+ All Categories
Home > Documents > ELEMENTAL GEOCHEMISTRY OF SHALES IN PENNSYLVANIAN A ...

ELEMENTAL GEOCHEMISTRY OF SHALES IN PENNSYLVANIAN A ...

Date post: 03-Apr-2022
Category:
Upload: others
View: 3 times
Download: 0 times
Share this document with a friend
179
ELEMENTAL GEOCHEMISTRY OF SHALES IN PENNSYLVANIAN CYCLOTHEMS. MIDCONTINENT NORTH AMERICA by WEE SENG TEO, B.PHARM., M.S. A DISSERTATION IN GEOSCIENCE Submitted to the Graduate Faculty of Texas Tech University in Partial Fulfillment of the Requirements for the Degree of DOCTOR OF PHILOSOPHY Approved August, 1991
Transcript

ELEMENTAL GEOCHEMISTRY OF SHALES IN PENNSYLVANIAN

CYCLOTHEMS. MIDCONTINENT NORTH AMERICA

by

WEE SENG TEO, B.PHARM., M.S.

A DISSERTATION

IN

GEOSCIENCE

Submitted to the Graduate Faculty of Texas Tech University in

Partial Fulfillment of the Requirements for

the Degree of

DOCTOR OF PHILOSOPHY

Approved

August, 1991

io\ ;'•—' '^ } 7"

c c /o 9 / 0 .

©1991, Wee Seng Teo.

ACKNOWLEDGEMENTS

I would like to thank the committee members Drs. Calvin G. Barnes, James E.

Barrick, B. L. Allen, Necip Guven, and Thomas M. Lehman for their encouragement,

advice, and guidance during the course of this smdy and in the preparation of this

dissertation. I am grateful to the Department of Geosciences for awarding me a Grover

E. Murray Scholarship and the Lewis G. Weeks Research Fellowship.

I would also like to thank these persons for their help in the specified projects

mentioned: James E. Barrick and Darwin R. Boardman (field work and sample

collection), Melanie Barnes and Bill Shannon (major and trace element analyses by

Inductive Coupled Plasma Emission Spectroscopy and Atomic Absorption

Spectroscopy, and ferrous iron determination). Nelson Rolong (total organic carbon

determination by titration), Chariie Landis of Arco Oil and Gas Company in Piano,

Texas (sulfur and total organic carbon determination by infrared spectral analysis), Mike

Gower (preparation of shale thin sections), and Alonzo D. Jacka, Thomas M. Lehman,

and Ali Trabelsi (petrography). The librarians of Texas Tech Library and Geoscience

reading room, and the office staff of Department of Geosciences kindly allowed me the

use of their facilities.

Let me express my appreciation to professors, non-teaching staff, and fellow

student colleagues who had made my sojourn in Lubbock, Texas, USA, a pleasant and

memorable experience.

u

CONTENTS

ACKNOWLEDGEMENTS ii

ABSTRACT v

FIGURES vii

CHAPTER

L INTRODUCTION 1

Objectives 1

Controls on Element Abundance in Shales 2

Pennsylvanian Cyclothems 8

IL LOCAL STRATIGRAPHY OF SECTIONS 18

North-Central Texas 18

Kansas and Oklahoma 26

Permian Bead Mountain Limestone 28

m. GEOCHEMICAL ANALYSIS 42

Methods and Analysis 42

Distribution of Elements in Stratigraphic Sections 48

IV. GEOCHEMICAL DISTRIBUTION OF ELEMENTS 68

Elements in Detrital Minerals 68

Elements Associated With Calcium Carbonate 73

Organic Carbon, Sulfur, Iron, and Manganese 76

Essential Transition Metals 80

V. DISCUSSION 104

VI. CONCLUSIONS 110

REFERENCES 113

m

APPENDICES

A. PETROGRAPHY OF SELECTED SAMPLES 123

B. GEOCHEMICAL DATA 124

C. STRATIGRAPHIC DISTRIBUTION 133

IV

ABSTRACT

Pennsylvanian cyclothemic marine shales present a wide range of depositional

environments that allow the study of depositional controls on distribution of certain

elements in shales. Samples were collected from upper Desmoinesian to lower Virgilian

units in north-central Texas, Kansas, and Oklahoma, The samples were analyzed for

Si, Ti, Al, Fe, Mn, Mg, Ca, Na, K, P, Sc, V, Cr, Co, Ni, Cu, Zn, Be, Sr, Ba, Zr, Y,

Rb, S, and total organic carbon (TOC). X-ray diffraction showed that illite, kaolinite,

and quartz were the predominant minerals. The weathering index and the chemical

index of alteration both indicate that the source minerals of the shales were highly

weathered. Thin sections reveal the presence of red brown aggregates of clay, organics,

and oxides, gray clay aggregates, and quartz grains. Abundances of Mn and Fe are

quite variable (except for Mn in calcareous shales, and Fe in pyritic shales).

Core shales, deposited during maximum transgression, may be high or low TOC

shales depending on the original sedimentary redox conditions. In high TOC core

shales (TOC/Al ratio above 1.2), abundances of V, Zn, and Cr correlate strongly with

TOC. Sulfur correlates strongly with Fe. In low TOC core shales (TOC/Al ratio below

1.2), abundances of V, Zn, and Cr do not correlate with TOC. In some low TOC core

shales, Zn, Cr, Ni, and Cu increase in maximum transgressive intervals and decrease

stratigraphically upwards due to dilution by deltaic clays. Outside shales, deposited

during regression, are normal to marginal marine shales with low TOC. Carbonate-

related elements (Ca, Sr, Zn, Mn, P, Y, Ni) are more abundant where the shale contains

more calcareous skeletal material. Marginal marine shales show widely variable TOC

and elemental composition.

This study indicates that the main factors controlling the distribution of elements

in cyclothemic shales are (1) the degree of weathering before deposition, (2) redox

condition in the depositional environment, (3) settling time of the clay and organic matter

through the water colunm, (4) conditions conducive to the formation and deposition of

carbonates, (5) the composition of the organic matter, and (6) dilution by fine-grained

terrestrial sediments.

VI

FIGURES

1. Basic vertical sequence of an individual Pennsylvanian cyclothem 13

2. Cross-section showing low stand of sea level during regression and high stand of sea level during transgression in west-facing tropical epicontinental sea 14

3. Upper Pennsylvanian paleogeography of the Midcontinent Region of the United States showing position of paleoequator and areas of orogenic belts 15

4. Paleogeographic map showing probable fades relations of Upper Pennsylvanian Midcontinent sea during deposition of offshore shale along Midcontinent outcrop at phase of maximum transgression 16

5. Paleogeographic map showing probable facies relations of Upper Permsylvanian Midcontinent sea during deposition of upper part of regressive limestone, and locally base of nearshore shale along Midcontinent outcrop about midway through phase of late regression 17

6. Distribution of Pennsylvanian strata in the Midcontinent Region of the United States 30

7. Eustatic sea-level curve for part of Pennsylvanian sequence in north-central Texas outcrop (right) and biostratigraphic correlation with curve forMidcontinent outcrop (left) 31

8. Onshore-offshore model for Pennsylvanian community succession related to water depth and overlying water masses 32

9. Depositional model incorporating deltaic influx into regressive phase of eustatic model for Pennsylvanian cyclothems in Texas and Oklahoma 33

10. Stratigraphic profile of East Mountain (EM) section 34

11. Stratigraphic profile of Dog Bend (DB) section (Lower Salesville) 35

12. Stratigraphic profile of 3027 section (Upper Salesville) 36

13. Stratigraphic profile of UPS section (Upper Salesville) 37

14. Stratigraphic profile of Colony Creek (CCB) section at Brad 38

15. Stratigraphic profile of Type Lost Branch (LBK) section in Kansas 39

16. Stratigraphic profile of Lower Tackett Shale (TRR) at Tulsa Railroad Yard in Oklahoma 40

• • vii

17. Stratigraphic profile of Permian P180 section (Bead Mountain Limestone) 41

18. Graphs of LECO total organic carbon versus titration total organic carbon 56

19. Graph of Munsell value number N versus total organic carbon (weight per cent) for all samples 57

20. Stratigraphic distributions of zinc, chromium, nickel, vanadium, copper, cobah, and total organic carbon in East Mountain (EM) section 58

21. Stratigraphic distributions of zinc and Munsell value number N in EM, 3027, and UPS sections 59

22. Stratigraphic distributions of calcium, yttrium, manganese, phosphorus, strontium, zinc, and nickel in Dog Bend (DB) section 60

23. Stratigraphic distributions of zinc, chromium, nickel, vanadium, copper, cobalt, and total organic carbon in UPS section 61

24. Stratigraphic distributions of zinc, chromium, nickel, vanadium, copper, cobalt, and total organic carbon in 3027 section 62

25. Stratigraphic distributions of calcium, yttrium, manganese, phosphonis, strontium, zinc, and nickel in Colony Creek (CCB) section 63

26. Graphs of sulfur versus total iron and sulfur versus total organic carbon 64

27. Stratigraphic distributions of zinc, vanadium, chromium, nickel, copper, and total organic carbon in Type Lost Branch (IBK) section 65

28. Stratigraphic distributions of vanadium, chromium, zinc, nickel, copper, and total organic carbon in Lower Tackett Shale (TRR) section 66

29. Stratigraphic distributions of calcium, yttrium, manganese, phosphorus, strontium, zinc, and nickel in P180 section 67

30. Frequency histograms of silicon and titanium abundances for all samples 86

31. Frequency histograms of scandium and zirconium abundances for

all samples 87

32. Graphs of vanadium versus potassium and zirconium versus silicon 88

33. Graph of weathering index for all samples 89 • ••

Vlll

34. Graphs of chemical index of alteration (CIA) for all samples 90

35. Graphs of potassium versus aluminum and rubidium versus potassium for all samples 91

36. Frequency histograms of potassium and rubidium abundances for all samples 92

37. Eh-pH diagram showing stability field of cobah sulfide 93

38. Eh-pH diagram showing stability field of copper sulfide 94

39. Eh-pH diagram showing stability field of iron sulfide 95

40. Eh-pH diagram showing stability field of manganese sulfide 96

41. Eh-pH diagram showing stability field of nickel sulfide 97

42. Eh-pH diagram showing stability field of zinc sulfide 98

43. Graphs of total iron versus total organic carbon 99

44. Graphs ofmanganese versus total organic carbon 100

45. Graph of vanadium versus total organic carbon for Lower Tackett (TRR) and Type Lost Branch (LBK) shales 101

46. Graphs of vanadium, zinc, chromium, nickel, copper, and cobalt versus total organic carbon for low TOC shales 102

47. Graphs of vanadium, zinc, chromium, nickel, copper, and cobalt versus total organic carbon for high TOC shales 103

48. Stratigraphic distribution of total organic carbon (TOC) in Upper Salesville (3027 and UPS), Lower Salesville (DB), and P180 sections 134

49. Stratigraphic distribution of total organic carbon (TOC) in East Mountain (EM) and Colony Creek at Brad (CCB) sections 135

50. Stratigraphic distribution of total organic carbon (TOC) in Tulsa RR (TRR) and Lost Branch (LBK) sections 136

51. Stratigraphic distribution of scandium and titanium in Upper Salesville (3027 and UPS), Lower Salesville (DB), and PI80 sections 137

52. Stratigraphic distribution of scandium and titanium in East Mountain (EM) and Colony Creek at Brad (CCB) sections 138

53. Stratigraphic distribution of scandium and titanium in Tulsa RR (TRR) and Lost Branch (LBK) sections 139

ix

54. Stratigraphic distribution of silicon and zirconium in Upper Salesville (3027 and UPS), Lower Salesville (DB), and P180 sections 140

55. Stratigraphic distribution of silicon and zirconium in East Mountain (EM) and Colony Creek at Brad (CCB) sections 141

56. Stratigraphic distribution of siHcon and zirconium in Tulsa RR (TRR) and Lost Branch (LBK) sections 142

57. Stratigraphic distribution of barium and berylUum in Upper Salesville (3027 and UPS), Lower Salesville (DB), and PI80 sections 143

58. Stratigraphic distribution of barium and beryllium in East Mountain (EM) and Colony Creek at Brad (CCB) sections 144

59. Stratigraphic distribution of barium and beryllium in Tulsa RR (TRR) and Lost Branch (LBK) sections 145

60. Stratigraphic distribution of potassium and rubidium in Upper Salesville (3027 and UPS), Lower Salesville (DB), and P180 sections 146

61. Stratigraphic distribution of potassium and rubidium in East Mountain (EM) and Colony Creek at Brad (CCB) sections 147

62. Stratigraphic distribution of potassium and rubidium in Tulsa RR (TRR) and Lost Branch (LBK) sections 148

63. Stratigraphic distribution of magnesium and sodium in Upper Salesville (3027 and UPS), Lower Salesville (DB), and PI80 sections 149

64. Stratigraphic distribution of magnesium and sodium in East Mountain (EM) and Colony Creek at Brad (CCB) sections 150

65. Stratigraphic distribution of magnesium and sodium in Tulsa RR (TRR) and Lost Branch (LBK) sections 151

66. Stratigraphic distribution of calcium and strontium in Upper Salesville (3027 and UPS), Lower Salesville (DB), and PI80 sections 152

67. Stratigraphic distribution of calcium and strontium in East Mountain (EM) and Colony Creek at Brad (CCB) sections 153

68. Stratigraphic distribution of calcium and strontium in Tulsa RR (TRR) and Lost Branch (LBK) sections 154

69. Stratigraphic distribution ofmanganese and total iron in Upper Salesville (3027 and UPS), Lower Salesville (DB), and PI80 sections 155

70. Stratigraphic distribution ofmanganese and total iron in East Mountain (EM) and Colony Creek at Brad (CCB) sections 156

71. Stratigraphic distribution of manganese and total iron in Tulsa RR (TRR) and Lost Branch (LBK) sections 157

72. Stratigraphic distribution of phosphorus and yttrium in Upper Salesville (3027 and UPS), Lower Salesville (DB), and PI80 sections 158

73. Stratigraphic distribution of phosphorus and yttrium in East Mountain (EM) and Colony Creek at Brad (CCB) sections 159

74. Stratigraphic distribution of phosphorus and yttrium in Tulsa RR (TRR) and Lost Branch (LBK) sections 160

75. Stratigraphic distribution of nickel and vanadium in Upper Salesville (3027 and UPS), Lower Salesville (DB), and P180 sections 161

76. Stratigraphic distribution of nickel and vanadium in East Mountain (EM) and Colony Creek at Brad (CCB) sections 162

77. Stratigraphic distribution of nickel and vanadium in Tulsa RR (TRR) and Lost Branch (LBK) sections 163

78. Stratigraphic distribution of chromium and zinc in Upper Salesville (3027 and UPS), Lower Salesville (DB), and P180 sections 164

79. Stratigraphic distribution of chromium and zinc in East Mountain (EM) and Colony Creek at Brad (CCB) sections 165

80. Stratigraphic distribution of chromium and zinc in Tulsa RR (TRR) and Lost Branch (LBK) sections 166

81. Stratigraphic distribution of cobalt and copper in Upper Salesville (3027 and UPS), Lower Salesville (DB), and P180 sections 167

82. Stratigraphic distribution of cobalt and copper in East Mountain (EM) and Colony Creek at Brad (CCB) sections 168

83. Stratigraphic distribution of cobalt and copper in Tulsa RR (TRR) and Lost Branch (LBK) sections 169

XI

CHAPTER I

INTRODUCTION

Objectives

Sedimentary rocks cover 66 percent of the surface of the continent, and because

65 percent of all sedimentary rocks are shales (Ehlers and Blatt, 1982), the sttidy of

shales is important for understanding a large portion of the earth's sedimentary rocks.

The physical and chemical conditions of the depositional environment could cause the

distribution of elements to vary in marine shales to the extent that particular associations

of elements tend to characterize particular sedimentary conditions. The chemical

variations may provide clues about the prevailing environmental conditions in the area of

deposition. Therefore, the study of shale geochemistry may be helpful in reconstructing

environmental conditions. Transition metals associated with organic matter may vary

according to whether the organic matter is terrestrial or marine. Settling time, which

depends upon size and density of settling particles, controls the efficiency of adsorption

of elements by organic and clay particles (Vine and Tourtelot, 1970; Yin et al., 1989).

Redox potential of the sediment affects accumulation of organic matter, metal sulfides,

and metal oxides. The rate of terrestrial sediment influx into the depositional basin

affects the relative abundances of the marine organic matter and the transition metals

associated with it.

Environmental conditions may enrich some elements in shales and make their

mining and extraction economical (Krauskopf, 1955). The rare transition metals most

characteristic of shales with high organic carbon content are vanadium, nickel, cobalt,

copper, and molybdenum (Krauskopf, 1956), which are economically important metals.

Shales could act as sources of pollution if any undesired elements they contain

leak out with flushing water. Generally, shales could act as pollutant sinks if they

1

adsorb undesired elements with which they come into contact, and serve as hydraulic

barriers. The shales are then useful for environmental protection (Vine and Tourtelot,

1970).

Sediments forming shales are usually deposited with varying amounts of organic

matter. Gray shales contain about 1% organic carbon and black shales contain about 3%

organic carbon (Vine and Tourtelot, 1970). Black shales are source rocks for

hydrocarbons because of the amount of initial organic matter in them (Twenhofel, 1939;

Blatt, Middleton, and Murray, 1980); the hydrocarbons can be distilled from them

(Krauskopf, 1967).

Most of the known marine oil source beds are in shales that were originally

deposited during transgression and under anaerobic conditions (DeMaison and Moore,

1980). Knowing the geochemistry of transgressive and regressive shales should help

narrow the possible exploration areas in the search for hydrocarbons and for transition

metals of economic value.

The objectives of this study are to determine the stratigraphic abundance and

distribution of elements in Pennsylvanian cyclothemic shales, to infer the sedimentary

conditions in the environment of deposition, and to estimate the degree of weathering of

the source materials. Deep-water to shallow-water shales were selected as

representatives of different depositional environments. The characteristic Hving

communities are known and can be used as a background correlation for the

geochemical stratigraphy. The elemental associations found in these shales may help in

defining geochemical characteristics of cyclothems.

Controls on Element Abundance in Shales

Changes in the elemental geochemistry of shales from one stratigraphic layer to

the next provide clues to the provenance, depositional, and diagenetic history of the

2

shale. Althougih the concentration of a single element may contribute information about

some aspects of provenance, deposition, or diagenesis, multi-element analysis

constrains the models used to explain the accumulation, preservation, or depletion of

elements as present in the shale (Shaw et al., 1990).

The presence or absence of resistant detrital minerals in fine-grained terrigenous

clastic sediments will affect the abundance of certain elements, in particular, aluminum,

silicon, titanium, and zirconium. The proportion of detrital quartz and detrital heavy

minerals, such as rutile, ilmenite, and zircon, in the sediment will significantly affect the

silicon, titanium, and zirconium abundances. The distribution of resistant detrital

minerals is a consequence of sedimentological processes that distribute the grains by

size and density.

The clay minerals in shales, mostly detrital in origin, are weathering products of

silicate minerals. Other common elements in clay minerals are sodium, potassium,

calcium, magnesium, and iron (Murray, 1954). Clay minerals in marine shales are

mainly kaolinite and illite, whereas chlorite, montmorillonite-smectite, and mixed-layer

clays occur in lesser amounts (Murray, 1954; Degens et al., 1957; Potter et al., 1963;

Coveney and Martin, 1983; Schultz, 1987). Illite is the predominant component of

marine shales more so than in nonmarine shales (Keith and Degens, 1959); illite is more

stable in marine sediments and kaolinite is more stable in freshwater sediments (Weaver,

1967).

The types of clay minerals affect distribution of elements in shales (Degens et

al., 1957). The elements could be an essential part of the clay minerals or could be

associated with them. Elements associated with clay minerals could occur as

isomorphous substitutions in the crystal structure and/or exchangeable cations adsorbed

on the mineral surface (Chester, 1965; Cody, 1971; Paropkari, 1990). Strucmral

substinitions include Fe2+ for Mg2+, Fe3+ for Al3+ (Chester, 1965; Paropkari, 1990),

Co2+ or Ni2+ for Fe2+ or Mg2+ (Nicholls and Loring, 1962), and V3+ for Al3+

(Rankama and Sahama, 1950). In the clay interlayer, Ba2+ may sustimte for K" in

illite (Murray, 1954; Nicholls and Loring, 1962; Cubitt, 1979) and Mg2+ and Na+ for

Ca2"'" in calcium-rich terrigenous clays (Weaver 1967).

Elements dissolved in sea water may be scavenged by clay minerals, iron

oxides, manganese oxides, and organic matter (Krauskopf, 1956; Vine and Tourtelot,

1970; Shaw et al., 1990). Insoluble metal oxides and hydroxides may be adsorbed onto

clay platelet surfaces (Chester, 1965; Cody, 1971).

The presence of disseminated organic matter in fine-grained terrigenous clastic

sediments affects elemental abundances in three different ways: by bringing the

elements physically with them into the sediment, by chelating the elements from the

surrounding pore water when in the sediment, and indirectly by changing the redox

potential which affects the dissolution or precipitation of elements (Goldberg, 1954;

Krauskopf, 1955, 1956; Degens et al., 1957; Brongersma-Sanders, 1969; Holland,

1979; Dabard and Paris, 1986).

Increased sedimentation of organic matter due to increased biological

productivity will increase sedimentation of metals connected with the organic matter

(Shaw et al., 1990). Metallic elements associated with organic matter are originally part

of the organic matter or are scavenged by adsorption or chelation by the organic matter

(Krauskopf, 1955). Magnesium, iron, copper, nickel, and vanadium are contained,

originally or by scavenging, in porphyrins and are enriched where oiganic matter

accumulates (Keith and Degens, 1959; Brumsack, 1989).

Degens et al. (1957) found that marine organic matter contains more nickel and

vanadium than terrestrial organic matter, which contains more zinc and copper than

marine organic matter. Vanadium is an essential element for the metabolism of algae

(Amon and Wessel, 1953), and the vanadium concentration in some marine plzmts and 4

animals is several thousand times that in the sea water (Krauskopf, 1956; Manskaya and

Drozdava, 1968). Therefore the more organic-rich the shale, the more vanadium it

contains (Wenger and Baker, 1986). Quinby-Hunt et al. (1988) reported that in marine

black shales, the vanadium concentration is correlated with organic carbon, that the

amount of vanadium is an indication of the level of marine productivity, and that it is

also an indication of low redox potential (low Eh). Coal, a terrestrial deposit, has low

content of vanadium (Wenger and Baker, 1986).

