

thrusting of the northward advancing subduction complex (Houseknecht and Kacena 1983, and

Dickinson, 1974, 1976). This led to the widespread development of normal faults that strike

parallel to the Ouachita Mountains. Houseknecht (1983) maintain that these normal faults broke

the undeformed continental crust and the shelf-slope rise geometry that was prevalent along the

passive continental margin since Precambrian/Cambrian time, and that this resulted in an abrupt

increase in both subsidence and sedimentation.

This down-thrown normal faulting is syndepositional, and contributed to the abrupt

thickenings of the sediments toward the south. The abrupt increase of the strata southward are

syndepositional with the subsidence as well as the normal faults. The thickening represents the

transition from fluvial, shallow marine facies to deep marine sedimentation from north to south.

It also marks the beginning of the development of the foreland basin.

Page 34: Diagenesis and Reservoir Characterization of the

19

Figure 8: Fan models proposed by Shanmugam and Moiola (1988) showing (A) middle

fan setting (channel dominated) for the DeGray section and outer fan setting (lobe dominated),

(B) longitudinal Fan system in a regional tectonic framework (Shanmugam, 2006)

The fifth stage is composed of rocks of the Upper Atoka Formation and Desmoinesian

Strata. It is composed of sandstones, shales and coals of shallow marine origin. In the late

Atokan-Desmoinesian, thrusting of the Ouachita orogeny onto North America continued and

predominated as the subduction pushed northward against the strata deposited during Devonian

to early Atoka time (Houseknecht and Kacena, 1983). The peripheral foreland basin was formed

N

N

OklahomaArkansas

Texas

Mississippi

Louisiana

Distal fan Medial fan Proximal fan

Several kilometers

Page 35: Diagenesis and Reservoir Characterization of the

20

at this stage due to the Ouachita thrust that culminated in the deposition of shallow marine,

deltaic, and fluvial deposits (Dickenson, 1974, and Houseknecht and Kacena, 1983).

During late Atokan time, the closure of the Arkoma Basin was complete at the southern

side of the basin (Viele, 1973). The deposits at this stage were mostly deltaic coals, which

prograded southward across the muddy open shelf (Zachry and Sutherland, 1984). Deposition at

the western end of the basin was composed of alternating shale and sandstone layers in a

transgressive open shelf (Zachry and Sutherland, 1984).

2. 2 Regional Stratigraphy of the Atoka Formation

The Atoka Formation is informally divided into lower, middle, and upper Atoka

Formation (Tables 1 & 2). The division is based on the differences in sedimentary responses to

tectonic processes that prevailed during the formation of the basin (Buchanan and Johnson, 1968;

Haley and Hendricks, 1972; Fritsche, 1981; Thomas, 1983; Zachry, 1983; and others).

2. 2. 1 The Lower Atoka Formation

The Lower Atoka Formation lies unconformably over the Morrowan Series (Buchanan

and Johnson, 1968). It is composed of eight sandstone units (Tables 1 & 2) separated by thin-

layers of shale. The Spiro Sandstone is the basal sandstone unit of Lower Atoka. The Spiro unit

is widespread within the Arkoma Basin (Berry and Trumbly, 1968). This sandstone unit is fine-

grain, calcitic, submature, and fossilized. The rocks of the Lower Atoka Formation are

depositionally related to those of the Morrowan Series. They are believed to be deposited by

southward progradation of high destructive deltaic systems and the related shallow marine

processes (Scull, 1961; Berry and Trumbly, 1968; Uszynski; 1982; and Stephens; 1985). The

thickness of the lower Atoka Formation is uniform and consistent across the basin (Zachry,

1983).

Page 36: Diagenesis and Reservoir Characterization of the

21

2. 2. 2 The Middle Atoka Formation

Four to five sandstone units (Table 1) separated by shale intervals characterize the middle

Atoka Formation (Zachry, 1983). It is overlain by the Carpenter “B” Unit of the upper Atoka

Formation and underlain by the Sells Sandstone Unit, which is the top of lower Atoka

Formation. The unit thickens from about 1500 feet southward around the Mulberry fault system

to over 4000 feet in the central part of the basin. The outcrops of the lower sandstone units of the

middle Atoka Formation are found north of Cass fault and are buried north of the Mulberry fault

systems (Zachry, 1983).

The deposition of the middle Atoka Formation in the southern area, south of the growth

faults, is created through submarine fan systems in a deep-marine setting. The sediments are

transferred and transported in a north-south direction, and longitudinally in an east-west, and

deposited as turbidities (Zachry, 1983; Stephens, 1985; Williams, 1983; and Shanmugam, 2006).

The deposits at the northern and central parts of the basin result from the processes of deltaic

systems that prograded slowly on a subsiding shallow marine shelf (Stephens, 1985; and

Williams, 1983).

The comparison of the middle and lower Atoka Formation indicates that the Middle

Member displays higher variability in thickness, lithology, and also contain higher amount of

shale (Zachry, 1983). Zachry (1983) and Fritsche (1981) described lobate and elongate

geometries of the Middle Atoka Formation. The abrupt southward increase in the thickness of

the unit suggests that the rate of subsidence and sedimentation increased in the middle Atokan

time (Zachry, 1983). The nomenclatures used for the sandstone units of the middle Atoka

Formation are the same as that of the Petroleum Operators in the study area (Tables 1 & 2).

Page 37: Diagenesis and Reservoir Characterization of the

22

In this study, the sandstones of the middle Atoka Formation are divided into four

sandstone units separated by shaly units. The four units are the Borum, Turner, Nichol, and the

Basham sandstone units. Most of the wells used for this study did not have complete penetration

of the Borum sandstone unit. The Borum, Turner, Nichol, and the Basham sandstones units are

the parallel equivalence of the Morris, Areci, Bynum, Freiburg and Casey sandstone units used

in previous studies (Table 2). Basham sandstone unit is parallel equivalent of Morris and Areci

sandstone units used in previous studies (Table 2).

2. 2. 3 The Upper Atoka Formation

This unit is composed of three Sandstone Units that thicken southward (Buchanan and

Johnson, 1968; Zachry, 1983). Toward the south, deposits of the Upper Atoka Formation are

deposited in a prograding high destructive deltaic system (Zachry, 1983; Stephens, 1985). In the

west, they are deposited in a shallow marine, inner and middle shelf of the deep-marine

environments (Scull, 1961). There is gradual replacement of the deep-water deposits of the

Upper Atoka Formation by the deltaic system of the overlying Hartshorne sandstones unit. The

gradual replacements of the upper Atoka sandstones by the deltaic deposits of the Hartshorne

Sandstone create localized erosional surfaces (Thomas, 1983).

Coal bed deposits cap the alternating deposits of shales and sandstones within the upper

Atoka Formation (Zachry, 1983; Gatson, 1985; and Scull, 1961). These coals are presence within

shale beds and within few feet of the sandstone units (Haley, 1961). Six coal layers were

identified in the Knoxville Quadrangle, and a very extensive one near Centerville in the Yell

County that is east of this study area (Merewether, 1967).

Page 38: Diagenesis and Reservoir Characterization of the

23

CHAPTER

Well Log Correlation

Forty eight wells with well logs were used for the lithology correlation. Three types of

well logs were used. They include the Gamma Ray, Resistivity and the Conductivity Logs

(Figure 3).

3.1 The Correlation of Sandstone Units

Well logs in the area have a variety of log motifs, ranging from blocky, thin and

symmetrical, thin-bedded to crescent-shaped signatures. Others include successions of fining

upward and coarsening upward sequences. Most of the blocky signatures in the northern part of

the area are serrated compared to the massive type-signatures at the southern portion (Figure 9).

The blocky log signatures usually have an abrupt top and bottom. Blocky signatures with

gradational tops are common at the northern portion of the study area.

Top of the middle Atoka Formation is marked by low resistivity shale. It is identified on

the well log by the blocky signature of the Conductivity Log, with an excursion toward the track

on the log (Figure 4). It can be recognized and correlated laterally across the study area. The

depth covered in this study is between 2500 ft. to 8200 ft.

Based on the log motifs, the reservoirs were differentiated into two broad groups. They

include the amalgamated reservoirs and the overbank reservoirs (Figure 12). Blocky gamma

signatures (serrated and non-serrated) are used to identify amalgamated reservoirs. Apart from

the blocky signatures, other log types are the identification features of the overbank reservoirs,

and they are dominated with thin-bedded signatures.

Page 39: Diagenesis and Reservoir Characterization of the

24

Figure 9 Types of well log motifs depicting depositional environments in the area.

Amalgamated reservoirs are channelized deposits for the most part. They include the proximal

and medial submarine fans. Overbank deposits, on the other hand, include levee-overbank

deposits, crevasse splays, and the distal lobes. The two terminologies are for explanatory

purposes in this study. Detailed description of the two reservoir units is included later in this

chapter.

Individual bed thicknesses range from a few centimeters to about 150 feet in the area.

The western part of the study area is predominantly composed of crescent-shaped, thin and

Blocky &

serrated

Blocky & massive

Fining upward

successions

Thin &

symmetrical

Crescent shape

Thin beds

PROMINENT LOG SIGNATURES

Tracks

Holcomb 1-36 Carter 2

Page 40: Diagenesis and Reservoir Characterization of the

25

symmetrical signatures. Other signatures include the blocky signatures (serrated) and the

successions of coarsening upward signatures. These signature-types grade into blocky (massive)

signatures, thin-bedded signatures, and the successions of fining upward sequences toward the

central portion of the area. Toward the northeastern portion, the log type changes again to

serrated blocky types, and to thin and symmetrical motifs. All the type of log signatures in the

area is shown in Figure 9.

Individual beds are amalgamated. They have a great degree of lateral continuity in the

north and at the central portion of the area. Gross sands in the northern portion are high, but

lateral connectivity of the individual beds is poor compared to the central portion. There is a

general thinning of individual beds away from the central portion of the area as seem on the well

logs. The gross sand bodies change between the four different stratigraphic sand units and

between the north and south of the area. There is an inconsistent and isolated thickening and

thinning of beds throughout the basin. This makes the correlation between wells difficult.

3.2 The Depositional Model of the Middle Atoka Formation

The deposits of the middle Atoka Formation are described as turbidites, deposited by the

submarine fan systems in a slope-basin deep-water environment (Figure 11). This is based on the

type of the well log signatures. According to Normark (1978), Walker, (1979), and Shanmugam,

(2006), turbidite systems are described in terms of the submarine fan model.

The sediments deposited into the basin are derived from the north, east, and the south of

the study area. This is based on the series of isopach maps, and the tectonic provenance studies.

Sediment failures at the continental shelf-margin gave rise to the gravity-driven processes, which

transported the sediments away from the shelf regions, into the deep-water environments, where

Page 41: Diagenesis and Reservoir Characterization of the

26

they were deposited as turbidites. Gravity-transport processes include mass-transport and

turbidity currents. Further description of these terms is not made in this study.

The failure of the shelf-margin sediments could be triggered by any of the following

events according to Shanmugam (2006): 1) Eustatic changes in sea level, 2) submarine volcanic

activity, 3) earthquakes, such as the 1929 Grand Banks tremor off the U.S. east coast, 4) over-

steepening of the submarine slope, such as near the mouth of the Magdalena River, Colombia, 5)

high sedimentation, such as in the Mississippi delta-front setting, Gulf of Mexico, 6) tsunamis, 7)

storm waves, and 8) biologic erosion of submarine canyon walls.

Figure 10 showing the conceptual submarine fan depositional model of the study area

(Kim et al, 2012).

Levee-overbank

Crevasse splay

Feeder

Channel

Non-channelized

(Medial)

Distal lobe

Channelized

(Proximal)

100’s

of km

Page 42: Diagenesis and Reservoir Characterization of the

27

The submarine fan systems are formed when changes of the slope gradient occur near the

basin floor. When this happens, the transport energy of the medium drops, channels splits and

then deposit their loads as submarine fans (Figure 11). Several episodes of this nature produce a

lobe-shifting system of submarine fans (Figure 10). The proximal and medial ends of the fans are

the sandier part of the submarine fans. On both sides of the channel, fine-grained materials in

suspension are deposited as mud-dominated levee-overbank and crevasse splay systems (Figure

10). Beyond the proximal and medial depositional lobes are the distal fans. The distal fans are

dominated by finer-grained sediments (Figure 10).

Figure 11. Simplified depiction of a slope-basin environment

3.3 The Depositional Environments

The sandstone reservoirs in the study area are mud-dominated sandstones. As said earlier,

they are grouped into amalgamated deposits and overbank deposits. This division is based on the

well log signatures. Amalgamated deposits are identified on the well logs by blocky gamma-ray

signatures that have abrupt bases. These blocky gamma signatures are both massive and serrated,

and channelized and non-channelized deposits. The overbank deposits, on the other hand, are

Shelfal margin Continental slope Deep-water

Several kilometers

Page 43: Diagenesis and Reservoir Characterization of the

28

recognized by every other signature-type except the blocky gamma-ray motifs. The overbank

deposits are mostly non-channelized.

3.3.1 Amalgamated Deposits

The amalgamated deposits have very good vertical and lateral beds continuity. They are

better developed in Nichol and Basham sandstone units. Within the Nichol sand unit, they occur

as blocky and massive sand bodies, characterized by abrupt tops and bottoms as suggested by the

well logs. The amalgamated deposits with the abrupt tops and bottoms are very prominent at the

central portion of the study area and occur as isolated sand bodies with gradational upper tops

and abrupt bottoms in the northern portion.

