You are on page 1of 103

ATTRIBUTE SUPPORTED SEISMIC

GEOMORPHOLOGY

AND RESERVOIR CHARACTERIZATION OF THE

GRANITE WASH, ANADARKO BASIN, TEXAS

By

GABRIEL EDUARDO GAVIDIA

Bachelor of Science in Geology

Arkansas Tech University

Russellville, Arkansas

2010

Submitted to the Faculty of the


Graduate College of the
Oklahoma State University
in partial fulfillment of
the requirements for
for the Degree of
MASTER OF SCIENCE
July, 2012
ATTRIBUTE SUPPORTED SEISMIC

GEOMORPHOLOGY

AND RESERVOIR CHARACTERIZATION OF THE

GRANITE WASH, ANADARKO BASIN, TEXAS

Thesis Approved:

Dr. Estella Atekwana

Thesis Adviser

Dr. Jamie Rich

Dr. James Puckette

Dr. Sheryl A. Tucker

Dean of the Graduate College


.

ii
TABLE OF CONTENTS

Page
TABLE OF CONTENTS ............................................................................................. iii
LIST OF FIGURES ..................................................................................................... iv

Chapter

I. Introduction ................................................................................................................1
Objectives ................................................................................................................3
Location ...................................................................................................................4
Data ..........................................................................................................................5
Previous Work .........................................................................................................7

II. TECTONIC AND STRATIGRAPHIC FRAMEWORK ..........................................9

III. SEISMIC INTERPRETATION .............................................................................17

IV. PRE-INVERSION ANALYSIS ............................................................................24


Cross plot analysis Cherokee Wash .......................................................................27

V. SEISMIC INVERSION ..........................................................................................37


Model Based Inversion ..........................................................................................38
Methodology ..........................................................................................................42
Quality check .........................................................................................................56
Inversion results .....................................................................................................58

VI. MULTI ATTRIBUTE DISPLAY AND RESERVOIR


CHARACTERIZATION .............................................................................................68
Similarity................................................................................................................68
Geobody multi attribute display and seismic geomorphology .............................71
Porosity and lithology prediction ...........................................................................75

VII. DISCUSSION ......................................................................................................80


VIII.CONCLUSION ....................................................................................................92
REFERENCES 94

iii
LIST OF FIGURES

Figure Page
1.1 Location……………………………………………………………………………….4
1.2 Processing sequence…………………………………………………………………...5
1.3 3D seismic survey area………………………………………………………………..6
2.1 Anadarko basin including study area………………………………………………...11
2.2 Simplified cross section, Anadarko basin……………………………………………12
2.3 Anadarko basin generalized stratigraphic column…………………………………...13
2.4 Structural elements of west-central USA…………………………………………….14
2.5 Granite Wash isopach………………………………………………………………..16
3.1 3D seismic survey area, including inversion wells…………………………………..17
3.2 Inline 183, including interpreted horizons…………………………………………...18
3.3 Cross line 186 including interpreted horizons……………………………………….20
3.4 Granite Wash G time structure map………………………………………………….21
3.5 Granite Wash Caldwell time structure map………………………………………….22
3.6 Granite Wash time thickness map……………………………………………………23
4.1 3D seismic survey, including inversion: wells………………………………………25
4.2 AI and gamma-ray logs from inversion: wells, Granite Wash interval……………...26
4.3 AI and gamma-ray logs from inversion: wells Cherokee Wash……………………..28
4.4 Cross plots from physical properties: Well D………………………………………..30
4.5 Cross plots from physical properties: Well A………………………………………..31
4.6 Cross plots from physical properties: Well B………………………………………..32
4.7 Cross plots from physical properties: Well C………………………………………..33
4.8 Combined cross plot…………………………………………………………………35
4.9 Combined cross plot sandstones……………………………………………………..36
5.1 Wavelets and frequency spectrum…………………………………………………...42
5.2 Normal incidence synthetic seismogram: Well A postack migrated volume………..43
5.3 Normal incidence synthetic seismogram: Well B postack migrated volume………..44
5.4 Normal incidence synthetic seismogram: Well C postack migrated volume………..45
5.5 Normal incidence synthetic seismogram: Well D postack migrated volume………..46
5.6 Normal incidence synthetic seismogram: Well A prestack migrated volume……….47
5.7 Normal incidence synthetic seismogram: Well B prestack migrated volume……….47
5.8 Normal incidence synthetic seismogram: Well C prestack migrated volume……….48
5.9 Normal incidence synthetic seismogram: Well D prestack migrated volume……….48
5.10 Inverted trace at well location: Well A postack migrated volume………………….50
5.11 Inverted trace at well location: Well B postack migrated volume………………….51
5.12 Inverted trace at well location: Well C postack migrated volume………………….51
5.13 Inverted trace at well location: Well A postack migrated volume………………….52

iv
5.14 Inverted trace at well location: Well A prestack migrated volume………………...53
5.15 Inverted trace at well location: Well B postack migrated volume………………….54
5.16 Inverted trace at well location: Well C postack migrated volume………………….54
5.17 Inverted trace at well location: Well D postack migrated volume………………….55
5.18 Blind well test location……………………………………………………………..56
5.19 Blind well test plot………………………………………………………………….57
5.20 Cross line 113 within the postack migrated inverted AI volume…………………...59
5.21 Cross line 113 within the prestack migrated inverted AI volume………………….60
5.22 AI attribute surface extracted from the postack migrated inverted volume………...62
5.23 AI attribute surface extracted from the prestack migrated inverted volume……….63
5.24 Comparative figure between AI attribute surfaces…………………………………64
5.25 AI attribute surfaces amplitude cross plots…………………………………………66
5.26 Comparative figure between inverted AI traces frequency content………………...67
6.1 Similarity geobody…………………………………………………………………...70
6.2 Co-rendered similarity and AI geobody……………………………………………..73
6.3 Co-rendered similarity and AI geobody including interpreted depositional
environments…………………………………………………………………………74
6.4 Interpreted arbitrary line including the gamma-ray and
porosity logs from wells D, F, G, H, and I…………………………………………...77
6.5 Co-rendered AI and similarity geobody including arbitrary line…………………….78
6.6 Co-rendered AI and similarity geobody including completed horizontal wells……..79
7.1 Interpreted line across AI attribute surface…………………………………………..81
7.2 Comparative figure between Cherokee time thickness map and AI surface………...83
7.3 Coherent-energy amplitude gradient surface………………………………………...85
7.4 Co-rendered coherent-energy amplitude gradient and AI surface…………………...86
7.5 Co-rendered coherent-energy amplitude gradient and AI surface interpreted……….87
7.6 Co-rendered coherent-energy amplitude gradient and AI surface interpreted
including well control……………………………………………………………….89
7.7 Cherokee Wash depositional model………………………………………………….91

v
INTRODUCTION

Submarine fans constitute major hydrocarbon reservoirs in various parts of the

world. Consequently, a clear understanding of their geometry and facies relationships is

critical for exploring and exploiting these deposits effectively. Historically,

Pennsylvanian alluvial fan deposits (Granite Wash) from the Texas Panhandle have been

produced in numerous fields including Burns Flat, Elk City, and Allison Ranch, but were

considered uneconomic in many other areas. Due to recent developments in horizontal

drilling and hydraulic fracturing, the Granite Wash has become an important target for oil

and natural gas exploration. Geologically, several studies including Dutton (1984),

Dutton and Land (1986), Parks (2011) and, Bouma (2000) generally agree in dividing the

fan complex into an upper, or proximal, fan composed of gravels and conglomeratic

facies, a middle fan composed of coarse clastic material, and the distal fan with finer

clastic materials that prograde into the basin to merge with shales of deeper marine

origin.

However, geophysically the Granite Wash remains a difficult unit for seismic

interpretation, specifically within areas proximal to the source. Due to the rapid change in

lithofacies, thickness, and discontinuity of beds, the identification of specific alluvial fan

depositional environments and reservoir facies from seismic data is not well documented.

From a geophysical perspective, there is no published research in this area, and all

inquiry is likely highly proprietary to the oil and gas industry.