Decaying organic matter lowers the redox potential of the sediment and causes

the environment to be reducing. After the oxygen is used up in oxidizing the organic

matter, the remaining organic matter then accumulates in a low redox potential (low Eh).

Thus, the organic matter creates its own environment with low redox potential and

enhances its own accumulation (Krauskopf, 1967; Tourtelot, 1979).

Iron and manganese abundances in natural environments are influenced by redox

potential (Kiumbein and Garrels, 1952). Iron oxides and manganese oxides tend to

remain as insoluble oxides at high redox potential in well-oxygenated waters. Dissolved

manganese and iron are precipitated hydrogenously in oxygenated sediments, the higher

the redox potential the more likely they would be precipitated (Bemer, 1971).

Therefore, high manganese and iron concentrations in shales could indicate aerobic

conditions during deposition (Quinby-Hunt et al., 1988; Coveney et al., 1989). The

concentrations of iron and manganese will start to decrease as redox potential decreases,

with manganese concentration decreasing faster than the iron concentration.

Comparisons of iron and manganese concentrations have been used by numerous

workers to infer increasing or decreasing reducing condition (Krauskopf, 1967; Wenger

and Baker, 1986; Quinby-Hunt et al.,1988; Coveney et al., 1989; Wilde et al., 1989).

Colloidal hydrated manganese oxides and colloidal hydrated iron oxides can

coprecipitate (Arrhenius, 1959; Chester, 1965) and the precipitate may absorb transition 5

metals such as cobalt, chromium, copper, nickel, vanadium, and zinc (Goldberg, 1954;

Krauskopf, 1955; Keith and Degens, 1959; Chester, 1965; Yin et al., 1989; Paropkari,

1990; Shaw et al., 1990). Hydrated manganese oxides and hydrated iron oxides could

also adsorb metal phosphates (Chester, 1965; Paropkari, 1990) and all coprecipitate.

Formation of authigenic minerals like sulfides in the area of deposition may

influence element abundance in shales (Degens et al., 1957; Brongersma-Sanders,

1969). Sulfur abundance as sulfides is dependent upon redox potential (Krumbein and

Garrels, 1952). The sources of sulfur for formation of sulfides are decaying organic

matter and dissolved sulfates (Baas Becking et al., 1960; Vine and Tourtelot, 1970;

Bemer, 1971). The source of iron for pyrite formation is assumed to be hydrated iron

oxides (Keith and Degens, 1959).

In the sediment, decaying organic matter uses up oxygen and in the resulting low

redox potential conditions (Krauskopf, 1967; Tourtelot, 1979) anaerobic bacteria reduce

sulfur from oiganic matter and dissolved inorganic sulfate to hydrogen sulfide (Keith

and Degens, 1959; Baas Becking et al., 1960). The reducing environment causes ferric

iron in oxides to become ferrous iron. The Fe2"'" combines with the S2- in H2S to form

iron monosulfide FeS which combines with elemental sulfur to form pyrite FeS2 (Doner

and Lynn, 1989). During the formation of iron sulfide, transition elements such as

copper, zinc, nickel, and cobalt may be incorporated into the sulfide (Yin et al., 1989).

Shales with high amounts of organic carbon contain various amounts of sulfide

minerals such as pyrite (FeS2), sphalerite (ZnS), chalcopyrite (CuFeS2), marcasite

(FeS2), and covellite (CuS) (Brongersma-Sanders, 1966; Coveney and Martin, 1983).

Nicholls and Loring (1962) reported that some sulfides of nickel and vanadium could

also be precipitated, but according to Krauskopf (1956), sulfides of nickel and

vanadium are too soluble to be present in great concentration. Vanadium is more likely

to form metallo-organic complexes because its sulfide is unstable (Brumsack, 1989). 6

Where the sediment surface layer is aerobic, the iron and manganese difluse

from the underlying anaerobic sediment and precipitate as oxides in the aerobic surface

layer (Bemer, 1971). However, for very anaerobic sediments and where sulfur is

present, iron is locked in pyrite but manganese remains mobile and diffuses away

(Wilde etal., 1989).

Calcite and dolomite occur in minor amounts in Pennsylvanian gray shales and

in some black shales. Most calcite occurs as skeletal detritus, but calcite and dolomite

may also occur as cement in organic-rich shales (Trabelsi, 1990). Strontium and

manganese occur in amounts up to a few weight percent in aragonite and calcite.

Strontium, barium, lead, and uranium are preferentially incorporated into aragonite,

whereas magnesium, manganese, iron, nickel, and phosphorus tend to occur in calcite

(Cubitt, 1979; Norman and Deckker, 1990).

Apatite also occurs in minor amounts mostly as skeletal grains (conodonts, fish

debris) and pelloids and nodules (Siy, 1988). Phosphorus concentration in sediment

increases with water depth (Brongersma-Sanders, 1969). Due to high biological

productivity of overlying waters, the sediment has higher phosphorus, copper, and

oi;ganic carbon contents (Ingall and Cappellen, 1990; Toyoda and Masuda, 1990).

Increased biological productivity in overlying waters will increase the amount of

organic debris with their associated transition elements settling onto the sediment at the

bottom of the water column, but the organic matter and their transition elements must be

preserved in the sediment for them to accumulate (Brongersma-Sanders, 1969; Shaw et

al., 1990). DeMaison and Moore (1980) reported that organic matter accumulation in

sediments is not related to levels of marine biological productivity in overlying waters.

The organic matter could be lost by oxidation in aerobic to dysaerobic sediments and

their associated transition elements are mobilized and lost from the sediments by

diffusion. After deposition, transition elements may migrate from the minerals or 7

organic compounds with which they are associated with, that is, they may change their

host phases (hydrated oxides, organic matter, clays, or sulfides) while in the sediment

but they are still present in the proportion of their original abundances (Krauskopf,

1955; Coveney etal., 1987).

Pennsylvanian Cyclothems

Heckel (1979) discussed the evolving concepts of Pennsylvanian cyclothemic

sedimentation and concluded that the primary cause of the cyclothems can be attributed

to glacial-eustatic events. Although some aspects of his interpretation continue to be

disputed, the glacial-eustatic model is accepted by most workers and has been applied to

the cyclothems studied in this report. The following description and interpretation of

Pennsylvanian cyclothemic sequences (Figures 1 and 2) are based on the model of

Heckel (1977, 1980, 1989).

During the Pennsylvanian, the equator passed through the Midcontinent Region

of the United States creating a tropical to subtropical climate (Figure 3). Orogenic

events to the modem east (Allegheny Orogeny) and the south (Ouachita Orogeny)

supplied large quantities of terrigenous elastics to the broad shallow cratonic

Midcontinent Region. Recurrent episodes of glaciation in Gondwanaland (Crowell,

1978) created eustatic sea-level rises and falls at intervals of 100,000 to 400,00 years

(Heckel, 1989). These eustatic sea-level changes repeatedly flooded and exposed large

areas of the gently sloping Midcontinent region, giving rise to a series of alternations of

marine and nonmarine deposits called cyclothems, each representing a single eustatic

cycle.

During low stands of sea level, large areas of the craton were exposed and rivers

incised charmels into older shelf deposits (Brown, 1989), forming the unconformity that

separates adjacent cyclothems. When sea level started to rise, a rapid transgression 8

ensued that may be marked by a thin transgressive limestone deposited in deepening

water. This is typically a thin skeletal calcilutite deposited below effective wave base,

but locally includes a basal calcarenite deposited in shallower water early in the

transgression.

A geographically widespread offshore shale formed during maximum

transgression (Figure 4). This is the "core shale** of Heckel (1977), and has been

interpreted as a marine-condensed interval by Brown (1989). In many cyclothems the

water became deep enough for a thermocline to develop which inhibited the

replenishment of oxygen to the bottom. Organic-rich black to dark gray shale with few

or no benthic organisms accumulated in the resulting anaerobic to dysaerobic conditions.

Clastic sedimentation was extremely slow during the deposition of the core shale facies

due to impoundment of sediments nearshore after the rapid rise of sea level.

In areas distant from orogenic source regions, regressive limestones were

deposited in shallowing water (Figure 5). Regressive limestones are typically thick

marine skeletal calcilutites, the base of which was deposited below wave base. They

grade upward into skeletal calcarenite with abraded grains, algae, and cross-bedding,

evidence of traction transport in shallow water. The tops of the regressive limestones

often contain features indicative of peritidal deposition (algal laminations and fenestral

fabric) or diagenetic features formed by meteoric water.

The nearshore (**outside**) shales represent a variety of nearshore marine and

terrestrial deposits on the shelf, deposited at lower stands of sea level. Near to orogenic

source areas, thick sections of prodeltaic shale prograded over the regressive limestone,

in some cases preventing formation of the regressive limestone and resting directly on

the core shale (Boardman and Malinky, 1985). In many places, the prodeltaic deposits

grade upward into delta-front and delta-plain sandstones and coals. Within some

outside shales, paleosols have been identified (Schutter and Heckel, 1985; Goebel et al.,

1989).

This glacial-eustatic model of Pennsylvanian cyclothems explains the occurrence

of black, thin, widespread and extensive, nonsandy, shales that formed in starved,

anaerobic, deep-water settings and are underiain and overlain by offshore, fully marine

limestones (e.g., Heckel, 1977; Tourtelot, 1979; Brown, 1989). The model excludes

the less laterally extensive black shales that formed in shoreline environments, such as

embayments, lagoons, or marshes (Heckel, 1977). Other workers have proposed an

alternative model, wherein the laterally extensive Pennsylvanian black shales

accumulated in near-shore settings. The near-shore model for Mecca Quarry-type shales

infers rapid deposition of organic material, usually of terrestrial origin, in shallow water

as the epeiric seas transgressed rapidly, accumulating debris from coastal peat swamps

(2 angerl and Richardson, 1963; Coveney et al., 1989). Even in the more offshore

settings of Heebner-type shales, the presence of terrestrial organic matter at the base of

the core shale has been used to infer shallow-water deposition and rapid burial (Wenger

and Baker, 1986).

Previous Geochemical Studies of Pennsylvanian Shales

Papers by many workers on sedimentology and geochemistry of Pennsylvanian

cyclothems can be found in part three of the official reports on the Ninth Intemational

Congress on Carboniferous Stratigraphy and Geology edited by Belt and Macqueen

(1979).

Murray (1954) divided Pennsylvanian cyclothems of Indiana and Illinois into

marine, brackish-water, and nonmarine shales and sttidied their clay mineralogy using

X-ray diffraction, differential thermal analysis, and chemical analysis. Illite, kaolinite,

chlorite, and colloidal-size quartz are predominant but the amount varies considerably. 10

Illite content is high in marine shales. It is not possible to distinguish nonmarine,

brackish, and marine shales from the aluminum, iron, titanium, magnesium, sodium,

and potassium contents. Vanadium abundance is high where the organic carbon content

is high. Glass et al. (1956) correlated clay mineralogy between clays in Pennsylvanian

sandstones and clays in the interbedded shales. Sandstones and the mudstones from

different depositional basins have a similar amount of kaolinite and illite. The heavy

detrital minerals are mostly zircon, tourmaline, and mtile.

Degens et al. (1957) differentiated marine from freshwater Permsylvanian shales

by examining the trace element content together with clay mineralogy. Illite, kaolinite,

and chlorite are present in the clay mineral fraction of the shales. Marine shales have

more vanadium, nickel, and mbidium and less lead, zinc, and copper when compared to

freshwater shales. Nicholls and Loring (1962) reported on mineralogy and major and

trace element analyses of Carboniferous mudstones including coal seams in Britain.

Redox conditions and acidity could be inferred from the presence or absence of sulfides

and carbonates. Vine and Tourtelot (1970) examined major and minor elements

associated with detrital, carbonate , and organic fractions in Ordovician to Tertiary black

shale. The elements aluminum, titanium, zirconium, and scandium are associated with

the detrital fraction. Calcium, magnesium, manganese, and strontium are associated

with the carbonate fraction. Molybdenun, vanadium, zinc, nickel, chromium, and

copper are associated with the organic carbon fraction.

Cubitt (1979) reported that abundances of certain elements in Kansas Upper

Paleozoic shales are positively or negatively associated with quartz, potassium feldspar,

calcite, dolomite, illite, chlorite, and organic matter contents. The elements associated

with the carbonate fraction are calcium, magnesium, manganese, and strontium, and

those of the detrital fraction are silicon, aluminum, iron, and zirconium. The black

shales are enriched in vanadium, zinc, chromium, copper, and nickel. Cubitt and 11

Merriam (1979) found that the Pennsylvanian core shales, which are black shales, are

enriched in molybdenum, lead, chromium, copper, nickel, vanadium, and zinc due to

low redox potential of the original sediments.

Wenger and Baker (1986) described organic geochemistry of Pennsylvanian

cyclothems in Kansas and Oklahoma, They found that vanadium and nickel increased

with increasing anaerobic conditions. Oiganic carbon showed rapid increase

stratigraphically suggesting a coimection with initial rapid marine transgression resulting

in increased productivity and organic preservation in flooded coastal swamps. They

attributed the slow decline of organic carbon to deeper submergence of the coastal

swamps such that productivity declined slowly.

Schultz (1987) determined the mineralogy of Heeber-type Pennsylvanian shales

in Kansas. The more fossiliferous and silty shales have more carbonate minerals, but

the proportions of clay mineral contents are similar. Schultz (1989) distinguished

among aerobic, restricted, and inhospitable conditions by using the extent of pyritization

in Kansas black shales.

Coveney and co-workers recognized Heebner-type and Mecca-type

Pennsylvanian black shales on the basis of sedimentation rates, type of organic matter,

area of deposition, and elemental composition (Coveney and Martin, 1983; Coveney et

al., 1987, 1989; Coveney and Glascock, 1989). Heebner-type shale formed offshore in

deep waters under slow sedimentation and contains low concentrations of molybdenum,

vanadium, and uranium. Its organic matter is mainly marine. Phosphate, calcite, and

dolomite are abundant and the content of coal is low. Mecca-type shale formed

nearshore in shallow waters under rapid sedimentation and contains high concentrations

of molybdenum, vanadium, and uranium. Its oiganic matter is mostly of terrestrial

origin. The phosphate, calcite, dolomite, and kaolinite contents are high.

12

4 —

BMIC Cydoih«m (•implinMi nrMgacydothMn) In IUnii» low outcrop ball

UUwiogy

Cray lo graan, locally rad, sandy ahala, with aiiuiona, aparta tosaila

OapoaJilonal Envkonmaru

Naarthora

f

II M M

k I 2 M

OKahora

I *

111 I r I u £

J o s ill I 31

o a

I I i

Phaaa of dapoahioA

If II

—I Oalrhal Influx altar carbonala ahoal lormad

I —2 Dairltal influx balora ahoal conditioTM laachad

SI J;

I?

Lamlnalad unfotaililaroua bifdaaya calcUulita lo ootita

Locally croaabaddad skalatal calcaranita with marina bktia

Gray ahaly tlialatal calcUuilta wiilt aburtdant marina bloia

Gray-brown ahala with abur«danl la aparaa marina biota

5:

Black fiaaila ahala with phoaphala. paUgIc bloia

Dania, dark akalatal calcllulka with marina bioia local calcaranita al bata

Sandy ahala with marina biota

Gray le brown aandy ahala wkh local coal, aandatona

Figure 1. Basic vertical sequence of an individual Pennsylvanian cyclothem. Modified from Heckel (1977). In north-central Texas additional carbonate beds may be present in the outside shales, and deltaic sands may also occur in the outside shales (Ebanks et al., 1979).

13

A. Low Sea-level Stand (only small wind-driven vertical cells) WEST EAST

Open Ocean

Approx. depth (m) ^

0 =-

Position of periodic upwelling

Epicontinental Sea (HOOOs km in extent)

prevailing wind

7 TtMrmocliM ..... *v ,

2 0 0 - ^ '

carbonate + Iigtit-colored detritus

200 m

y^ tf" rir »'""<>"i V 2 i t v i i , I cf - \ j r PQ.-poor water Approximote

V^TpP* ^ position of i* V present Upper

y\4^ Pennsylvonlon jjF MId-Contlnent r^nami

^ (Konsos-Iowa) G«."?^°' , outcrop position of

B. High Sea-level Stand (large quosl-estuarlne cell) i upweiiing ««°, < < <- prevailing wind < < «

CoW.Ot-ooo^'*^* 200 —

blocK organic-rich fine detritus • phosp"^

< Prevailing winds

Oxygen-rich water

y Oxygen-poor water

Anoxic water

ocean currents

100

200

— 300

— 400

I I settling material _ 500 m

Figure 2. Cross-section showing low stand of sea level during regression and high stand of sea level during transgression in west-facing tropical epicontinental sea (Heckel, 1977).

14

present "X ..•••' . ..i"' \ •* i

<^ A \ A /COLO...'<.NEBRASKA1 \ \ ' A A': I ': •* ': .•:' UIKIM ^ .

IO»^-V '-/{

a * a * * a * *

V ^ A A . A A \ / ,•'••• \ \ YK"'""- X ^

;fA.>,TEXAJXTA' ta •:'.A

On^_6u°

km 500 • Appalachian Mtns.

Figure 3. Upper Pennsylvanian paleogeography of the Midcontinent Region of the United States showing position of paleoequator and areas of orogenic belts (Heckel, 1980).

15

MAXIMUM TRANSGRESSION •20'

./•••••••• > ^ w P 3 ^ a ^/• / . j

: B L A C K ^ 5 ^ S H A L E :

^ rupw«lllng :

Onl'l'ou^

km 900 S \ / A A A A A A A A A M A A A A A A A A A / ^

•• ••' / Appolachian Mtns. • v. / ,-• DLL I

Figure 4. Paleogeographic map showing probable facies relations of Upper Pennsylvanian Midcontinent sea during deposition of offshore shale along Midcontinent outcrop at phase of maximum transgression (Heckel, 1980).

16

20*

v LATE REGRESSION

<r . * • -

present A /• 1- I X

e»enL\ A - - A A/T-.--2>t/.:<>.Xpv'-'.i I \

}=rr'— ve_r —

;/«Poi,tf', \

TRADE

WIND

>r .A />^^ A>y:-.l- I i ' t , T - ^ T : " ^ J - ^ ' ' ' > ^ U ; - D :r f _o

r-. A^^:f.^,^Sio" ^LETAC^^.IT^^?y4^ \ i ^ - r r r BELT

/ "^AAAAnC" ./" t A v'l 'n.'i'^o''u° X*"

r t -•• • •• " A ' s C»» . . -

••ly

DOLDRUMS

I RAIN . ' , AA 0°

0 M 500

V'-'" '-i? K / D L L I

Figure 5. Paleogeographic map showing probable facies relations of Upper Pennsylvanian Midcontinent sea during deposition of upper part of regressive limestone, and locally base of nearshore shale along Midcontinent outcrop about midway through phase of late regression (Heckel, 1980).

17

CHAPTER n

LOCAL STRATIGRAPHY OF SECTIONS

Shale samples were collected from the different geographic areas shown on the

map in Figure 6.

North-Central Texas

The majority of samples were collected from four cyclothemic intervals in north-

central Texas, that range in age from latest Desmoinesian to earliest Virgilian (Figure 7).

The intervals sampled were chosen because each is well-exposed in relatively fresh

roadcuts, permitting detailed sampling with little risk of sample contamination. The

stratigraphy and paleontology of the Middle and Upper Pennsylvanian cyclothems in

north-central Texas have been discussed in numerous works, the recent ones are those

by Boardman et al. (1989a, 1989b).

The pattem of deposition that characterizes Middle and Upper Pennsylvanian

cyclothems in north-central Texas (Boardman and Malinky, 1985) differs slightly from

the typical Kansas-type sequence. A plexus of terrestrial deposits containing freshwater

or terrestrial fossils and representing overbank deposits, marsh deposits, and paleosols

occurs at the base of the cycle. The terrestrial strata are directly overlain by a variety of

shallow marine deposits, the most typical of which may be either a greenish-gray

calcareous shale containing an open marine filter-feeding benthic association, or a thin

limestone, less than 0.3 m thick, bearing a fauna similar to that of the open marine shale.

Over these beds lie either a black, fissile phosphatic, organic-rich shale, a dark

gray-black, pyritic, bioturbated clay shale, or a medium to dark gray bioUirbated shale.

This lithofacies contains a sparse fauna dominated by pelagic organisms with essentially

no benthic elements. Ammonoids, conodonts, Dunbarella bivalves, and conularids 18

characterize the fauna of the black fissile shale lithofacies, which is fully developed in

only a few north-central Texas cycles. This association represents an anaerobic

environment apparently developed by the encroachment of the oxygen-minimum zone

from the nearby Midland Basin into the broad, gently sloping shelf area (Boardman and

Malinky, 1985; Brown, 1989)(Figure 8).

The dark gray-black clay shales, which directly overlie or may occur in place of

the black fissile shale lithofacies, contain a fauna dominated by ammonoids, conodonts,

nuculoid bivalves, and archaeogastropods. This association is dominated by detrinis

feeders, scavengers, and camivores and was designated by Boardman et al. (1984) as

the Sinuitina Conmiunity. Most skeletal remains are preserved by pyrite or limonite

after pyrite. The members of the Sinuitina Community are of small size, suggesting

mass mortality among junveniles along with possible stunting due to lowered oxygen

levels associated with a dysaerobic environment (Figure 8).

The dysaerobic interval characterized by the Sinuitina Conmiunity is overlain by

medium to dark gray shales that contain a fully aerobic community having the same

basic composition and trophic structure as the Sinuitina Community. However,

members of this association, the Treptospira Community of Boardman et al. (1984) are

full-sized, and preserved by calcite (Figure 8).

The black to dark gray shales described above correspond to the "core shale*' of

the Kansas cyclothem as interpreted by Heckel (1977). These are the maximum

trangressive deposits of the Texas cyclothems which accumulated in anaerobic and

dysaerobic environments, permitting preservation of organic matter in offshore clay

shales.