The abrupt contacts of the amalgamated deposits are attributed to the erosive activities of

the deep-marine currents and/or sudden and rapid deposition from the sand influxes resulting

from flood events. The amalgamated deposits are characterized by very poor to moderate sorting.

They are clean sands and usually overlain and underlain by the shaly overbank facies. The ratio

of sand/mud is high as suggested from the well logs. The presence of interbedded shales is more

noticeable at the north than in the south. This interbedded shale may serves as flow barriers, and

at the same time as a sealing layer. They have the potential to occlude vertical and lateral flow

continuity. On the whole, amalgamated deposits have great potential for significant hydrocarbon

production in the study area.

3.3.1.1 Amalgamated Proximal Deposits

The amalgamated proximal submarine fans are recognized on the well log by blocky

gamma-ray signatures that have abrupt tops and bases. The amalgamated proximal submarine

fans are elongate, and are distributed within the central portion of the study area as shown on the

gross sands isopach of the Nichol sandstone unit (Figure 24). They are best developed in the

Page 44: Diagenesis and Reservoir Characterization of the

29

Nichol sandstone unit, have relative poor development in Basham unit, and are absent in the rest

of the other two units (Borum and Turner sandstone units). They have a very wide lateral

continuity and good vertical stacking of sand bodies. They are underlain and overlain by shales

which may represent seal rocks at both ends. The proximal sand bodies may still be sealed

internally by the interstitial shale layers. Other forms of stratigraphic traps that are not seen on

the well logs may result from the switching of the submarine fan lobes as shown in Figure 10.

3.3.1.2 Amalgamated Medial Deposits

These groups are recognized on the well logs by blocky gamma-ray signatures that have

abrupt bases and gradational tops. The amalgamated medial fans are relatively thick, isolated,

and composed of intermittent alternation of sandstones and shales as supported by the well logs.

Individual bed amalgamation is poor vertically and horizontally compare to proximal sand

bodies. Medial sand bodies are more common at the northern portion of the area. The Basham

sandstone unit is typically made up of the medial submarine fan type for the most part.

3.3.2 Overbank Deposits

The overbank deposits form moderate to well-sorted reservoirs. The ratio of sand/mud is

low to moderate depending on the sandstone units. Thin-bedded deposits overlap all the reservoir

types in the area.

3.3.2.1 Levee-Overbank Deposits

Levee-overbank deposits form the sandier type of the overbank deposits. They are

recognized on the well logs by their crescent-shaped signatures. They form laterally to the

channel (Figure 10). The levee-overbank reservoirs are sparse and uneven. They are difficult to

correlate between wells, and formed compartmentalized reservoirs as suggested by their lateral

discontinuities on the well logs. Slatt et al, (2004) suggested that horizontal wells offer the best

Page 45: Diagenesis and Reservoir Characterization of the

30

potential to achieve maximum production from the levee-overbank reservoirs because of the high

degree of compartments, and the lack of good lateral bed connectivity.

3.3.2.1 Crevasse Splays

Crevasse splays, on the other hand, are recognized on the well logs as thinning and fining

upward sequences. They are composed of mixed silt-shale sediments as suggested by the well

logs. According to Walker (1979) thinning and fining upward sequences of sands formed as a

result of the infillings of the distributary channels that are incised into the levee-overbank

deposits. Accordingly, there is a gradual abandonment of the channels and gradual filling by

thinner and finer siltstones and shales fractions. They are rare in the area and when present, are

usually discontinuous laterally and vertically. Individual beds are thin, from a few centimeters to

two feet.

3.3.2.3 Distal Lobes

Beyond the proximal and medial submarine fans are the distal fans. They are also rare

and not well-developed. They are distinguished on the well logs by the vertical stacking

sequences according to Slatt el al, (2004). They are thin individual amalgamated sand bodies.

They are both channelized and non-channelized fan complexes in the study area. When they are

channelized, they formed lobes that are characterized by the successions of fining upward

patterns. This type is only noticeable in the southeastern portion of the study area. When they

formed non-channelized deposits, they form successions of coarsening upward sequences on the

well logs. This type too is present in isolation at the northwestern side of the area. According to

Slatt et al, (2004), distal lobes form a stratigraphic trap as a result of the down-dip termination of

the proximal and medial sands caused by the lateral continuous shaly layers of the distal lobes

that were not eroded before subsequent deposition of another layers. Laterally continuous layers

Page 46: Diagenesis and Reservoir Characterization of the

31

of shales of the distal lobes may partly be responsible for the lack of productivity of the Holcomb

well. This is because deposits that have the successions of fining upward signatures are present

on the log obtained from the Holcomb well within the interval of the Turner sandstone unit.

3.4 Stratigraphic Units

Vertical sequences of sand bodies that are closely spaced and separated by thick marine

shale intervals were correlated as genetically related sandstone deposits that resulted from one

depositional cycle according to Allen (1975). In addition, blocky gamma-ray signatures that have

abrupt bases are also mapped as genetically related sand bodies. On this basis, four informal

stratigraphic units were identified and mapped. They include the Borum, Turner, Nichol and

Basham sandstone units in ascending order (Table 2). The division of the reservoirs into four

different stratigraphic units is based mostly on the changes in the characters of the gamma ray,

and then the resistivity and conductivity logs.

The characters and the thickness of these sand bodies vary vertically and laterally

between wells across the study area. These variations make the stratigraphic correlation very

difficult even over a short distance. The abrupt regional north-south thickenings of the

sedimentary fills of the middle Atoka Formation were not observed at the study area. The basin

dips in a north-south direction, and series of east-west anticlines and synclines can be observed

along the east-west direction within the central portion (Figure 14).

Page 47: Diagenesis and Reservoir Characterization of the

32

Figure 12. Well log signatures depicting the depositional environments (after Miall,

2008).

Amalgamated

deposits

Overbank

deposits

Overbank deposits

Amalgamated

deposits

Holcomb Hales

Two type of sandstone reservoirs

Page 48: Diagenesis and Reservoir Characterization of the

33

Figure 13: Showing clean sands (API 75) for the four stratigraphic units

TMA

Hattabaugh Holcomb

Basham sands

Nichol sands

Turner sands

Borum sands

Page 49: Diagenesis and Reservoir Characterization of the

34

Figure 14: Contour map showing the location of the six wells used for study thin sections.

The top of the contour is the subsea, which is total depth minus the elevation (ground surface).

Figure 15: North-South cross section A-A’

Carter

Schultz-Teeter - H

Sanderson - H

Hale

Holcomb

State

Subsea (SS)

Page 50: Diagenesis and Reservoir Characterization of the

35

Figure 16: East-West cross-section of selected wells A-A’

Figure 17: Index map showing the two cross sections

3.4.1 Borum Sandstone Unit

The Borum sandstone unit is overlain and underlain by the Turner sandstone unit and the

lower Atoka Formation, respectively. Laterally persistent, thick shale units separated it from the

overlying sandstone unit. Seventeen wells have full penetration through the Borum sandstone

Page 51: Diagenesis and Reservoir Characterization of the

36

unit. There is a bit of an uncertainty about the distribution and the geometry of the Borum

sandstone unit. The most common well log signatures in this unit are the thin, symmetrical and

crescent log motifs. Other type of log signatures includes a few blocky types, and the successions

of fining upward sequences at the southeastern side of the area. The upward fining successions

are rare, isolated and unevenly distributed. The lithology is mainly interbedded sandstone,

siltstone and shale. Lateral bed connectivity is generally poor across the area.

Figure 18 Interval isopach map of Borum sandstone unit

The interval depth of the unit ranges from 6192 ft. to about 8200 ft. at townships T6N

R29W and T5N R29W, respectively. The thickness of this unit is up to 1750 ft. Individual bed

thickness ranges from a centimeter or less to over 10 ft. The interval isopach map of this unit

shows that the sediment dispersal is in the north-south trend, specifically at the township

boundary between T6N R231W and T6N R30W, and at township T5N R28W.

Page 52: Diagenesis and Reservoir Characterization of the

37

3.4.2 Turner Sandstone Unit

Stratigraphically, the Turner sandstone unit is above the Borum sandstone interval. It is

overlain by the Nichol sandstone unit. Most wells penetrated the entire Turner Sandstone

interval. The thickest part of this interval is located at the northern portion. It is up to about 2000

ft. as shown in Figure 19. There is no observed general thickening trend. The depth at which the

unit is encountered is between 5263 ft. to 8181 ft.

Figure 19. Isopach map of Turner sandstone unit

The most common log signatures of the Turner sandstone unit include the thin, crescent,

and blocky log motifs. Blocky signatures are generally serrated within this interval. In addition,

the crescent-log motifs are poorly developed, usually thin and serrated. Individual bed thickness

ranges from a centimeter to about twenty feet. Blocky and interbedded units of sandstones and

shales are prominent in the Tyler and Hattabaugh well logs situated in townships T6N-R31W &

T5N R29W. Successions of fining upward patterns are observed on some of the well logs at

township T5N R29W. Combinations of the amalgamated reservoirs and overbank reservoirs are

suggested by the log signature types. Distal lobes are suggested by the succession of fining

Page 53: Diagenesis and Reservoir Characterization of the

38

upward log patterns. The interval isopach map shows that the sediment dispersal within this

interval is haphazard, but generally in north-south direction across the area (Figure 18).

3.4.3 Nichol Sand Unit

Figure 20. Interval isopach map of Nichol sandstone unit

The Nichol sandstone unit stratigraphically is above the Turner sandstone unit and lies

below the Basham sandstone unit. The interval thickness is up to 1790 ft. The amalgamated

deposits are well-developed and have good lateral connectivity. They are more prominent and

laterally correlatable within the central portion of the townships T6N R29W and T5N R29W.

The unit is separated from the Turner sandstone unit by thin shale beds, and separated by a very

thick shaly unit from the overlying Basham sandstone unit. The bases and tops are usually

abrupt.

The most common log type is blocky and usually massive as suggested by the well logs.

The amalgamated deposit is the dominant reservoir in this unit, especially at the south-eastern

portion of the study area. The log type changes from the blocky motifs at the eastern area to thin,

symmetrical, and crescent-shaped signatures toward the western area. The changes in the log

Page 54: Diagenesis and Reservoir Characterization of the

39

signatures suggest changes in the depositional environments; from the amalgamated systems of

deposition in the east to overbank sedimentation in the west. The interval isopach map of this

Unit shows haphazard sediment dispersal trends (Figure 20). The isolated thickenings of the

interval suggest multiple accommodation spaces.

3.4.4 Basham Sandstone Unit

Figure 21. Interval thickness of Basham sandstone unit

This sandstone unit is above and below the Nichol sandstone and the Carpenter “B” units

of the middle and upper Atoka Formation, respectively. It is separated vertically from the

overlying and underlying sandstone units by thick shaly units that are laterally continuous across

the study area. Lithologically, it is much more similar to the Borum and Turner sandstone units.

All well log types (blocky, crescent-shaped, thin and symmetrical signatures) are present in the

Basham sandstone unit. At the northern portion, serrated and blocky signatures, and the upward

fining signatures are dominant. Thin and symmetrical signatures are common at the southern

portion of the area. Crescent-log signatures are common and evenly distributed within the unit.

Successions of coarsening upward signatures are sparse and occur in isolation. Varieties of well

Page 55: Diagenesis and Reservoir Characterization of the

40

log signatures are suggest diverse depositional environments and systems. The thickness of

Basham sandstone unit is up to 1800 ft. as shown in Figure 21. The trend of the sediments

dispersal is haphazard as reflected on the interval isopach map (Figure 21).

3.5 Evaluation of the Sandstone Reservoir Units

The individual clean sand bodies of each of the stratigraphic units were sum up to

obtained the respectively gross sands. The gross sands were then used to generate gross thickness

maps to determine the geometry and the distribution of reservoir of the four stratigraphic units.

The thicker portions on the isopach maps of the gross sandstones are indicative of amalgamated

deposits. Conversely, as the gross sands become thinner, the overbank deposition becomes

dominant and the ratio of sand/shale decreases

3.5.1 Borum Sandstone Reservoir

Figure 22: The isopach map of Borum gross sands

As stated earlier, seventeen wells penetrated this unit. The thickness of the gross sand is

up to 300 ft. They are well developed at the eastern part of the area, at townships T6N-R29W

and T5N-R29W. The thickenings extended across the eastern boundary into the western portion

of the area (Figure 22). The Borum sandstone unit is sand-dominated reservoir. The ratio of

Page 56: Diagenesis and Reservoir Characterization of the

41

sand/shale sediment is low at the extreme southeast and southwest. It is higher toward the north

and at the central portion. Both amalgamated and overbank reservoirs are identified in this unit.

Amalgamated reservoirs are common at the central portion of the area. Away from the central

portion of the study area, the overbank sedimentation increases at both directions.

3.5.2 Turner Sandstone Reservoir

Figure 23: The isopach map of Turner gross sands

The reservoir of the Turner sandstone unit is isolated and unevenly distributed across the

area (Figure 23). The thickness of the gross sand is up to 256 feet with elongate and radial

geometry observed. Overbank reservoir is dominant in this interval. It is shaly, and poor to well-

sorted reservoir. The ratio of sand/shale is low. Individual beds are thin commonly a few

centimeters to two feet. The thickness of the individual bed varies laterally. Individual beds are

laterally discontinuous and pinch out within short distances. The upper contact of the Turner

sandstone unit is sharp to gradational. Their bases are gradational, usually made up of succession

of fining upward sequences and coarsening upward sequences in most cases.