1
This project will examine the petrophysical and seismic expression of the Granite

Wash through the combination of seismic attributes and acoustic impedance (AI)

inversion. The project will extend the scientific knowledge of the Granite Wash

geomorphology and specific depositional environments in order to improve success in oil

and gas exploration and exploitation.

2
General Objectives

The objective of this study is to assess the advantage of 3D seismic attributes and

poststack seismic inversion to define the distribution of porous and permeable alluvial fan

facies and depositional environments within the Granite Wash of the Texas Panhandle.

The following specific objectives are proposed to systematically approach the application

of 3D seismic attributes to identify the distribution of porous and permeable rock.

Specific Objectives

1. Determine if seismic inversion for acoustic impedance can be used to differentiate

principal lithofacies in the Granite Wash.

2. Compare the results of inverting prestack migrated data versus postack migrated

data.

3. Use inversion for acoustic impedance to define a porosity range within the

sandstone and conglomeratic facies in the Granite Wash.

4. Use seismic attributes such as similarity in combination with AI to evaluate the

reservoir geomorphology and depositional environments.

3
Location

The study area is located in Wheeler and Hemphill counties, Texas. The specific

study area is not identified to respect the request from the data donor who has scientific

and economic interest in the study.

Figure 1.1 Locations of Wheeler and Hemphill Counties in the Texas Panhandle.

4
Data

The seismic survey is composed of two 3D stacked volumes (prestack migrated

and postack migrated) and is located in Wheeler and Hemphill counties, Texas. The

survey was acquired and processed by a third party in 2011. The survey area is 28 square

miles and roughly square in shape. The implemented processing sequence preserves the

original reflection amplitudes of the seismic to be input to the inversion; and 3D spherical

spreading corrections and surface-consistent gain corrections are applied. The data are

shifted to a datum of 3000 feet elevation to facilitate velocity analysis and migration. Pre-

stack Kirchhoff migration and postack Kirchhoff migration was applied to the data to

collapse the diffractions and properly position the events (Figure 1.2).

Figure 1.2. Processing sequence for the data.

5
Four wells were selected for the inversion, their relative spatial locations are

shown on the map as the letters A, B, C and, D (Figure 1.3). These wells were chosen

because P-wave sonic and density logs were available. The wells were corrected for

erroneous readings and borehole effects and they all tie to the seismic. An additional 53

wells within the survey were used to corroborate the results.

Figure 1.3. 3D seismic survey area, including the location of wells used for inversion.

6
Previous Work

Submarine alluvial fan systems are known for being hydrocarbon reservoirs.

Pennsylvanian alluvial fans in the Anadarko Basin have been known to produce in fields

such as Stiles Ranch, Britt Ranch, and Elmore field. However, before recent

developments in horizontal drilling and hydraulic fracturing, extensive areas of the

Granite Wash were considered noneconomic.

One of the first studies, Edwards (1958), discussed the facies of Pennsylvanian

rocks along the north flank of the Wichita Mountains. Edwards stated that Granite Wash

clastic material sourced from the mountains formed a range of rocks including

conglomerates, sandstones, shales, and detrital limestone. According to Edwards (1958),

the Granite Wash and finer clastic sediments were deposited as a series of coalescing

alluvial fans and deltas. The Mobeetie field in Wheeler County was of early interest to

the industry and Sahl (1970) studied the Granite Wash at the Mobeetie Field, establishing

important age correlations, lithofacies, and porosity. This was followed by Dutton (1982)

who examined the stratigraphy and sedimentology of the Granite Wash and Dutton

(1984), which detailed depositional environments for the Granite Wash, including: fan

plain distal fan, main channels, and prodelta. Other important records on the Granite

Wash include Land (1985) who studied the diagenesis and paragenesis in Pennsylvanian

rocks, including the Granite Wash and Carroll (1986) who concluded that the sediment

pulses coming out from the mountains creating fans were attributed to individual

7
drainage basin erosion cycles. Carroll (1986) also concluded that the best reservoirs occur

in the braided channels and proximal deltas.

Thompson (2007) assessed the reservoir properties and distribution of the

Cherokee Granite Wash of Roberts and Hutchison Counties, Texas and Parks (2011)

studied the depositional setting of the “Upper Granite Wash” (Marmaton and Cabaniss)

in the Buffalo Wallow Field, Hemphill County, Texas.

Papers that discussed the application of seismic techniques in the Granite Wash include

Burnett (2002) who noted the difficulty in characterizing the Granite Wash due to the

lack of seismic resolution, and Valerio (2006), who applied spectral decomposition and

spectral inversion on the Pennsylvanian Granite Wash of Texas, and was successful in

extracting greater detail at the high end of the frequency spectrum. Martinis (2010)

performed a series of footprint suppression algorithms in order to improve the signal to

noise ratio in a 3D volume from the early 1990’s. Martinis (2010) was able to use

discreet seismic attributes and seismic inversion for acoustic impedance to better

characterize the Granite Wash of Hartley County, Texas.

8
CHAPTER II

TECTONIC AND STRATIGRAPHIC FRAMEWORK

The Anadarko Basin is one of the deepest basins within the continental United

States, and contains in excess of 40,000 feet of Paleozoic sedimentary rock. The basin is

located in western Oklahoma and in the northeast corner of the Texas Panhandle (Figure

2.1). It is bordered by the Nemaha Ridge on the east, the ancient Amarillo-Wichita

Mountains on the south, the Hugoton embayment on the west, and the transition to the

central Kansas uplift on the north (Hill and Clark, 1980).

The axis of the basin is 12-37 miles north of the Wichita uplift; up to 40,000 feet

of upper Cambrian-Permian strata are preserved in portions of the basin (Ham et al.,

1964; Ham and Wilson, 1967; Ham, 1969; Adler, 1971). An additional 20,000 feet of

layered upper Proterozoic-Middle Cambrian metasedimentary, plutonic, and volcanic

rocks are estimated to underlie the Paleozoic sedimentary rocks along the southern edge

of the basin (Ham et al., 1964; Ham and Wilson, 1967; Adler, 1971). Siliciclastic rocks

dominate the upper part of the Paleozoic section in the deep Anadarko basin, whereas

carbonates are more common in the lower part of the section (Figure 2.2). The

stratigraphic section on the northern shelf is relatively thin (4,900-9,800 feet) and is

composed primarily of lower Paleozoic carbonates (Carter, 1998).

9
According to Johnson et al. (1988), southern Oklahoma has been structurally

quiescent since the Permian, except for slight eastward epeirogenic tilting due to the

Laramide orogeny. Permian strata are currently being eroded from most of the basin

surface, but Mesozoic and Cenozoic units probably covered the area. Triassic and

Jurassic red beds are found in the subsurface in the Oklahoma and Texas panhandles and

may have extended eastward into western Oklahoma before late Jurassic and Cenozoic

erosion (Cunningham, 1961). Scattered Cretaceous outliers in western Oklahoma indicate

that Cretaceous strata originally were deposited over the region (Adler, 1971; Johnson et

al., 1988). Late Miocene to Pliocene fluvial deposits of the Ogallala Formation are up to

900 feet thick and cover most of the western one-third of the basin (Cunningham, 1961).

Figure 2.3 represents the stratigraphic chart of the Pennsylvanian and Permian strata in

the Anadarko Basin.

10
Figure 2.1. Major tectonic features of Oklahoma and part of Texas Panhandle, with the
approximate position of the study area shown.

11
Figure 2.2. Simplified cross section of the Anadarko basin perpendicular to the basin axis
(Johnson, 1989)

12
Figure 2.3. Generalized column showing informal stratigraphic nomenclature employed by the
oil and gas industry to classify the Pennsylvanian strata for the Anadarko Basin the Granite Wash
interval is outlined by the red box (After Johnson and Cardott, 1992)

During the middle Mississippian and Pennsylvanian, the Anadarko Basin

underwent rapid subsidence as the adjacent Amarillo-Wichita uplift was elevated. Uplift

continued through the early Permian and, during this time, thick wedges of coarse-

grained, terriginous clastic deposits accumulated along each of the fault-bounded uplifts.