Above the Texas core shale interval any of a variety of lithofacies may occur

(Figure 9). The most common situation is where the core shale is overlain by a medium

to light gray shale containing a mixed marine association that is adapted to higher rates

19

of clastic influx. This shale is overlain by gray silty to sandy carbonaceous shale which

is ahnost devoid of marine benthic fossils, the prodeltaic facies. Higher in the section a

sequence of distal deltaic, proximal deltaic, and distributary channel facies, as described

by Erxleben (1975) is present. In areas lacking active deltaic progradation, the core

shale is directly overlain by a light gray to brown offshore open marine shale containing

a sequence of communities dominated by filter-feeding organisms. This open marine

shale may be overlain by a thick carbonate, the upper portion of which bears abundant

phylloid algae. The highest deposits include a variety of marginal marine shales that

grade into terrestrial deposits (Boardman and Malinky, 1985).

East Mountain Section

The East Mountain section (Figure 10) lies on the south side of East Mountain,

within the city of Mineral Wells, Texas (UTM GRID 14SNM58341363096). It is the

type locality of East Mountain Shale Member of the Mineral Wells Formation as

designated by Plummer and Moore (1922). Three transgressive-regressive cycles of

sedimentation occur in the East Mountain section (Merrill et al.,1987; Boardman et al.,

1989d). The lower two cycles are latest Desmoinesian in age, whereas the uppermost

one is earliest Missourian (Boardman et al., 1989d). The middle cycle containing the

East Mountain black shale bed was sampled for this study (EM section, Figure 10).

The uppermost part of the underlying cycle (UNIT 1) is gray to black coaly

shale that contains abundant plant compressions and one species of agglutinated

foraminifera. This unit has been interpreted to represent a marginal swamp

envirormient. A thin, positionally transgressive limestone (UNIT 2) forms the base of

the second cycle. The limestone is a sandy, limonitic wackestone to packstone and

bears an abundant marine fauna of ostracodes, microgastropods, and crinoid and

echinoid debris.

20

The core of the cycle is the East Mountain black shale bed which is a slightly

phosphatic fissile black shale (UNIT 3). The black shale contains the dysaerobic

benthic association of the 5iDu/tiha Community of Boardman and Malinky (1985) and

Kammer et al. (1986), which here also includes ammonoids, bivalves, and trilobites.

The black shale is characterized by abundant conodonts (>1000/kg at the base) of the

Gondolella-Idioprionodus biofacies, which has been interpreted to be a product of a

more offshore, dysaerobic environment by several workers (e.g., Heckel and

Baesemann, 1975; Heckel, 1977,1980, 1986; Boardman and Malinky, 1985; Yancey

and McLerran, 1988). An ahemate depositional model for Pennsylvanian black shales,

including the East Mountain black shale, has been presented by Merrill et al. (1987) and

Merrill and Grayson (1989). These workers place the dark gray to black shales into an

"organic-rich** marginal marsh envirormient where extremely high organic productivity

encouraged the abundant Gondolella-Idioprionodus biofacies to develop.

The succeeding dark gray shale (UNIT 4) contains the r/eptosp/'/a-Ammonoid

Community of Boardman et al. (1984) and a conodont fauna similar in species

composition to that of the black shale, but less diverse and abundant. This unit

accumulated in an offshore, slightly more oxygenated environment than the black shale

(Boardman et al., 1989d). Gray, poorly fossiliferous shale (UNIT 5), representing

distal prodeltaic settings, appears higher in the section, and the regressive sequence is

capped by interdeltaic shale and marginal marsh deposits (not shown in Figure 10).

Samples from the East Mountain section (EM section, Figure 10) were selected

to investigate two problems. Four samples (EM-95 to 99) from the top of the

underlying regressive interval were taken to characterize the elemental composition of

beds interpreted by all authors to represent the marginal marsh environment. Data from

these samples could be compared with samples from the maximum transgressive beds of

the overlying cycle (EM-101 to 113), which have been interpreted either as marsh 21

deposits or far offshore deep marine deposits. The sequence of samples from the base

of the East Mountain black shale and higher (EM-101 to 127) pennits analysis of

geochemical changes as the shales become lighter in color, apparently as a result of

increasing oxygenation of the depositional environment.

Dog Bend Limestone (Lower Salesville Formation)

The Dog Bend Limestone (Figure 11) is the name that has been applied to the

limestone-shale-limestone sequence in the lower part of the Salesville Fomiation

(Plummer, 1929; Shelton, 1958). Boardman and Heckel (1989) interpreted this marine

horizon to represent the expression of the second Missourian eustatic event in the north-

central Texas section. The exposure (DB section, Figure 11) sampled for this sUidy is a

roadcut, 3.4 miles (about 5.4 kilometers) south of Palo Pinto, Texas, on the east side of

Texas Highway 4 (STOP 4B of Boardman et al., 1989a; UNM GRID:

14SNM56734362161).

Slightly fossiliferous, silty gray shale (UNIT 1; Figure 11) underlies the lower

limestone of the Dog Bend at this section. The lower limestone (UNIT 2) is a massively

bedded fossiHferous packstone to grainstone that includes intraclasts and ooids among

its grains. It is interpreted to represent the basal transgressive unit of the cycle.

The dark gray fossiliferous, slightly phosphatic shale (UNIT 3) that rests

directly on the lower limestone is the maximum transgressive unit. It contains an

abundant conodont fauna of the Idioprioniodus-Idiognathodus biofacies in addition to

ammonoids, ostracodes, forams, corals, chonetid brachiopods, gastropods and

bivalves. It is equivalent to the offshore Treptospira Community of Boardman et al.

(1984). It is overlain by a somewhat thicker interval of gray fossiliferous shale (UNIT

4) characterized by a more diverse filter-feeding marine association that includes corals,

productid brachiopods, bryozoans, and crinoids, in addition to ostracodes and forams. 22

The upper limestone (UNIT 5) is a massively bedded wackestone to packstone, bearing

abundant corals and phylloid algae. It is overlain by highly fossiliferous marine gray

shale.

A series of samples were taken from the shale between the Umestones at this

section (DB-101 to 113) to analyze the elemental changes in a regressive shale sequence

that is bounded by carbonates and which was deposited in oxygenated environments.

Upper Salesville Black Shale

The upper Salesville black shale unit represents the expression of the third

Missourian marine eustatic event in north-central Texas (Boardman and Heckel, 1989).

Two sections of the upper Salesville black shale (3027 section. Figure 12; and UPS

section, Figure 13) were sampled in order to ascertain the nature and magnitude of

geochemical differences between nearby sections representing the same eustatic event.

The first upper Salesville section (3027 section. Figure 12) is located on the west

side of Texas Highway 3027 about 2.6 miles (about 4.2 kilometers) northwest of

Mineral Wells, Texas (STOP 2 in Boardman et al., 1989a; UTM GRID

14SNM58194^63425). The base of the transgressive-regressive cycle is the Devil's

Hollow Sandstone of Cleaves (1975) and is designated as SS2 on the Geological Atlas

of Texas, Abilene Sheet. This calcareous sandstone (UNIT 1) is massively bedded,

highly bioturbated, and bears a few marine fossils at the top. An extremely thin (5 cm)

gray-green shale (UNIT 2) rests directly on the sandstone, which in turn is overlain by

black fissile shale that represents the core of the cycle. The base of this black shale is

sandy and contains a layer of small phosphate nodules (UNIT 3). Abundant conodonts

of the offshore Gondolella-Idioprioniodus biofacies are present, but few other fossils

occur. The black shale that occurs slightly higher (UNIT 4) is less phosphatic and

sandy, and ammonoids and the Treptospira megafaunal association occurs in addition to 23

conodonts of the Gondolella-Idioprioniodus biofacies. The black shale grades upward

into dark shale (UNIT 5), which contains a comparable megafaunal association, but less

abundant and diverse conodonts. The upper part of the section is a thick interval of gray

fossiliferous marine shale.

Samples were taken from the black shale interval above the Devil's Hollow

Sandstone, from the maximum transgressive anaerobic to dysaerobic core sh2ile into the

dysaerobic lithofacies. Evidence of chemical alteration attributed to weathering

(discoloration along fracttires) limited sampling to the lower part of the regressive shale

sequence.

The second upper Salesville section (UPS section. Figure 13) is exposed along

the west side of Texas Highway 337, about 2.6 miles (about 4.2 kilometers) west of the

3027 section (UTM GRID 14SNM57867363156; STOP 6 in Boardman et al., 1989a).

The Devil's Hollow Sandstone (UNIT 1) forms the base of the transgressive-regressive

cycle and is identical with that at the outcrop at section 3027. The overlying black

fissile, phosphatic shale is nearly identical with that at section 3027, both in lithologic

and faunal characters. Samples were taken from the lower black shale into the overlying

transitional dark gray, more fossiliferous shale. Discoloration along fractures, attributed

to recent weathering, limited sampling to only the lower part of the shale section.

Colony Creek Section at Brad

The Colony Creek Shale (Figure 14) forms the lower part of the Caddo Creek

Formation, the upper part comprising the Home Creek Limestone. The Colony Creek

overlies the Ranger Limestone at the top of the Brad Formation. The Colony Creek

Shale lies near the MissourianA irgilian boundary, depending on the level used to define

the boundary in the North American Midcontinent region. Following the suggestion of

Boardman, Barrick, and Heckel (1989c), who placed the base of the Virgilian at the 24

Little Pawnee Shale in Kansas, the Colony Creek is considered to represent the

expression of the first major Virgilian eustaric event in north-central Texas. The section

(CCB section. Figure 14) along US Highway 180, west of Brad, Texas, was sampled

in this smdy (UTM 14SNM54362362305; STOP 13 of Boardman et al., 1989a). This

section was chosen not only for its excellent exposure, but also because of the study on

oxygen and carbon isotopes in shell material from this section published by Adlis et al.

(1988).

The uppermost Ranger limestone (UNIT 2) rests on a thick interval of highly

fossiliferous gray marine shale with abundant brachipods and bryozoans (UNIT 1).

The uppermost Ranger limestone is a thickly bedded, highly fossiliferous crinoidal

packstone to grainstone that represents the transgressive interval of the Colony Creek

cycle. UNIT 3, overlying the limestone, is a dark gray, highly fossiliferous shale that

contains abundant Cmrithyris brachiopods and a moderately abundant and diverse

conodont fauna of the offshore Idioprioniodus-Idiognathodushiofacies. It is overlain

by a darker gray, slightly phospatic shale (UNIT 5) that is characterized by limonite

concretions and a fauna of ammonoids, Trepospira gastropods, and moderately

abundant conodonts. This association represents the aerobic Trepospira benthic

association of Boardman et al. (1984). The dark gray shale of UNIT 4 is overlain by a

thicker interval of brownish-gray, slightly silty shale (UNIT 5) with a molluscan-

dominated benthic assemblage. Locally, brachiopods, bryozoans, and crinoids are

common in this unit, which may have formed near to a prograding delta lobe. Sharply

overlying UNIT 5 is a blackish, carbonaceous shale with common plant compressions,

probably representing a type of interdeltaic marsh. Dark gray highly fossiliferous shale

with a molluscan-dominated assemblage (UNIT 8) overlies the carbonaceous shale.

Adlis et al. (1988) studied the stratigraphic variation in carbon and oxygen

isotopes preserved in brachiopod shales at this section of the Colony Creek Shale.

25

These authors recorded a decrease in the delta ^^C values from 2.9-3.6 per thousand in

the lower 2 meters of the shale (Unit 3 and the lower part of UNIT 4) to 2.7-2.9 per

thousand in the 3 to 7 m interval (upper part of UNIT 4 into the middle of UNIT 5).

The delta ISQ values also show a shift from about -2.7 per thousand to about -3.0 per

thousand at the 3 meter mark. Although the shifts in delta l^o values were relatively

sli^t, Aldis et al. (1988) attributed the change to wanner water temperamres resulting

from a shallower water setting in the upper part of the section. By comparison with

modem analogues, an approximate maximum depth of 70 m was estimated for the core

of the cycle.

The Colony Creek section at Brad (CCB section, Figure 14) was sampled to

concentrate on three aspects. Detailed sampling at the base of the core shale section was

to determine elemental changes from maximum transgressive dysaerobic to areobic

settings. Samples in the interval of UNIT 3 through UNIT 5 parallel the isotopic

sampling of Aldis et al. (1988) to see if any elemental changes coincide with the isotope

stratigraphy. The blackish coaly shale of UNIT 6 was sampled to provide geochemical

information on organic-rich marsh deposits.

Kansas and Oklahoma

Two sections (LBK section. Figure 15; and TRR section. Figure 16) of

Permsylvanian shales in southeastern Kansas and northeastem Oklahoma were collected

to obtain samples from highly fissile, organic-rich, black shale typical of cyclothems in

the northern Midcontinent region.

Type Lost Branch in Kansas

The Lost Branch Shale (Figure 15) is the name proposed by Heckel (1986;

1991, in press) for strata in the upper part of the Holdenville Formation in southeastem 26

Kansas and northeastem Oklahoma. The Lost Branch Shale overlies the Lenapah

Limestone and lies beneath the Hepler Sandstone, which forms the base of the

Missourian Pleasanton Group. At its type section along the Lost Branch of Pumpkin

Creek in southeastem Kansas ( NWl/4, sec. 10, T. 33 S., R. 18 E.), the Lost Branch

Shale includes the Dawson Coal (UNIT 1) and its underclay near the base. Fifteen

centimeters of dark gray shale (UNIT 2) bearing only a few invertebrate fossils overlie

the coal. The base oftheNuyaka Creek black shale bed (UNTT 3; Bennison, 1984)

rests with a knife-sharp contact on this gray shale. The Nuyaka Creek black shale

comprises 45 cm of black, highly fissile, conodont-rich, phosphatic shale that represents

the core shale of the highest Desmoinesian eustatic cycle. Nearly 4 m of gray shale with

abundant and diverse marine invertebrates (UNIT 4) rests with a sharp contact on the

Nuyaka Creek black shale. The Lost Branch Shale is interpreted to be equivalent to the

East Mountain black shale in north-central Texas (Boardman and Heckel, 1989).

The Nuyaka Creek black shale was sampled (LBK section. Figure 15) because it

is an excellent example of a highly fissile, organic-rich, black shale (Mecca Quarry-type

shale) that overlies a coal. The presence of sharp lithologic contacts at the base and top

of the black shale permits examination of the degree of mobility of elements out of black

shales into adjacent beds.

Lower Tackett Section at Tulsa Railroad

The Tulsa Railroad section (Figure 16) is an outcrop exposing a succession of

gray and black shales that overlie the Checkerboard Limestone in Tulsa, Oklahoma,

(E/2, sec. 22, T. 19 N., R. 12 E.). The stratigraphy of the Tackett shales that overlies

the Checkerboard Limestone (UNIT 1, Figure 16) is unusual because the core shales of

two cyclothems are superimposed and the intervening limestones and shales appear to be

absent. The lower black shale, the Lower Tackett black shale, is correlated with the

27

Mound City Shale in Kansas, and the upper black shale, the Upper Tackett black shale,

is correlated with Hushpuckney Shale of Kansas, the second and third Missourian

eustatic cycles according to Boardman and Heckel (1989). This atypical succession and

similar sections in the area of Tulsa may be due to the presence of a local basinal area

during the early Missourian, distant from sources of terrigenous elastics, which was not

completely exposed, even during low stands of sea level.

Although the Upper Tackett black shale is strongly weathered at the Tulsa

Railroad section (UNIT 6), the Lower Tackett black shale is freshly exposed and shows

litde sign of weathering. The blocky dark gray shale immediately below the Lower

Tackett black shale (UNIT 3) is apparently unfossiliferous and is separated from the

Checkerboard Limestone by a thin shale that may represent an underclay (UNIT 2). The

Lower Tackett black shale is a platy to fissile black shale that contains no calcareous

fossils and is phosphatic in only the upper 30 cm. It rests with a sharp contact on the

gray shale of UNIT 3, and is overlain, perhaps unconformably, by a thin interval of

calcareous gray shale (UNIT 5).

Like the Nuyaka Creek black shale, the Lower Tackett shale at this section was

sampled (TRR section, Figure 16) to determine the geochemical characteristics of a

highly organic, northern Midcontinent, black shale.

Permian Bead Mountain Limestone

A series of seven samples were taken from an organic-rich shale interval m a

section of the Lower Permian Bead Mountain Limestone west of Albany, Texas (PI 80

section. Figure 17). The Bead Mountain Limestone, part of the Wichita-Albany Group,

consists of a series of alternating limestone and marly shales that attain an average

thickness of 75 feet in Shackelford County and adjacent areas. It is shown to be early

Leonardian in age, as part of the Belle Plains Formation, on most correlation charts 28

(e.g., Dunbar, 1960). Although detailed paleontologic and petrographic studies

apparently have not been published, most authors consider the Bead Mountain, like

other thin carbonate bodies in the Wichita-Albany Group, to represent extremely

shallow-water, peritidal, environments. The dark organic-rich shales present in these

limestones have been reported as lignites by some workers (S.A.S.G.S. Guidebook,

1963, p. 63-64).

The outcrop (P180 section, Figure 17) sampled for this sttidy, 4.5 miles (7.2

kilometers) west of Albany, Texas, in the upper part of the Bead Mountain Limestone,

is exposed along US Highway 180, (N32*' 42' 30", W99'' 22' 25", Shackelford County,

Texas). Just below the series of thickly bedded limestones that top the exposure is a 40

cm dark shale interval forming a reentrant sandwiched between gray fossiliferous

limestones (UNITS 1 and 5). The basal 7 cm of the shale is black in color (Unit 2) and

is succeeded by 20 cm of gray to brown shale (Unit 3). At the base of the gray shale is

a thin (1-2 cm) light brown oxidized zone. At the top of the shale interval lies 10 cm of

black shale (UNIT 4). The shale in each of the three units appears to differ only in

color. The shales are thinly bedded, but are more cmmbly than fissile when extracted

from the outcrop. No skeletal remains of any organisms were observed in the shale.

The dark shale in the Bead Mountain Limestone was included in the smdy for

two reasons. First, it represents a thin organic shale included wholly within a

carbonate-dominated sequence, unlike any of the Pennsylvanian shales in this smdy.

Second, all evidence suggests that the shale formed in a marginal marine situation where

there should be no influence of low salinity waters.

29

Figure 6. Distribution of Pennsylvanian strata in the Midcontinent Region of the United States (Cocke, Boardman, and Mapes, 1989). Samples were collected from the areas shown as dots.

30

MIDCONTINENT SOUTH

TRANSGRESSION *• Bosin I Low Shelf Mid Shelf High Shelf

ROulcrep Limit

NORTH

D«tPttl-i»ol>f toeUt ol llmll TEXAS

TRANSGRESSION

•toMM-• HI*

- j m a - ' t ' ^Smm^f^^yhrTrrfFrn

Figure 7. Eustatic sea-level curve for part of Pennsylvanian sequence in north-centrkl Texas outcrop (right) and biostratigraphic correlation with curve for Midcontinent outcrop (left, derived from Heckel, 1986, as modified by Heckel, in press, with new data on conodonts from J. E. Barrick). From Boardman and Heckel (1989).

31

bO

S

8-

s Si o

! - < C: ON O OO • ON

o_r

32

o

.O • -

33

EAST MOUNTAIN (EM)

4 m

3 -

2 -

1 -

0 m

O

a: O Li.

IS

iS

East Mountain black shale bed

127

- 125

-123

-121

- 1 1 9

-117

gray shale

dark gray shale

113

109

105 103 101

black to dark gray fissile shale;

slightly phosphatic

y fossil, wackestone i B B B B B B B a B B B B B a a a a

99 98 97 96 95

gray clay shale with abundant plant detritus

UNITS

UNIT 4

UNIT 3

UNIT 2

UNIT1

Figure 10. Stratigraphic profile of East Mountain (EM) section. Numbers adjacent to the stratigraphic column designate samples analyzed.

34

DOG BEND LIMESTONE (DB) (LOWER SALESVILLE)

3 A m

2 -

1 -

0 m

fe cc O LJ-

UJ

UJ —I < CO

DOG BEND LIMESTONE

rs T * T

rri i±r

x^

I I £ ^

r ^

O i=z iS^ i S I I I o

I ' I ' ' r*T

r i - r i . X I ja

light gray shale

gray fossiliferous limestone

- 1 1 3

UNITS

UNITS

I . I . I ,

O O

O d III

cc

cx

111 gray shale,

with calcareous fossils

•109

• 107

-105 -103 .101

dark gray shale

light gray shale

UNIT 4

UNIT 3

UNIT 2

UNIT1

Figure 11. Stratigraphic profile of Dog Bend (DB) section (Lower Salesville). Numbers adjacent to the stratigraphic column designate samples analyzed.

35

3 m

2 -

1 -

0 m

Z o I-< a: O Li_ UJ

> CO UJ . J < CO

1 .

UPPER SALESVILLE SHALE (3027)

upper Salesville

black shale

dark gray shale

35

black shale

h 3 1

27

DEVIL'S HOtUDW

SANDSTONE

^—^^ 1 gray-green shale

••^••.•y.-:-:-:<-

23

UNITS

UNIT 4

19

15 phosphatic black shale

highly bioturbated sandstone

UNIT 3

UNIT 2

UNIT1

Figure 12. Stratigraphic profile of 3027 section (Upper Salesville). Numbers adjacent to the stratigraphic column designate samples analyzed.

36

UPPER SALESVILLE SHALE (UPS)

3 m

2 -

1 _

0 m

O fe CC

s UJ

UJ _ l < CO

upper Salesville

black shale

DEVIL'S HOLLOW

SANDSTONE • • • • • " • ' • • • " • • • • "

- 3 5

1-31

27

23

19

15

1 1

gray shale

dark gray shale with brown

alteration streaks

phosphatic black shale

highly bioturbated sandstone

UNIT 4

UNIT 3

UNIT 2

UNIT1

Figure 13. Stratigraphic profile ofUPS section (Upper Salesville). Numbers adjacent to the stratigraphic column designate samples analyzed.

37

COLONY CREEK SHALE (CCB)

12 m

1 0 -

8 -

6 -

4 -

0 m

fe CC

UJ UJ cc o 8 9 O

QQ

COLJONY CREEK SHALE

uppermost Ranger

J 27 *lark gray shale

-122 1-121

T25^26 ^T7T^124 black coaly shale

•119

•117

brownish-gray shale with marine

calcareous fossils

• 115

.113

•111

•109

•107 •105

gray-black shale slightly phosphatic

TT5T -102 .101

dark gray shale

gray packstone to wackestone

gray shale numerous calcareous

marine fossils

UNIT 7

UNITS

UNITS

UNIT 4

UNIT 3

UNIT 2

UNIT1

Figure 14. Stratigraphic profile of Colony Creek (CCB) section at Brad. Numbers adjacent to the stratigraphic column designate samples analyzed.