Page 57: Diagenesis and Reservoir Characterization of the

42

3.5.3 Nichol Sandstone Reservoir

Figure 24: The isopach map of Nichol gross sands

The thickness of the Nichol’s gross sand is up to 300 feet. This unit has the best

developed amalgamated deposit. They are connected vertically and horizontally especially at the

central portion, and at the extreme eastern part of the study area (Figure 24). The reservoir is

sand-dominated and composed mostly of the amalgamated proximal submarine fans. The ratio of

sand/shale sediment is high. Individual beds are in the range of a few centimeters to as high as

150 ft. Blocky well logs with abrupt bases and tops are the dominant type in this unit. The unit is

poorly sorted and possesses more dissolution cavities. The geometry of the reservoir is elongate

and is distributed within the central portion as shown in Figure 24.

Page 58: Diagenesis and Reservoir Characterization of the

43

3.5.4 Basham Sandstone Reservoir

Figure 25: The isopach map of Basham Gross Sands

The Basham sandstone unit is composed of an amalgamated reservoir and is usually

interbedded layers of sandstones and shales. The upper contacts are mostly gradational contacts.

The degree of beds amalgamation is poor to moderate. Lateral bed connectivity is poor when

compared to the Nichol sandstone unit, but much better than the other two units. The thickness of

the gross sand is up to 280 feet. The Basham sandstone unit is sand-dominated reservoirs. The

ratio of sand-shale fractions is high. The amalgamated medial deposits are dominant in this unit.

Individual beds are thin, ranging from a centimeter to about 8 feet. The geometry of the reservoir

is elongate to radial, isolated and unevenly distributed across the area (Figure 25).

The elongate portions (Figures 22, 23, 24, and 25) of the four stratigraphic units were

extracted from each gross isopach maps using a cutoff thickness of 150 ft.

Page 59: Diagenesis and Reservoir Characterization of the

44

Figure 26: The isopach map showing sediments dispersal patterns.

Then, the extracted elongate portions were superimposed on each other using the principle of

superposition, in order to determine the sweet spots, the direction of the sediments dispersal, and

the depositional systems in the study area (Figure 26). The sweet spots are located at the eastern

side and within the central portion of the area (Figure 26). The sediments dispersals into the

basin were from the northeastern side and the southern portion of the study area. Submarine fan

systems were interpreted as a result of the fan shifting pattern of depositions (Figure 26).

Page 60: Diagenesis and Reservoir Characterization of the

45

CHAPTER 4

PETROGRAPHY STUDY

Fifty thin-sections made from well cuttings of four vertical and two horizontal wells were

analyzed for the petrographic study. The sandstone point-counting method was utilized to

document the composition and the tectonic provenance of the middle Atoka Formation.

4.1 Composition

Point-counting of the framework grains was in accordance with Folk’s method (Folk,

1968). The components of the point counts were divided into framework grains, cementing

materials, accessory minerals, porosity, and the “others”. When the percentages of the

components of framework grains were plotted on the QFL diagram (Figure 26), the compositions

of the middle Atoka sandstone is quartzarenites (41%), sublitharenites (39%), litharenites (3%),

and subarkoses (17%) in Figure 27.

Figure 27: Folk’s sandstone classification (Folk, 1980). Q-pole includes all types of

detrital quartz grains (mono and poly crystalline) except cherty fragments. F-pole includes all

Q

LF

10

25

SublithareniteSubarkose

Quartzarenite

Arkose Lithicarkose LithareniteFeldspathic

litharenite

Page 61: Diagenesis and Reservoir Characterization of the

46

types of detrital feldspar grains (mono and poly crystalline rock fragments). L-pole includes all

other rock fragments, including chert. Granite and gneiss fragments are not included.

Figure 28: QFL Diagram of the Middle Atoka Formation

4.2 Framework Grains

These include quartz, feldspar and lithic fragments. Framework grains average 69% of

the modal composition of sandstones as shown in table 5.

4.2.1 Quartz

Quartz is the major framework grain in the section. It averages 87% of the framework

grains. Under the microscope, low birefringence, lack of cleavage, and extinction patterns were

used to identify it. Extinction pattern was undulose to straight under cross polarized light. Two

types of quartz, monocrystalline and polycrystalline quartz were recorded. Monocrystalline

quartz is the most common type of quartz. It has an average of 92%. The numbers of quartz with

Lithic Fragements0 10 20 30 40 50 60 70 80 90 100

Quartz Arenites

0

10

20

30

40

50

60

70

80

90

100

Feldspar

0

10

20

30

40

50

60

70

80

90

100

Schult-teeter

Holcomb

Sanderson

State

Carter

Hale

F

QQ

FL

Page 62: Diagenesis and Reservoir Characterization of the

47

straight extinctions are more common. The percentages of undulose and straight extinction were

not determined.

Polycrystalline quartz is less common. It has an average of 8%. Polycrystalline quartz

displayed a varying degree of high birefringence under cross polarized light by the individual

grains. Plates 1, 2 and 3 shows monocrystalline and polycrystalline quartz grains under the

microscope.

Plate 1: Photomicrographs of (A) Partial to complete grain dissolution, and advance

quartz overgrowths. (B) Quartz grains with straight extinction, cherty fragment, and

polycrystalline quartz grains. (C) Grain edge dissolution, and fractured grains. (D) Quartz grain

exhibiting undulose extinction and straight extinction.

Polycrystalline Quartz

Undulose Extinction

Chert Fragments

Undulose extinction

Straight Extinction

Advance quartz overgrowth

Grain dissolution

Grain Fractures

Grain edge dissolution

A B

C D

Page 63: Diagenesis and Reservoir Characterization of the

48

Most of the quartz grains exhibit straight extinction (Plate 1). They are free of inclusions

and lack vacuoles. Some are fractured and cloudy in appearance (Plate 2). Grain-edge dissolution

of quartz overgrowths is rare to common (Plate 3). Pressure-induced dissolution due to

compaction has resulted in some suturing between the grains. Polycrystalline quartz ranges from

grains with straight jagged borders to grains with elongate minerals.

Plate 2: Photomicrographs of: (A) Well-compacted grains cemented by quartz cement.

Grain boundaries are from straight, angular, rounded to well-rounded grains (B) Grain

replacement, alteration, and dissolution (C) Matrix, polycrystalline quartz grains, and lithic

fragments. (D) Clay-filled pores.

Quartz cementation, well-compacted Grain replacement, alteration, & dissolution

Polycrystalline, Lithic fragments, matrix

Grains alteration to clay

A B

C D

Page 64: Diagenesis and Reservoir Characterization of the

49

4.2.2 Feldspar

Orthoclase, microcline and plagioclase feldspar are grouped together as feldspar. They

average 4% of the framework grains. Orthoclase is by far the most common type. There are

traces of feldspar recognized as partially replaced feldspar grains (Plate 2). Some feldspar grains

exhibit straight and untwined extinction pattern. Orthoclase exhibits a blocky outline with

undulose extinction. Some of the feldspar grains have undergone partial to complete dissolution,

and while others have been completely altered to clay minerals (Plate 2, 3 and 4). Selective

replacement of feldspar is common (Plate 2).

4.2.3 Lithic Fragments

Lithic fragments average 9% of the framework grains. They include all rock fragments

and cherty fragments. No further differentiation of the lithic fragments is made in this study.

Most of the rock fragments are sedimentary materials and a few isolated cherty fragments. They

are sub-rounded to rounded grains.

Page 65: Diagenesis and Reservoir Characterization of the

50

Plate 3: Photomicrograph of: (A) Very fine, well to poorly sorted grains, composed of

monocrystalline quartz grains with straight extinction, mudstone, matrix, fractured grains, and

muscovite. (B) Advance quartz overgrowths, with quartz grains exhibiting undulatory and non-

undulatory extinction. (C) Healed fractures by secondary quartz cementation. (D) Coating of

grains before quartz cementation.

4.3 Accessory Minerals

Accessory minerals include biotite and muscovite. They average 3% of the modal

components of the sandstones. Muscovite is by far the most abundant accessory mineral. It

makes up 99% of the accessory grains. It is present in virtually all of the samples. Others that are

present in traces include pyrite and biotite.

Moderately sorted

Muscovite

Mudstone

MatrixMono-quartz

Fractured grain Advance quartz overgrowth

Undulose & straight extinction

Quartz edge

dissolution

Fracture healed by quartz cementation Clay coatings before quartz overgrowth

A

C

B

D

Page 66: Diagenesis and Reservoir Characterization of the

51

Plate 4: Photomicrograph of: (A) Clay linings prevented quartz cementation, but also

occluded pore throats. (B) Partial dissolution of grains, and clay-filled voids. (C) Grain to grain

contact, isolated dissolution voids, and leached grain. (D) Poorly interconnected pores.

4.4 Cementing Materials

Cementing minerals resulted from the diagenetic alteration of detrital material including

cements that were carried in suspension. They occur as clay cements and quartz overgrowths on

the most part. Others that occur in traces include feldspar overgrowths, calcite cements, and

dolomite. They constitute an average of 17% of the sandstone modal composition.

4.4.1 Quartz and Feldspar Overgrowths

Quartz and feldspar overgrowths average 39% and 0.4% respectively of the cementing

materials. Quartz overgrowths occur mostly as cementing materials on grains boundaries (Plates

Clay-cemented grains, no quartz overgrowth Clay-filled voids

Partial Dissolution

Grain to grain contacts

Isolated dissolution structures

Poorly interconnected pores

Clay mineral occluded porosity

Effective grain coatings

Leached remnant

A

C

B

D

Page 67: Diagenesis and Reservoir Characterization of the

52

1, and 3). Sometimes the grains and overgrowth are separated by clay rims (Plate 4). Secondary

quartz overgrowths were also observed as fracture fillings in detrital quartz grains (Plate 3).

These fractures were probably deformed by compaction and were later healed by secondary

quartz cementation. Sometimes overgrowths showed signs of partial edge dissolution (Plate 3).

Plate 5: 10 X Photomicrographs of: (A) Poorly-sorted grains and a deformed muscovite

fragment. (B) Plant fragments. (C) Mudstone with intercalated chips of shale. (D) Broken

muscovite with sharp and angular edges.

4.4.2 Clay Minerals

Clay mineral has an average of 59% of the cementing materials. Clay mineralization in

the reservoir occurred in three ways: laminar clays (disperse between interbedded units of

sandstones and shales), dispersed clays (pore-lining and pore-filling clays between grains), and

Deformed muscovite

Poorly sorted

Plant fragments

Poor plate

Mudstone Broken muscovite

Angular edge

A B

CD

Page 68: Diagenesis and Reservoir Characterization of the

53

structural clays (selectively replaced grains). Laminar and dispersed clays are both pre- and syn-

depositional. Structural clays, on the other hand, are post-depositional.

Pore-lining clays were identified as alteration products (rims) around detrital grains (Plate

4). In some instances, they also appeared as cement infilling the intergranular pore throats (Plate

4-A).

The composition of the sandstones that are classified as “others” makes up 5% of the

composition. The constituent minerals in this category include matrix, pyrite, and those minerals

that cannot be identified under the microscope.

4.5 Grain Sizes

Grain size is determined by measuring randomly selected grains along their long axes.

Quartz grains were used for the measurement as a result of their stability and abundance. The

measured size of the grains ranges from 0.008 mm to 0.82 mm (Tables 3 to 9). The measured

grains fall within silt and sand fractions. They range from very fine silt to coarse sand size

fractions according to Wentworth (1922). The average of the total grains is 0.2 mm.

Page 69: Diagenesis and Reservoir Characterization of the

54

Plate 6: 20 X photomicrograph of: (A) Sharp edges of polycrystalline quartz grains. (B)

Well-sorted, rounded to well-rounded grains. (C) Poorly sorted grains. (D) Poorly-sorted grains

with quartz overgrowths, grains sutures, and dissolution cavity.

4.6 Texture

Sorting is poor, moderate and well-sorted. Roundness and sphericity were made by visual

estimation (Plate 4). Grain contacts are sutured and broken at times (Plate 4 and 6). Muscovite is

broken, bent, and angular (Plate 5). Quartz and feldspar grains are angular to rounded grains. The

edge of the detrital grains may be sutured or overgrown by quartz. Some thin sections have some

plant fragments (Plate 5).

Polycrystalline grains Well-sorted, well rounded &

compacted grains

Poorly sorted, angular to rounded grains Poorly sorted

Grain edge dissolution

A

C

B

D

Page 70: Diagenesis and Reservoir Characterization of the

55

4.7 Diagenesis

Diagenesis Early Late

Compaction

Pore-filling and pore-lining clays

Dissolution

Quartz overgrowths

Secondary cementation

(secondary clay and quartz cements)

Figure 29: Paragenetic Sequence

The paragenetic sequence was established from the petrographic studies. Some of the

significant diagenetic processes that have impacted the reservoir tremendously include

compaction, shale diagenesis, quartz overgrowths, clay cements, and dissolution.