During the Pennsylvanian Period, the west-central United States contained several

similar fault-bounded uplifts and adjacent basins (Figure 2.3). The Amarillo-Wichita,

Sierra Grande, Apishapa, Uncompahgre, Frontrange, and Pedernal Uplifts were all active

during Pennsylvanian time (Dutton, 1984)

13
Figure 2.4. Structural elements of west-central USA during middle Pennsylvanian
(Desmoinesian) (Dutton, 1984)

Uplift associated coarse-grained rocks have been described as fan-delta deposits.

Fan-delta deposits are recognized in the subsurface as coarse-grained conglomeratic

clastics that were deposited adjacent to an elevated source area and that interfinger with

deeper marine sediments (Dutton, 1984). In general, fan deltas which prograde onto a

shelf environment show a complete upward-coarsening sequence (Brown, 1979; Wescott

and Ethridge, 1980). In a typical sequence, prodelta shale and siltstone are overlain by

distal-fan sandstone and shale; these deposits are in turn superimposed by the braided-

channel-fill sandstones and conglomerates as the fan progrades.

14
According to Dutton (1984), the Amarillo-Wichita uplift was the primary source

area for the Granite Wash and supplied sediment typically composed of nonresistant

grains, such as feldspar and rock fragments, which survive because of short transport

distances. The basement rock exposed in the Amarillo-Wichita uplift belongs to the

Wichita igneous province. Rock types in this province are predominantly Granite (56%),

diabase (14%), and granodiorite (12%) (Dutton, 1984). The composition of the source

terrain is reflected in the sediments which eroded from the uplifts; Granite and

granodiorite fragments, perthite, quartz, and plagioclase are the most common detrital

constituents of the Granite Wash (Dutton, 1984). The regional extent and total thickness

of the Granite Wash in Atokan (Lower Pennsylvanian) through Middle Pennsylvanian

(Desmoinesian) strata are shown by a Granite Wash thickness map (Figure 2.5)

15
Figure 2.5. Thickness of Granite Wash in parts of the Texas Panhandle (after Dutton,
1984)

16
CHAPTER III
SEISMIC INTERPRETATION

Nine main horizons were interpreted using seismic data including: The Granite

Wash G, The Granite Wash F, The Granite Wash E, The Granite Wash D, The Granite

Wash C, The Granite Wash B, The Granite Wash A, The Granite Wash Cherokee, and

The Granite Wash Caldwell (informal industry names) all within the Middle

Pennsylvanian (Desmoinesian). Proper well to seismic ties were made from wells A, B,

C, and D (Figure 3.1) to ensure that the interpretation was based on solid seismic to

stratigraphy ties.

Figure 3.1. Map showing location of the 3D seismic survey including the wells that tie the
seismic volume to known stratigraphy.

17
Figures 3.2 and 3.3 are seismic lines that display the Granite Wash section with

the corresponding interpreted informal stratigraphic horizons recognized by the oil and

gas industry in the area of the study.

Figure 3.2. Nine interpreted horizons within the Granite Wash section, Wheeler and Hemphill
counties, Texas.

18
Figure 3.3 is a cross line from east to west showing the nine interpreted horizons

within the Granite Wash section. Figure 3.3 also illustrates the structural and stratigraphic

complexity of the area with a major fault that displays high angle reverse movement. This

fault trends from north to south to form the well-known anticline of the Buffalo Wallow

field.

The hanging wall of the fault in Figure 3.3 is to the west and the footwall to the

east. Thinning older Washes on the up side are evidence of syntectonic deposition.

The time structure maps in figures 3.4, 3.5 and 3.6 represent the bottom of the

Granite Wash, the top, and the thickness from the bottom to the top of the Granite Wash.

Figure 3.4 represents the time structure map from the Granite Wash G. Pictured as

a red line is the high angle reverse fault that trends from north to south. The hanging wall

of the fault is to the west of the survey, while the footwall is to the east. The red arrow

pointing to the northeast represents the direction in which the sediment was being

deposited as it was eroded and transported from the southwest. Figure 3.5 represents the

time structure map for the top of the Granite Wash.

Figure 3.6 represents a time thickness map for the whole Desmoinesian Granite

Wash. This thickness map demonstrates how the sediment was deposited as alluvial fans

with the thicker proximal intervals close to the source (southwest) and thinner toward the

distal facies (northeast). Also the reverse fault to the west demonstrates its influence in

19
how the sediment was being deposited or accommodated in the basin. One can also see

the syntectonic nature of the Washes as they were being deposited; this process is

demonstrated as the Washes become thinner along the up dip direction (west and north).

Figure 3.3. Eleven interpreted horizons, the entire Granite Wash interval and, a major fault. Up
dip thinning of the units toward the west (W) is interpreted as evidence of syndepositional
tectonics.

20
Figure 3.4. Time structure map from Granite Wash G. The red line represents the main fault
within the survey and the red arrow represents the paleo-direction the sediment followed as it was
being transported from the southwest to the northeast.

21
Figure 3.5. Time structure map for the top of the Granite Wash, which is known in local informal
nomenclature as the Caldwell Wash.

22
Figure 3.6. Time-thickness map for the Granite Wash interval. The thickest area lies to the south
of the survey. The sediment accommodation in this area is concordant with the accumulation of
typical alluvial fan wedges that become thinner as the fan prograded toward deeper parts of the
basin. The red arrow represents the direction of progradation.

23
CHAPTER IV
PRE-INVERSION ANALYSIS

Prior to inversion, well logs must be examined for suitable relationships between

measured impedance (density times P-wave velocity) and other desirable properties, such

as porosity, gamma-ray, and neutron porosity. Cross plots between these physical

properties should produce clear relationships between acoustic impedance, porosity, and

lithology. Also, the geology of the area should be considered; in this case high-porosity

arkosic reservoirs may be confused for radioactive shale if only the gamma-ray is

considered to differentiate lithology.

Completing a pre-inversion analysis (well log cross plots) will prevent incorrect

assumptions concerning the results the inversion may yield. This process should be

completed in all areas being considered for an inversion, in order to determine what the

inversion may yield from a rock property standpoint (Latimer and Davison, 2000).

It is also important to analyze the section above and below any major geologic

change in order to correctly understand the relationship between the impedance log and

any other rock properties. A cross plot derived from the entire length of the log suite

would lead you to believe that inverting the data for acoustic impedance will be fruitless

(Latimer and Davison 2000). In this study each of the Washes was analyzed for suitable

rock properties by comparing rock properties using cross plots.

24
Within the seismic survey area, four wells (A, B, C and, D) were used for the

inversion procedure and thereby analyzed prior to inversion (Figure 4.1).

Figure 4.1. 3D seismic area, including the location of wells used for inversion.

An important aspect of the wells that were utilized for inversion is that they all

fulfill specific conditions regarding borehole logging. The wells include curves such as P-

wave sonic, bulk density, neutron porosity, resistivity, and caliper. The logs have also

been edited to account for erroneous readings and borehole effects, and they all tie to the

3D seismic volume. Figure 4.2 is a display of gamma-ray and calculated acoustic

impedance across the Granite Wash interval. Notice how zones of low acoustic

25
impedance correspond with areas of high gamma-ray readings. This empirical

relationship already allows the interpreter to differentiate lithology based on acoustic

impedance.

Figure 4.2. Inversion wells, including gamma-ray log to the left and calculated acoustic
impedance log (scale from right to left) to the right. Notice how areas with high gamma-ray
readings (black) correspond with areas of low acoustic impedance (AI).

26
Cross Plot Analysis Granite Wash Cherokee

Several Granite Wash intervals were analyzed for suitable relationships between

AI and porosity, including the Cherokee, A, B, C, D, E, F, and G Washes (informal

industry names). Of these intervals, the Cherokee Granite Wash stands out as having the

best petrophysical response and also because of the availability of completed horizontal

wells within this unit to validate the results.

Figure 4.3 displays sections of the Cherokee Wash including the gamma-ray log

curve, calculated AI, and the density porosity curve. Figure 4.4 is a tridimensional cross

plot from Well D with calculated AI on the x-axis, density porosity along the y-axis and a

gamma-ray color filter which discriminates sandstones for having less than 75 API units

and shales for having more than 75 API units. Because of the geology in this area, which

includes sandstones of arkosic composition, porous and permeable units could be

overlooked due to their high radioactivity. Therefore, a 75 API (gamma-ray log) unit

mark was chosen to identify sandstone as opposed to the conventional 70 API units to

compensate for the abundance of radioactive feldspathic minerals within the reservoir

rock. Along with the cross plot, the Granite Wash Cherokee is displayed to the left with

the respective AI and porosity logs and a color table which discriminates sandstones

(yellow) and shales (blue).