38

TYPE LOST BRANCH (LBK) KANSAS

Figure 15. Stratigraphic profile of Type Lost Branch (LBK) section in Kansas. Numbers adjacent to the stratigraphic column designate samples analyzed.

39

LOWER TACKETT SHALE (TRR) OKLAHOMA

4 m

3 -

2 -

1 m

UPPER TACKEFT BLACK SHALE

LOWER TACKETT BLACK SHALE

CHECKERBOARD LIMESTONE

black fissile shale

calcareous gray shale

39 37 3 5 33 29

.25

black fissile shale

gray shale

underclay?

fossiliferous gray limestone

UNITS

UNITS

UNIT 4

UNIT 3

UNIT 2

UNIT1

Figure 16. Stratigraphic profile of Lower Tackett Shale (TRR) at Tulsa Railroad Yard in Oklahoma. Numbers adjacent to the stratigraphic column designate samples analyzed.

40

P 180 0.6-m

0 .5 -

0.4 -

0.3 -

0.2 -

0.1 -

0.0-

cc O >-z

<

i o

cr

2 z

i IS CD

oc X^

T * T

X ^ J C

1

O

^

O S

r'=P X I X I

1.1 ,1

- 8

C: ^ ? S'

fossiliferous gray limestone

•10

9

black shale

gray shale, with yellow-brown stains

black shale

fossiliferous gray limestone

UNITS

UNIT 4

UNIT 3

UNIT 2

UNIT1

Figure 17. Stratigraphic profile of Permian PI 80 section (Bead Mountain Limestone). Numbers adjacent to the stratigraphic column designate samples analyzed.

41

CHAPTER m

GEOCHEMICAL ANALYSIS

Methods and Analysis

Sample Collection and Preparation

The rock samples were collected at intervals of less than 10 cm from fresh

outcrops. Plastic tools were used to separate the samples from the outcrop to avoid

metallic contamination. The collected samples were sealed in plastic bags for transport

to the laboratory. The rock samples were broken into smaller pieces using ceramic

mortar and pestle, and ground to about 200 mesh (75 micron) powder in a ceramic

shatter box. The powder form was used for chemical and X-ray analysis.

Mineralogy and Petrology

X-ray diffraction analysis of clay minerals was carried out with a Phillips

diffractometer using copper K-alpha radiation. The scanning speed and range were V

two theta per minute and 2°-65*', respectively. X-ray diffraction analysis showed that

quartz, kaolinite, and illite were present in all the samples; in addition, samples LBK-8,

10, 15, and TRR-25, 40 also contained pyrite. Amorphous substances were present in

the coal sample LBK-8. On the whole, there was insignificant variations in the relative

amounts of quartz, kaolinite, and illite in all the samples.

Thin sections of some of the rocks were made and examined under the

petrographic microscope. The shales are silty, some with quartz grains ranging from 10

to 30 microns, and others from 20 to 50 microns. Mica flakes are present in both gray

shales and black shales. The mica flakes are small, short, pieces, few in number, and

arc scattered. They are not observed in some samples probably because they are too

small and too widely disbursed. There are many dark brown aggregates of 42

clay + organics + oxides occurring as irregular spherical or elongated particles. These

aggregates of clay + organics + oxides are present in all the samples. They are most

likely composed of mixtures of illite + kaolinite, organic matter, and iron + manganese

oxides. Calcitic fossils occcur in most ofthe samples but vary in abundance. The

abundance of calcitic fossils varies inversely with the size ofthe silt fraction in the

samples. Grayish clay aggregates are made up of oval to spherical masses. These gray

aggregates are distributed irregularly. They are probably composed of kaolinite + illite

clay; kaolinite probably predominates because low birefringence was observed. Pyrite

is seen only in the black shale samples from sections TRR and LBK. Further details of

petrography are given in Appendix A.

Total Oiganic Carbon (TOC)

Various methods of total organic carbon (TOC) detennination are given by

Jeffrey and Hutchison (1981) and Johnson and Maxwell (1981). Comparisons of

methods to determine TOC are discussed by Leventhal and Shaw (1980). Three

different methods were used to measure TOC in this study.

Total oiganic carbon (TOC) can be determined by loss on ignition (LOI) at a

selected temperature. Details are given by Dean, 1974, (LOI at 550X); Ball, 1964,

(LOI at 375°C and at 850*'C); and Keeling, 1962, (LOI at 375°C). For this work, the

method of Dean (1974) was followed. Four grams ofthe sample were dried at 100°C

for one hour and the resulting weight loss was taken as the moisture content (Appendix

B). The samples dried at 100°C were subsequently heated at 550°C for one hour. The

loss on ignition at this temperature was converted to total organic carbon using Dean's

(1974) graph (Appendix B).

Gaudette et al. (1974) and Prince (1955) give details of wet combustion

measurement of TOC. For this work, the wet combustion method of Prince (1955) 43

was followed. The oiganic carbon in the sample was oxidized by acidified potassium

dichromate and the remaining unused dichromate was determined by back titration with

ferrous ammonium sulfate. The TOC was calculated from the amount of dichromate

needed to oxidize it (Appendix B).

Determination of organic carbon by infrared spectral analysis was done by Arco

Oil and Gas Company using a Leco carbon analyzer.

For the low organic carbon samples, TOC below 5 weight percent, the titration

method gives TOC values about 1.5 weight percent lower than that given by LOI

method (Figures 48 and 49 in Appendix C). For the high organic carbon samples, TOC

above 5 weight percent, the titration method gives higher TOC values than the LOI

method (Figure 50 in Appendix C). The higher the TOC, the larger is the difference in

TOC between the two methods. The TOC content variations are rcflected by both

methods for these low and high organic carbon shales. Only in samples from PI80

section do the titration and LOI methods yield a few TOC values that have opposite

trends, that is, TOC value is lower by the titration method but higher by the LOI

method.

The LECO method gives TOC values comparable to those ofthe titration method

(Figure 18). For TOC values of less than 2 weight percent, the results ofthe two

methods differ by about 0.2 weight percent; for TOC values of more than 5 weight

percent, the difference is between 1 to 4 weight percent. The higher the TOC value, the

larger the difference. The values of TOC used to calculate TOC/Al ratios in this sttidy

are those obtained by titration. The TOC weight percent of LBK-18 obtained by titration

is unusually low comparcd to that obtained by LECO analysis. Consequently, the TOC

weight percent of LBK-18 was adjusted upward according to the value expected on the

basis of LECO analysis.

44

Color and Total Organic Caibon

Chroma, hue, and value are determined by comparing a fresh, dry whole rock

sample with the Munsell color chart (Appendix B). The graph of Munsell value number

N versus TOC weight percent is shown in Figure 19. The lower the value number N ,

the darker is the sample.

At low color value, the TOC varies widely. Therefore, one cannot visually

judge oiganic caibon content for black shales when the value number N is low. For the

gray shales in this smdy, the TOC varies within narrow limits and an estimation of TOC

can be made by visual inspection ofthe degree of blackness (the value number N). In

dark-colored marine shales, the degree of blackness maybe due more to the size and

distribution ofthe organic debris than to the actual concentration ofthe organic carbon

(Degens etal., 1957).

Discussions of color of shales are given by Myrow (1990); Blatt et al. (1980);

Potter et al. (1980); and Twenhofel (1939). The degree of blackness of gray to black

shales with TOC values below 5 weight percent is related to the TOC content according

to Potter et al. (1980) and Myrow (1990). However, the relationship is not strong

according to Blatt et al. (1980) due to the presence of dark-colored minerals like iron

sulfide, and the nature ofthe organic matter.

Ferrous Iron

Ferrous iron determination methods are given by Jeffrey and Hutchison (1981);

John and Maxwell (1981); Von Amd (1968); and Nicholls (1960). In this smdy, the

ferrous iron analysis was done according to the method of Von Amd (1968), The

ferrous ions in the sample were oxidized by acidified ammonium metavanadate ions and

the excess vanadate was determined by back titration with ferrous ammonium sulfate.

The concentration of ferrous ions was calculated from the amount of vanadate needed to 45

oxidize them to ferric ions (Appendix B). For shales with high TOC, the titration end-

point tends to be masked by the blackness ofthe suspension. The method of Nicholls

(1960) is able to overcome this difficulty by extracting the indicator into an organic

solvent. All titration methods use an oxidizing agent, and any sulfur compounds present

would also be oxidized giving higher ferrous iron values. An additional analytical

uncertainty is that the ferrous iron could be oxidized by the air during processing for

titration with resultant lower ferrous iron values. In Appendix B some parts in the

ferrous iron column were left blank because negative values were obtained when weight

percent FeO was subtracted from weight percent total iron (Fe203 + FeO). Most

probably, some organic carbon, in addition to any sulfur present, was oxidized by the

reagent giving false higher value of FeO.

Sulfur

Details of sulfur analysis are given by Canfield et al. (1986); Jeffrey and

Hutchison (1981); and Johnson and Maxwell (1981). Infrared spectral analysis for

sulfur was done for some ofthe samples in this study by Arco Gas and Oil Company

using a LECO sulfur analyzer.

Major and Trace Elements

Inductive coupled plasma (ICP) emission spectroscopy was used to determine

the abundance of selected elements. The ICP equipment used was Leeman model

plasma-spec 40. Rubidium was analyzed by Perkin-Ebner atomic absorption

spectroscope model 3030.

A 0.2 gram powdered sample was mixed with 1.2 grams of sodium metaborate

flux. The mixture was fused in a carbon crucible at 1000°C for 20 minutes. The molten

rock and flux were dissolved in 50 milliliters 5 percent HCl. The stirring time for 46

dissolution was 20 minutes. This solution was analyzed for trace elements. Twenty

milHliters ofthe original solution was mixed with 50 milliliters of 5 percent HCl and the

resulting dilution analyzed for major elements Results ofthe analyses are Hsted in

Appendix B. As the major elements of LBK-18 totalled above 150 percent probably due

to a dilution error, the major and minor element data of this sample were adjusted so that

the total percent ofthe major elements equal to the average ofthe total percent ofthe

major elements of LBK 13, 15, 16,17,and 19 (excluding LOI and H2O).

Presence of components other than the one being analyzed may affect the

analysis of the sample. Thisiscalledthematrixeffect (Hume, 1973). Sample CCB-

109 was used as an internal standard and given the code "BMS". The USGS internal

standards used were 1633A, SGR, and SCO (Seward, 1986; Johnson and Maxwell,

1981; and Flanagan, 1976).

Experimental uncertainty is calculated by dividing the sample standard deviation

ofthe intemal standard (BMS) by the mean and expressing it as a percentage. The

uncertainties for the data obtained by inductive coupled plasma spectroscopy are as

follows: < ±3% (Si02, Ti02, AI2O3, Zr, total Fe203); < ±6% (MgO, CaO, K2O, Sc,

V, Be, Ba); < +9% (Na20, P2O5, Cr, Zn, Sr); < ±14% (MnO, Co, Ni, Cu, Y).

Presentation of Data

In this study, the elemental concentrations were first converted to atomic

concentrations and then divided by atomic concentration of aluminum to give the element

atomic ratio with respect to aluminum. Aluminum is used as a conserved element

(Salomons and Forstner, 1984). Aluminum is chosen because the increase or decrease

of clay content is reflected by the increase or decrease of alummum. This normalization

with aluminum will offset the effect caused by different rates of sedimentation of clay

minerals. Comparison of shale samples using this aluminum-normalized ratio should 47

display enrichment, depletion, or no change, in elemental contents between samples

relative to clay content.

Aluminum, titanium, and scandium have been used as conserved elements for

normalization of data for comparison. For study of carbonate components of sediments,

normalization with strontium or calcium has been used. Shaw et al. (1990) used

element/titanium ratios for smdying elemental distribution with depth of sediment.

Chakrapani and Subramanian (1990) used element/aluminum ratios to sttidy

downstream changes of elemental contents in river bed sediments. Norman and De

Deckker (1990) used element/strontium ratios to show physical mixing of detrital clay

and biogenic caibonate components. To avoid interference by carbonate enrichment,

they used element scandium ratios to smdy stratigraphic distributions of detrital

elements.

Distribution of Elements in Stratigraphic Sections

East Mountain Section (EM)

The dark gray to black shales at East Mountain contain relatively little total

oiganic carbon (TOC). All samples contain less than 2 weight percent TOC; the TOC/Al

ratio is less than 0.5. Three samples in UNIT 3 (EM-103, 105, 109), the maximum

transgressive black shale, show a TOC/Al ratio of near 0.4, twice that of most other

samples. Sulfur values are extremely low for the four samples analyzed from UNIT 3,

ranging from 0.05 to 0.25 weight percent. The marginal marine samples (EM-95 to 99)

have a slightly higher water content than the noimal marine samples.

The most striking aspect of elemental distribution in the East Mountain section is

the sharp increase in abundance of some transition metals at the base of UNIT 3, the

maximium transgressive deposit, followed by a decline in abundance higher in the

section (Figure 20). The Zn/Al xlO^ ratio rises from about 4 in the underiying marginal 48

marine beds (EM.95 to 99) to near 14 in samples EM-101 and 103, then falls gradually

to 5 in the stratigraphically highest sample, EM-127. The Munsell value number N

reflects the sedimentation process, which is picked up by the decreasing zinc content

rather than reflecting the oiganic carbon content (Figure 21). Nickel displays a pattem

similar to that of zinc, rising from a Ni/Al xlO^ ratio of 1 to a maximum near 7, before

erratically falling to a value between 3 and 4. Chromium, cobalt, manganese,

magnesium, and copper show a similar pattem of abundance. Vanadium and beryllium

mcrease in abundance at the base of UNIT 3, but do not show any pattem in

stratigraphically higher samples. Iron (total) does not show any regular stratigraphic

change in abundance. However, the Fe2"'"/(Fe2++Fe3+) ratio is less than 0.2 in the

maiginal marine shale, then jumps to over 0.6 in UNIT 3 (EM-105 to 109).

Barium is slightly more abundant in the marginal marine shales, than in the dark

normal marine shales. Calcium, magnesium, phosphorus, and yttrium show a peak in

abundance at EM-103, but otherwise display little change. Sodium and strontium have a

higher and more erratic abundance in samples EM-95 to 103, than in the overlying

samples.

Dog Bend Section (DB)

All samples from the shale (UNITS 3 and 4) between the two limestone beds of

the Dog Bend contain less than 0.5 weight percent TOC, and have a TOC/Al ratio less

than 0.2. A number of elements display a similar pattem of abundance (Figure 22).

The lowermost sample, DB-101, just above the lower limestone, and the highest

sample, DB-113, just below the upper limestone, contain the greatest amount of

calcium, barium, strontium, and manganese which are about twice as much as in the

middle ofthe the shale. Nickel, zinc, magnesium, and yttrium show a small decrease in

49

the middle ofthe section when compared to the top and base ofthe section. Only the

lowermost sample contains more iron (total), chromium, and phosphonis.

Upper Salesville Sections (UPS and 3027)

The quantity of TOC in samples from UPS and 3027 is below one weight

percent. 3027-11 has the lowest value at 0.1 weight percent, and UPS-15 the highest at

one weight percent. Sulfur analyses were available for only section 3027, and show a

decline from 0.06 weight percent in sample 3027-13, to 0.03 weight percent in sample

3027-23.

The samples from the two sections ofthe Upper Salesville black shale start at the

base ofthe marine shale that overlies a transgressive sandstone. The abundance of some

transition elements in UPS tend to decline steeply from a maximum value at the base

(Figure 23). Chromium, cobalt, nickel, and copper in UPS section fall to less than one-

half the maximum value by the top of UNIT 2. Zinc and magnesium show smaller

declines in content. Sodium is more abundant in the lower three samples than in the

higher samples. A pronounced peak in the distribution of several elements occurs in

sample UPS-23: calcium, strontium, phosphoms, yttrium, manganese, and zinc.

In section 3027, the basal thin green shale (sample 3027-11 in UNIT 2) differs

from the overlying dark shales in UNIT 3 by containing a greater abundance of silicon,

zirconium, calcium, barium, strontium, manganese, copper, vanadium, phosphorus,

and yttrium. The Fe2V(Fe2++Fe3+) ratio of 0.9 for 3027-11 is the highest ofthe

Upper Salesville samples, being about twice the next highest.

Zinc, chromium, cobalt, nickel, and iron (total) are more abundant in UNIT 3,

the lower black shale, than in UNIT 4 (Figure 24). The Fe2+/(Fe2++Fe3+) ratio rises

from about 0.2 in UNIT 3 to about 0.4 in UNIT 4.

50

Copper shows a Cu/Al xlO^ ratio jump from 0.5 to 5.6 in going from sample

3027-19 at the top of UNIT 3 to sample 3027-23 in the base of UNIT 4. Copper also

has an extremely high abundance peak in the highest level sampled, 3027-38, near the

top of UNIT 4. No other elements show comparable variations at these stratigraphic

levels.

Calcium, magnesium, strontium, and manganese show a strong peak indicating

a maximum abundance at 3027-35, whereas phosphoms, and yttrium show a weaker

peak.

Colony Creek Section at Brad (CCB)

Total organic carbon is less than 0.5 weight percent in samples from UNITS 3,

4, 5, and 7. The black coaly shale of UNIT 6 contains between 2 and 4 weight percent

TOC. The TOC/Al ratio remains near a value of 0.05 through UNITS 3,4 and 5, and in

UNIT 6 it varies from 0.6 to 1.1. Ofthe five samples analyzed for sulfur from UNITS

3 and 4 (CCB-102 to 109), most contained around 0.03 weight percent sulfur, but

sample CCB-103 peaks at 0.14 weight percent.

Silicon, titanium, zirconium, and sodium, and less so nickel, are more abundant

in the coaly shale of UNIT 6. Zinc, cobalt, beryllium, and yttrium show a peak at CCB-

126 at the top of UNIT 6, but otherwise show dissimilar distribution patterns for the rest

ofthe section. Potassium, mbidium, vanadium, iron (total), and magnesium show

similar pattem, lower concentration in UNIT 6 than in the underiying shales.

Unlike the black shale sequences at East Mountain and the two upper Salesville

sections, transition metals in Colony Creek shales do not show a pattem of high

abundance at the base and a decrease higher in the section. From UNIT 3 to UNIT 5

calcium, manganese, and strontium have a similar distribution pattem, that is, the

abundances are high at the base of UNIT 3, CCB-101, low through the rest of UNITS 3 51

and 4, and higher in UNIT 5 (Figure 25). In UNIT 6, these three elements fall to the

level of UNIT 4. Zinc also is somewhat more common in UNIT 5, than in other parts

ofthe section. At the base of UNIT 5 (CCB-113), sihcon, titanium, zirconium, and

sodium have a peak in abundance.

Type Lost Branch Section in Kansas (LBK)

The TOC in UNITS 2 and 4 are about 1 or 2 weight percent. From the base of

UNIT 3 (LBK-11) to the midddle (LBK-15 and 16), the TOC increases in abundance

from 5 to 19 weight percent, and then decreases from the middle of UNIT 3 to the top

(LBK-19) from 19 to 11 weight percent. Sulfur shows a similar increase and decrease,

with a maximum value of 4.3 weight percent in sample LBK-17 to lows of 2 weight

percent in sample LBK-11 at the base of UNIT 3, and 1 weight percent in sample LBK-

19, at the top of UNIT 3.

The S/Al ratio correlates positively with the TOOAl ratio and the trend of ratios

of LBK-13, 15, 17, and 19 passes through the x-axis at TOC/Al ratio of about 2 (Figure

26). The S/Al ratio correlates with the total Fe/Al ratio (Figure 26) and although the

correlation line ofthe two ratios does not pass through the origin it is parallel to the

stoichiometric FeS2 line. The stoichiometric FeS2 line is based on the theoretical ratio

in pyrite of two sulfur atoms to one iron atom.

From the base of black fissile shale UNIT 3, several elements increase steeply at

LBK-13, attain their greatest abundance in the upper part of UNIT 3, but decline at the

top ofthe UNIT (Figure 27). All these elements are significantly less in the gray shales

of UNIT 2 below, and UNIT 4 above.

Vanadium, zinc, copper, and to a lesser degree calcium, have prominent

abundance maxima in samples LBK-15 and 17, separated by a lower content in sample

LBK-16. From the base of UNIT 3, vanadium has nearly a tenfold increase in the 52

V/Al xlO^ ratio to a maxiumum of 120 in LBK-17, and copper has a threefold increase

to a Cu/Al xlO^ ratio maximum near 8 in LBK-17. Zinc shows an ninefold increase to a

maximum Zn/Al xlO^ ratio of over 90 in LBK-15 and 17.

The contents of nickel and magnesium also rise sharply from the base of UNIT 3

(from sample LBK-11), and attain a maxunum value in LBK-18 but have no significant

peak in LBK-15 or 17. The distribution of chromium and cobalt show a weak peak at

LBK-15; chromium attains its greatest abundance in LBK-17 and 18, and cobalt in

LBK-18 and 19. Beryllium also shows a slight increase from the base of UNIT 3, with

maximum abundances in LBK-17 to 19.

Sodium, potassium, mbidium, iron (total), barium, beryllium, scandium, and

zirconium abundances are exceptionally lower in LBK-8, the Dawson coal, than in the

associated shales. Magnesium, cobalt, and nickel are slightly less in LBK-8. The

concentrations of vanadium, zinc, copper, chromium, strontium, manganese, calcium,

phosphoms, yttrium, and silicon and titanium in the Dawson coal are compzirable to the

associated shales.

In this section, yttrium and phosphoms show the same distribution pattem with

peaks at LBK-8, 13,16, and 20 that are about 2 or 3 times higher than samples with low

abundances.

Lower Tackett Section at Tulsa Railroad (TRR)

The TOC varies from 2 to 5 weight percent for the samples TRR-25 to 37. The

basal two samples TRR-25 and 29 contain 2 to 3 weight percent TOC, and TRR-33 to

37 contain about 3.5 to just over 5 weight percent TOC. In contrast, TRR-39 and 40

have TOC values of about 10 weight percent.