The reservoir quality is the product of depositional and diagenetic processes. The source,

transport systems, and the depositional environments controlled the initial sediment texture,

composition, porosity, and permeability. These depositional characteristics together with early

sediment compactions, shale diagenesis, and the acid nature of the pore waters, resulted in the

dissolution of some grains, and the precipitation of pore-filling and pore-lining clay. Pore-filling

and pore-lining clays were identified as alteration products, such as rims around detrital grains

under the microscope (Plate 4). In some instances, they occur as the cement infilling the

intergranular pore spaces (Plate 4-A). These early diagenetic processes impacted negatively on

the reservoir quality by reducing the intergranular porosity and permeability of the reservoir.

Page 71: Diagenesis and Reservoir Characterization of the

56

As diagenesis continues with further burial and increased temperature gradient, intense

compaction and dissolution of grains occurred as a result the increased temperature gradients.

Further addition of pore waters that drives the diagenetic events occurs through dehydration; that

is, the pore waters derived from the destruction of organic matters, grain dissolutions, and the

destruction and alteration of earlier formed clay minerals. Grain rotation, slippage,

reorganization of grain fabrics, and plastic and brittle deformation of grains are some of the

effects of this stage. Plastic and brittle grain deformations were recognized by grain-edge

sutures, fractures, and distortion of grains under the microscope.

These diagenetic processes have further destroyed the reservoir quality, completely

occluding the porosity within the overbank reservoirs. Other diagenetic effects include grain

replacement, recrystallization, dissolution, intense and advanced quartz overgrowths and clay

cementation. This stage will persist toward an onset of metamorphism (Burley, 1987).

Figure 30: Clay minerals distribution in the reservoir rock (modified after Boyd et al,

1995).

Laminar clays Structural clays

Dispersed clays as pore coatingsDispersed Clays as pore fillings

Page 72: Diagenesis and Reservoir Characterization of the

57

4.8 Tectonic Provenance

Dickinson et al, (1983) demonstrated that standard QFL and QmFLt diagrams that are

used for plotting the framework modes of sandstones can also be used to determine the tectonic

provenances of sandstones. On the QFL and QmFLt ternary diagrams shown in Figures 30 and

31, Q includes both the monocrystalline quartz (Qm), and polycrystalline quartz (Qp). Feldspar

includes the totality of feldspar grains, which include the plagioclase, orthoclase, and microcline.

Lithic fragments are made of sedimentary, igneous, and metamorphic rock fragments, and

include the cherty fragments. Volcanic fragments are absent in all the samples. Cherty rock

fragments are scarce to absent. Lt is the totality of polycrystalline quartz fractions and the lithic

fragments (Dickinson, 1985).

Three main classes of tectonic provenances are defined by Dickinson ternary diagram

(1983 and 1985) as shown in Figures 30 and 31. They include: the continental sources (craton

interior and uplifted basement); the magmatic arc provenances (dissected and undissected arcs);

and the recycled orogeny (subduction complex, collision and the foreland uplift).

Page 73: Diagenesis and Reservoir Characterization of the

58

Figure 31. Triangular QmFLt plot showing mean framework grains derived from

different provenances: Qm is monocrystalline quartz grains; F is total feldspar grains; Lt is total

polycrystalline quartz, including lithic fragments (after Dickinson, 1985).

Figure 32 showing QFL distribution of sedimentary rocks in various tectonic regimes

(modified from http://csmres.jmu.edu/geollab/fichter/SedRx/Sedtect.html); Qt = Qm + Qp; Qm =

a

Qm

F Lt

Quartzose

recycled

Craton

Interior

Basement

uplift

Lithic recycled

Transitional

recycled

Transitional

continental

Transitional

arc

Undissected

Dissected

arc

Mixed

Q

FL

Undissected arcTransitional arc

Dissected arc

Recycled orogeny

Basement

uplift

Transitional

continental

Craton

interior

Page 74: Diagenesis and Reservoir Characterization of the

59

monocrystalline quartz; Qp = Polycrystalline quartz; F = feldspars; L = Lithics (after Dickinson,

1985).

When the sandstones in the area were plotted on the QFL and QmFLt diagrams (Figures

32 and 33), they plotted within the fields of craton interiors, recycled orogeny, and quartzose

recycled (Figures 30 and 31). According to Dickinson (1985), points that plotted within the field

of craton interior are derived from either interior basins, platforms, oceanic basins or a

combination. The areas that fit these descriptions on the geologic map (Figures 6 and 7) are the

continental north and the eastern part of the Arkoma Basin.

Figure 33. Triangular QFL plot of the middle Atoka Formation showing means of

framework grains derived from different types of provenances. The sandstone falls in the field of

continental block and recycled orogeny provenances.

Lithic Fragements0 10 20 30 40 50 60 70 80 90 100

Quartz Arenites

0

10

20

30

40

50

60

70

80

90

100

Feldspar

0

10

20

30

40

50

60

70

80

90

100

Schult-teeter

Holcomb

Sanderson

State

Carter

Hale

F

QQ

F

L

Page 75: Diagenesis and Reservoir Characterization of the

60

Figure 34. Triangular QmFLt plot middle Atoka Formation showing means of framework

grains derived from different types of provenances. The sandstone falls in the field of craton

interior and quartzose recycled.

In addition, the points that plotted within the fields of the recycled orogeny and quartzose

recycled were derived from uplifted terranes, fold belts or both. Uplifted terranes and fold belts

around the area are located south of the Arkoma Basin (Figures 6 and 7).

By integrating the results obtained from the tectonic provenances and the isopach map

that depicted the direction of the sediments dispersal (Figure 26), I concluded that the sandstones

of the middle Atoka Formation were derived from the northeastern and southern portions of the

Arkoma Basin. Because further studies were not made on lithic fragments, feldspar, and quartz

fractions, the certainty of the exact sources was not made.

4.9 Reservoir Quality

Point-counted porosity averaged 8% of the sandstone modal composition. Porosity

distribution is highly variable among the samples examined. The reservoir quality is also chaotic

and heterogeneous as determined from the textural studies and the statistical analysis. Scatter

Q

Lithic Fragments0 10 20 30 40 50 60 70 80 90 100

Monocrystalline Quartz

0

10

20

30

40

50

60

70

80

90

100

Feldspar

0

10

20

30

40

50

60

70

80

90

100

Sanderson

State

Carter

Hales

Holcomb

Schult-teeter

Qm

F Lt

Qm

Page 76: Diagenesis and Reservoir Characterization of the

61

plots of the framework grains and cementing materials against total porosity respectively, were

made to see if the trend of the reservoir quality can be determined or be predicted, and to assess

where major differences in pore volumes are occurring. All regression coefficients fell below

0.36, which suggest very weak correlations, and therefore, fall within the error bars.

Porosity is depositional related and is dependent on sediments sorting and grain size

distributions. The reservoir qualities were mainly destroyed by sediments compaction, clay and

quartz cementations. Secondary porosity, where it occurs, resulted from the dissolution of the

labile grains, probably from the interaction with migrating organic acid waters or as a result of

the increasing geothermal gradient. These diagenetic processes effects are both negative and

positive, either enhancing the porosity by dissolving mineral grains, or reducing the porosity by

stimulating the precipitation and growth of clay and quartz minerals. Quartz overgrowths and

clay cements are intense and advanced and variable within the two reservoirs.

Amalgamated reservoirs contain a higher amount of dissolution, clay cements, and lower

content of quartz overgrowths. Most of the dissolution cavities within these reservoir types are

filled with clay cements. Dissolution, when it is present in overbank reservoirs, is better

preserved than in amalgamated portions as result of the lower clay cements. In addition, the

degree of dissolution and clay cement is higher in the south than in the northern part of the study

area. Higher clay content in the south ensures that the dissolution is almost effectively occluded

by clay cements. This phenomenon leads to almost no net-gain in porosity in the south.

As said earlier, grain dissolution is significantly higher within the amalgamated portions

of the reservoirs, and dissolution of grains generates silica cements. These silica cements are

exported away from their sources (amalgamated reservoirs) into the overbank portions of the

reservoir, where they precipitate as quartz overgrowths. It seems like the amalgamated reservoirs

Page 77: Diagenesis and Reservoir Characterization of the

62

tend to act as the silica cement generators and the exporters, while overbank reservoirs act as the

recipients of the silica cements. These suggestions are based on the observations that higher

concentrations of dissolution and quartz overgrowths are found in the amalgamated portions and

the overbank reservoirs respectively.

One unique distinction between the amalgamated and overbank reservoirs is in the degree

of heterogeneity in each of the reservoirs. The amalgamated portions show more textural

heterogeneity in terms of the amount of dissolution cavities, clay cements, and quartz

overgrowths. The reservoir quality and the degree of heterogeneity of the amalgamated

reservoirs were controlled by the degree of sediments sorting and the grain size distributions.

The portion of the reservoirs that were less sorted corresponded with areas with better reservoir

quality. The overbank zones tend to be more homogeneous, with higher quartz overgrowths, less

dissolution cavities and clay cements. Generally, the overbank reservoirs were highly

compartmentalized than the amalgamated reservoir reservoirs, as suggested by the well log

signatures, and by the lateral discontinuities of individual beds; as observed from multi-well

correlation across the study area.

Two types of heterogeneities were observed in the study area. They are the basin and the

rock heterogeneity.

Basin heterogeneity occurred in the form of vertical and lateral heterogeneity. Vertical

heterogeneities were identified on the well logs by the degree of serrated well log signatures,

which reflect the amount of shales that are interspersed between clean sand bodies. On the other

hand, lateral heterogeneities reflect the extent of horizontal bed connectivities were made

apparent by multi-well correlation across the study area. Basin heterogeneity controls the degree

of compartments within each reservoir. Overbank reservoirs have much more compartments than

Page 78: Diagenesis and Reservoir Characterization of the

63

the amalgamated reservoirs. These types of heterogeneity have negative impact on the recovery

rate because it creates huge production differences between well-to-well and laterally across well

bores.

Rock heterogeneities were identified by the degree of diagenetic changes in the

composition and texture of the original rock facies. These changes were driven by the interaction

between pore-fluids, organic matters, and the framework grains within the reservoir, through

which cementing materials like clay cements and quartz overgrowths precipitated. This kind of

heterogeneity may complicate the completion of the horizontal well processes by introducing

unpredicted reservoir variability arising from the variation in cementations.

The increases of clay content toward the south and within the amalgamated reservoirs

cause these portions to be shalier. This has negative consequences on the reservoir. For instance,

wells in these areas may still be productive, but fracturing the formation effectively may become

more complicated, which may cause production and recovery rates to be well below those at the

northern portion of the area. Better lateral connectivity and thickness of the amalgamated

reservoirs may counter balance the negative impact of clay cements, and therefore, act as better

performance drivers within these areas all things being equal.

Cementations were placed into two categories based on the impact on the reservoir

quality. They include the rim cements and the occluding cements, which is in accordance with

the Pittman and Larese classification (1991).

Rim cements were syn-depositional in the area. They include clay coatings that

precipitated on the surfaces of the framework grains as interspersed between clean sandstones.

Where clay coatings were enveloped by quartz cementation, the dissolution of the framework

grains was nearly absent. Rim cements were more noticeable in overbank reservoirs. According

Page 79: Diagenesis and Reservoir Characterization of the

64

to Houseknecht (1982), even small amount of rim cements have great tendency of reducing

permeability without altering the porosity. Occluding cements were post-depositional. They

included the pore-filling clays and quartz overgrowths. They occluded cavities left by the

dissolution of the framework grains. They exhibit no preferential relationships to the framework

grains surfaces unlike the rim-types.

Cementations in the study area are capable of causing significant reduction in

hydrocarbon production. The negative impact of the cements will be more felt within the

overbank reservoir portions than in the amalgamated units. This is because of the fact that the

amalgamated reservoirs possessed more porosity. Therefore, hydrocarbon fluids will be able to

navigate around the isolated quartz overgrowths through these pores than in the tighter overbank

reservoirs.

Page 80: Diagenesis and Reservoir Characterization of the

65

CHAPTER 5

CONCLUSIONS

1. Regional and abrupt north-south thickenings of the sedimentary fill of the middle Atoka

Formation were not observed at the study area.

2. The basin dips in a north-south direction. Anticlines and synclines are observed within

the central portion along the east-west direction.

3. The reservoirs include the amalgamated reservoirs and the overbank reservoirs.

4. The amalgamated reservoirs are poor to moderately sorted reservoirs. They contains

much higher sand/shale ratio. The two types of the amalgamated reservoirs include

proximal and medial submarine fans.

5. The overbank reservoirs are moderate to well sorted reservoirs. The ratio of sand/shale

ratio is low. Three classes are recognized - the levee-overbanks, crevasse splays, and the

distal fan complex.

6. The amalgamated reservoir comprises poor to moderately sorted deposits, composed of

lower quartz overgrowths, higher amount of clay cements and dissolution cavities.

7. The overbank reservoir is moderate to well-sorted deposits, composed of higher quartz

overgrowths, lower amount of clay cement and dissolution cavities.

8. Clay cementation increases toward the south and decreases progressively northward.

9. Quartz overgrowths increase toward the north and decrease in the southern direction.

10. The amount of preserved dissolution is higher in the north than in the south due to the

lower amount of clay cements.

Page 81: Diagenesis and Reservoir Characterization of the

66

11. The middle Atoka Formation is composed of very fine to coarse quartzarenites,

subarkoses, sublitharenites, and litharenites. They were transferred and transported into

the basin from the north, east, and the southern portion of the Arkoma Basin.

12. The reservoirs were deposited in the deep-water environments by the submarine fan

systems.