27
Figure 4.4 represents the attribute cross plot from Well D; this cross plot allows

the differentiation of lithology within the interval. Most sandstones have AI values of

more than 39,400 [(g/cc)(ft/s)], where as most of the shales have AI values of less than

this threshold value. The cross plot also shows that sandstones with lower AI tend to have

higher porosity and that sandstones with lower porosity tend to have much higher AI

values.

Figure 4.3. Section of the Cherokee Wash unit including gamma-ray log to the left calculated AI
in the middle and density porosity to the right.

Figure 4.5 is a cross plot from Well A with the same parameters used for Well D.

In Well A the sandstone interval is thinner, the porosity is lower and, the clay content is

higher. As in Well D, sandstones with higher porosity have lower acoustic impedance

28
than those with low porosity. The red line (Figure 4.5) differentiates sandstone from

shale.

Figure 4.6 represents the cross plot from Well B using the same parameters

described previously. Well B has a different petrophysical response than the previous

two wells. In this area, the clay content of the Wash is much higher and the porous

interval is thinner (less than 10 ft), and the sandstone units are also more radioactive due

to the presence of feldspars. Because of the high radioactivity of the Cherokee Wash in

this area, a neutron porosity color bar was used to discriminate sandstone from shale.

Neutron porosity values higher than 14% are caused by shale effect, which in this case

allows the differentiation of shale from sandstone.

Figure 4.7 displays the attribute cross plot from well C. In this case the scattering

of the data is higher, but the overall data distribution follows the same trends as were

evident in plots of wells D and A.

29
Figure 4.4 AI vs density porosity cross plot from Well D.

30
Figure 4.5 AI vs density porosity cross plot from Well A.

31
Figure 4.6 AI vs density porosity cross plot from Well B.

32
Figure 4.7 AI vs density porosity cross plot from Well C.

33
Figure 4.8 represents the attribute cross plot for the previous four wells combined.

The relationships previously established are sustained in this figure; sandstones (red oval)

of higher porosity correspond with lower AI values. The cross plot also displays a

substantial overlapping of shale (black points) within the sandstone oval. This

overlapping is attributed to the high radioactivity caused by the abundance of feldspar in

the reservoir sandstones. Figure 4.9 represents the same combined cross plot from Figure

4.8, but in this case only the sandstones are displayed. Although the gamma-ray color bar

indicates that some of the points in the scatter correspond to shale (black), a neutron

porosity filter that deleted all values with neutron porosity higher than 14% (neutron

porosity shale effect) allows to display the sandstone reservoir including the sandstones

rich in feldspathic minerals.

Figure 4.9 shows that AI inversion can be used in the Granite Wash to

differentiate lithology and identify areas of higher porosity within the Granite Wash.

34
Figure 4.8 AI vs density porosity cross plot from Wells A, B, C, and, D. Note the overlapping of the data caused by the high
radioactivity of the reservoir rock

35
Figure 4.9 AI vs density porosity cross plot from Wells A, B, C, and, D. The data has been filtered by using the neutron’s porosity
shale effect to only display the sandstones.

36
CHAPTER V
SEISMIC INVERSION

Acoustic impedance (AI) is the product of rock density and P-wave velocity. This

means that AI is a rock property and not an interface property (e.g., seismic reflection

data). Acoustic impedance inversion is simply the transformation of seismic data into

pseudo-acoustic-impedance logs at every trace (Latimer and Davison, 2000).

Inverted seismic data contains all information that is present in the seismic data

without the complicating factors caused by wavelets, as well as the essential information

extracted from the log data. The AI volume is the result of the integration of data from

several different sources, usually seismic and, well log data (Latimer and Davison, 2000).

AI is closely related to lithology, porosity, pore fill, and other factors. It is

common to find strong empirical relations between AI and one or more rock properties.

In essence, a high-quality AI inversion provides more information than the original

seismic data (Latimer, 2000). During this research project, two data volumes with

different processing sequences from the same survey were inverted. The purpose of

inverting these two volumes was to assess the differences between inverted prestack

migrated data versus inverted postack migrated data.

37
MODEL-BASED INVERSION

The model-based inversion algorithm was chosen upon different inversion

techniques because it provided the closest approximation to the real AI log after

comparing the inverted trace with the original AI log.

In model-based inversion, geological information from logs is employed parallel

to the seismic data. This reduces the inherent none-uniqueness of seismic inversion.

Additionally, integration of well data guarantees the contribution of missing low

frequency content of seismic data into the final impedance model (Shamsa and Lines

2010).

Model-based inversion starts with a low frequency model of the P-impedance

(Hampson-Russell Software Strata Inversion Theory, 2012). This low frequency model is

perturbed until a good fit is obtained between the seismic data and the synthetic trace

computed by applying equations (1) and (2) where r represents the zero-offset reflectivity

of the earth, W the seismic wavelet, s the seismic trace, Z the acoustic impedance and (*)

the convolution operator.

1.

2. .

38
Equation (1) is rewritten as the following matrix equation:

3. [ ] [ ][ ]

In this equation, si represents the ith sample of the seismic trace and Wj represents the jth

term of an extracted seismic wavelet. This can be written in the shorter form (4)

(Hampson-Russell Software Strata Inversion Theory, 2012).

4.

Here, s is the seismic trace vector, W is the wavelet matrix, and r is the

reflectivity vector of equation (3). Note that the convolution operator is no longer needed

since convolution becomes multiplication when matrix operators are used. Equation (2)

can be linearized by noting that for small reflection coefficients (in the order of 0.1 or

less) we can write:

5.

Where and after taking the derivative of the

logarithm, the impedance can be written as (Hampson-Russell Software Strata Inversion

Theory, 2012).

)) )
6. )

39
By removing the dt term in equation (6) to leave only the differential, and using Δ instead

of d, we can combine equations (5) and (6) to give:

7. ( ) [ ( ) )]

Equation (7) can be written in matrix form as:

8. [ ] [ ]
[ ]

Where Lpi=ln(Zpi), and the matrix consisting of all 1’s and -1’s can be thought of as a

derivative matrix. This can be written in the shorter form

9.

where R is the reflectivity vector, D is the derivative matrix, and L is the log impedance

vector of equation (8).

Combining equations (4) and (9) gives us the forward model which relates the seismic

trace to the logarithm of the impedance:

10.

Model-based inversion uses the low frequency P-impedance model (described above),

generated from well data and horizons and the known seismic trace and wavelet plus

40
the random noise. This method solves for the reflectivity iteratively, looking for

differences between the real seismic trace and the synthetic formed from the model and

modifying the model to compensate (Hampson-Russell Software Strata Inversion Theory,

2012).

The approach is to minimize this function:

11. ) )

where s is the seismic trace, W is the extracted wavelet, r is the final reflectivity, M the

initial guess model impedance, H the integration operator which convolves with the final

reflectivity to produce the final impedance (Hampson-Russell Software Strata Inversion

Theory, 2012).

41
Methodology

The first step of the inversion process was to select the wells to be used. Four

wells were selected for inversion: A, B, C, and, D. These wells were chosen because they

all tie to the seismic volumes and also because of the high quality of the sonic and density

logs. Figures 5.2 through 5.9 represent the well to seismic ties from Wells A, B, C, and D

to the postack migrated and prestack migrated data volumes. In order to ensure the

quality of the well to seismic ties, two statistical wavelets corresponding to each of the

volumes were extracted using a search window (1000-2400 ms) that avoided shallow

seismic effects and focused on the target area. Figure 5.1 compares the statistical

wavelets used for the well to seismic ties along with the frequency spectrum from each of

the seismic volumes.

Figure 5.1. Statistical wavelets and frequency spectrum corresponding to the postack migrated
data and prestack migrated data volumes.

42
Figure 5.2. Normal incidence synthetic seismogram from Well A. Synthetic seismogram (blue
trace). Postack migrated volume.