Abundances of vanadium and chromium are high only at TRR-39 and 40 (Figure

28). The Ci/Al xlO^ ratio increases eight times from the underiying shale, and the V/Al 53

xlO^ ratio inceases four times. Zinc, copper, nickel, and magnesium show a sharp

increase at TRR-39, but fall to typical levels in TRR-40. Iron (total) and phosphonis are

significantly more abundant only in TRR-40, at the top ofthe black shale. Yttrium

content increases slightly in TRR-40.

Between TRR-29 and TRR-33, both low TOC shales, several elements show

distinct changes in distribution. From TRR-29 to TRR-30, calcium, manganese, and

strontium decrease about 50 percent, but chromium and zinc double in content, and

nickel, copper, and beryllium increase about 10%. Chromium, zinc, and beryllium

decease in TRR-35, but nickel and copper are more abundant.

PI80 Section (Permian Bead Mountain Limestone)

The black shale at the base (UNIT 2) and top (UNIT 4) ofthe shale sequence at

PI80 section contains from 1 to 2.5 weight percent TOC, whereas the intervening gray

shale (UNIT 3) contains 1 weight percent or less TOC. Because the weight percent of

aluminum oxide is lower in UNITS 2 and 4, the TOC/Al ratio is two to three times

greater than in UNIT 3. Calcium oxide, probably as calcium carbonate, forms 7 to 11.5

weight percent of samples from UNITS 2 and 4, and less than 0.6 weight percent of

samples in UNIT 3. The Ca/Al ratio in UNITS 2 and 4 is about ten times that in UNIT

3 (Figure 29).

Strontium shows higher abundance in UNITS 2 and 4, the Si/Al xlO^ ratio

increasing about four times that in UNIT 3. Magnesium increases at UNIT 2 only, the

Mg/Al ratio increasing one and a half times. Cobalt, nickel, yttrium, manganese, and

possibly iron (total) and phosphoms, show the same distribution pattem , with high

abundances at P180-5 in UNIT 2 and at P180-9 in UNIT 4. The exception to this

pattem is that iron (total) and phosphoms are not low at PI80-4. Silicon, zirconium,

titanium, sodium, mbidium, and vanadium show higher abundances at PI80-8, 9, and 54

10, with a peak at sample P180-8 just below UNIT 4. Potassium content does not

change throughout the section.

55

c t>

t> o.

o o o

c o

O O

2

1 . 9

1 . 8

1 . 7

1 . 6

1 . 5

1 .••

1 . 3

1 . 2

1 .1

1

0 . 9

0 . 8

0 . 7

0 . 6

0 . 5

O . *

0 . 3

0 . 2

O . I

o

-—

1 1

1 1

1 1

1 1

1 1

1 1

1 1

1 1

~~

D D

i ° D

1 1 1 1 1 1

D

1 1 1 1 1 1

O

a

1 1 1 1 1

D

1 1 1

0 . 2 0.4- 0 . 6 0 . 8 1 1 . 2 1.4.

T i t r a t i o n O r g a n i c C a r b o n ( p e r c e n t )

1 . 6 1 . 8

(per

cent

) C

arbo

n O

rgan

ic

LECO

2 4

2 2

2 0

1 8

1 6

1 4

1 2

1 0

8

6

4

2

0

LBK —1 8 O

LBK — 1 8

L B K - 1 7

a

L B K - 1 5 D

L B K - 1 3 LBK — 1 9 a

L B K - 1 1

n

1 2 1 6 4- 8

T i t r a t i o n O r g a n i c C a r b o n ( p e r c e n t )

a d j u s t e d LBK — 1 8 O o r i g i n o i L B K — 1 8

2 0 2 4

Figure 18. Graphs of LECO total organic carbon versus titration total organic caibon. Top: Samples in which the total oiganic carbon is less than two weight percent. Bottom: Samples in which the total organic carbon is more than two weight percent.

56

L 0) n £ c

0)

2 0 >

(0 c 3

6 -

5 -

4 -

2 -

- 1

-

-

-

-

T\

DQDD

luiuiu

nncD D

D DD OD D D D

D D D

1 1 1

LBK-8 D

1 1

0 20

Total Organic Carbon (weight percent)

• All samples

40

Figure 19. Graph of Munsell value number N versus total oiganic carbon (weight percent) for all samples. The lower the Munsell value number N, the darker the sample. The LBK-8 is the coal sample.

57

>

o c M

E M - 1 2 5 E M - 1 2 1 E M - 1 1 7 E M - 1 0 9 E M - 1 0 3

Z n / A l • 1 0 0 0 0 + Cr /A I ' 1 0 0 0 0 O Ni /Al - lOOOO

EM—98 T E M - 9 6 | E M - 9 9 E M - 9 7 EM —95

V / A l ' lOOOO

O .O

O O O O

O O

3

o

E M - 1 2 7 I E M - 1 2 3 I E M - 1 1 9 | E M - 1 1 3 | E M - 1 0 5 | E M - 1 0 1 | E M - 9 8 | E M - 9 6 I E M - 1 2 5 E M - 1 2 1 EM—117 EM—109 E M - 1 0 3 E M - 9 9 E M - 9 7 E M - 9 5

C u / A l "lOOOO Co /A I - lOOOO TOC/Al

Figure 20. Stratigraphic distributions of zinc, chromium, nickel, vanadium, copper, cobalt, and total organic carbon in East Mountedn (EM) section. Base of section is at right.

58

o

8 o E 3 C

E

c NI

1 5

1 4

1 3

1 2

1 1

1 0

9

8

7

6

5

4-

3

2

1

O

N 3 N 4

N 2 . 5

N 5

EM—127 I E M - 1 2 3 | E M - 1 1 9 | EM—113 | E M - 1 0 5 | E M - 1 0 1 | E M - 9 8 | EM —96 | EM—125 EM—121 EM—117 E M - 1 0 9 E M - 1 0 3 E M - 9 9 E M - 9 7 E M - 9 5

O O O O

E C

•£

u C

Ki

3 0 2 7 - 3 8 | 3 0 2 7 - 3 1 | 3 0 2 7 - 2 3 | 3 0 2 7 - 1 5 | 3 0 2 7 - 1 1 | U P S - 3 1 I U P s ' - 2 3 | U P s ' - l 5 | 3 0 2 7 - 3 5 3 0 2 7 - 2 7 3 0 2 7 - 1 9 3 0 2 7 - 1 3 U P S - 3 5 U P S - 2 7 U P S - 1 9 U P S - 1 1

Figure 21. Stratigraphic distributions of zinc and Munsell value number N in EM, 3027, and UPS sections. Base ofsection is at right. Top: EM section. Bottom: 3027 and UPS sections.

59

o o

c

o

8 O •_

>-

o o

D B - 1 1 3 DB—111 D B - 1 0 9 DB—107 DB—105 DB—103 DB—101

• C a / A l •+ Y / A I "lOOOO O Mn/A I " lOO A P/AI • 1 OO

O

<

c

(/)

D B - 1 1 3 D B - 1 1 1 DB—109 OB—107 DB—105 DB—103 D B - 1 0 1

D S r / A I ' lOOOO + 2 n / A I ' lOOOO O Ni /A l • lOOOO

Figure 22. Stratigraphic distributions of calcium, yttrium, manganese, phosphoms, strontium, zinc, and nickel in Dog Bend (DB) section. Base of section is at right.

60

>

c NI

UPS—35 UPS —31 UPS—27 U P S - 2 3 U P S - 1 9 U P S - 1 5 U P S - 1 1

n Z n / A l "lOOOO -t- C r /A I ' lOOOO O N i /A l * 1 0 0 0 0 A V / A l - 1 0 0 0 0

O P

O O O O

O O

3

o

U P S - 3 5 U P S - 3 1 U P S - 2 7 U P S - 2 3 U P S - 1 9 U P S - 1 5 UPS—11

C u / A l ' lOOOO + Co /A I ' 1 0 0 0 0 TOC/Al

Figure 23. Stratigraphic distributions of zinc, chromium, nickel, vanadium, copper, cobalt, and total organic carbon in UPS section. Base ofsection is at right.

61

o

< 3

o

o c M

1 1

10

9 -

8

7

6

5

4-

3

2

1

O 3027-

• Zn/Al 'lOOOO Cr/AI -lOOOO O Ni/Al -10000

X Cu/Al 'lOOOO

V/Al "lOOOO

1 .1

O O O O

O

1 -

0.9

0.8

0.7

0.6

0.5

0.4

0.3

0.2

0.1

o

Co/AI 'lOOOO TOC/AI

Figure 24. Stratigraphic distributions of zinc, chromium, nickel, vanadium, copper, cobalt, and total organic carbon in 3027 section. Base ofsection is at right.

62

o o

8"

c

>-

o

1 .3

1 .2 -

1 .1

1

0.9

0 . 8

0 .7

0 .6

0 . 5

0.4-

0 .3

0 .2

O.I

O C C B -

C a / A I + Y / A I " lOOOO O M n / A I " l O O P / A I " l O O

1 8

8 o o

c NI

1/1

S r / A I - lOOOO Z n / A l ' 1 0 0 0 0 N i / A l "lOOOO

Figure 25. Stratigraphic distributions of calcium, yttrium, manganese, phosphorus, strontium, zinc, and nickel in Colony Creek (CCB) section. Base ofsection is at right.

63

> c

NJ

P

o o o o

3

o d o o o

n Z n / A l ' lOOOO -t- V / A l ' lOOOO O Cr /A I • 1 0 0 0 0

4 0

3 5 -

3 0 -

2 5 -

2 0 -

15 -

1 0 -

5 -

D N i /A l * 1 0 0 0 0 + C u / A l ' 1 0 0 0 0 O TOC/Al

Figure 27. Stratigraphic distributions of zinc, vanadium, chromium, nickel, copper, and total organic carbon in Type Lost Branch (LBK) section. Base of section is at right.

65

1 0 0 00

00

— « •

^

Ni.

Zn.

Cr.

>

9 0

BO

7 0

6 0

5 0

AO

3 0

2 0

1 0

TRR—40 TRR—39 T R R - 3 7 TRR—35 T R R - 3 3 TRR—29 TRR—25

D V / A l ' lOOOO + Cr /A I » 1 0 0 0 0 O Z n / A l ' lOOOO £i N i /A l - 1 0 0 0 0

P

O O O O

3

o

4 . 5 -

3 . 5

0 . 5

2 . 5 -

1.5 -

TRR—40 T R R - 3 9 T R R - 3 7 T R R - 3 5 T R R - 3 3 T R R - 2 9 T R R - 2 5

D TOC/Al + C u / A l • 1 0 0 0 0

Figure 28. Stratigraphic distributions of vanadium, chromium, zinc, nickel, copper, and total organic carbon in Lower Tackett Shale (TRR) section. Base of section is at right.

66

o o

a.

8 •_

o o

P 1 8 0 — 1 0 P 1 8 0 - 9 P 1 8 0 - 8 P 1 8 0 - 7 P 1 8 0 - 6 P 1 8 0 - 5 P 1 8 0 — 4

• C o / A I Y / A I ' 1 0 0 0 0 Mn /A I ' l O O P/AI ' l O O

1 0

O

<

c NI

m

P I 8 O - 1 0 P 1 8 0 - 9 P 1 8 0 - 8 P 1 8 0 - 7 P 1 8 0 - 6 P 1 8 0 - 5 P 1 8 0 — 4

D S r / A I • 1 0 0 0 0 + Z n / A l - 1 0 0 0 0 O Ni /Al • 1 0 0 0 0

Figure 29. Stratigraphic distributions of calcium, yttrium, manganese, phosphorus, strontium, zinc, and nickel in PI 80 section. Base ofsection is at right.

67

coaly shales of UNIT 6 (CCB-123 to 126), also have higher concentrations of silicon

and zirconium. The upper three samples ofthe Permian shale (P180-8, 9, and 10) have

more silicon and zirconium than the lower four samples. The higher contents of silicon

and zirconium in these samples may be due to the presence of additional detrital quartz

and zircon. Slightly higher amounts of titanium in these samples suggest that titanium-

bearing minerals such as ilmenite and mtile may also be more common.

LBK-8, which is a coal, is strongly depleted in scandium and zirconium relative

to the marine shales of all the sections. When compared to the marine shales of all the

sections, LBK-8 has a typical concentration of titanium and slightly more silicon.

Kaolinite, the predominate clay mineral in coals, has a higher silicon/aluminum ratio

than illite. Terrestrial plants may take up titanium, and coal contains significant titanium,

the titanium being associated with the organic matter (Milnes and Fitzpatrick, 1989).

The low abundance of zirconium may be due to an absence of non-clay detrital particles.

Also, plants do not seem to accumulate zirconium (Milnes and Fitzpatrick, 1989).

Scandium in source rocks is weathered to scandium hydroxide which is more

soluble than aluminum hydroxide (Rankama and Sahama, 1950). Thus, aluminum is

less easily leached than scandium (Curtis, 1972). During weathering, clay minerals are

formed, and aluminum is one ofthe major elements of clays. While aluminum

accumulated as part ofthe clay minerals, the scandium was not deposited in the detrital

sediment in LBK-8. Scandium may correlate with total iron in shales because scandium

hydroxide coprecipitates with ferric hydroxide (Norman and Haskin, 1968), or the

scandium may replace ferrous ion (Curtis, 1972). The data in this sttidy show that

scandium does not correlate with total iron.

Most ofthe samples have Bc/Al xlO^ ratios varying from 0.80 to 0.90 (Figures

57, 58, and 59). The Dawson Coal (LBK-8) contains little beryllium (ratio less than

0.40), and the sample from the overiying shale (LBK-10) has a ratio near 0.70. The 69

marginal marine shale at East Mountain (UNIT 1) has a ratio less than around 0.80, as

does the sample from UNIT 7 (CCB-127) at the top ofthe Colony Creek Shale. In the

black shales at the Tulsa Railroad Cut and the Type Lost Branch slightly higher than

average abundances of beryllium (greater than Bo Al xlO^ ratio of 1.00) are present.

The greatest abundance of beryllium (Be/Al xlO^ ratio =1.25) occurs in one sample of

the carbonaceous shale near the top ofthe Colony Creek Shale (CCB-126 in UNIT 6).

Beryllium typically occurs in detrital clays where it substittites for silicon and

aluminum in clay minerals (Rankama and Sahama, 1950; Vine and Tourtelot, 1970).

The weathering and sedimentary chemistry of beryllium is similar to that of aluminum

(Goldschmidt, 1958; Kretz, 1972).

Degree of Weathering

The source rocks ofthe shale samples in this study were probably subjected to

intense weathering and erosion caused by heavy and frequent rainfall, because

Pennsylvanian North America was at the equator. The high equatorial temperature

facilitated chemical breakdown ofthe source rock minerals. The terrigenous clastic

sediments that travelled along the continental surface via mnoffs and rivers and reached

the marine depositional areas were already highly weathered before deposition. The

weathering index and the chemical index of alteration presented below are ways of

measuring the degree of weathering ofthe source rock minerals into mineral weathering

products before those products setded into the sediment and underwent transfomiation

into shale.

The degree of compositional mamrity ofthe shales in this smdy can be

approximated using a diagram (the weathering index graph) in which the ratio

[(CaO+Na20+K20)/(Al203+CaO+Na20+K20)] is plotted against the ratio

[(Si02+CaO+Na20+K20y(Al203+Si02+CaO+Na20+K20)l (Kronberg and Nesbitt,

70

1981). This index was origmally developed to detennine the degree of weathering of

minerals in soil profiles, but should be applicable to detrital materials in shales derived

from terrestrial weathering.

The y-axis on the weathering index graph (Figure 33) represents the breakdown

of feldspars and the accumulation of clay minerals (measured from 0 to 1). The x-axis

represents the accumulation of Si02 (measured from 0 to 1) or the accumulation of

AI2O3 (measured from 1 to 0) (Kronberg and Nesbitt, 1981). In plotting the

weathering index graph, CaO is not used because, in this sttidy, the amount of CaO in

silicates is not known. Most ofthe calcium in the shales in this sttidy appear to be

associated with carbonates, not silicates.

No significant differences in the weathering index exist among the shales in this

study. Almost all the samples plot around 0.19 on the y-axis and 0.85 on the x-axis

(Figure 33). This plot area is near to the compositions of illite, montmorillonite,

kaolinite, and quartz (Kronberg and Nesbitt, 1981) on the mineral distribution graph,

and is also near to the point on the weathering graph where the weathering is

approaching maximum. The coal (LBK-8) at the Type Lost Branch section, however, is

from a more highly weathered source, and, according to the weathering graph, has more

kaolinite and quartz.

Comparison of aluminum oxides with oxides of calcium, sodium, and potassium

provides a measure of chemical alteration that had occurred to produce the clay minerals

in shales. The chemical index of alteration (CIA) is based on the degree of change from

original feldspar mineral to final product, kaolinite clay (Nesbitt and Young, 1982):

CIA = [ AI2O3/(AI2O3 + CaO + Na20 + K2O ) 1 xlOO.

During chemical weathering, the calcium, sodium, and potassium of feldspars go into

solution while the aluminum and silicon form clays. The greater the degree of chemical

weathering, the less the abundances of calcium, sodium, and potassium, and the higher

71

the CIA value. In the CIA graphs (Figure 34), CaO has been omitted, as explained

above for the weathering index graph.

Although the samples are from different geographical areas and stratigraphic

levels, they have almost the same degree of chemical alteration (Figure 34). There are

slight differences in CIA values among each series of samples, but within each section

there is little change in CIA values. The samples have CIA values ranging from 77

percent to 85 percent, within the illite and montmorillonite range (Nesbitt and Young,

1982). The coal (LBK-8) plots at 97 percent which is significantly above the illite range

and is near to the kaolinite value of 100 percent.

The potassium atomic percent in the samples correlates with the aluminum

atomic percent (Figure 35). The positive correlation of potassium with aluminum may

be due to the possibility that both elements are constituents of illite (Murray, 1954;

Degens et al., 1957; Cubitt, 1979; Dabard and Paris; 1986). The potassium ion is held

tightly in the illite in ancient shales so it is a fixed ion that is not easily exchangeable with

other cations (Cody, 1971); potassium in the samples is therefore assumed to be largely

held in illite.

Rubidium content is rather constant for the samples (Figures 36, 60, 61, and

62). A plot of RVAl xlO^ versus K/Al (Figure 35) suggests that the mbidium is

substituting for potassium in illitic clays in the samples, assuming all the potassium is in

the illite. Rubidium is not found in independent mbidium minerals, but is taken up by

potassium-bearing minerals like illite during weathering. Rubidium easily substimtes

for potassium because the mbidium and the potassium ions have similar ionic radii and

similar ionic potential (Rankama and Sahama, 1950). Rankama and Sahama (1950)

reported a Rb:K ratio of 0.011 for argillaceous sediments. The samples in this smdy

have a Rb:K ratio of about 0.004. The low potassium and mbidium abundances in

LBK-8 coal are probably due to more intense weathering resulting in a higher proportion 72

of kaolinite. The same may be tme ofthe coaly shales of CCB-123 to 126 of UNIT 6

and the daric gray shale CCB-127 of UNIT 7.

Because there is little variation in the potassium content ofthe shales (Figures

36, 60, 61, and 62), the differences in the weathering index and chemical index of

alteration is due primarily to changes in sodium content. Sodium content appears to be

very erratic. Greater differences exist among the sections than among the samples in

each section (Figures 63, 64, and 65). The two sections ofthe Upper Salesville black

shale (3027 and UPS) display strikingly different pattems of sodium content. The

reason for the distribution is not clear. Perhaps the sodium was not retained uniformly

during dewatering ofthe shales.

Elements Associated With Calcium Carbonate

Many cyclothems have a basal transgressive unit that consist of a limestone or a

calcareous sandstone. Near the top ofthe cycles, a regressive shallow-water carbonate

unit may also occur. Calcareous skeletal debris is present in the shales to varying

degrees, commonly adjacent to the carbonate units, and may also be common in shales

that accumulated in shallow-water aerobic envirormients.

Shale samples adjacent to carbonate-rich units have higher concentrations of

calcium due to their greater content of calcium carbonate (Figures 66, 67, and 68). This

occurs in the Dog Bend section adjacent to the trangressive and regressive Hmestones,

above the transgressive limestones at the East Mountain section and the Colony Creeek

section, and above the trangressive calcareous sandstone at the upper Salesville sections

(UPS and 3027). Similariy, the shale adjacent to the carbonates in the Permian section

(PI80) also contains a higher calcium content.

The peak in calcium abundance in UNIT 5 at section CCB, (samples CCB-115

through 121), corresponds to a significantly greater concentration of skeletal debris as 73

determined by wet sieving ofthe shales. Other variations in calcium content in the

shales are not as easily explained. Isolated peaks in the upper Salesville Shale (UPS-23

and 3027-35) and flucttiations of calcium in sections TRR and LBK cannot be readily

attributed to the causes discussed above, but are also believed to be due to variations in

the proportion of calcium carbonate in the samples.

Strontium is a common trace element in carbonate units because it readily

substimtes for calcium in aragonite. In samples in this smdy, strontium abundance

(Figures 66, 67, and 68) is higher in every sample with elevated calcium, suggesting

that the strontium is associated directly with calcium carbonate in these samples. Several

other samples, however, possess a high concentration of strontium and a low

concentration of calcium. Samples from the carbonaceous marginal marine shale at East

Mountain (UNIT 1; EM-95 through 99) and the Colony Creek Shale (UNIT 6; CCB-

123 through 126) tend to have elevated amounts of strontium, especially as compared to

calcium. The shale below the marine black shale, and above the Dawson coal, at the

Type Lost Branch (LBK-10 in UNIT 2) also has high strontium and low calcium

abundances.

Variations in the abundance ofmanganese (Figures 69, 70, and 71) parallel

variations in calcium abundance in a majority of samples, and manganese is thought to

be held largely in carbonates, phosphates, and oxides. However, in the high TOC

samples at the Type Lost Branch (LBK-18, 19 , 20) manganese remains constant,

although calcium sharply decreases.