13. The geometry of the reservoir is elongate and radial depending on the stratigraphic units.

14. The Borum and Nichol sandstone units’ reservoirs are elongate and distributed within the

central portion of the study area.

15. The Turner and Basham sandstone units have both elongate and radial geometries. The

distribution of their reservoirs are isolated and unevenly spread across the area.

16. Wells in the south can still be productive. But the higher clay content will make

fracturing of the reservoir much more complicated than it is in the north. Production and

recovery rate will be well below wells in the northern portion.

Page 82: Diagenesis and Reservoir Characterization of the

67

REFERENCES

Ali, S.A., Clark, W.J., Moore, W.R., and Dribus, J.R., 2010. Diagenesis and Reservoir Quality.

Schlumberger Oilfield Review Summer, 22, No. 2.

Allen, J.R.L., 1984. Sedimentary Structures, their Character and Physical Basis, Unabridged

One-volume. Elsevier, Amsterdam, I, p. 593, and II, p. 663.

Allen, J.R.L., 1965. Late Quaternary Niger Delta and Adjacent Areas: sedimentary

Environments and Lithofacies. AAPG Bulletin, 49, p. 547–600.

Allen, J.R.L., 1970. Physical Processes of Sedimentation. American Elsevier Publishing

Company, Inc., New York, p. 239.

Allen, D.R., 1975. Identification of Sediments – Their Depositional Environment and Degree of

Compaction from Well Logs: In Developments in Sedimentology 18a, P. 349-401.

Berry, R.M. and Trumbly, W.D., 1968, Wilburton Gas Field Arkoma Basin, Oklahoma; In

Cline, L.M. (Ed), Guidebook – Geology of Western Arkoma Basin and Ouachita

Mountains, Oklahoma, Oklahoma City Geol. Society, P. 86-103

Bjorlykke, K., & Jahren, J., 2010. Sandstones and Sandstone Reservoirs. In K. Bjorlykke (Ed),

Petroleum Geoscience: From Sedimentary Environments to Rock Physics, P. 113-140.

Bjorlykke, K., & Nagy, J., 2010. Stratigraphy. In K. Bjorlykke (Ed), Petroleum Geoscience:

from Sedimentary Environments to Rock Physics, P. 213-234.

Bjørlykke, K., 1989. Sedimentology and Petroleum Geology. Berlin: Springer-Verlag, p. 363.

Boyd, A., Darling, H., Tabanou, J., Flaum, C., Klein, J., Sneider, R.M., Sibbit, A., and Singer,

J., 1995. The Lowdown on Low-Resistivity Pay. Oilfield Review. Vol. 7, Issue 3.

Bouma, A.H., Hollister, C.D., 1973. Deep Ocean Basin Sedimentation. In: Middleton, G.V.,

Bouma, A.H. (Eds.), Turbidites and Deep-Water Sedimentation, Anaheim, SEPM

Pacific Section Short Course. California, P. 79–118.

Bowen, D.W., Weimer, P., and Scott, A.J., 1993. The Relative Success of Siliciclastic Sequence

Stratigraphic Concepts in Exploration: Examples from Incised-Valley Fill and

Turbidite-Systems Reservoirs: AAPG Memoir No. 58, P. 15-42.

Branan, C.B. Jr., 1968. Natural Gas in Arkoma Basin of Oklahoma and Arkansas: In Bebee,

B.W. (Ed), Natural Gases of North America: A.A.P.G. Memoir No. 4, P. 985-1000.

Briggs, G., and D.H. Roeder, 1975, Sedimentation and Plate Tectonics, Ouachita Mountains and

Arkoma Basin: _In G. Briggs Et Al. (Eds.), Sedimentology of Paleozoic Flysch and

Page 83: Diagenesis and Reservoir Characterization of the

68

Associated Deposits, Ouachita Mountains - Arkoma Basin, Oklahoma: Dallas Geol.

Society, P. 1-22.

Buchanan, R.S, and Johnson, F.K., 1968. Bonanza Gas Field – A Model for Arkoma Basin

Growth Faulting: In Cline, L.M. (Ed), Guidebook – Geology of Western Arkoma

Basin and Ouachita Mountains, Oklahoma: Oklahoma City Geol. Society, P. 75-85.

Burley, S.D., Kantorowicz, J.D., and Waugh, B., 1987. Clastic Diagenesis: In Reservoir II

Sandstones, AAPG Treatise of Petroleum Geology Reprint Series; No. 4, P. 408-445.

Caplan, W.M., 1957. Subsurface Geology of Northwest Arkansas: In Arkansas Geology and

Conservation Comm., Info. Circ., 19, P. 8

Christoph, S., Houseknecht, D.W., and 2 and Jaques, R., 1993. Clay Mineralogy and Illite

Crystallinity of the Atoka Formation, Arkoma Basin, and Frontal Ouachita Mountains:

In Clays and Clay Minerals, Vol. 41, No. 6, 745-754, 1993.

Cronies, C., 1930. Geology of Arkansas Paleozoic Area, with Special Reference to Oil and Gas

Possibilities: In Arkansas Geol. Sur. Bulletin, Vol. 3, P. 457

Devlin, W.J., Rudolph, K.W., Shaw, C.A., and Ehman, K.D., 1993. The Effect of Tectonic and

Eustatic Cycles on Accommodation and Sequence-Stratigraphic Framework in the

Upper Cretaceous Foreland Basin Of Southwestern Wyoming. Spec. Publs Int. Ass.

Sediment, 18, P. 501-520.

Dickinson, W.R., 1974, Plate Tectonics and Sedimentation: In W.R. Dickinson (Ed.), Tectonics

and Sedimentation: Soc. Econ. Paleontologists Mineralogists, Spec. Pub. 22, P. 1-27.

Dickinson, W.R., 1976. Plate tectonic Evolution of sedimentary basins: AAPG Continuing

Education Course Notes series 1, P. 61-64.

Dickinson, W.R., and Suczek, C., 1978, Plate Tectonics and Sandstone Compositions: The

American Association of Petroleum Geologists Bulletin V. 63, No. 12, P. 2164-2182.

Dickinson, W. R., Beard, L. S., Brackenbridge, G. R., Erjavec, J. L., Ferguson, R. C., Inman, K.

F., Knepp, R. A., Lindberg, F. A. and Ryberg, P. T., 1983, Provenance of North

American Phanerozoic Sandstones in relation to Tectonic Setting. Geol. Soc. Amer.

Bull., Vol. 94, P. 222-235.

Dickinson, W.R., 1985. Interpreting Provenance Relation from Detrital Modes of Sandstones:

In Zuffa, G.G. (Ed.), Provenance of Arenites: NATO ASI Series, C 148, D. Reidel

Publishing Company, Dordrecht, 333–363.

Denison, R.E., 1989. Foreland Structure Adjacent to the Ouachita Foldbelt: In Hatcher, R. D.,

Thomas, W.A., and Viele, G.W. Eds. The Geology of North America, Vol. F-2 The

Appalachian-Ouachita Orogen in the United States, Pp. 681-688.

Page 84: Diagenesis and Reservoir Characterization of the

69

Fisher, D.M., and Brantley,S.L., 1992. Models of Quartz Overgrowth and Vein Formation:

Deformation and Episodic Fluid Flow in an Ancient Subduction Zone. Journal of

Geophysical Research, Vol. 97, No. B13, P. 20,043 - 20,061.

Folk, R.L., 1968. Petrology of Sedimentary Rocks: Austin, Hemphill Publishing Co., p. 44-50.

Folk, R.L. 1980. The Petrology of Sedimentary Rocks. Austin, Texas, Hemphill Publishing

Company. P. 182.

Fritsche, G.D., 1981. Subsurface Stratigraphy and Structure Related to Petroleum Occurrences

in the Middle Atoka Formation, Arkoma Basin: Unpublished Master’s Thesis,

University Of Arkansas, Fayetteville, Arkansas.

Galloway, W.E., 1998. Siliciclastic Slope and Base-Of-Slope Depositional Systems:

Component Facies, Stratigraphic Architecture, And Classification. AAPG Bulletin, 82,

P. 569–595.

Galloway, W.E., Hobday, D.K., 1983. Terrigenous Clastic Depositional Systems: Applications

to Petroleum, Coal, And Uranium Exploration, Springer-Verlag, New York, P. 423.

Gaston, T.A., 1985. Atoka Stratigraphy of the Eastern Arkoma Basin: Unpublished Master’s

Thesis, University of Arkansas, Fayetteville, Arkansas.

Gilbreath, L.B., and Haley, B.R., 1982. Restored Thickness of Cecil Series (Informal) in the

Atoka Formation, West Central Arkansas, Oil and Gas Chart 8, Ark. Deol. Comm.

Ginzburg, A., Mooney, W.D., Lutter, W., and Healy, J., 1983, Deep Structure of Northern

Mississippi Embayment: American Association of Petroleum Geologists

Bulletin, 67. 2031-2046.

Haley, B.R., 1961, Thickness Trends in the Hartshorne Sandstone and the Mcalester Formation

in Northwestern Arkansas: U.S. Geol. Surv., Prof. Paper 424C, P. 80-81.

Hansen, S.M., Fett, T., 2000. Identification and Evaluation of Turbidite and other Deepwater

Sands Using Open Hole Logs and Borehole Images. In: Bouma, A.H., Stone, C.G.

(Eds.), Fine-Grained Turbidite Systems, AAPG Memoir 72 And SEPM Special

Publication, 68, P. 317–337.

Harris, G.A., 1983. Sedimentology and Depositional History of a Deltaic Lower Atoka

(Pennsylvanian) Sandstone, Northwestern Arkansas. Unpublished Master’s Thesis,

University Of Arkansas, Fayetteville, Arkansas.

Henbest, L.G., 1953. Morrow Group and Lower Atoka Formation of Arkansas: AM. Assoc. of

Petroleum Geologist Bull. Vol. 37, No. 8, P. 1935-1953.

Page 85: Diagenesis and Reservoir Characterization of the

70

Hendricks, T.S., 1937, Pennsylvanian Sedimentation in Arkansas Coal Field: American Assoc.

Petrol. Geol. Bull., V. 21, P. 1403-1421.

Hendricks, T.A., and B. Parks, 1950, Geology of the Fort Smith District: U.S. Geol. Surv. Prof.

Paper 221-3, p. 67-94.

Houseknecht, D.W., 1984, Influence of Grain Size and Temperature on Intergranular Pressure

Solution, Quartz Cementation, and Porosity in Quartzose Sandstone: Jour. Sed.

Petrology, V. 54, P. 348-361.

Houseknecht, D.W., and L.A. Hathon, 1987a, Relationships among Thermal Maturity,

Sandstone Diagenesis, and Reservoir Quality in Pennsylvanian Strata of the Arkoma

Basin (Abs.): American Assoc. Petrol.

Houseknecht, D.W., 1986, The Atoka Formation of The Arkoma Basin: The Atoka Formation

of The Arkoma Basins: Tectonics, Sedimentology, Thermal Maturity, Sandstone

Petrology. Tulsa Geological Society - Short Course Notes April, 1987.

Houseknecht, D.W., and Kacena, J.A., 1983, Tectonic and Sedimentary Evolution of The

Arkoma Foreland Basin: In D.W. Houseknecht (Ed.), Tectonic-Sedimentary Evolution

of The Arkoma Basin: Soc. Econ. Paleontologists Mineralogists Midcontinent Section,

V. 1, P. 3-33.

Houseknecht, D. W., and Weaverling, P.H., 1983, Early Paleozoic sedimentation in Reelfoot

Rift: Abstract, 12th Annual Meeting of the Eastern Section, American Association of

Petroleum Geologists, Carbondale, I L.

Houseknecht, D.W., and S.M. Matthews, 1985, Thermal Maturity of Carboniferous Strata,

Ouachita Mountains: American Assoc. Petrol. Geol. Bull., V. 69, P. 335-345.

Kim, D., Yang, C.A.I., and Kim, J., 2012. Geological Modeling with Seismic Inversion for

Deepwater Turbidite Fields Offshore Northwestern Myanmar. In Search and

Discovery Article No. 40877.

Krumbein, W.C., Sloss, L.L., 1963. Stratigraphy and Sedimentation, 2nd Edition, W.H.

Freeman and Company, San Francisco, p. 660.

Khidir, A., and Catuneanu, O., 2009. Predictive Diagenetic Clay-Mineral Distribution in

Siliciclastic Rocks as a Tool for Identifying Sequence Boundaries in Non-Marine

Successions: The Coalspur Formation, West-Central Alberta: Geologos, 2009, 15 (3–

4): 169–180.

Kupecz, J.A., 1995. Depositional Setting, Sequence Stratigraphy, Diagenesis, and Reservoir

Potential of a Mixed-Lithology, Upwelling Deposit: Upper Triassic Shublik

Formation, Prudhoe Bay, Alaska. AAPG Bulletin, Volume 79, Issue 9. Pages 1301 –

1319.

Page 86: Diagenesis and Reservoir Characterization of the

71

LeBlanc, R.J., 1987. Geometry of Sandstone Reservoir Bodies: In Clastic Diagenesis: In

Reservoir II Sandstones, AAPG Treatise of Petroleum Geology Reprint Series; No. 4,

P. 191-211.

Leblanc, R.J., 1987. Distribution and Continuity of Sandstone Reservoir – Part I: In Clastic

Diagenesis: In Reservoir II Sandstones, AAPG Treatise of Petroleum Geology Reprint

Series; No. 4, P.212-228.