43
Figure 5.3. Normal incidence synthetic seismogram from Well B. Synthetic seismogram (blue
trace). Postack migrated volume.

44
Figure 5.4. Normal incidence synthetic seismogram from Well C. Synthetic seismogram (blue
trace). Postack migrated volume.

45
Figure 5.5. Normal incidence synthetic seismogram from Well D. Synthetic seismogram (blue
trace). Postack migrated volume.

After the proper time to depth relationship between the wells and the postack migrated

volume was established, the same procedure was applied to the prestack migrated data.

Figures 5.6 through 5.9 represent the well to seismic ties from Wells A, B, C and, D for

the prestack migrated volume.

46
Figure 5.6. Normal incidence synthetic seismogram from Well A. Synthetic seismogram (blue
trace). Prestack migrated volume.

Figure 5.7. Normal incidence synthetic seismogram from Well B. Synthetic seismogram (blue
trace). Prestack migrated volume.

47
Figure 5.8. Normal incidence synthetic seismogram from Well C. Synthetic seismogram (blue
trace). Prestack migrated volume.

Figure 5.9. Normal incidence synthetic seismogram from Well D. Synthetic seismogram (blue
trace). Prestack migrated volume.

48
Following the well to seismic ties, and due to differences in phase between the

prestack migrated volume and the postack migrated volume, two sets of seismic horizons

were interpreted correspondingly. These horizons correspond to the following units

within the Granite Wash: Granite Wash Caldwell, Granite Wash Cherokee, GRWA,

GRAB, GRWC, GRWD, GRWE, GRWF and, GRWG (informal industry names).

After the two sets of horizons were interpreted and imported into the inversion

software the next step was to create an initial guess model for impedance. The initial

guess model was created by using a high frequency cut of 20-25Hz, the computed AI logs

from the wells, and the interpreted horizons. The interpreted horizons acted as a variable

check-shot at every trace allowing the impedances between two picked events to be

stretched and squeezed accordingly with the interpreted horizons.

After the initial guess model was created, the only step left to do was to run the

inversion and this was performed in two steps. The first step was to perform the inversion

at the well location to confirm the inversion parameters and to allow the program to

perform the optimum scaling. Figures 5.10 through 5.13 represent the inversion analysis

at the well location. The inversion analysis window (figures 5.10 – 5.17) shows a number

of curves, including the real log which has been filtered with a bandpass of 60 – 75Hz

(blue), the initial guess model (black), and the inverted trace (red). As one can appreciate,

the inverted trace closely correlates to the original log in all of the wells and the error

between the synthetic traces and the original seismic is low. The analysis window

49
confirms that the inversion is good enough to proceed. Finally the volume was inverted

using a time interval between 1500ms and 2400ms.

Figure 5.10. Inversion analysis window for the postack migrated data volume at the Well A
location showing the frequency filtered original AI log (blue), the initial guess low frequency
model (black), and the inverted trace (red). Also, note the small error between the synthetic trace
and the real seismic traces.

50
Figure 5.11. Inversion analysis window for the postack migrated data volume at Well B location
showing the frequency filtered original AI log (blue), the initial guess low frequency model
(black), and the inverted trace (red).

Figure 5.12. Inversion analysis window for the postack migrated data volume at Well C location
showing the frequency filtered original AI log (blue), the initial guess low frequency model
(black), and the inverted trace (red).

51
Figure 5.13. Inversion analysis window for the postack migrated data volume at Well D location
showing the frequency filtered original AI log (blue), the initial guess low frequency model
(black), and the inverted trace (red).Notice how the inverted trace correlates the original log.

In the same manner as explained above, the same procedure was applied to the

prestack migrated data. Figures 5.14 through 5.17 show the inverted trace at the well

location for the prestack migrated data volume.

52
Figure 5.14. Inversion analysis window for the prestack migrated data volume at the Well A
location showing the frequency filtered original AI log (blue), the initial guess low frequency
model (black), and the inverted trace (red).

53
Figure 5.15. Inversion analysis window for the prestack migrated data volume at the Well B
location showing the frequency filtered original AI log (blue), the initial guess low frequency
model (black), and the inverted trace (red).

Figure 5.16. Inversion analysis window for the prestack migrated data volume at the Well C
location showing the frequency filtered original AI log (blue), the initial guess low frequency
model (black), and the inverted trace (red).

54
Figure 5.17. Inversion analysis window for the prestack migrated data volume at the well D
location showing the frequency filtered original AI log (blue), the initial guess low frequency
model (black), and the inverted trace (red).

55
Quality Check

In order to ensure the quality of the final inversion product, a blind well test was

performed over the two inverted volumes. In this procedure, one of the wells previously

used in the inversion (Well D) is removed from the process. The inversion was then

performed using Wells A, B, and C, obtaining two inverted volumes from the three wells.

Finally an inverted AI trace is extracted from the well path of Well D and plotted against

its original AI log. Figure 5.18 show the location of the blind well.

Figure 5.18. Location of the blind well used to test the postack and prestack inversion.

Figure 5.19 is a plot comparing the original log (blue) and the inverted trace (red) from

the prestack migrated data and the postack migrated data. The original log was filtered

56
using a band pass frequency of 60 – 75Hz to approximate the frequencies from the

seismic data. Figure 5.19 allows the interpreter to compare how close the inverted trace

and the original log correlate; in this case, the resemblance between both is strong and

allows confidence in the process.

Figure 5.19. Blind well test plot from Well D. Blue line corresponds to the original log and red
line represents the inverted trace.

57
Inversion Results

Figures 5.20 and 5.21 represent the postack migrated inverted volume and the

prestack migrated inverted volume accordingly. Both of the inverted volumes allow the

interpreter to distinguish the different lithofacies within the Granite Wash. The gamma-

ray log from Well A corroborates the results from the inversion. Because the postack

migrated inverted volume has lower frequency content (figure 5.1), the results show

better lateral coherency. This lower frequency content allows easier interpretation and

characterization of the continuity of the different lithofacies when using an interpretation

window.

58
Figure 5.20. Cross line 113 within the postack migrated inverted impedance volume showing the Desmoinesian portion of the
Granite Wash with its corresponding interpreted horizons. Notice the lateral continuity of the interpreted horizons when compared
with the prestack migrated volume.

59
Figure 5.21. Cross line 113 within the prestack migrated inverted volume showing the Desmoinesian portion of the Granite Wash
with its corresponding interpreted horizons. Notice the decrease of lateral continuity when compared with the postack migrated
inverted volume.

60
Figures 5.22 and 5.23 show two attribute surfaces that were extracted using a 15-

millisecond search window below the top of the Cherokee Wash. Both of the attribute

surfaces show a series of sinusoidal features that were interpreted as channels. As

mentioned before, the attribute surface extracted from the postack migrated inverted

volume shows higher lateral coherency than the same surface extracted from the prestack

migrated inverted volume. However, although the attribute surface extracted from the

postack migrated inverted volume has higher lateral coherency due to the lower

frequency content, the attribute surface extracted from the prestack migrated inverted

data shows better vertical and lateral resolution.

Figure 5.24 is an enlarged image of the area highlighted in figure 5.23. The

surface extracted from the prestack migrated inverted volume shows higher vertical and

lateral resolution. As seen in the arrows in figure 5.24, the edges between areas of

contrasting impedance are improved due to the higher frequency content from the

prestack migrated data. This allows the interpreter to define areas with higher spatial

accuracy within the target interval.

61
Figure 5.22. AI surface extracted from the postack migrated inverted volume using a 15ms search window below the Cherokee
Wash top; the features inside the box represent areas that show higher lateral coherency.

62
Figure 5.23. AI surface extracted from the prestack migrated inverted volume by using a 15ms search window below the
Cherokee Wash top. The area inside the box shows higher vertical resolution when compared with the surface extracted from the
postack migrated inverted volume.

63
Figure 5.24. Comparative figure between the prestack and postack migrated inverted volumes. Notice how the higher frequency
content from the prestack inverted data increases the lateral and vertical resolution of the surface.