Magnesium (Figures 63, 64, and 65) does not show a clear correlation with

calcium. Magnesium abundances are probably related to other chemical phases in

addition to calcium carbonate. Little variation in the Mg/Al xlO* ratio occurs in the low

TOC samples, ranging from 0.10 to 0.15. Minor peaks occur in the maximum

transgressive shale at the East Mountain section, the Upper Salesville sections (UPS and 74

3027), and the Colony Creek section. One sample in the Upper Salesville 3027 section

(3027-35) has an unusually high ratio, 0.40. The marginal marine shale at East

Mountain (UNIT 1) and the carbonaceous shale at the top ofthe Colony Creek shale

(UNIT 6 and 7) have lower ratios around 0.05 to 0.07. In the high TOC shales, (LBK

and TRR sections) accumulation of magnesium seems to be to the same degree as in the

low TOC shales.

The abundance of barium (Figures 57, 58, and 59) in the shales is relatively

constant, with most values ofthe Ba/Al xlO^ ratio falling between 7.5 to 8.5. Three

samples in which barium is higher are samples with peaks in calcium abundance (DB-

101 and 113; CCB-119). The other samples with high calcium lack a corresponding

increase in barium. The Dawson coal (LBK-8) has an unusually low barium content.

Inorganic barium occurs in illite, substituting for the potassium (Murray, 1954; Nicholls

and Loring, 1962; Cubitt, 1979), and biogenous barium occurs in carbonates (Goldberg

and Arrhenius, 1958; Chester, 1965).

In samples with low values of TOC, changes in phosphoms abundance (Figures

72, 73, and 74) generally parallel changes in calcium abundance. The similarity of

phosphoms highs and lows with those of calcium suggests that phosphoms may be

connected with calcium due to the presence of apatite [Ca3(P04)2] or a carbonate-

phosphate phase. In the high TOC samples, phosphoms does not correlate with

calcium. Samples of marginal marine shale at East Mountain (EM-95 to 99 in UNIT 1)

and at the Colony Creek section (CCB-123 to 127 in UNITS 6 and 7) contain slightly

more than half the phosphoms found in the samples of marine shale at these sections.

The Y/Al xlO* ratios for most samples range from around 0.7 to 1.0 (Figures

72, 73, and 74). The yttrium highs and lows generally follow those of phosphorus,

suggesting that the yttrium is occurring in the phosphates (Rankama and Sahama,

75

1950). There is no phosphorus peak corresponding to the extremely high yttrium peak

in CCB-126, the carbonaceous marginal marine shale.

Organic Carbon, Sulfur, Iron, and Manganese

Organic Carbon

In sediments with high organic matter content, oxygen is depleted during

decomposition of organic matter producing anaerobic conditions. No oxygen is then

available to oxidize settling organic matter which then accumulates in the sediment.

Essential transition metals (discussed under the subheading "Essential Transition

Metals**), which are transition elements essential for life and are associated with organic

matter, accumulated in proportion to the organic matter. Thus in high TOC shales, the

abundance of essential transition metals tend to correlate with the abundance of TOC.

Also, the low redox potential (low Eh) in the environment of deposition ofthe high

TOC shales was low enough to produce H2S, with resultant formation of pyrite. The

high TOC shales in this sttidy have TOC/Al ratios of above 1.2.

Using a TOC/Al ratio above 1.2 as a criterion for high TOC shales (Figure 50),

the Kansas shales LBK-13, 15, 16, 17, 18, 19, and the Oklahoma shales TRR-39, 40,

are high TOC marine shales. Their essential transition metal contents tend to correlate

with TOC content. The coal LBK-8 is a high TOC shale as its TOC/Al ratio is above

1.2. However, despite its high TOC content, the coal has a low abundance of essential

transition metals. What is applicable to high TOC marine shales, namely high

abundances of essential transition metals are associated with high TOC and vice versa, is

not applicable to high TOC coal.

The envirormient of deposition of sediments with low organic matter content

probably ranged from dysaerobic to near aerobic. The extent of oxidation ofthe organic

matter varied because ofthe flucttiation ofthe redox potential. Thus in low TOC shales 76

the abundances of essential transition metals do not correspond with the increase or

decrease of TOC due to the variability of environmental conditions.

Using TOC/Al ratio below 1.2 as a criterion for low TOC shales (Figures 48 and

49), all the Texas marine samples (EM, CCB, DB, UPS, 3027) are assigned to this

group. The marginal marine carbonaceous shales of Colony Creek at Brad CCB-123 to

127 of UNITS 6 and 7, and the marginal marine shales ofthe Permian PI 80 section

have variable TOC/Al ratios. The Kansas samples LBK-10, 11, 20, and the Oklahoma

samples TRR-25, 29, 33, 35, 37, can be considered low TOC shales, because they have

TOC/Al ratios of below 1.2. LBK-11, with TOC/Al ratio of 1.4, is on the borderline

between high and low TOC shales.

Sulfur

In the low TOC samples, the S/Al values do not correlate with total Fc/Al values

or with TOC/Al values. However, in the high TOC samples (LBK-11, 13,15,17,18,

and 19), the S/Al values correlate well with total Fe/Al values. The slope ofthe graph is

generally parallel to the stoichiometric FeS2 (pyrite) slope (Figure 26). The negative y-

intercept and the positive x-axis intercept indicate that not all the total iron is in the

pyrite. Some iron may be the constituent of organic matter or substituted in clay

minerals, or may be adsorbed onto the organic matter or onto the clay minerals. In high

TOC samples, pyrite was identified in thin sections and by X-ray diffraction.

In the high TOC samples, the S/Al values also correlate well with TOC/Al

values. The regression line for LBK-13, 15, 17, and 19 values (Figure 26) passes

through the x-axis suggesting that pyrite formation depends upon a certain level of TOC

accumulation to cause a favorable low redox condition

The degree of pyritization (DOP) is a way of determining the level of reaction of

unbound reactive iron with hydrogen sulfide (Leventhal and Taylor, 1990). The DOP is 77

the ratio of pyrite iron to total iron, the total iron being the acid-soluble iron plus pyrite

iron (Raiswell and Bemer, 1985; Schultz, 1989). The degree of pyritization is an

indication of oxygenation level at the sediment-water interface. A DOP of 0.45 divides

aerobic and dysaerobic environments, and a DOP of 0.75 divides dysaerobic and

anaerobic conditions (Raiswell et al., 1988; Schultz, 1989). For marine sediments, if

percent pyrite sulfur correlates positively with percent organic carbon and the graph

passes through the origin, then pyrite formation depends on organic carbon

accumulation (Raiswell and Bemer, 1985). If the graph has a positive intercept on the

pyrite sulfur axis, the relation of DOP versus organic carbon could distinguish whether

pyrite formation is limited by organic carbon or by unbound reactive iron (Raiswell and

Bemer, 1985). As acid-soluble iron was not determined in this smdy, the DOP value

could not be obtained.

The pE is the negative logarithm ofthe electron concentration in moles per liter.

The Eh (redox potential) is a measure of a solution's ability to supply electrons for

chemical reaction. Eh and pE are related by the equation Eh = [(2.303 RT)/F] pE

(Kraskopf, 1955, 1967; Brookins, 1988), where R is the gas constant (0.001987

kcal/mole/degree K), T is the absolute temperature, and F is the Faraday constant

(23.06 kcaWolt-gram equivalent). For 25**C, T = 298.15°K, and Eh = 0.059 pE.

The Eh of a solution at non-standard condition is given by the Nemst equation

Eh = E'*+[(2.303 RTy(nF)]log{([Y]y[ZF)/([B]b[D]d)}, where E° is the standard

potential, and n is the number of electrons given by one atom to another (Kraskopf,

1955, 1967; Brookins, 1988). The symbols B and D are the reactants, Y and Z are the

products, in the general chemical equation bB+dD = yY+zZ. The Eh-pH stability field

diagrams of selected metal sulfides are shown in Figures 37, 38, 39,40, 41,42.

78

Iron and Manganese

The stratigraphic distributions of iron (total) and manganese are shown in

Figures 69, 70, and 71. In low TOC shales (East Mountain, Colony Creek, upper and

lower Salesville, and PI80), total iron and ferrous iron abundances pooriy reflect the

redox conditions of marginal and normal marine.depositional environments. The total

iron content is erratic and does not show a relationship with manganese content. The

ferrous iron/total iron ratio does not have any correlation with stratigraphy. The

manganese content of these shales parallels variations in the calcium abundance. The

manganese may be incorporated into the sediment as carbonate, phosphate, or hydrated

oxide. The increase in manganese content with the abundance of skeletal calcium

carbonate in these sections suggests oxidizing conditions. Many ofthe carbonate

skeletal microfossil grains show thin oxide coatings in these shales. Both the

manganese and the skeletal calcium abundances may reflect the dysaerobic-near aerobic

transition zone ofthe depositional environment better than the total iron abundance.

Both iron (total) and manganese contents are slightly lower in the marginal

marine shales than in the normal marine samples of East Mountain and Colony Creek.

Other than that, the iron (total) and manganese contents are variable and erratic. The

ferrous iron/total iron ratios in the marginal East Mountain samples (EM-95 to 99) are

about half that ofthe normal marine East Mountain samples (Appendix B). The Colony

Creek marginal marine carbonaceous shale samples (CCB-123, 125, and 126 in UNIT

6) have about twice the ferrous iron/total iron ratios ofthe normal marine Colony Creek

samples (Appendix B). Marginal marine East Mountain shales have less ferrous iron

than normal marine East Mountain shales because they were deposited under higher

redox conditions. The carbonaceous Colony Creek shales, due to higher TOC content,

were deposited under lower redox conditions than the normal marine Colony Creek

79

shales and the ferrous iron content is correspondingly higher. The normal marine East

Mountain samples show no significant variation in total iron content.

The iron (total) and manganese abundances in the high TOC shales are about the

same level as in the low TOC shales. There is no correlation of iron or manganese with

TOC in the low TOC shales (Figures 43 and 44). For the high TOC shales, total iron

shows correlation with increase of TOC, whereas the manganese shows no association

with TOC (Figures 43 and 44). When the TOC content was high, the redox potential

was low, and iron was probably locked in the sulfide phase as pyrite. Manganese

sulfide is stable within a very narrow range of Eh (redox potential) and pH, whereas

iron sulfide is stable over a greater range (Garrels and Christ, 1965; Brookins, 1988).

Thus in the depositional environment ofthe high TOC shales, the manganese ions had

diffused away leaving behind the iron which was immobilized in the pyrite solid phase.

The manganese in the high TOC and the low TOC shales is probably a

constiment ofthe organic matter. The slightly higher abundance ofmanganese in some

samples ofthe low TOC shales are associated with carbonate, phosphate, or oxide

phases as explained above. In the low TOC shales, the coal LBK-8 and the CCB

carbonaceous samples have a higher, but still narrow range of total iron abundances.

For the low and high TOC shales, perhaps most ofthe iron was associated with organic

matter when they came into the area of deposition. For the high TOC shales, after

deposition in the sediment, the iron switched from association with organic matter to

association with sulfur; thus the iron, organic carbon, and sulfur seem to correlate with

each other.

Essential Transition Metals

Nutrient elements are chemical elements essential for biological systems to carry

out the life processes of maintenance, growth, and reproduction. Shaw (1960) grouped

80

nutrient elements into essential alkali and alkaline earth metals, essential transition

metals, and essential non-metals. Therefore, essential transition metals are transition

metals that are nutrient elements. The essential transition metals discussed here are

vanadium, chromium, cobalt, nickel, copper, and zinc. The essential transition metals

iron and manganese are discussed under the subheading "Organic Carbon, Sulfur, Iron,

and Manganese.**

Essential transition metals are concentrated to levels exceeding oceanic

abundances in the tissues of marine plankton (Manskaya and Drozdova, 1968;

Krauskopf, 1956). Upon the death ofthe organisms, the metals are remmed to the

environment, either in the form of soluble complexes or locked in particulate organic

detritus. Where the TOC of shales is high, the abundances of essential transition metals

are expected to be correspondingly high. In low TOC samples, the content of essential

transition metals will be lower, but other factors may permit enrichment of some of these

metals above anticipated levels. Clay minerals and iron/manganese oxides may

preferentially scavenge these metals from sea water when plankton productivity is high

and rates of sediment deposition and burial are low. Analyses of shales from outer shelf

settings show that low TOC sedunents may be enriched in some metals like zinc and

chromium (Yin et al., 1989).

Vanadium

The vanadium content is very low in low TOC shales (Figures 75 and 76) and

high in all the high TOC shales (Figure 77), except for the Dawson coal (LBK-8). For

TOC/Al ratios of less than 2, the samples show V/Al xlO^ ratios of around 6 to 10. For

TOC/Al ratio of more than 2 in marine shales, there is a linear relationship (Figure 45).

For each 1 unit increase in TOC/Al ratio, the V/Al xlO^ ratio increases by about 25. The

strong linear relationship [r=0.98, n=7 (LBK-13, 15, 16, 17, 19, TRR-39, 40)]

81

between vanadium and TOC suggests that the vanadium content of these Pennsylvanian

marine shales can be used to approximate the TOC content. For the high TOC marine

samples, vanadium is strongly associated with organic matter because it is part ofthe

organic matter or is scavenged by adsorption and chelation (Krauskopf, 1955, 1956;

Keith and Degens, 1959).

In the low carbon shales, the vanadium shows little relationship to TOC, and the

vanadium may be occurring in illite (Coveney and Martin, 1983). In the EM, CCB,

3027, DB, UPS, and PI80 samples, the vanadium shows a weak relationship with

potassium (Figure 32). As potassium is assumed to be in the illite, the vanadium may

be substittiting in the illite in the shales (Murray, 1954; Krauskopf, 1955; Nicholls and

Loring, 1962; Chester, 1965; Coveney and Martin, 1983; Coveney etal., 1987).

Although most vanadium in these shales may have originated with organic matter, some

vanadium may have come with detrital illite or may have migrated into the illite lattice

after deposition (Nicholls and Loring, 1962). The low vanadium-low potassium points

in the graph (Figure 32) are those of CCB-122 to 127 and those of P180. In high TOC

marine shales, the quantity of vanadium attached to the organic matter overwhelms the

effect of vanadium attached to the illite (Figure 45).

Despite its high TOC, the Dawson coal (LBK-8) contains significantly less

vanadium than the high TOC marine shales and about half the vanadium in the low TOC

shales. Similar low values of vanadium in coals has been documented by other workers

(e.g., Swaine, 1983, table II). Compared to marine organic matter, terrestrial organic

matter contains relatively little vanadium (Manskaya and Drozdova, 1968; Keith and

Degens, 1959). The interpreted low illite content (see position of LBK-8 in Figure 32

top graph and Figure 33) ofthe clay in the coal, would also explain why the coal has

less vanadium than the shales.

82

Essential Transition Metals in Low TOC Shales

The majority of samples in this smdy are low TOC shales, where the quantity of

oiganic matter will have a small effect on the abundances of essential transition metals.

There is no correlation between abundances of essential transition metals and organic

carbon in low TOC shales (Figure 46).

Chromium. Chromium shows relatively little variation in low TOC shales

(Figures 78 and 79), with the Ci/Al xlO^ ratios varying from 4.5 to 6. At most

sections, the chromium is more abundant at the base ofthe dark gray to black shale that

represents the maximum transgressive interval. At the Upper Salesville section UPS,

there is a steep drop ofthe Cr/AI xlO^ ratio from 15 to 5 higher in the section. At the

Upper Salesville section 3027 and East Mountain secion, the Cr/AI xlO^ ratio drops

from around 6 to 7 in the maximum transgressive shale to 4 higher in the section. At the

Dog Bend and the Colony Creek sections, chromium abundance drops only slightly.

The marginal marine shale at PI 80 shows the lowest ratio of between 2.5 to 3.

Zinc. In general, the marine shales have Zn/Al xlO^ ratios varying widely from

5 to 13 (Figures 78 and 79). The abundance of zinc varies in a manner similar to that of

chromium, but shows a greater range of values. The Zn/Al xlO* ratio decreases from a

maximum value in the maximum transgressive shale interval to lower values higher in

the section. In the Upper Salesville at UPS and 3027 and the East Mountain sections,

zinc content drops sharply. In the Colony Creek at Brad and the Dog Bend sections,

zinc shows a slight decrease. There is more zinc in the upper part ofthe Colony Creek

shale section at Brad and the top sample ofthe shale at the Dog Bend section. These are

the samples that contain higher abundances of elements (calcium and strontium)

associated with carbonates. Two samples with unusually high concentrations of zinc

occur associated with the carbonaceous shale at section CCB (samples 122 and 126).

83

The marginal marine shale at section P180 has low Zn/Al xlO^ ratios of about 3

to 4, comparable to those ofthe marginal marine shale at East Mountain (EM-95 to 99),

which has values of about 3 to 5.

Nickel. Nickel has a somewhat mixed distribution in the low TOC shale

samples, with most values ofthe Ni/Al xlO^ ratio falling between 2 and 4 (Figures 75

and 76). Peaks of nickel occur stratigraphically just above the peaks of chromium and

zinc in the maximum transgressive shale in some sections (Upper Salesville shale at

UPS and 3027, and at East Mountain). In the Colony Creek section at Brad, the

abundance of nickel is only slightly greater in the maximum transgressive shale. Other

significant peaks in nickel distribution occur in samples bearing high calcium, strontium,

and manganese in the Dog Bend section and PI80 section. In the carbonaceous shales

near the top ofthe Colony Creek at Brad (UNIT 6), nickel is also high in some samples,

but near average in others.

In the shale at the top ofthe Colony Creek at Brad (UNIT 7) and the marginal

marine shale in the lower part ofthe East Mountain section (UNIT 1) nickel abundance

falls to its lowest values, Ni/Al xlO^ ratios of less than 1.5, compared to typical values

greater than 2.

Cobalt Most values ofthe Co/AI xlO^ ratio range between 0.4 and 0.8 (Figures

81 and 82). A minor peak occurs in the maximum transgressive shale at the Upper

Salesville sections 3027 and UPS and at East Mountain section. Little regular

stratigraphic variation exists higher in these sections or in the marine shales in Colony

Creek section or the Dog Bend section. Greater abundances of cobalt occur at the top of

UNIT 6 in the carbonaceous shale in the Colony Creek section, and in two samples at

PI 80 (PI 80-5 and 9). The marginal marine shales at the base ofthe East Mountain

section (UNIT 1) have slightly lower Cc/Al xlO^ ratios, less than 0.4.

84

Copper. The Cu//Al xlO^ ratios range between 0.5 to 1.5 for most marine shale

samples (Figures 81 and 82). Minor peaks occur in the maximum transgressive shale at

the East Mountain, UPS, 3027, and perfiaps the Colony Creek sections. Unusually

high values occur in two samples at the 3027 section, where the Cu/Al xlO^ ratio rises

to above 5.5 and 14 at samples 3027-23 and 38, respectively.

Essential Transition Metals in High TOC Shales

There is correlation between the abundances of essential transition metals and

organic carbon in high TOC shales (Figure 47). In the high TOC shales, (LBK and

TRR sections) accumulation of cobalt (Figure 83) seems to be ofthe same degree as in

the low TOC shales (Figures 81 and 82). Nickel and copper abundances (Figures 77

and 83) in the high TOC shales are two to four times larger than in low TOC shales

(Figures 75, 76, 81, and 82). Vanadium, zinc, and chromium (Figures 77 and 80) in

the high TOC shales show abundances that are five to nine times higher than in the low

TOC shales (Figures 75, 76, 78, and 79). Above a certain threshold of TOC/Al ratio

(between 1 and 2), nickel and copper correlate weakly with the TOC , and vanadium,

zinc, and chromium correlate strongly with the TOC (Figure 47).

Although the organic carbon content ofthe shale (TRR-40) at the top of TRR

section shows no change compared to the lower adjacent shale layer, the abundances of

copper, zinc, and nickel decrease, vanadium remains the same, and chromium increases

(Figure 28). From the base to the middle of LBK section, organic carbon and essential

transition metals generally increase, and from the middle to the top ofthe section, they

decrease (Figure 27). At LBK-16, organic carbon, copper, and nickel abundances

hardly change compared to the adjacent layers above (LBK-17) and below (LBK-15)

but zinc, vanadium, and chromium increase.

85

c V 3 CT V

32

30

28

26

24

22

20

18

16

14-

12

10

8

6

4

2

O —r-o I I I I r

0.5 I 1 I 1.5 I 2 I 2.5 0.25 0.75 1.25 1.75 2.25 2.75

Si l icon/Aluminium 3.75

I -^.5 I 4.25 4.75

c V 3 cr V

40

35 -

30

25 -

20 -

15

10 -

5 -

'o I O d04 I 0.d08 I 0.612 I 0.016 I 0.02 I 0.024 I 0.028 | 0.032 | 0.036 | 0.04-0.002 0.006 0.01 0.014 0.018 0.022 0.026 0.03 0.034- 0.038

Ti ton ium/Alu min ium

Figure 30. Frequency histograms of silicon and titanium abundances for all samples. Top: silicon. Bottom: titanium.

86

c v 3 O" V

5 0

4 5

4-0

3 5

3 0

2 5

2 0

15 I -

10

b&A&i

O.I 0 . 2 0 .3 0 . 4 0 .5 0 . 6 0 .7 0 .8 0 .9

S c a n d i u m / A l u m i n i u m * 1 0 0 0 0

1.1 1 .2 1 .3

4 0

3 5 -

3

3 0

2 5 -

2 0 -

15 -

10 -

5 -

2 3 4 5 6 7

Z i r c o n i u m / A l u m i n i u m "lO.OOO

1 0

Figure 31. Frequency histograms of scandium and zirconium abundances for all samples. Top: scandium. Bottom: zirconium.

87

E 3 C

'E _3

E 3

o c

11

1 0

9

8

7

6

5

4-

3

2

1

O

LBK—8

1 i

a^i

<^^

o^o

a

0 . 0 4

a All lov

0 . 0 8 O.I 2

P o t a s s i u m / A l u m i n u m

0 . 1 6 0 . 2 0 . 2 4

TOC s a m p l e s + C C B - 1 2 2 t o 1 2 7 o P 1 8 0 - 4 t o 1 0

X T R R - 2 5 to 3 5 V L B K - 8 . 10 , a n d 2 0 EM—95 to 9 9

8 o o

E 3 C

E _3 < E 3

'c O

• All samples

S i l i c o n / A l u m i n u m

C C B - 1 2 3 t o 1 2 6 O EM —101

V 3 0 2 7 - 1 1 A L B K - 8 X C C B - 1 1 3

Figure 32. Graphs of vanadium versus potassium and zirconium versus silicon. Top: Graph of vanadium versus potassium for all low TOC samples. The LBK-8 is the coal sample. Bottom: Graph of zirconium versus silicon for all samples.