Lillie. R.J.. Nelson, K.D.. De Voogd, B.. Brewer, J.A., Oliver, J.E.. Brown, L.D.. Kaufman, S.

and Viele, G.W., 1983. Crustal Structure of Ouachita Mountains. Arkansas: A Model

Based on Integration of COCORP Reflection Profiles and Regional Geophysical Data,

Am. Assoc. Pet. Geol. Bull.. 67: 907-931.

Lowe, D.R., 1997. Reinterpretation of Depositional Processes In a Classic Flysch Sequence

(Pennsylvanian Jackfork Group), Ouachita Mountains, Arkansas and Oklahoma:

Discussion. AAPG Bulletin, 81, P. 460–465.

Lumsden, D.N., E.D. Pittman, and R.S. Buchanan, 1971, Sedimentation and Petrology of Spiro

and Foster Sands (Pennsylvanian), Mcalester Basin, Oklahoma: American Assoc.

Petrol. Geol. Bull., V. 55, P. 254-266.

Merewether, E.A., 1961. Thickening of Atoka Formation in the central part of Arkansas Valley:

In U.S Geol. Survey Prof. Paper 424 –C, p. 85-87.

Merewether, E.A., 1972. Geology of Knoxville and Delaware Quadrangle, Jonhson and Logan

Counties and vicinity, Arkansas, U.S.G.S. Prof. Paper 657-B, p.18.

Merewether, E.A., 1967. Geology of Knoxville and Delaware Quadrangle, Jonhson and Pope

Counties, Arkansas, Information Circular 20-E, Ark. Geol. Comm. P.55

Merewether, E.A., and Haley, B.R., 1972. Geology of Delaware Quadrangle, Logan County and

vicinity, Arkansas, Information Circular 20-A, Ark. Geol. Comm. P.39.

Miall, A.D., 1999. In Defense of Facies Classifications and Facies Models. Journal Of

Sedimentary Research, 69, P. 2–5.

Miall, A.D., 2008. The Southern Midcontinent, Permian Basin, and Ouachitas. In: The

Sedimentary Basins of the United States and Canada, A.D. Miall, ed., Elsevier,

Amsterdam, pp. 297-327.

Mickus, K.L., and Keller, G.R., 1992, Lithospheric Structure of the South-Central United

States: Geology, vol. 20, p. 335–338.

Page 87: Diagenesis and Reservoir Characterization of the

72

Mitchum Jr., R.M., Sangree, J.B., Vail, P.R., Wornardt, W.W., 1990. Sequence Stratigraphy in

Late Cenozoic Expanded Sections, Gulf Of Mexico. In: Proceedings of 11th Annual

Research Conference, Gulf Coast SEPM Foundation, 237–256.

Moiola, R.J., Shanmugam, G., 1984. Submarine Fan Sedimentation, Ouachita Mountains,

Arkansas and Oklahoma. Transactions-Gulf Coast Association Of Geologists

Societies, 34, 175–182.

Morris, R.C., 1974. Carboniferous Rocks of the Ouachita Mountains, Arkansas: A Study of

Facies Patterns along the Unstable Slope and Axis of a Flysch Trough. In: Briggs, G.

(Ed.), Carboniferous of the Southeastern United States, Geological Society of America

Special Paper 148, P. 241–279.

Middleton, G., and M. Hampton, 1976, Subaqueous Sediment Transport and Deposition by

Sediment Gravity Flows: In D. Stanley and D. Swift (Eds.), Marine Sediment

Transport and Environmental Management: Wiley, New York, P. 197-218.

Moiola, R.J., and G. Shanmugam, 1984, Submarine Fan Sedimentation, Ouachita Mountains,

Arkansas and Oklahoma: Trans. Gulf Coast Assoc. Geol. Society, V. 34, P. 175-182.

Mollison, R.A., 1983. Petrology and Depositional Environments of The Pitkin Formation

(Chesterian), Washington and Madison Counties, Northwest Arkansas: Unpublished

Master’s Thesis, University of Arkansas, Fayetteville, Arkansas.

Mutti, E., 1977. Distinctive Thin-Bedded Turbidite Facies and Related Depositional

Environments in The Eocene Hecho Group (Southcentral Pyrenees, Spain).

Sedimentology, 24, P. 107–131.

Mutti, E., Normark, W.R., 1987. Comparing Examples of Modern and Ancient Turbidite

Systems: Problems and Concepts. In: Leggett, J.R., Zuffa, G.G. (Eds.), Marine Clastic

Sedimentology: Concepts and Case Studies, Graham and Trotman, London, P.1–37.

Mve, O.P.N., 2005. Petrology, Geochronology and Provenance of the Lainsburg and Tanqua

Karoo Submarine Fan Systems, ECCA Group, South Africa: Unpublished Master’s

Thesis, University Of Stellenbosch, South Africa.

Naidu, A.S., Burrell, D.C., and Hood, D.W., 1971. Clay Mineral Composition and Geologic

Significance of Some Beaufort Sea Sediments. Journal of Sedimentary Petrology,

Vol., 41, No. 3, P. 691-694.

Neethling, J.M., 2009. Stratigraphic Evolution of the Channel-Lobe Transition Zone:

Unpublished Master’s Thesis, University of Stellenbosch, South Africa.

Nicholas, R. L., and D. E. Waddell, 1989, The Ouachita System in the Subsurface of Texas,

Arkansas, and Louisiana, in R. D. Hatcher, Jr., W. A. Thomas, and G. W. Viele, eds.,

Page 88: Diagenesis and Reservoir Characterization of the

73

The Appalachian-Ouachita Orogene in the United States: The Geology of North

America: Boulder, Colorado, Geological Society of America, v. F-2, p. 661–672.

Normark, W.R., 1978. Fan Valleys, Channels, and Depositional Lobes on Modern Submarine

Fans: Characters for Recognition of Sandy Turbidite Environments. AAPG Bulletin,

62, P. 912–931.

Normand, D.L., 1986, Petrography and Diagenetic History of the Alma Sandstone, Upper Atoka

Formation, in the Arkoma Basin of Arkansas, Unpublished Master’s Thesis,

University of Arkansas, Fayetteville, Arkansas.

Ole J. Martinsen, O.J., 2010. Deep-Water Sedimentary Processes and Systems: The Role of

Internal Vs. External Controls on Lithology Distribution and Stratigraphy. Search and

Discovery Article #5026 (2010); Adapted from 2009-2010 AAPG Distinguished

Lecture.

Owoyemi, A.O.D., 2004. The Sequence Stratigraphy of Niger Delta, Delta Field, Offshore

Nigeria: Unpublished Master’s Thesis, Texas A&M University, Texas.

Pettijohn, F.J., 1975. Sedimentary Rocks, 3rd Edition. Harper & Row, Publishers, New York,

p. 628.

Pettijohn, F.J., Potter, P.E., 1964. Atlas and Glossary of Primary Sedimentary Structures.

Springer, Berlin.

Pittman, E.D., 1987. Clastic Diagenesis: In Reservoir II Sandstones, AAPG Treatise of

Petroleum Geology Reprint Series; No.4; P. 447-461.

Pittman, E. D., and R. E. Larese, 1991, Compaction of Lithic Sands: Experimental Results and

Applications: AAPG Bulletin, v. 75, p. 1279–1299.

Pitman, J.K., Goldhaber, M.B., and Spöetl, C., 1997. Regional Diagenetic Patterns in the St.

Peter Sandstone (Evolution of Sedimentary Basins—Illinois Basin; A): Implications

for Brine Migration in the Illinois Basin: U.S. Geological Survey Bulletin 2094-A.

Posamentier, H.W., and James, D.P., 1993. An Overview of Sequence-Stratigraphic Concepts:

Uses and Abuses. Spec. Publs Int. Ass. Sediment, 18, P. 3-18.

Reading, H.G., Collinson, J.D., 1996. Clastic Coasts. In: Reading, H.G. (Ed.), Sedimentary

Environments: Processes, Facies, and Stratigraphy, Blackwell Science, Oxford, P.

154–231.

Reading, H.G., Richards, M., 1994. Turbidite Systems in Deep-Water Basin Margins Classified

by Grain Size and Feeder System. AAPG Bulletin, 78, P. 792–822.

Page 89: Diagenesis and Reservoir Characterization of the

74

Ricci Lucchi, F., 1975. Depositional Cycles in Two Turbidite Formations of Northern

Apennines (Italy). Journal of Sedimentary Petrology, 45, P. 3–43.

Rieke, H.H., 1980. Geologic Overview – Coal and Coalbed Methane Resources of the Arkoma

Basin, Arkansas and Oklahoma: In C.T. Rightmire, G.E. Eddy, and J.N. Kirr, Eds.,

Coalbed Methane Resources of the United States: AAPG Studies in Geology 17, P.

135-161.

Ronck, J.L., 1997, Structural Geology of Thrust Faulting in the Wister Lake Area of the Frontal

Ouachita Mountain, Arkoma Basin, Southeastern Oklahoma: Unpublished Master’s

Thesis, Oklahoma State University, Oklahoma.

Ryu1, C., and Niem, A.R., 1999. Sandstone Diagenesis, Reservoir Potential, and Sequence

Stratigraphy of the Eocene Tyee Basin, Oregon: In Journal of Sedimentary Research,

Vol. 69, No. 2, 1999, P. 384–393.

Scull, B.J., 1961. A comparison of Plio-Mio Sediments of the Gulf Coast with the Atoka

Sediments of the Arkoma Basin, Univ. of Okla., 9. 129-175.

Scruton, P.C., 1950, The Petrography and Environment of Deposition of the Warner, Little

Cabin, and Hartshorne Sandstones in Northeastern Oklahoma: American Jour.

Science, V. 248, P. 408-426.

Shanmugam, G., 2003. Deep-Marine Tidal Bottom Currents and their Reworked Sands in

Modern and Ancient Submarine Canyons. Marine and Petroleum Geology, 20, 471–

491.

Shanmugam, G., 2006. Deep-Water Processes and Facies Models: Implications for Sandstone

Petroleum Reservoirs (Handbook of Petroleum Exploration and Production 5).

Shanmugam, G., Clayton, C.A., 1989. Reservoir Description of a Sand Rich Submarine Fan

Complex for a Steam Flood Project: Upper Miocene Potter Sandstone, North Midway

Sunset Field, California. AAPG Bulletin, 73, 411.

Shanmugam, G., Moiola, R.J., 1988. Submarine Fans: Characteristics, Models, Classification,

and Reservoir Potential. Earth-Science Reviews, 24, 383–428.

Shields, J.P., 1961, Distribution and Evidence of Normal Faulting in the East-Central Portion of

the Arkoma Basin, In Nicholson, Ed., The Arkoma Basin: Proceedings of the Seventh

Biennial Geological Symposium, The University of Oklahoma, School of Geology,

P.31-38.

Slatt, R.M., Forgotson, Jr. J.M., and Saleh, A.A., 2004. Petroleum Geology of the Deepwater

Jackfork Group and Atoka Formation with a Primer on the Petroleum Geology of

Deepwater Depositional Systems: Based On A Workshop Co-Sponsored By PTTC's

Page 90: Diagenesis and Reservoir Characterization of the

75

South Mid-Continent Region PTTC and The School of Geology and Geophysics,

University of Oklahoma on August 18, 2004 in Norman, Oklahoma.

Spotl, C., Houseknecht, D.W., and Longstaffe, F.J., 1994. Authigenic Chlorites in Sandstones as

Indicators of High-Temperature Diagenesis, Arkoma Foreland Basin, USA. Journal of

Sedimentary Research Vol. A64, No. 3, P. 553—566.

Stephens, J.F., 1985. The Alma Sandstones of the Atoka Formation: A Stratigraphic

Investigation in the Arkoma Basin, Arkansas. Unpublished Master’s Thesis,

University Of Arkansas, Fayetteville, Arkansas.

Stow, D.A.V., Piper, D.J.W. (Eds.), 1984. Fine-Grained Sediments: Deep-Water Processes and

Facies. Geological Society Special Publication No. 15, Pp. 659.

Sutherland, P.K., and W.L. Manger (eds.), 1977, Upper Chesterian-Morrowan Stratigraphy and

The Mississippian-Pennsylvanian Boundary In Northeastern Oklahoma and

Northwestern Arkansas: Oklahoma Geol. Surv., Guidebook P. 18, 185.

Sutherland, P.K., and W.L. Manger (eds.), 1979, Mississippian-Pennsylvanian Shelf-to-Basin

Transition, Ozark and Ouachita Regions, Oklahoma and Arkansas: Oklahoma Geol.

Surv., Guidebook 19, 81 P.

Sutherland, P.K., and Manger, W.L., 1984. The Atoka Series (Pennsyvanian) and its Boundarie:

A Symposium. Oklahoma Geological Survey, Bulletin 136.

Swanson, D.C., 1987. Deltaic Deposits in The Pennsylvanian Upper Morrow Formation of The

Anadarko Basin: In Clastic Diagenesis: In Reservoir II Sandstones, AAPG Treatise of

Petroleum Geology Reprint Series; No. 4, Pp. 73-109.

Taylor, T.R., Giles, M.R., Hathon, L.A., Diggs, T.N., Braunsdorf, N.R., Birbiglia, G.V.,

Kittridge, M.G., Macaulay, C.I., and Espejo, I.S., 2010. Sandstone Diagenesis and

Reservoir Quality Prediction: Models, Myths, and Reality. AAPG Bulletin, V. 94, No.