64
Figure 5.25 allows further comparison of the slight differences between the

prestack and postack migrated inverted volumes. In this cross plot, AI amplitude values

extracted from the postack migrated inverted AI surface are plotted versus those from the

prestack migrated inverted AI surface. Notice the somewhat linear relationship between

the two extracted AI surfaces and how close the equation of the line comes to being one.

Figure 5.26 demonstrates that the amplitude differences between the two surfaces are

small and therefore, the minor differences between both inverted volumes can be

attributed to the higher frequency content found in the prestack migrated volume. This

higher frequency content increases the temporal resolution of the extracted attribute

surface, thereby improving the detection of smaller depositional changes within the fan

complex. Moreover, temporal and spatial resolutions are not independent of each other.

Improving one automatically increases the other, as a result, the reduction in the Fresnel

zone due to the higher frequency allows the detection of changes in reflectivity that occur

over a shorter distance. This reduction in the Fresnel zone due to the higher frequency

content in the prestack migrated volume explains the improved spatial resolution in the

AI surface extracted from the prestack migrated inverted volume.

65
Figure 5.25. Extracted pseudo-acoustic impedance logs (traces) from prestack and postack migrated inverted volumes at well D
(right). Cross plot corresponding to the Cherokee Wash between postack migrated inverted AI surface and prestack migrated
inverted AI surface.

66
Figure 5.26 compares two extracted inverted AI traces from Well D. Notice how the green trace

(postack migrated) appears smoother than the red trace (prestack migrated) due to its lower

frequency content.

Figure 5.26. A comparison of prestack and postack migrated inverted AI traces extracted from
the path of Well D. Note the higher frequency content in the red curve (prestack migrated).

In addition to having higher frequency content, parameters regarding the prestack

Kirchhoff migration algorithm, such as the space and time aperture (migration radius) of

the migration operator and a better velocity model, could have contributed to improving

the resolution of the prestack migrated inverted volume.

67
CHAPTER VI
MULTI ATTRIBUTE DISPLAY AND RESERVOIR CHARACTERIZATION

SIMILARITY

Similarity is a seismic attribute that measures the coherency between waveforms

or traces (Chopra and Marfurt, 2008). When seen on a processed section, the seismic

waveform is a response of the seismic wavelet convolved with the geology of the

subsurface. That response changes in terms of amplitude, frequency, and phase,

depending on the acoustic impedance contrast and thickness of the layers above and

below the reflecting boundary. In turn, acoustic impedance is affected by lithology,

porosity, and density of the subsurface layers. Consequently, the seismic waveforms that

we see in processed sections differ in lateral character – that is, strong lateral changes in

AI contrast give rise to strong lateral changes in waveform character. Geologically,

highly coherent seismic traces indicate laterally continuous lithologies. Abrupt changes in

waveform can indicate changes such as faults, fractures, and depositional features

(Chopra and Marfurt, 2008).

Figure 6.1 represents a similarity attribute surface extracted using the same 15

millisecond window below the Cherokee Wash top. The similarity surface allows the

interpretation of areas with higher seismic coherency (white) and areas of lower seismic

coherency (black) among the seismic traces. Geologically, higher energy depositional

environments are recognized for having chaotic depositional patterns and stratification.

68
On the other hand, passive depositional environments are known for being well

stratified and continuous. Geophysically, similarity allows the interpreter to identify areas

of higher and lower seismic coherency. In turn, high energy depositional environments

like those expected in the proximal and middle facies of a submarine alluvial fan will

reflect a seismic signal with less lateral coherency. In an opposite manner, the seismic

signal reflected from a quiet depositional environment will have higher continuity and

coherency between the seismic traces.

69
Figure 6.1. Image of the similarity geobody that represents the Cherokee Wash. Areas with high similarity (white)
among seismic traces represent areas with continuous lithology, and areas with low similarity (dark) indicate abrupt
changes in lithology. The red arrows correspond to the interpreted channels

70
GEOBODY MULTI ATTRIBUTE DISPLAY
SEISMIC GEOMORPHOLOGY

The combination of different attributes can greatly facilitate and improve the

interpretation of different depositional environments and lithofacies. In this case, inverted

AI allows the interpreter to differentiate specific qualities from the rock such as clay

content and porosity, while, similarity allows the identification of specific changes in

depositional patterns that affect the coherency of the seismic signal.

In order to combine the two attributes, a three dimensional geobody was created.

This geobody represents a time interval (isochron) between the top of the Cherokee Wash

and a phantom horizon created 15ms below the Cherokee Wash. This 15ms thickness

body was created in order to focus the attribute window over the cleaner sandstones of

the upper Cherokee Wash. Figure 6.2 represents the Cherokee Wash geobody with the

co-rendered AI from the prestack migrated volume and similarity.

In figure 6.3 the combination of acoustic impedance and similarity allows the

interpreter to identify two parts of a submarine fan complex. The first part, in the south

and southeast portion of the study area, corresponds to the upper fan depositional

environment composed of near-source-braided stream channels that acted as sediment

feeders for the fan complex. The second part, located near the middle, is interpreted as

the middle fan depositional environment. In this area, as sediment encountered a lower

71
gradient on the fan surface, the feeder channels branched into smaller channel

systems to form the characteristic alluvial fan shape.

72
Figure 6.2. Co-rendered AI and similarity geobody. Note how the areas of low similarity (black) have been made transparent by
modifying the color bar opacity curve, allowing the interpreter to see the changes in AI on the reservoir rock.

73
Figure 6.3. Co-rendered acoustic impedance and similarity with interpreted depositional environments. The upper fan feeder
channels show bifurcations that form the mid to lower facies of the fan complex.

74
POROSITY AND LITHOLOGY PREDICTION

Figure 6.4 is an arbitrary cross-section generated from the prestack migrated

inverted volume, which allows visual correlation of the gamma-ray and density porosity

curves from Wells D, F, G, H and I with the AI attribute. This correlation indicates that

Cherokee Wash clean sandstones with porosities around 10% normally have impedance

values approximately 39,000-40,000[(g/cc)(ft/s)]. Dirty sandstones of the upper and

lower Cherokee Wash with porosities between 2-5% have impedances higher than 40,000

[(g/cc)(ft/s)], shales from the middle and lower Cherokee Wash and overlying Caldwell

Wash with porosities of less than 2% have impedance values of less than 39,000

[(g/cc)(ft/s)]. This interpretation is consistent with the cross plots from Chapter Four.

Based on the cross plot from Chapter Four, impedance greater than 39,400

[(g/cc)(ft/s)] is more likely related to sandstones, whereas impedance lower than this

value is likely related to silts and shales. Visual correlation of these values with the

gamma-ray and density porosity logs in figure 6.4 corroborates the results from Chapter

Four.

Figure 6.5 represents the co-rendered AI and similarity volume including the

arbitrary line from figure 6.4. Notice how the arbitrary line intentionally cuts across the

interpreted channelized features interpreted earlier. Also the AI color bar has been

75
modified to only display values greater than 38,000 [(g/cc)(ft/s)] which correspond to

sandstones. Figure 6.6 further corroborates the results from the AI inversion. Notice the

abundance of completed horizontal wells in the area.

76
Figure 6.4. Arbitrary line showing the AI inverted from the prestack migrated data, including the gamma-ray log (left) and
density porosity (right).

77
Figure 6.5. Image showing the AI and similarity geobody including the arbitrary line and the wells displayed on the section
constructed from the wells.

78
Figure 6.6. Image of the co-rendered AI and similarity geobody including the horizontal wells completed on the area. The
presence of horizontal wells is an indirect affirmation of the better reservoir properties predicted by the attributes.

79
CHAPTER VII
DISCUSSION

Performing inversion for AI in the Granite Wash of the Texas Panhandle makes it

possible to better define the fan morphology and lithofacies of the Cherokee Wash.

Sandstones with relative high impedance are typical of reservoirs in the Mid-continent

(Peddy et al., 1995; Eissa and Castagna, 2003). The geologic mechanism for high

impedance sandstones at depth is differential compaction (Neidell and Berry, 1989).

Consequently, the overall high impedance units more than 39,400 [(g/cc)(ft/s)] are

anticipated to correlate with high sandstone/shale ratios, and local decreases in

impedance units are interpreted as areas with increased porosity. A key feature in Figure

7.1 is that it shows the distribution of the higher AI areas (green, yellow and, red) on an

attribute surface from the Cherokee Wash. Within these areas of higher AI, one can

consistently find areas with decreased AI that regularly correspond to increased porosity.