88

0 CM

+ 0 CM D z + 0 CM

<

\

0 CM

+ 0 CM D

z

1

0.9

0.8

0.7

0.6

0.5

0.4

0.3

0.2

0.1

Gibtsite

-e-0 0.2 0.4

Na & K-Felcispars Ca-Felcispars • •

llite D

All Samples

Montmorillonite •

Kaolinite LBK-8

D

0.6 0.8

(Si02+Na20+K20)/(AI203+Si02+Na20+K20)

Quartz

1

Figure 33. Graph of weathering index for all samples. The LBK-8 is the coal sample.

y-axis = [(CaO+Na20+K20y(Al203+CaO+Na20+K20)] X-axis = [(Si02+CaO+Na20+K20y(Al203+Si02+CaO+Na20+K20)l (Kronberg and Nesbitt, 1981).

Note that CaO values are not used in the calculations of weathering index for the shale samples (see text).

89

c «> 3

«>

3 4

3 2 -

3 0 -

2 8 -

2 6

2 4

2 2

2 0

1 8

1 6

1 4

1 2

1 0

8

6

4

2

O o T 1^ T -r 8 I 16 1 2 4 I 3 2 I 4 0 | 48 | 56 | 64 | 72 | 80 | 88 | 96

1 2 2 0 2 8 3 6 4 4 5 2 6 0 6 8 7 6 8 4 9 2 TOO C h e m i c a l I n d e x o f A l t e r a t i o n ( C I A )

< •

o_ c o \^ o

1 — V 5

o X V -o c

"5 u 'E t> .c o

1 0 0

9 8

9 6

9 4

9 2

9 0

8 8

8 6

8 4

8 2

8 0

7 8

7 6

74^

7 2

7 0

- " ^ ^ ^ ^ ^ ^

1 , , E M

t S ? r ^ K C S L E , i ^ ^ P t ^ ^

c * L r M ^ ^ • ^ D

P

1 CCB 1 3027 E M C C B

^ n \? ^ M C^Sn D ^ ^

rf!^^ 1

1 DB 1 UPS 3 0 2 7 D B

r ^ C l l ^ ^ I3aJ7«=3XJ

,£1 o p^

C ^ 9 V •

1 P180 1 TRR U P S P 1 8 0

r i ^ oR5^J^ L ^ r r M P

1 LBK T R R

D

I LBK

Figure 34. Graphs of chemical mdex of alteration (CL\) for all samples. Top: frequency histogram. Bottom: distribution by sample and section; base ofsection is at right.

CIA = [Al2Oy(Al2O3+CaO+Na2O+K2O)lxl00 (Nesbitt and Young, 1982).

Note that CaO values are not used in the calculations of CL\ for the shale samples (see text).

90

c V 0) a.

E o

E 3

O

0 . 1 3

0 . 1 2 -

O. I 1

O.I I -

0 . 0 9

0 . 0 8

0 . 0 7

0 . 0 6

0 . 0 5

0.04-

0 . 0 3

0 . 0 2

0 . 0 1

O _

LBK—8 D

0 .2

Aluminum (a tomic percent )

• All samples

0 . 4 0 . 6

O

8 o E 3 C '£ _3

E 3

'•U

'.O 3

6 -

5 -

4 -

2 -

^SQ CP D

LBK—8 O

_ L JL.

0 . 0 4

All somples

0 . 0 8 O.I 2 0 . 1 6 0 . 2

P o t a s s i u m / A l u m i n u m

+ C C B - 1 2 3 to 1 2 7 LBK —8

0 . 2 4

Figure 35. Graphs of potassium versus aluminum and mbidium versus potassium for all samples. Top: potassium versus aluminum (atomic percent). Bottom: mbidium versus potassium (normzdized with aluminum).

91

t> 3 a

4 5

4-0 -

3 5 -

3 0 -

2 5 -

2 0 -

15 -

1 0 -

5 -

0 . 0 4 I 0 . 0 8 I 0 1 2 I O.I 6 | 0 . 2 | 0.24- | 0 . 2 8 0 . 0 2 0 . 0 6 O.I 0.14. 0 . 1 8 0 . 2 2 0 . 2 6

Po tass ium/A lumin ium

c t> 3 a V

4 5

4 0

3 5

3 0

2 5

2 0

1 5

1 0

tggg^ 0 . 5 1.5 2 2 . 5 3 3 . 5 4 4 . 5

Rub id ium/A lumin ium "lOOOO

5 .5 6 . 5

Figure 36. Frequency histograms of potassium and mbidium abundances for all samples. Top: potassium. Bottom: mbidium.

92

1 1 1

SYSTEM C o - S - C - 0 - H

25*»C. 1 bar

10 12 14

Figure 37. Eh-pH diagram showing stability field of cobalt sulfide (Brookins 1988). The solubility product of cobalt sulfide is (Dean, 1985; Krauskopf, 1967): Mineral Fonnula Solubilitv Pmdnrt jaipurite alpha-CoS 10-20.4 jaipurite beta-CoS 10-24.7

93

SYSTEM C u - C - S - 0 - H 25°C, 1 bar

12 14

Figure 38. Eh-pH diagram showing stability field of copper sulfide (Brookins, 1988). The solubility product of copper sulfide is (Dean, 1985; Krauskopf, 1967): Mineral Formula Solubilitv Product chalcocite Cu2S 10-47.6 covellite CuS 10-35.2

94

UJ

-0.2 —

-0.4 —

-o.e

-0.8 I

Figure 39. Eh-pH diagram showing stability field of iron sulfide (Brookins, 1988). The solubility product of troilite (FeS) is 10-17.2 ny^^ 1985. Krauskopf, 1967). ' '

95

SYSTEM Mn-C-S-O-H 25°C, 1 bar

12 14

Figure 40. Eh-pH diagram showing stability field ofmanganese sulfide

ThcTsolubility product ofmanganese sulfide is (Dean, 1985; Krauskopf, 1967): Mineral Fonnula Solubility Product

alabandite (amorphous, pink) MnS ^ 12 A alabandite (ciystalline, green) MnS 10- * •<>

96

0.8 -

SYSTEM Ni - 0 - H - S 25°C, 1 bar

6 8 pH

X o

10 12 14

Figure 41. Eh-pH diagram showing stability field of nickel sulfide (Brookins. 1988). The solubility product of nickel sulfide is (Dean, 1985; Krauskopf, 1967): Mineral Fonnula Solubilitv Produrt millerite alpha-NiS 10-18-5 millerite beta-NiS 10-24.0 millerite gamma-NiS 10-25.7

97

SYSTEM Z n - 0 - H - S - C 25°C, 1 bar _

Figure 42. Eh-pH diagram showing stability field of zinc sulfide (Brookins. 1988). The solubility product ofzinc sulfide is (Dean, 1985; Krauskopf, 1967): Mineral Formula Solubilitv Produrt sphalerite alpha-ZnS 10-23.8 wurtzite beta-ZnS 10-21.6

98

E 3 C

'E _3 <

M O

C O

2 "5

0 . 3 2

0 . 3

0 . 2 8

0 . 2 6

0 . 2 4

0 . 2 2

0 . 2

0 . 1 8

0 . 1 6

0 . 1 4

0 . 1 2

O.I

0 . 0 8

0 . 0 6

0 . 0 4

0 . 0 2

O

-

:

1 1

1 1 1

1

-

-

-

o

O

D

n

1 1

D °

1 1 1

D

in

1

O

]

a

1 1

D

1 1

0 . 2 0 . 4 0 .6 0 .8

Total Organic Carbon / Aluminum

1 .2

E 3 C

'E

<

in

c o

o o

0 . 4

0 . 3 5

0 . 3

0 . 2 5

0 . 2

0 . 1 5

O . I

0 . 0 5

D

D

n

^^

1 i

a

a

1 1

a

1

D

1 1 1 1 i

L B K - 8 D

1 1 1 1 1 1

4 6 8 1 0

Total Organic Carbon / Aluminum

1 2 1 6

Figure 43. Graphs oftotal iron versus total organic carbon. Top: low TOC shales. Bottom: high TOC shales.

99

F

Alu

min

ui

V M t>

o c (7 IS

0 . 0 1

0 . 0 0 9

0 . 0 0 8

0 . 0 0 7

0 . 0 0 6

0 . 0 0 5

0 . 0 0 4

0 . 0 0 3

0 . 0 0 2 -

0 . 0 0 1 -

-

-

-

D

O

D

a

1 1 1 1 1

D

n

a

1

a 1

D

1 1 D 1 1

0 . 2 0 . 4 0 .6 0 . 8

Total Organic Carbon / Aluminum

1 .2

E 3 C

'E 3

V V) V c o o> c o

2

0 . 0 0 3 2

0 . 0 0 3

0 . 0 0 2 8

0 . 0 0 2 6

0 . 0 0 2 4

0 . 0 0 2 2

0 . 0 0 2

0 . 0 0 1 8

0 . 0 0 1 6

0 . 0 0 1 4

0 . 0 0 1 2

0 . 0 0 1

0 . 0 0 0 8

0 . 0 0 0 6

0 . 0 0 0 4

0 . 0 0 0 2

O

-

-

-

-

_

1 1

1

1

• a

• D

a

D

a D

1 1

a

D D

D

1 1

n

1

D

D

1 1 1 1 1

L B K - 8 O

I J 1 1 1 1

4 6 8 1 0

Total Organic Carbon / Aluminum

1 2 1 6

Figure 44. Graphs ofmanganese versus total organic carbon. Top: low TOC shales. Bottom: high TOC shales.

100

0 o 0 o

E 3 C

I < \ E 3 '•D

D C D >

130

120

110

100

90

80

70

60

50

40

30

20

10

0

LBK-17 +

LBK-15 +

LBK-16 +

LBK-13 +

TRR-39 & 40 LBK-18 + +

LBK-19 +

LBK-8 . + .

0 8 10 12 14 16

Total Organic Carbon / Aluminum

D Low TOC samples + High TOC samples

Figure 45. Graph of vanadium versus total organic carbon for Lower Tackett (TRR) and Type Lost Branch (LBK) shales. The low TOC samples include those of EM, CCB, 3027, UPS, DB, and P180 sections. The LBK.8 is the coal sample.

101

o

O

8 o

o o 3

o

18

1 7

16

15

1 4

1 3

12

1 1

1 0

9

8

7

6

5

4

3

2

1 —

O

16

15

14

13

12 -

1 1

10

9

8

7

6

5

4

3

2 -

a <J> ^

0 . 2 0 . 4 0 . 6 0 . 8 1 .2

Total Organic Carbon / Aluminum

V / A l ' lOOOO + Z n / A l - lOOOO O Cr /A I - lOOOO

D

a D

Pa ^ +Jr^»r ° tSb

t o <? o o o -I 1 I

0 . 2 0 . 4 0 . 6 0 . 8 1 .2

Total Organic Carbon / A luminum

D N i /A l "lOOOO -t- C u / A l " 1 0 0 0 0 O Co /A I "lOOOO

Figure 46. Graphs of vanadium, zinc, chromium, nickel, copper, and cobalt versus total organic carbon for low TOC shales. The low TOC shale samples are from EM, CCB, 3027, UPS, DB, and P180 sections.

102

o c M

1 30

120

1 10

100

90

80

70

60

50

40

30

20

10

0

-

-

-

-

-

-

"~

§ 8 1

• -•-0

1

0

D

-t-

1

D

+

0

1

CO

8 1 0

L B K - 8 ' S L.

1 2 1 4 1 6

V / A l Total Organic Carbon / Aluminum

• 1 0 0 0 0 + Z n / A l - 1 0 0 0 0 O Cr /A I " 1 0 0 0 0

O O

O

3 o

4 0

3 5 -

3 0 -

2 5 -

2 0 -

15 -

10 -

5 -

4 6 8 1 0

Total Organic Carbon / A luminum

1 2 1 4 1 6

Ni /A l "lOOOO C u / A l • 1 0 0 0 0 C o / A I • 1 0 0 0 0

Figure 47. Graphs of vanadium, zinc, chromium, nickel, copper, and cobalt versus total organic carbon for high TOC shales. The high TOC shale samples are from LBK section.

103

CHAPTER V

DISCUSSION

The abundances of vanadium, zinc, and chromium have been used to infer the

level of redox potential (Eh) in depositional environments because these elements

accumulate in low redox environments characterized by accumulation of oiganic matter.

Therefore a strong correlation of vanadium, zinc, and chromium abundances with TOC

abundance suggests that the original sediment was deposited in a low redox environment

that allowed preservation of organic matter. Offshore deep-water shales (core shales of

cyclothems) were formed from sediments deposited during maximum transgression.

Organic matter accumulation and preservation during maximum transgression may have

been enhanced by lower redox potential and/or by higher rates of organic sedimentation.

The presence of a thermocline would lower oxygen concentration at the bottom ofthe

water column and enhance development of a lower redox potential. The relative rate of

sedimentation of organic material would also be greater due to lower rate of terrestrial

clastic influx. The accumulated organic carbon may lower the redox potential to a level

favorable for the formation of hydrogen sulfide. Then, iron can be locked in pyrite but

not manganese. The strong association between iron and sulfur is because ofthe

presence of pyrite. Data for the marine shales in this study show that beyond a

threshold concentration of TOC (TOC/Al ratio > 2), vanadium, zinc, and chromium

correlate strongly with TOC. Nickel and copper also correlate with the TOC although

their increase is not as pronounced as that of vanadium, zinc, and chromium. Iron

abundance is high, and correlates with TOC and with sulfur, whereas manganese is low

and does not correlate with TOC. The inference is that these high TOC shales were

deposited in an original sedimentary condition that was anaerobic.

104

For the high organic carbon shales, a slight change in sedimentary redox

condition might affect the content of essential transition metals without affecting the

organic carbon content. The abundances of essential transition metals might decrease or

remain the same compared to the adjacent shale layeis.

Some core shales are low in TOC; their essential transition metals may increase

at maximum transgression even though the organic caibon content shows no

corresponding increase in the same stratum. The essential transition metals decrease as

regression proceeds. The decrease does not correspond with oiganic carbon content.

The relatively small amount of organic matter preserved in these core shales may be due

to loss of organic matter where the redox potential is high enough for more extensive

oxidation ofthe organic matter. The higher redox potential could be due to the absence

of a thermocline. Following rapid transgression, abundances of essential transition

metals in low TOC core shales increase due to longer settling time allowing settling clay

and oiganic particles to scavenge the essential transition metals with greater efficiency.

As deltas prograde, clastic influx dilutes the sediment causing a relative decrease in the

essential transition metals. In low TOC core shales (TOOAl ratio < 1.2) of EM, 3027,

and UPS sections, abundances of essential transition metals zinc, nickel, and chromium

increase steeply from base then decrease gradually. The essential transition metals do

not correlate with the TOC because of unequal accumulation or loss ofthe essential

transition metals and the organic matter during deposition. Although the underlying

sediment may be anaerobic, some oiganic matter may have been oxidized at a dysaerobic

sediment surface layer and so the organic carbon and the essential transition metals do

not accumulate to any great extent. Core shales with sudden increase from base and

gradual decrease towards upper outside shales of some essential transition metals are

from deep water with dysaerobic sediments.

105

Transgressive and regressive limestone beds are normal feamres of many

cyclothems. Although one or both ofthe transgressive and regressive carbonate strata

may be visibly absent in some cyclothems, higher concentrations of caibonate-related

elements may mark the positions in the cyclothems where limestone beds would be

expected. The Dog Bend section, with transgressive and regressive limestone beds,

shows the expected higher concentrations of caibonate-related elements calcium,

strontium, zinc, manganese, phosphorus, yttrium, and nickel in the upper and lower

part ofthe section. The Colony Creek section, with transgressive limestone but no

regressive limestone, shows higher abundances of caibonate-related elements calcium,

strontium, manganese, and zinc in the upper outside shales compared to its core shales.

The upper outside shales contain marine calcareous fossils. In low TOC regressive

shales, the strong association of calcium, strontium, manganese, nickel, phosphoms,

and yttrium suggests that the environment of deposition was near aerobic and was

conducive to the formation of carbonates, phosphates, and hydrated oxides.

Iron and manganese abundances are expected to be high in aerobic sediments (as

insoluble oxides) and low in anaerobic sediments. Iron and manganese are essential

transition metals and their contents would be affected by organic carbon content. In the

low TOC marine shales in this study, the iron and manganese abundances are erratic,

vary widely, and do not correlate with TOC.

Marginal marine shales are from depositional environments that receive marine

oiganic particles from overlying waters, and terrestrial organic and clastic particles from

land. As the marginal marine sediment is open to influence from sea and land, there

may be no distribution pattem of organic carbon and essential transition metals. The

redox condition in the sediment may vary as the oxygen level in the overlying water

varies. The marginal marine shales of CCB have about 2 percent more organic carbon

than the normal marine shales in the same section. The marginal marine shales of EM 106

have similar levels of organic caibon content as the nonnal marine shales. In CCB

marginal marine shales, the abundances of essential transition metals are erratic whereas

in EM maiginal marine shales the abundances tend to be similar to those of its normal

marine shales. In both EM and CCB sections, there is no correlation between organic

carbon and essential transition metals. Terrestrial oiganic matter is more refractory and

has a lower essential transtion metal content than marine organic matter. Deposition of

varying mixtures of marine and terrestrial organic matter produce flucttiating amounts of

oiganic caibon and essential transition metals with no correlation between the

abundances. Fluctuating redox conditions could also cause fluctuating abundances and

no correlation. There is no observed pattem of elemental distribution that could clearly

differentiate maiginal marine shales from normal marine shales.

Terrigenous clastic sediments are composed of weathering products which result

from different stages of weathering and are sorted during transportation by size and

density. The concentrations of elements in terrigenous clastic sediments can reveal the

stages of weathering and the degree of sorting. The detrital elements in the shales in this

study show similar abundances within each section and among different sections.

Heavy mineral elements (titanium and zirconium), and clay mineral elements (aluminum,

siHcon, potassium, mbidium, scandium, and beryllium), do not vary much in content.

Determinations of degrees of chemical changes from feldspars to clay minerals indicate

that the source rocks ofthe samples were weathered to the same degree. Although the

terrigenous clastic sediments forming the shales in this study were deposited in marginal

to normal marine environments, in anaerobic to dysaerobic to near aerobic

environments, and in widely separated geographic locations, they were uniformly

weathered and sorted before sedimentation.

The sedimentary conditions of Pennsylvaniem cyclothemic shales can be deduced

by correlation of organic carbon with essential transition metals, by correlation of iron 107

with sulfur, by abmpt increase and gradual decrease from base to top of some essential

transition metals but without correlation with organic caibon, and by the presence of

caibonate-related elements.

The correlation of TOC with vanadium, zinc, and chromium, could be used to

decide whether there should be further prospecting for metals in high TOC shales. If

there is correlation, then not only vanadium, zinc, and chromium, but other essential

transition metals like nickel, copper, cobalt, and molybdenum might be concentrated in

some parts ofthe high TOC shales to make mining economical. If the high TOC shales

have TOC correlating with sulfur, metal sulfides of iron, lead, zinc, copper, and silver

might be present in economic concentrations. Deteimining which metal correlates with

sulfur would help decide which metal to look for. Further exploration would help

decide if mining for that metal is economically feasible.

Petroleum is a diagenetic product of marine oiganic matter. When TOC

correlates with essential transition metals, the organic carbon most likely comes from

marine organisms rather than from terrestrial organisms. The high TOC marine shales,

in which the TOC correlates with essential transition metals, could be hydrocarbon

source beds and could be distilled to recover the contained hydrocarbon. The

abundances of TOC and essential transition metals could be used in prospecting for

hydrocarbon in high TOC shales.

Marine organic matter in shales scavenge essential transition metals. Low TOC

marine shales could be powdered and used as an adsorbent for these metals. Liquid

industrial waste which contains essential transition metals could be passed through

powdered low TOC shales kept at low redox potential. The marine oiganic matter in the

shales would scavenge the metals. Other transition metals which are not essential

transition metals might be scavenged too. The effluent would not have the metals as

contaminant, and after oxygenation in air could be dischaiged into rivers or the sea. 108

Perhaps if the transition metals collected in the powdered low TOC shale filter reach a

certain level of concentration, they could be economically recovered and recycled for use

by industry.

Economic concentrations of essential metals and hydrocarbons are more likely in

high TOC shales, where the TOC correlates with the essential transition metals. Low

TOC shales could be used to remove essential transition metals from industrial waste

water before the waste is dischaiged into the environment.

109

CHAPTER VI

CONCLUSIONS

The Pennsylvanian marginal and normal marine sections smdied seem to have

similar terrigenous clastic source materials as indicated by weathering index, chemical

alteration index, and abundances of detrital elements. Some normal marine sections

(TRR and LBK) consist of high TOC shales and were deposited in low redox

environments. The abundances of essential transition metals in these sections correlates

with the organic carbon abundances. Pyrite is present, and the iron correlates with

sulfur.

Although the normal marine section LBK shows oiganic carbon and essential

transition metal abundances increasing from base to middle ofsection, and then

decreasing from middle to top ofsection, the normal marine section TRR does not show

this pattem. The middle part of LBK section and the top of TRR section show that in

high organic caibon shales essential transition metal might vary even if organic carbon

does not. The variation might be due to a very slight increase in redox potential which

does not affect the organic carbon content. Other normal marine sections (EM, CCB,

DB, and 3027, UPS) are low TOC shales and were deposited in dysaerobic

environments. The essential transition metals do not correlate with organic carbon.

Some shales ofthe low TOC normal marine sections (CCB and DB) were

deposited in dysaerobic environments that were near aerobic as shown by higher

concentrations of caibonate-related elements when compared to the other shales in the

same section.

Some normal marine sections (EM, 3027, and UPS) have higher contents of

essential transition metals at the base ofthe section. The abundances of these essential

transition metals decrease from the base to the top ofthe section. The depth of water 110

column in these sections at maximum transgression was such that settling particles had

time to scavenge a larger proportion ofthe essential transition metals in the overlying

waters before reaching the sediment.

The maiginal marine shales of CCB section show wide variations of abundances

of oiganic caibon and essential transition metals because of varying redox conditions

and dilution of marine organic matter by varying amounts of terrigenous clastic influx.

However, the marginal marine shales of EM section show little variation of organic

carbon and essential transition metal contents.