8 Pp. 1093–1132.

Taylor, B.L., 2007. Petrography And Diagenesis of The Eocene Point Of Rocks Sandstone,

Mckittrick Oil Field, Kern County, California: Unpublished Master’s Thesis,

California State University, Bakersfield.

Thomas, D.A., 1983. Middle Atokan Stratigraphy of the Southern Arkoma Basin. Unpublished

Master’s Thesis, University Of Arkansas, Fayetteville, Arkansas.

Triplehorn, D.M., 1979. Clay Mineral Diagenesis in Atoka (Pennsylvanian) Sandstones,

Crawford County, Arkansas.

Page 91: Diagenesis and Reservoir Characterization of the

76

Uszynski, B.J., 1982, Stratigraphic Framework and Depositional Systems: Upper Bloyd and

Lower Atoka (Pennsylvanian) Strata, Arkoma Basin of Central Arkansas: Fayetteville,

University of Arkansas, Unpublished M.S. Thesis.

Vail, P.R., 1987. Seismic-Stratigraphy Interpretation Using Sequence Stratigraphy. Pt. 1:

Seismic Stratigraphy Interpretation Procedures. In: Bally, A.W. (Ed.), Atlas of Seismic

Stratigraphy, AAPG Studies in Geology 27, vol. 1, p. 1–10.

Vail, P.R., Wornardt, W.W., 1990. Well Log-Seismic Sequence Stratigraphy: An Integrated

Tool for the 90’s. In: Sequence Stratigraphy as an Exploration Tool, Concepts and

Practices in the Gulf Coast, 11th Annual Research Conference, Gulf Coast SEPM

Foundation, pp. 379–388.

Vail, P.R., Mitchum Jr., R.M., Thompson III, S., 1977. Seismic Stratigraphy and Global

Changes of Sea Level, Part 4: Global Cycles of Relative Changes of Sea Level. In:

Payton, C.E. (Ed.), Seismic Stratigraphy – Applications to Hydrocarbon Exploration,

AAPG Memoir 26, p. 83–97.

Viele, G.W., 1973, Structure and Tectonic History of the Ouachita Mountains, Arkansas: In

K.A. Dejong and R. Scholten (Eds.), Gravity and Tectonics: Wiley, New York, P. 361-

377.

Viele, G.W., 1979. Geological Map and Cross Section, eastern Ouachita Mountains, Geological

Society of America Map and Chart Series MC-28F; Map Summary: Geological

Society of America Bulletin, Vol. 90, P. 1096-1099.

Viele, G.W., and Thomas, W.A., 1989. Tectonic Synthensis of the Ouachita Orogenic Belt: in

Hatcher, R.D., Jr., Thomas, W.A., and Viele, G.W., eds., The Appalachian-Ouachita

Orogen in the United States: Boulder, Colorado, Geological Society of America, The

Geology of North America, Vol. F-2, P. 695-728.

Visher, G.S., S. Saitta B., and R.S. Phares, 1971, Pennsylvanian Delta Patterns and Petroleum

Occurrences in Eastern Oklahoma: American Assoc. Petrol. Geol. Bull., v. 55, p.

1206-1230.

Walderhaug, O., Bjørkum, P.A., and Aase, N.E., 2006. Kaolin-Coating of Stylolites, Effect on

Quartz Cementation and General Implications for Dissolution at Mineral Interfaces.

Journal of Sedimentary Research, vol. 76, p. 234-243.

Walker, R.G., 1979, Shallow marine sands: in R.G. Walker (ed.), Facies Models: Geoscience

Canada Reprint Ser. 1, P. 75-89.

Walper, J.L., 1977, Paleozoic Tectonics of the Southern Margin of North America: Gulf Coast

Assoc. Geol. Socs. Trans., vol. 27, p. 230-241.

Page 92: Diagenesis and Reservoir Characterization of the

77

Waugh, B., 1970. Petrology, Provenance and Silica Diagenesis Of The Penrith Sandstone

(Lower Permian) of Northwest England. Journal of Sedimentary Petrology, Vol: 40,

Issue: 4, p. 1226-1240.

Wentworth CK. 1922. A Scale of Grade and Class Terms for Clastic Sediments. Journal of

Geology 30: 377–392.

Williams, P.S., 1983. Lithostratigraphy and Depositional Systems of the Middle Atoka

Formation, Arkoma Basin: Unpublished Master’s Thesis, University of Arkansas,

Fayetteville, Arkansas.

Wilson, M.D., and Pittman, E.D., 1987. Clastic Diagenesis: in Reservoir II Sandstones, AAPG

Treatise of Petroleum Geology Reprint Series; No. 4; p. 463-487.

Wright, W.R., 2011. Depositional History of the Atokan Succession (Lower Pennsylvanian) in

the Permian Basin: Bureau of Economic Geology Jackson School of Geosciences, The

University of Texas at Austin, Texas.

Zachry, D.L., 1979, Early Pennsylvanian Braided Stream Sedimentation, Northwest Arkansas:

in N.J. Hyne (ed.), Pennsylvanian Sandstones of the Mid-Continent: Tulsa Geol, Soc,

Spec. Pub. 1, p. 269-281.

Zachry, D.L., 1983, Sedimentologic Framework of the Atoka Formation, Arkoma Basin,

Arkansas: _in_ D.W. Houseknecht (ed.), Tectonic-Sedimentary Evolution of the

Arkoma Basin: Soc. Econ. Paleontologists Mineralogists Mid-continent Section, v. 1,

p. 34-52.

Zachry, D.L., and Haley, B.R., 1975. Stratigraphic Relations between the Bloyd and Atoka

Formations (Pennsylvanian) Of Northern Arkansas: Contribution to Geology of the

Arkansas, Ozarks, P. 96-105.

Zachry, D.L., and Sutherland, P.K., 1984. Stratigraphy and Depositional Framework of the

Atoka Formation (Pennsylvanian), Arkoma Basin of Arkansas and Oklahoma: In

Sutherland, P.K., and Manger, W.L. (eds.), The Atoka Series (Pennsylvanian) and its

Boundaries – a Symposium: Oklahoma Geological Survey Bulletin 1336, P. 9-17.

Zachry, D.L., and Pierre-Orly, D., 2008. Stratigraphic and Sedimentologic Analysis of the

"Middle Bloyd" Sandstone (Morrowan) in Outcrops, Mount Judea and Lurton

Quadrangles, Newton County, Northern Arkansas. Unpublished Master’s Thesis,

University of Arkansas, Fayetteville, Arkansas.

Page 93: Diagenesis and Reservoir Characterization of the

78

APPENDICES

Page 94: Diagenesis and Reservoir Characterization of the

79

0 2 4 6 8

Schult-Teeter 3-18H

Holcomb 1-36

Hale 5-17

Carter 2-26

State 2-22

Sanderson 3-9H

Secondary

Primary

Porosity

Figure 35 showing the percentage abundance of the primary and secondary porosities from

modal analysis of the samples obtained from the six wells in the area. Porosity is generally

higher in the north as shown in the bar chart. Carter 2-26 well has the highest values among the

vertical wells. The values obtained from the Sanderson 3-39H may be biased due to the poor

nature of the thin sections.

0 20 40 60 80 100

Scultz-Teeter 3-18H

Holcomb 1-36

Sanderson 3-9H

State 2-22

Carter 2-26

Hale 5-19

Lithic fragments

Feldspar

Quartz

Framework Grains

Figure 36 showing the composition of the framework grains of the middle Atoka Formation from

the six wells. Quartz has the highest composition.

Page 95: Diagenesis and Reservoir Characterization of the

80

0 20 40 60 80 100

Sanderson 3-9H

State 2-22

Carter 2-26

Hale 5-19

Holcomb 1-36

Scultz-Teeter 3-18H

Dolomite

Calcite cements

Feldspar overgrowths

Quartz overgrowths

Clay cements

Cementing materials

Figure 37 showing percentage abundance of the component of cementing materials. Carter 2-26

and Schultz-Teeter 3-18H are both in the northern portion of the area. Clay cements and quartz

overgrowths are generally higher in the south and north respectively (Figure 35).

0 20 40 60 80 100 120

Sanderson 3-9H

State 2-22

Carter 2-26

Hale 5-19

Holcomb 1-36

Scultz-Teeter 3-18H

Biotite

Muscovite

Acessoryminerals

Figure 38 showing the distribution of the accessory minerals with muscovite averaging about

99% of the accessory minerals. Biotite occurs in trace amount.

Page 96: Diagenesis and Reservoir Characterization of the

81

0 20 40 60 80

Monocystalline

Polycrystalline

Felspar

Lithic Fragments

Clay Cement

Quarts Overgrowths

Felspar Overgrowths

Calcite cement

Dolomite

Muscovite

Bitoite

Others

Primary

Secondary

8010

7890

7770

7620

6288

Depth (ft.)

Figure 39 showing percentage abundance of the thin sections modal analysis of the samples

obtained from the various depths of the Sanderson 3-9H well.

Page 97: Diagenesis and Reservoir Characterization of the

82

0 10 20 30 40 50 60 70

Monocystalline

Polycrystalline

Felspar

Lithic Fragments

Clay Cement

Quarts Overgrowths

Felspar Overgrowths

Calcite cement

Dolomite

Muscovite

Biotite

Others

Primary

Secondary

7130

7040

6110

5630

5990

5570

Depth (ft.)

Figure 40 showing the percentage abundance of the thin sections modal analysis of the thin

sections obtained from the various depths of the State 2-22 well.

Page 98: Diagenesis and Reservoir Characterization of the

83

0 10 20 30 40 50 60

Monocystalline

Polycrystalline

Felspar

Lithic Fragments

Clay Cement

Quarts Overgrowths

Felspar Overgrowths

Calcite cement

Muscovite

Biotite

Others

Primary

Secondary

4080

3910

3870

3830

3800

3680

3670

3660

3640

3630

3620

Depth (ft.)

Figure 41 showing the distribution of the thin sections modal analysis obtained from the various

depths of the Carter 2-26 well.

Page 99: Diagenesis and Reservoir Characterization of the

84

0 20 40 60 80

Monocystalline

Polycrystalline

Felspar

Lithic Fragments

Clay Cement

Quarts Overgrowths

Felspar Overgrowths

Calcite cement

Dolomite

Muscovite

Biotite

Others

Primary

Secondary

8110

7080

6900

6970

8350

7180

7780

7330

7570

9134

6750

6950

Depth (ft.)

Figure 42 showing the distribution of the thin sections modal analysis obtained from the various

depths of the Hale 5-19 well.

Page 100: Diagenesis and Reservoir Characterization of the

85

0 10 20 30 40 50 60 70

Monocystalline

Polycrystalline

Felspar

Lithic Fragments

Clay Cement

Quarts Overgrowths

Felspar Overgrowths

Calcite cement

Dolomite

Muscovite

Biotite

Others

Primary

Secondary

7609

7588

7573

7518

7509

7075

7027

7018

Depth (ft.)

Figure 43 showing the percentage abundance of the thin sections modal analysis obtained from

the various depths of the Holcomb 1-36 well.

Page 101: Diagenesis and Reservoir Characterization of the

86

0 20 40 60 80

Monocystalline

Polycrystalline

Felspar

Lithic Fragments

Clay Cement

Quarts Overgrowths

Felspar Overgrowths

Calcite cement

Dolomite

Muscovite

Biotite

Others

Primary

Secondary

7470

7080

6000

Depth (ft.)

Figure 44 showing the percentage abundance of the thin sections modal analysis obtained from

the various depths of the Schultz-Teeter 3-18H well.