Furthermore, the interpreted section within figure 7.1 displays the AI (right) and porosity

(left) logs from Well D. Notice how the area with lower AI towards the top of the unit

corresponds with porosities close to 12%. This area of higher porosity also corresponds

with AI amplitude values between 38,000 -39,000 [(g/cc)(ft/s)] previously established as

values of high porosity for sandstones within the Cherokee Wash.

80
Figure 7.1. Interpreted cross line (top) extracted from the prestack migrated inverted volume
along with the corresponding AI surface (bottom). Note the areas on the attribute surface that
were previously interpreted as braided stream channels. The interpreted line on top shows how
the zones with lower AI correspond to higher porosity.

Figure 7.2 is a time thickness map between the top of the Cherokee Wash and the

underlying GRWA. Note how the thicker intervals directly coincide with the interpreted

upper and middle fan facies. The thicker areas in the isochron map also correspond to

81
higher impedance zones with predicted higher sandstone/shale ratios. Theoretically, the

areas with higher impedance (higher sandstone/shale ratio) should show evidence of

differential compaction and so be thicker. The comparison between the isochron map and

the impedance surface confirms this prediction.

82
Figure 7.2. Isochron map to the left and AI to the right. Note how the areas of higher AI correspond with the thicker intervals in
the Cherokee Wash.

83
Figure 7.3 represents an energy-weighted coherent-amplitude gradient attribute

surface also extracted using a 15ms search window below the Cherokee Wash. In

summary, the gradients are large when there is rapidly varying, high-amplitude coherent

energy and the gradients are small when the reflectivity is smoothly varying, low energy,

or incoherent (Chopra and Marfurt, 2008). Gradients of coherent energy are useful in

delineating thin channels as gradients emphasize subtle lateral changes in tuning (Chopra

and Marfurt, 2008). In this case, the areas of low coherent energy (black) correspond with

the areas of higher impedance described earlier. Similarly, the coherent energy attribute

also delineates the edges of the fan complex.

Figure 7.4 is a co-rendered attribute surface combining AI and energy-weighted

coherent-amplitude gradient. Note how the areas of low coherency correspond with the

higher impedance areas. Additionally, the energy-weighted coherent-amplitude gradient

better delineates channels thinner than the tuning thickness; these channels, or braided

channel systems, are expressed as slight changes in lateral amplitude rather than wave

form.

Figure 7.5 is an interpreted AI and coherent-amplitude gradient surface. Note the

interpreted upper, middle, and lower fan depositional environments.

84
Figure 7.3. Coherent-amplitude gradient surface. Note how the attribute enhances the contrast between streams at the fan edge.

85
Figure 7.4. Co-rendered AI and coherent-amplitude gradient. The yellow arrows follow what is interpreted as a braided channel
system.

86
Figure 7.5. Interpreted co-rendered AI and coherent-amplitude gradient. Note the discrete depositional environments: Upper fan,
middle fan and fan, edges.

87
The well control shown in figure 7.6 inside the survey confirms the previous

interpretations. The wells denoted by the black arrows follow the path of the largest

interpreted braided stream channel system. Within the channel system, well log

signatures for the gamma-ray (left) are very clean (less than 70 API units). Furthermore,

the density porosity log (right) shows porosities up to 12% in this area. On the other

hand, the wells indicated by the orange arrows all lie in areas interpreted to have higher

shale/sandstone ratios commonly found along the distal facies of the lower fan as

sediment prograted toward deeper marine environments. The logs in this area show a

greater clay content and lower porosity.

88
Figure 7.6. Co-rendered AI and coherent-amplitude gradient along with gamma-ray and density porosity logs from selected wells
within the survey. Note how the wells highlighted by the black arrows correspond with the interpreted braided stream channel
system of higher porosity and cleaner sands. The wells denoted by the orange arrows show higher shale/sandstone ratios and
lower porosity as one would expect by the seismic attributes response.

89
The proposed facies distributions suggested here are based in the morphologic

subdivisions suggested by Bouma, (1985), Mutti (1977), and Walker (1978). Within the

Cherokee Wash the upper fan is characterized by three main braided stream channels that

act as conduits for the sand and gravel moving out toward the fan. These main channels

are probably prone to the deposition of coarse gravel facies that became conglomerates.

As the channels braid and switch position, the sand bodies appear to coalesce to

form the middle fan depositional environment. Likely the facies to be deposited in the

middle fan channels are pebbly massive sandstones which result from a lower gradient

depositional environment.

The lower fan area of topographically smoother and lower gradient regions are

interpreted as zones of lower depositional energy (lower AI and higher amplitude

coherency). The depositional facies for this area are interpreted to be shales interrupted

periodically by sand-rich turbidity currents. Figure 5.7 is a depositional model created

from the interpretation of the seismic attributes. Furthermore, the well control within the

survey area confirms the interpretations based on the seismic attributes.

90
Figure 7.7. Submarine fan depositional model relating, fan geomorphology, and depositional environment.

91
CHAPTER VIII
CONCLUSION

The pre-inversion analysis in the Cherokee Wash established that most sandstones

have AI values of more than 39,400 [(g/cc)(ft/s)], and that most of the shales have AI

values of less this threshold value. The pre-inversion analysis also demonstrates that

sandstones with lower AI 39,400 – 42,000 [(g/cc)(ft/s)] tend to have higher porosity and

that sandstones with lower porosity tend to have much higher AI values > 42,000

[(g/cc)(ft/s)].

After inverting the prestack and postack migrated volumes, the petrophysical

parameters previously established in the pre-inversion analysis were used to differentiate

lithology and confirmed using the well control. Furthermore, comparing the results of

inverting prestack migrated data versus postack migrated data established that the higher

frequency content from the prestack migrated data increased the temporal and spatial

resolution of the inverted volume. In addition, it is further suggested that the residual

velocity analysis and the azimuthal velocity correction in the prestack migrated data

produced a better velocity model that allowed for a more accurate location of events

during the migration process, also improving the spatial resolution.

The use of geometric attributes is demonstrated to be a powerful tool in

delineating and separating discrete depositional environments within the Cherokee Wash.

92
Moreover, by combining the different geometric attributes with the inverted AI it was

possible not only to build a geomorphological model for the reservoir, but also to

delineate the lithological heterogeneity and reservoir properties within the Cherokee

Wash.

93
REFERENCES

Asquith, G., and D Krygowski., 2004, Basic Well Log Analysis. AAPG Methods in
Exploration No. 16. 244 p.

Adler, F., 1971, Anadarko basin and central Oklahoma area, in future petroleum
provinces of the United States – their geology and potential. AAPG Memoir 15, v. 2, p.
1061-1070

Benabentos, M., M. Silva, F. Ortigosa, and V. Mercado, 2007, Reservoir characterization


in Burgos Basin using simultaneous inversion: The Leading Edge, v. 26, no. 5,p. 556–
561.
Bouma, H., 2000, Coarse-grained and fine-grained turbidite systems as end member
models: applicability and dangers. Department of Geology and Geophysics, Louisiana
State University, Baton Rouge, Louisiana. 351 p.
Bouma, H., and W. R. Normark, N. E. Barnes., 1985, Submarine Fans and Related
Turbidite Systems. Springer-Verlag New York.
Carter, L., Kelley, S., Blackwell, D., and Naeser., 1998, Heat flow and thermal history of
the Anadarko Basin, Oklahoma. AAPG Bulletin, v. 82, No 2. P. 291-316.
Cunningham, B. J., 1961, Stratigraphy of Oklahoma – Texas Panhandles: oil and gas
fields of the Texas and Oklahoma panhandles: Amarillo, Panhandle Geological Society,
p. 45-60.
Cardott, J., and Michael Lambert, 1982, Thermal Maturation by Vitrinite Reflectance of
Woodford Shale, Anadarko Basin, Oklahoma. AAPG Bulletin, v.69, p.1982-1998.
Calderon, J. E., and J. Castagna, 2007, Porosity and lithologic estimation using rock
physics and multi-attribute transforms in Balcon Field, Colombia: The Leading Edge, v.
26, no. 2, p. 142–150.
Castagna, J. P., and H. W. Swan, 1997, Principles of AVO crossplotting: The Leading
Edge, Interpreter’s Corner, v. 16, no. 4, p. 337–342.
Chopra, S., and Kurt J. Marfurt., Seismic attributes for prospect identification and
reservoir characterization. SEG Geophysical Developments Series No.11, 2008.
Dubucq, D., S. Busman, and P. Van Riel, 2001, Turbidite reservoir characterization:
Multi-offset stack inversion for reservoir delineation and porosity estimation; A Gulf of
Guinea example: SEG, Expanded Abstracts, v. 20, no. 1, p. 609–612.