Page 102: Diagenesis and Reservoir Characterization of the

87

Depth (ft): 70808 6000 7470

Cuttings Condition (mm): Very Poor Good Poor

Grain Size Average (mm): 0.14 0.23 0.32

Grain Size Range (mm): 0.1-0.18 0.03-0.42 0.03-0.60

Sorting: Well Sorted Moderately Moderately

Rock Name (Folk): Subarkose Quartzarenite Subarkose

Framework Grains

Quartz 56 59 71

Monocystalline 53 52 69

Polycrystalline 3 7 2

Felspar 8 1 2

Lithic Fragments 6 7 5

Cementing Materials 11

Clay Cement 10 5 5

Quarts Overgrowths 4 5 6

Felspar Overgrowths 0 0 0

Calcite cement 0 0 0

Dolomite 0 0 0

Accessory Minerals 4

Muscovite 4 2 4

Biotite 0 0 0

Others 6 12 0

6 9 7

Porosity 6 7 4

Primary 0 2 3

Secondary

TOTAL (%): 100 100 100

Thin Section Analysis

Schultz teeters 3-18H

Sabestian/Logan Counties

Table 3. Petrographic Data of Samples of Schultz-Teeters 3-18H

Page 103: Diagenesis and Reservoir Characterization of the

88

Depth (ft): 7609 7588 7573 7518 7509 7075 7027 7018

Cuttings Condition (mm): Good Good Good Good Good Good Good Good

Grain Size Average (mm): 0.26 0.13 0.17 0.16 0.21 0.2 0.21 0.1

Grain Size Range (mm): 0.04-0.52 0.04-0.24 0.08-0.32 0.08-0.26 0.08-0.32 0.04-0.36 0.04-0.52 0.068-0.14

Rock Name (Folk): Subarkose

Framework Grains

Quartz 57 61 57 53 61 50 50 49

Monocystalline 52 59 55 48 61 50 48 49

Polycrystalline 5 2 2 5 0 0 2 0

Felspar 7 3 4 4 5 3 5 4

Lithic Fragments 3 3 1 2 1 1 1 1

Cementing materials 13 19 27 26 9 26 25 33

Clay Cement 11 9 15 18 5 11 8 29

Quarts Overgrowths 2 11 12 6 4 14 15 4

Felspar Overgrowths 0 0 0 1 0 1 2 0

Calcite cement 0 0 0 1 0 0 0 0

Dolomite 0 0 0 0 0 0 0 0

Accessory minerals

Muscovite 2 1 1 2 4 3 3 9

Biotite 0 0 0 0 0 0 0 0

Others 9 1 1 3 2 9 12 2

Porosity 9 11 9 10 18 8 4 2

Primary 3 3 3 3 2 1 1 2

Secondary 6 8 6 7 16 7 3 0

TOTAL (%): 100 100 100 100 100 100 100 100

Quartzarenites Subarkoses

Thin Section Analysis

Holcombe 1-36

Sabestian/Logan Counties

Table 4. Petrographic Analysis of Thin-Sections obtained from Holcomb 1-39 well

Page 104: Diagenesis and Reservoir Characterization of the

89

Depth (ft): 6950 6750 9134 7570 7330

Cuttings Condition (mm): Poor Poor Very Poor Very Poor Very Poor

Grain Size Average (mm): 0.07 0.21 0.17 0.16 0.09

Grain Size Range (mm): 0.036-0.12 0.12-0.32 0.08-0.0.26 0.08-0.24 0.08-0.16

Pore Throat Average (mm): 0 0.016 0.012 0 0.004

Rock Name (Folk): Sublithics Subarkose Sublithics Q/arenite Sublithics

Framework Grains (%):

Quartz 70 50 54 44 52

Monocystalline 56 40 49 33 44

Polycrystalline 14 10 5 11 8

Felspar 1 7 5 1 1

Lithic Fragments 13 3 5 1 8

Cementing Materials 6 19 14 35 19

Clay Cement 6 3 13 24 15

Quarts Overgrowths 0 16 1 11 4

Felspar Overgrowths 0 0 0 0 0

Calcite cement 0 0 0 0 0

Dolomite 0 0 0 0 0

Accessory Minerals

Muscovite 0 3 1 0 0

Biotite 0 0 0 0 0

Others 10 1 7 2 2

Porosity 0 17 14 17 18

TOTAL (%): 100 100 100 100 100

Thin Section Analysis

Hale 5-17H

Sabestian/Logan Counties

Page 105: Diagenesis and Reservoir Characterization of the

90

Table 5. Petrographic Analysis of Thin-Sections obtained from Hale 5-17H Well

Depth (ft): 7780 7180 8350 6970 6900 7080 8110

Cuttings Condition (mm): Very Poor Very Poor Very Poor Very Poor Good Poor Poor

Grain Size Average (mm): 0.1 0.16 0.1 0.15 0.12 0.16 0.08

Grain Size Range (mm): 0.08-0.12 0.008-0.24 0.08-0.2 0.08-0.22 0.05-0.55 0.03-0.44 0.02-0.062

Pore Throat Average (mm): 0.11 0.01 0.01 0.012

Rock Name (Folk): Q/arenite

Framework Grains (%):

Quartz 64 64 37 54 70 66 66

Monocystalline 64 54 32 46 68 60 66

Polycrystalline 0 10 5 8 2 6 0

Felspar 5 1 5 1 2 2 1

Lithic Fragments 1 8 9 1 3 2 2

Cementing Materials 23 9 21 20 15 17 5

Clay Cement 23 9 16 20 5 6 5

Quarts Overgrowths 0 0 5 6 10 11 6

Felspar Overgrowths 0 0 0 0 0 0 0

Calcite cement 0 0 0 0 0 0 0

Dolomite 0 0 0 0 0 0 0

Accessory Minerals

Muscovite 0 0 0 0 0 2 0

Biotite 0 0 0 0 0 0 0

Others 2 3 0 0 1 2 16

Porosity 5 15 28 18 9 9 4

TOTAL (%): 100 100 100 100 100 100 100

Thin Section Analysis

Hale 5-17H continues…

Sabestian/Logan Counties

Sublithics Quartzarenites

Table 6. Petrographic Analysis of Thin-Sections obtained from Hale 5-17H well

Page 106: Diagenesis and Reservoir Characterization of the

91

Depth (ft): 3620 3640 3660 3670 3680 3800

Cuttings Condition (mm): Poor Poor Poor Poor Poor Poor

Grain Size Average (mm): 0.17 0.18 0.28 0.14 0.13 0.2

Grain Size Range (mm): 0.08-0.28 0.13-0.22 0.06-0.6 0.06-0.2 0.04-0.20 0.08-0.32

Pore Throat Average (mm): 0.001 0.001 0.012 0.01 0.008 0.01

Rock Name (Folk): Q/arenites Sublitharenites

Framework Grains

Quartz 48 54 59 53 66 49

Monocystalline 48 54 44 52 31 23

Polycrystalline 0 0 15 1 35 26

Felspar 1 0 0 1 1 2

Lithic Fragments 8 8 11 15 7 14

Cementing Materials 19 25 19 12 9 12

Clay Cement 4 11 15 10 8 6

Quarts Overgrowths 15 14 4 2 1 6

Felspar Overgrowths 0 0 0 0 0 0

Calcite cement 0 0 0 0 0 0

Accessory Minerals

Muscovite 6 2 4 0 3 6

Biotite 0 0 0 0 0 0

Others 2 6 3 9 6 3

Porosity 16 5 4 10 8 14

TOTAL (%): 100 100 100 100 100 100

Thin Section Analysis

Carter 2-26

Sabestian/Logan Counties

Sublitharenites

Table 7. Petrographic Analysis of Thin-Sections obtained from Carter 2-26 well

Page 107: Diagenesis and Reservoir Characterization of the

92

Depth (ft): 3830 3870 4080 3630 3910

Cuttings Condition (mm): Good Very Poor Poor Good Poor

Grain Size Average (mm): 0.12 0.16 0.21 0.33 0.3

Grain Size Range (mm): 0.04-0.24 0.008-0.32 0.04-0.42 0.03-0.82 0.02-0.73

Pore Throat Average (mm): 0.006 0.006 0.01 0.002 0.05

Rock Name (Folk):

Framework Grains

Quartz 59 61 63 60 59

Monocystalline 31 39 43 50 52

Polycrystalline 28 22 20 10 7

Felspar 2 0 4 0 1

Lithic Fragments 14 12 6 6 7

Cementing Materials 10 9 19 22 20

Clay Cement 2 3 15 6 5

Quarts Overgrowths 8 9 4 16 15

Felspar Overgrowths 0 0 0 0 0

Calcite cement 0 0 0 0 0

Accessory Minerals

Muscovite 0 0 0 3 1

Biotite 0 0 0 0 0

Others 1 1 1 1 3

Porosity 14 13 7 8 9

TOTAL (%): 100 100 100 100 100

Thin Section Analysis

Carter 2-26 continues…

Sabestian/Logan Counties

Table 8. Petrographic Analysis of Thin-Sections obtained from Carter 2-26 well

Page 108: Diagenesis and Reservoir Characterization of the

93

Depth (ft): 5990 5630 7040 7130 5570 6110

Cuttings Condition (mm): Very Poor Very Poor Poor Very Poor Poor Poor

Grain Size Average (mm): 0.22 0.16 0.16 0.23 0.19 0.14

Grain Size Range (mm): 0.1-0.39 0.08-0.24 0.12-0.21 0.12-0.44 0.01-0.48 0.02-0.55

Pore Throat Average (mm): 0.016 0.016 0.012 0.015

Rock Name (Folk): Q/arenite

Framework Grains

Quartz 60 55 64 65 63 66

Monocystalline 60 55 64 65 60 65

Polycrystalline 0 0 0 0 3 1

Felspar 8 5 1 0 0 1

Lithic Fragments 11 15 9 20 5 4

Cementing materials 12 19 6 4 12 13

Clay Cement 6 19 6 4 7 8

Quarts Overgrowths 3 0 0 0 5 5

Felspar Overgrowths 0 0 0 0 0 0

Calcite cement 3 0 0 0 0 0

Dolomite 0 0 0 0 0 0

Accessory minerals

Muscovite 0 0 0 0 2 3

Biotite 0 0 0 0 0 0

Others 1 1 2 1 15 9

Porosity 8 5 18 10 3 4

TOTAL (%): 100 100 100 100 100 100

Thin Section Analysis

State 2-22H

Sabestian/Logan Counties

Sublitharenites

Table 9. Petrographic Data of Samples of State 2-22H Well

Page 109: Diagenesis and Reservoir Characterization of the

94

Depth (ft): 8010 7770 7620 6288 7890

Cuttings Condition (mm): Very Poor Very Poor Very Poor Poor Poor

Grain Size Average (mm): 0.24 0.01 0.19 0.22 0.19

Grain Size Range (mm): 0.2-0.28 0.008-0.012 0.14-0.24 0.03-0.67 0.02-0.81

Rock Name (Folk):

Framework Grains

Quartz 50 67 68 68 51

Monocystalline 50 67 66 66 50

Polycrystalline 0 0 0 2 1

Felspar 4 1 1 1 1

Lithic Fragments 4 1 1 1 2

Cement 18 22 27 17 13

Clay Cement 9 15 8 8 6

Quarts Overgrowths 9 7 11 9 5

Felspar Overgrowths 0 0 0 0 0

Calcite cement 0 0 0 0 0

Dolomite 0 0 8 0 2

Accessory minerals

Muscovite 4 0 0 1 5

Bitoite 0 0 0 0 0

Others 20 6 1 10 21

Porosity 0 3 4 2 7

TOTAL (%): 100 100 100 100 100

Thin Section Analysis

Sanderson 3-9H

Sabestian/Logan Counties

Subarkose Quartzarenites

Table 10. Petrographic Data of Samples of Sanderson 3-9H Well

Page 110: Diagenesis and Reservoir Characterization of the

95

Wells Basham Nichol Turner Borum

TANNER 'B' 57 163 61 0

Ft Chaffee 171 131 200 359

GIPSON 0 0 0 0

EVELYN "G" 41 196 245 318

LOWE 69 65 73 0

TATUM 147 118 45 41

James 86 127 53 0

HOWARD 163 98 102 41

FERREL 171 135 86 20

WILD HOG 114 69 90 0

Anderson R 102 102 69 0

Sanatorium 212 37 24 0

Gilker James 65 90 49 0

GASAWAY 106 65 82 0

ROBERTS JACK 65 53 41 0

Wooten 53 16 12 0

JONES 155 37 241 0

LEWIS 216 135 180 0

TEXAS JACK 208 147 151 0

Wooten EJ 135 69 86 0

BUCCELLA 69 143 49 0

HARRELL 73 241 0 0

AINSWORTH 94 78 270 20

HANCOX 155 114 0 0

BARNES 86 102 220 167

YADON 0 90 139 0

HERMIE 131 139 167 0

HOPEWELL 151 253 41 33

PAUL X 286 180 90 147

GIBSON PEARL 106 233 147 274

LOWE 65 245 24 261

TYLER 110 65 269 41

TURNER 110 216 0 0

WHITNEY 225 200 98 294

GREGORY 98 294 61 0

HALE 135 143 135 57

STATE 139 106 131 0

HATTABAUGH1-33 163 196 241 0

HATTABAGAUH 143 114 49 94

HOLCOMB 192 131 98 0

CARTER 188 94 24 0

STATE 122 159 159 0

HALE 151 269 61 0

Sterling 163 94 139 41

Cate B 114 110 33 102

Gross Sands

Table 11. Gross sands of the four Stratigraphic Units

Page 111: Diagenesis and Reservoir Characterization of the

96

Quartz (%) Felspar (%) Lithic Fragments (%) Depth (ft)

80 11 9 70808

81 11 8 6000

91 2 7 7470

85 10 5 7609

91 5 4 7588

93 1 6 7573

89 7 4 7518

90 7 3 7509

92 5 3 7075

88 9 3 7027

89 8 3 7018

84 8 8 8010

98 1 1 7770

98 1 1 7620

97 1 2 6288

95 2 3 7890

75 10 15 5990

73 7 20 5630

86 1 13 7040

72 0 28 7130

75 0 25 5570

93 0 7 6110

83 4 13 3620

86 0 14 3640

84 0 16 3660

79 2 19 3670

98 1 1 3680

75 4 21 3800

78 2 20 3830

83 0 17 3870

86 6 8 4080

90 0 10 3630

88 1 11 3910

84 1 15 6950

83 11 6 6750

82 9 9 9134

97 1 2 7570

84 2 14 7330

91 7 2 7780

87 1 12 7180

72 9 19 8350

97 1 2 6970

93 2 5 6900

95 2 3 7080

96 1 3 8110

87 4 9

Framework Grains

Sanderson 3-9H

State 2-22H

Carter 2-26

Wells

Hale 5-17H

Total %

Holcombe 1-36

Schultz-Teeters 3-18H

Table 12: Percentages of the framework grain components for all the depths

Page 112: Diagenesis and Reservoir Characterization of the

97