94
Dutton, S., 1984, Fan-delta Granite Wash of the Texas Panhandle. Oklahoma City
Geological Society. Short course 1984.
Dutton, S., and Land., 1985, Meteoric Burial Diagenesis of Pennsylvanian Arkosic
Sandstones, Southwestern Anadarko Basin, Texas. AAPG Bulletin v.69, No 1, p. 22-38.
Eissa, M. A., and J. P. Castagna, 2003, Case Study: AVO analysis in high impedance
Atoka Sandstone (Pennsylvanian), North Arkoma Basin, McIntosh County, Oklahoma:
The Leading Edge, 22, p. 988-997.
Heller, P., and William R. Dickinson., 1985, Submarine Ramp Facies Model for Delta-
Fed, Sand-Rich Turbidite Systems. AAPG Bulletin v. 69. No. 6.
Ham, W. E., R. E. Denison, and C. A. Merritt, 1964, Basement rocks and structural
evolution of southern Oklahoma: Oklahoma Geological Survey Bulletin 95, 302 p.
Hampson – Russell., 1999, Theory of the strata program. Hampson- Russell Software
Services Ltd.
Ham, W. E., 1969, Regional geology of the Arbuckle Mountains, Oklahoma. Oklahoma
Geological Survey Guidebook 17, 52p.
Ham, W. E., and J. L. Wilson, 1967, Paleozoic epirogeny and orogeny in the central
United States. American Journal of Science, v. 265, p. 332-407.
Hill, G., and Clark., 1980, The Anadarko basin – a regional petroleum accumulation—a
model for future exploration and development. Shale Shaker, v. 31, No 3, p. 36-49.
Hill, S. J., 2005, Inversion-based thickness determination: The Leading Edge, v. 24, no.5,
p. 477–480.
Johnson, K. S., T. W. Amsden, R. E. Denison, S. P. Dutton, A. G. Goldstein, B. Rascoe,
Jr., P. K. Sutherland, and D. M. Thompson, 1988, Southern mid- continent region, in L.
L. Sloss, ed., Sedimentary cover – North American craton, U. S.: the geology of North
America: Geological Society of America, V. D-2, p. 307-359.
Latimer, R. B., 2005, Uses, abuses, and examples of seismic-derived acoustic impedance
data: What does the interpreter need to know, AAPG/SEG Distinguished Lecture:
http://www.seg.org/SEGportalWEBproject/portals/SEG_Online.portal?_
nfpb=true&_pageLabel=pg_gen_content&doc_URL=prod/SEG-Education/Ed-
Presentation-Library/Dl-Presentations/fall2005/presentation.htm, accessed December
2009.

95
Latimer, R. B., and Rick Davidson., 2000, An interpreter’s guide to understanding and
working with seismic-derived acoustic impedance data. The Leading Edge, v. 19, p. 242-
256, doi: 10.1190/ 1.1438580.
Latimer, R. B., P. van Riel, 1996, Integrated seismic reservoir characterization and
modeling: A Gulf of Mexico 3D case history: Paper presented at the Gulf Coast Society
SEPM Research Conference: http://www.fugro-
jason.com/readingroom/techpapers/GOM_1996_GCSSEPM.pdf, accessed November
2009.
Mosher, C., Timothy H. Keho, Arthur B. Weglein and Douglas J. Foster., 1996, The
Impact of Migration on AVO. GEOPHYSICS, VOL. 61, NO.6.
Mutti E., 1975, Distinctive thin-bedded turbidite facies and related depositional
environments in the Eocene Hecho Group (south-central Pyrenees, Spain). Instituto di
Geologia dell’ Universita di Torino, Palazzo Carignano, 10123 Torino, Italy.
Neidell, N., and Neal Berry, 1989, Documenting the sand/shale crossover.
GEOPHYSICS, VOL.54, NO.11.
Parks M., 2011, Depositional Setting of Desmoinesian (Marmaton and Canabiss) Granite
Wash Buffalo Wallow Field Hemphill County, Texas. Unpublished MS. Thesis;
Oklahoma State University, Stillwater, Oklahoma 74078.
Russell, B., 1988, Introduction to Seismic Inversion Methods. SEG Course Notes Series,
No. 2.
Singh, Y., 2007, Lithofacies detection through simultaneous inversion and principal
component attributes: The Leading Edge, v. 26, no. 12, p. 1568–1575.
Walker, R., 1978, Water Sandstone Facies and Ancient Submarine Fans: Models for
Exploration Stratigraphic Traps: AAPG Bulletin, v. 62, No. 6, p. 932-966.
Yilmaz, Ӧ., 2001, Seismic Data Analysis: processing, inversion, and interpretation of
seismic data. SEG Investigations in Geophysics, Volume I.A

96
VITA

Gabriel E. Gavidia

Candidate for the Degree of

Master of Science

Thesis: ATTRIBUTE SUPPORTED SEISMIC GEOMORPHOLOGY


AND RESERVOIR CHARACTERIZATION OF THE GRANITE WASH,
ANADARKO BASIN, TEXAS

Major Field: Geology

Bibliographical:

Personal: Born in Merida, Venezuela, April 11, 1984, the son of Francisco
Gavidia and Gisela Garcia.

Education: Graduated from Liceo Romulo Gallegos, Merida, Venezuela, June,


2002; received Bachelor of Science degree in Geology from
Arkansas Tech University, in August 2010; completed
requirements for Master of Science at Oklahoma State University
in June, 2012.

Experience: Research Assistant, Oklahoma State University 2010 to 2012;


summer internship with EOG Resources 2011; Employed as a
Geophysicist I at Noble Energy Inc.

Professional Memberships: AAPG, SEG, GSOC, GSA.


Name: Gabriel Gavidia Date of Degree: July, 2012

Institution: Oklahoma State University Stillwater, Oklahoma

Title of Study: ATTRIBUTE SUPPORTED SEISMIC GEOMORPHOLOGY


AND RESERVOIR CHARACTERIZATION OF THE GRANITE WASH,
ANADARKO BASIN, TEXAS

Pages in Study: 96 Candidate for the Degree of Master of Science

Major Field: Geology

Scope and Method of Study: Hydrocarbon bearing alluvial fan deposits from the Texas
Panhandle Granite Wash were studied using geometric attributes and postack migrated
seismic inversion techniques.

Findings and Conclusions: Seismic attributes and seismic inversion have become
increasingly useful for characterizing hydrocarbon-bearing reservoirs. These tools allow
the seismic interpreter to delineate specific depositional patterns and their corresponding
geomorphology. This study implements and evaluates these techniques on a geologic unit
that has historically been difficult to characterize seismically, the Pennsylvanian Granite
Wash of the Texas Panhandle. In addition, this study compares how various seismic
processing sequences impact postack seismic inversion.
Seismic similarity, computed from postack migrated data, and energy-weighted coherent-
amplitude gradient, computed from prestack migrated data, delineated the
geomorphology of the Granite Wash reservoir in the study area. Inverted acoustic
impedance (AI) computed from the seismic amplitude volumes combined with density
and sonic logs provided an excellent means for mapping reservoir heterogeneity.
Previous wells drilled in the area targeted channel facies that correspond to the upper and
middle sections of the submarine fans (Washes). This study demonstrates that the
combination of seismic attributes and inversion for AI greatly facilitates the location and
interpretation of hydrocarbon-bearing reservoir rock within the upper and middle fan
depositional environments.

ADVISER’S APPROVAL: Estella Atekwana

You might also like