In the low TOC marginal and normal marine shales, the total iron and manganese

abundance variations are too erratic to be of use for inferring sedimentary redox

conditions.

The elemental geochemistry of Pennsylvanian cyclothemic shales fumishes

information regarding sedimentary redox potential in different stages ofthe cyclothemic

sequence and the relative influx of terrestrial inorganic and marine organic particles.

Core shales may be deposited under anaerobic to dysaerobic environments. Core shales

deposited under dysaerobic conditions may show increased abundances of essential

transition metals at maximum transgression. Outside shales may be deposited under

dysaerobic to near aerobic environments and may be influenced by terrestrial clastic

influx. Outside shales may show varying abundances of essentizd transition metals or

may show little variation. Low organic carbon core shales and outside shales may be

useful as metal pollutant sinks. Elemental geochemistry may be used to search for high

organic carbon core shales which may be sources of oil.

For a preliminary survey of marine cyclothemic shales, analysis oftotal organic

caibon, aluminum, zinc, vanadium, chromium, iron (total), sulfur, calcium, strontium,

manganese, and phosphoms, could be carried out. The aluminum analysis is for

normalization of abundances ofthe other elements. Redox conditions are indicated by

111

whether zinc, vanadium, and chromium correlate with TOC. Correlation of iron and

sulfur could mean the presence of metal sulfide minerals besides pyrite. Concomitant

increase of calcium, strontium, zinc, manganese, and phosphorus indicates near aerobic

conditions. This initial and limited geochemical survey provides a quick characterization

of shales.

112

REFERENCES

Adlis, D .S. , Grossman, E. L., Yancey, T. E. , McLerran, R. D. , 1988, Isotope stratigraphy and paleodepth changes of Pennsylvanian cycUcal sedimentary deposits, Palaios, v. 3, pp. 487-506.

Amon, D. I . , and Wessel G., 1953, Vanadium as an essential element for green plants, Namre, v. 172, pp. 1039-1040.

Anhenius, G. O. S., 1959, Sedimentation on the ocean floor, /n Abelson, P. H., editor, Researches in Geochemistry, John Wiley & Sons, Inc., New York.

Baas Becking, L. G. M., Kaplan, I. R., and Moore, D., 1960, Limits ofthe namral environments in terms of pH and oxidation-reduction potential, Journal of Geology, v. 68, pp. 243-284.

Ball, D. F., 1964, Loss-on-ignition as an estimate of organic matter and organic carbon in non-calcareous soils, Journal of Soil Science, v. 15, pp. 84-92.

Belt, E. S., and Macqueen, R. W., editors, 1979, Intemational Congress on Caiboniferous Stratigraphy and Geology, v 3, Part 3: Sedimentology and Geochemistry, Southern Illinois University Press.

Bennison, A. P., 1984, Source Bed Field Conference, Guide Book: Upper Pennsylvanian Source Beds of Northeastem Oklahoma and Adjacent Kansas, Tulsa Geological Society Field Trip.

Bemer, R. A., 1971, Principles of Chemical Sedimentation, McGraw-Hill, New York.

Bemer, R. A., 1984, Sedimentary pyrite formation: An update, Geochimica et Cosmochimica Acta, v. 48, pp. 605-615.

Blatt, H., Middleton, G., and Murray, R., 1980, Origin of Sedimentary Rocks, Prentice-Hall, Inc.

Boardman, D. R., II, Barrick, J. E., Cocke, J., and Nestell, M. K., editors, 1989a, Middle and Late Pennsylvanian Chronostratigraphic Boundaries in North-Central Texas: Glacial-Eustatic Events, Biostratigraphy, and Paleoecology, Part I, Texas Tech University Studies in Geology 2.

Boardman, D. R., II, Barrick, J. E., Cocke, J., and Nestell, M. K., editors, 1989b, Middle and Late Pennsylvanian Chronostratigraphic Boundaries in North-Central Texas: Glacial-Eustatic Events, Biostratigraphy, and Paleoecology, Part n, Texas Tech University Studies in Geology 2.

Boardman, D. R., II, Barrick, J. E., Heckel, P. H., 1989c, Proposed redefinition of the Missourian-Virgilian stage boundary (Late Pennsylvanian), Midcontinent North America, Geological Society of America, Abstracts with programs, v. 21, p. A168.

113

Boardman, D. R., H, Barrick, J. E., Heckel, P. H.,and Nestell, M. K., 1989d, Upper Pennsylvanian chronostratigraphic subdivisions ofthe North American midcontinent, in Boardman, D. R., II, Barrick, J. E., Cocke, J., and Nestell, M. K., editors, Middle and Late Pennsylvanian Chronostratigraphic Boundaries in North-Central Texas: Glacial-Eustatic Events, Biostratigraphy, and Paleoecology, Part II, Texas Tech University Sttidies in Geology 2, pp. 1-16.

Boardman, D. R., n, and Heckel, P. H., 1989, Glacial-eustatic sea-level curve for eariy upper Pennsylvanian sequence in north-central Texas and biostratigraphic correlation with curve for midcontinent North America, /n Boardman, D. R., II, Barrick, J. E., Cocke, J., and Nestell, M. K., editois. Middle and Late Pennsylvanian Chronostratigraphic Boundaries in North-Central Texas: Glacial-Eustatic Events, Biostratigraphy, and Paleoecology, Part II, Texas Tech University Sttidies in Geology 2, pp. 63-73.

Boaniman, D. R., II, and Malinky, J. M., 1985, Glacial eustatic control of Virgilian cyclothems in north central Texas, American Association of Petroleum Geologists, Southwest Section Transactions, pp. 13-23.

Boardman, D. R., II, Mapes, R. H., Yancey, T. E., and Malinky, J. M., 1984, A new model for the depth-related allogenic community succession within North American Pennsylvanian cyclothems and implications on the black shale problem, inHyne, N. J., editor. Limestones ofthe Midcontinent, Tulsa Geological Society Special Publication 2, pp. 141-182.

Brongersma-Sanders, M., 1966, Metals of Kupferschiefer supplied by normal sea water, Geologische Rundschau, v. 55, pp. 365-375.

Brongersma-Sanders, M., 1969, Origin of trace-metal enrichment in bimminous shales, inHobson, G. D., and Speers, C. C , editors. Advances in Organic Geochemistry, Third Intemational Congress Proceedings, Oxford, Pergamon Press, pp. 231-236.

Brookins, D. G., 1988, Eh-pH Diagrams for Geochemistry, Springer-Veriag, Beriin.

Brown, L. F., 1989, A sequence-stratigraphic and systems-tract model for the Virgilian and Wolfcampian Series, Eastern Shelf and Adjacent Midland Basin, Texas, i/rBoardman, D. R., II, Barrick, J. E., Cocke, J., and Nestell, M. K., editors., 1989, Middle and Late Pennsylvanian Chronostratigraphic Boundaries in North-Central Texas: Glacial-Eustatic Events, Biostratigraphy, and Paleoecology, Part II, Texas Tech University Smdies in Geology 2.

Bmmsack, H., 1989, Geochemistry of Recent TOC-rich sediments from the Gulf of California and the Black Sea, Geologische Rundschau, v. 78, pp. 851-882.

Canfield, D. E., Raiswell, R., Westrich, J. T., Reaves, C. M., and Bemer, R. A., 1986, The use of chromium reduction in the analysis of reduced inoiganic sulfur in sediments and shales. Chemical Geology, v. 54, pp. 149155.

Chakrapani, G. J., and Subramanian, V., 1990, Preliminary smdies on the geochemistry ofthe Mahanadi River basin, India, Chemical Geology, v. 81, pp. 241-253.

114

Chester, R., 1965, Elemental geochemistry of marine sediments, /n Riley, J. P., and Skirrow, G., editors. Chemical Oceanography, v. 2, Academic Press.

Cleaves, A. W., m, 1975, Upper Desmoinesian-Lower Missourian depositional systems (Pennsylvanian), North-Central Texas: unpublished Ph. D. dissertation. The University of Texas at Austin, 256 pages.

Cocke, J. M., Boardman, D. R., H, Mapes, R. H., 1989, Stratigraphic distribution and facies control of Late Pennsylvanian coral assemblages in the North American Midcontinent, /nBoardman, D. R., II, Barrick, J. E., Cocke, J., and Nestell, M. K., editors. Middle and Late Pennsylvanian Chronostratigraphic Boundaries in North-Central Texas: Glacial-Eustatic Events, Biostratigraphy, and Paleoecology, Part II, Texas Tech University Smdies in Geology 2, pp. 249-27o.

Cody, R. D., 1971, Adsorption and the reliability of trace elements as environment indicators for shales. Journal of Sedimentary Petrology, v. 41, pp.461-471.

Coveney, R. M., and Glascock, M. D., 1989, A review ofthe origins and metal-rich Pennsylvanian black shales, central USA, with an inferred role for basinal brines. Applied Geochemistry, v. 4, pp.347-367.

Coveney, R. M., Leventhal, J. S., Glascock, M. D., and Hatch, J. R., 1987, Origin of metals and organic matter in the Mecca Quarry Shale Member and stratigraphically equivalent beds across the midwest. Economic Geology, v. 82, pp. 915-933.

Coveney, R. M., and Martin, S. P., 1983, Molybdenum and other heavy metals ofthe Mecca Quarry and Logan Quarry shales. Economic Geology, v. 78, pp. 132-149.

Coveney, R. M., and Shaffer, N. R., 1988, Sulfur-isotope variations in Pennsylvanian shales ofthe midwestem United States, Geology, v. 16, pp. 18-21.

Coveney, R. M., Watney, W. L., and Maples, C. G., 1989, Rates and durations for accumulation of Pennsylvanian black shales in the midwestem United States, in Franseen, E. K., and Watney, W. L., editors. Sedimentary Modelling: Computer Simulation of Depositional Sequences, Subsurface Geology Series 12, Kansas Geology Survey.

Crowell, J. C , 1978, Gondwanan glaciation, cyclothems, continental positioning and climatic change, American Joumal of Science, v. 728, pp. 1345-1372.

Cubitt, J. M., 1979, Geochemistry, mineralogy, and petrology of upper Paleozoic shales of Kansas, Kansas Geological Survey Bulletin, no. 217, 117 pages.

Cubitt, J. M., and Merriam, D. F., 1979, Geochemical distributions in Pennsylvanian cyclic sediments ofthe midcontinent (USA), in Belt, E. S., and Macqueen, R. W., editors, 1979, Intemational Congress on Carboniferous Stratigraphy and Geology, v 3, Part 3: Sedimentology and Geochemistry, Southern Illinois University Press.

115

Curtis, C. D., 1972, Scandium: element and geochemistry, /i7Fairt)ridge, R. W., The Encyclopedia of Geochemistry and Environmental Sciences, Dowden, Hutchinson and Ross, Inc.

Dabard, M., and Paris, F., 1986, Palaeontological and geochemical characteristics of Silurian black shale formations from the central Britanny domain of die Armorican Massif (northwest France), Chemical Geology, v. 55, pp. 17-29.

Dean, J. A., editor, 1985, Lange's Handbook of Chemistry, McGraw-Hill, Inc.

Dean, W. E., 1974, Determination of carbonate and organic matter in calcareous sediments and sedimentary rocks by loss on ignition: Comparison with other methods, Joumal of Sedimentary Petrology, v. 44, pp. 242-248.

Degens, E. T., .Williams, E. G., and Keith, M. L., 1957, Environmental smdies of Carix)niferous sediments. Part I: Geochemical criteria for differentiating marine from fresh-water shales, American Association of Petroleum Geologist Bulletin, v. 41, pp. 2427-2455.

DeMaison, G. J., and Moore, G. T., 1980, Anoxic environments and oil source bed genesis, American Association of Petroleum Geologists Bulletin, v. 64, pp. 1179-1209.

Doner, H. E., and Lynn, W. C , 1989, Carbonate, halide, sulfate, and sulfide minerals, in Dixon, J. B., and Weed, S. B., editors. Minerals in Soil Environments, Soil Science Society of America.

Dunbar, C. O., 1960, Correlation ofthe Permian formations of North America, Bulletin ofthe Geological Society of America, v. 71, pp. 1763-1806.

Ebanks, W. J., Jr., Lawrence, L. B., Heckel, P. H., O'Conner, H. G., Sanderson, G. A., West, R. R., and Wilson, F. W., 1979, The Mississippian and Pennsylvanian (Carboniferous) Systems in the United States - Kansas, United States Geological Survey Professional Paper 1110-Q, 30 pages.

Ehlers, E. G., and Blatt, H., 1982, Petrology: igneous, sedimentary, and metamorphic, W. H. Freeman and Company.

Erxleben, A. W., 1975, Depositional systems in Canyon Group (Pennsylvanian System), North-Central Texas: Bureau of Economic Geology, University of Texas at Austin, Report of Investigations 82, 76 pages.

Flanagan, F. J., editor, 1976, Descriptions and Analysis of Eight New USGS Rock StandanJs, Geological Survey Professional Paper 840, United States Department ofthe Interior.

Garrels, R. M., 1960, Mineral Equilibria at Low Temperature and Pressure, Harper and Brothers, New York.

Garrels, R. M., and Christ, C. L., 1965, Solutions, Minerals, and Equilibria, Haiper and Row, New York.

116

Gaudette, H. E., Flight, W. R., Toner, L., Folger, D. W., 1974, An inexpensive titration method for the determination of oiganic carbon in recent sediments, Joumal of Sedimentary Petrology, v. 44, pp. 249-253.

Glass, H. D., Potter, P. E., and Siever, R., 1956, Clay Mineralogy of some basal Pennsylvanian sandstones, clays, and shales, American Association of Petroleum Geologists Bulletin, v. 40, pp.750-754.

Goebel, K. A., Bettis, E. A., Jr. and Heckel, P. H., 1989, Upper Pennsylvanian paleosol in Stranger Shale and underiying latan Limestone, Southwestern Iowa, Joumal of Sedimentary Petrology, v. 59, in press.

Goldberg, E. D., 1954, Marine geochemistry, 1. Chemical scavengers ofthe sea, Joumal of Geology, v. 62, pp. 249-265.

Goldberg, E. D., and Arrhenius, G. O. S., 1958, Chemistry of Pacific pelagic sediments, Geochimica et Cosmochimica Acta, v. 13, pp. 153-212.

Goldschmidt, V. M., 1958, Geochemistry, Oxford University Press.

Heckel, P. H., 1977, Origin of phosphatic black shale facies in Pennsylvanian cyclothems of midcontinent North America, American Association of Petroleum Geologists Bulletin, v. 61, pp. 1045-1068.

Heckel, P. H., 1979, Changing concepts of midcontinent Pennsylvanian cyclothems, North America, in Belt, E. S., smd Macqueen, R. W., editors, Intemational Congress on Carboniferous Stratigraphy and Geology, v. 3, Part 3, Sedimentology and Geochemistry, Southem Illinois University Press.

Heckel, P. H., 1980, Paleogeography of eustatic model for deposition of midcontinent Upper Permsylvanian cyclothems, in Fouch, T. D., and Magathan, E. R., editors. Paleozoic Paleogeography of West-Central United States, Society of Economic Paleontologists and Mineralogists, Rocky Mountain Section, Paleogeography Symposium I, pp. 197-215.

Heckel, P. H., 1986, Sea-level curve for Pennsylvanian eustatic marine transgressive-regressive depositional cycles along midcontinent outcrop belt. North America, Geology, v. 14, pp. 330-334.

Heckel, P. H., 1989, Current view of midcontinent Pennsylvanian cyclothems, in Boardman, D. R., II, Barrick, J. E., Cocke, J., and Nestell, M. K., editors, Middle and Late Pennsylvanian Chronostratigraphic Boundaries in North-Central Texas: Glacial-Eustatic Events, Biostratigraphy, and Paleoecology, Part II, Texas Tech University Sttidies in Geology 2, pp. 17-34.

Heckel, P. H., and Baesemann, J. F., 1975, Environmental interpretation of conodont distribution in Upper Pennsylvanian (Missourian) Megacyclothems in eastern Kansas, American Association of Petroleum Geologists Bulletin, v. 59, pp. 486-509.

Holland, H. D., 1979, Metals in black shales - A Reassessment, Economic Geology, v. 74, pp.1676-1680.

117

Hume, D. N.,1973, Pitfalls in the determination of environmental trace metals, in AJiuja, S., Cohen, E. M., Kneip, T. J., Lambert, J. L., and Zweig, G., editors, Chemical Analysis ofthe Environment and other Modem Techniques, Progress m Analytical Chemistry, v. 5, Plenum Press, New York.

Ingall, E. D. and Cappellen, P. V., 1990, Relation between sedimentation rate and bunal of organic phosphoms and organic carbon in marine sediments, Geochimica et Cosmochimica Acta, v. 54, pp. 373-386.

Jeffrey, P. G., and Hutchison, D., 1981, Chemical Methods of Rock Analysis, Pergamon Press.

Johnson, W. M., and Maxwell, J. A., 1981, Rock and Mineral Analysis, John Wiley and Sons.

Kammer, T. W., Brett, C. E., Boardman, D. R., H, and Mapes, R. H., 1986, Ecologic stability ofthe dysaerobic biofacies during the Late Paleozoic: Lethaia, v. 19, pp. 109-121.

Keeling, P. S., 1962, Some experiments on the low-temperamre removal of carbonaceous material from clays. Clay Minerals Bulletin, v. 5, pp. 155-158.

Keith, M. L., and Degens, E. T., 1959, Geochemical indicators of marine and fresh­water sediments, in Abelson, P. H., editor. Researches in Geochemistry, John Wiley and Sons, New York.

Krauskopf, K. B., 1955, Sedimentary deposits of rare metals. Economic Geology, 50th Anniversary volume, part I, pp. 411463.

Krauskopf, K. B., 1956, Factors controlling the concentrations of thirteen rare metals in sea-water, Geochimica et Cosmochimica Acta, v. 9, pp. 1-32B.

Krauskopf, K. B., 1967, Introduction to Geochemistry, McGraw-Hill, Inc.

Kretz, R., 1972, Beryllium: Element and geochemistry, inFairbridge, R. W., The Encyclopedia of Geochemistry and Environmental Sciences, Dowden, Hutchinson and Ross, Inc.

Kronberg, B. I., and Nesbitt, H. W., 1981, Quantification of weathering, soil geochemistry and soil fertility, Joumal of Soil Science, v. 32, pp. 453-459.

Kmmbein, W. C , and Garrels, R. M., 1952, Origin and classification of chemical sediments in terms of pH and oxidation-reduction potentials, Joumal of Geology, v. 60, pp. 1-33.

Leventhal, J. S., and Shaw, V. E., 1980, Oiganic matter in Appalachian Devonian black shale: (I) Comparison of techniques to measure organic carbon, (II) Short range organic carbon content variations, Joumal of Sedimentary Petrology, v. 50, pp.77-81.

Leventhal, J., and Taylor, C , 1990, Comparison of methods to determine degree of pyritization, Geochimica et Cosmochimica Acta, v. 54, pp. 2621-2625.

118

Manskaya, S. M., and Drozdova, T. V., 1968, Geochemistry of Organic Substances, Pergamon Press.

Merrill, G. K., and Grayson, R. C , 1989, Conodont paleoecology in the Type East Mountain Shale, north-central Texas, in Boardman, D. R., Barrick, J. E., Cocke, J., and Nestell, M. K., editors, 1989, Middle and Late Pennsylvanian Chronostratigraphic Boundaries in North-Central Texas: Glacial-Eustatic Events, Biostratigraphy, and Paleoecology, Part II, Texas Tech University Studies in Geology 2, pp. 235-248.

Merrill, G. K., Grayson, R. C , and Mosley, J. L., 1987, Resmdy ofthe localities and conodont faunas of Stauffer and Plummer (1932), in Grayson, R. C, Merrill, G. K., and Miller, J. F., Early and Late Paleozoic conodont faunas ofthe Llano Uplift Region, central Texas - Biostratigraphy, Systemic Boundary Relationships, and Stratigraphic Importance, Fieldtrip Guidebook for 21st Annual Meeting South-Central Section ofthe Geological Society of America, pp. 23-47.

Moussavi-Harami, R., and Brenner, R. L., 1984, Deltaic sedimentation on a carbonate shelf: Stanton Formation (Upper Pennsylvanian), southeastem Kansas, American Association of Petroleum Geologists Bulletin, v. 68, pp. 150-163.

Milnes, A. R., and Fitzpatrick, R. W., 1989, Titanium and zirconium minerals in Dixon, J. B., and Weed, S. B., editors, Minerals in Soil Environments, Soil Science Society of America.

Murray, H. H., 1954, Genesis of clay minerals in some Pennsylvanian shales of Indiana and Illinois, Clays and Clay Minerals, Proceedings ofthe Second National Conference, pp. 47-67.

Myrow, P. M., 1990, A new graph for understanding colors of mudrocks and shales, Joumal of Geology Education, v. 38, pp. 16-20.

Nesbitt, H. W., and Young, G. M., 1982, Eariy Proterozoic climates and plate motions inferred from major element chemistry of lutites. Nature, v. 299, pp. 715-717.

Nicholls, G. D., 1960, Techniques in sedimentary geochemistry: (2) Detennination of the ferrous iron contents of carbonaceous shales, Joumal of Sedimentary Petrology, v. 30, pp. 603-612.

Nicholls, G. D., and Loring, D. H., 1962, The geochemistry of some British Carboniferous sediments, Geochimica et Cosmochimica Acta, v. 26, pp. 181-223.

Nonnan, M. D., and De Deckker, P., 1990, Trace metals in lacustrine and marine sediments: A case sttidy from the Gulf of Carpentaria, Northern Australia, Chemical Geology, v. 82, pp. 299-318.

Norman, J. C , and Haskin, L. A., 1968, The geochemistry of Sc: A comparison to the rare earths and Fe, Geochimica et Cosmochimica Acta, v. 32, pp. 93-108.

119

Paropkan, A. L., 1990, Geochemistry of sediments from the Mangalore-Cochin shelf and upper slope off southwest India: Geological and environmental factors controlling dispersal of elements. Chemical Geology, v. 81, pp. 99-119.

Plummer, F. B., 1929, Geologic Map of Palo Pinto County, (revised, 1930), Umversity of Texas Bureau of Economic Geology (in cooperation with the American Association of Petroleum Geologists).

Plummer, F. B., and Moore, R. C , 1922 (1921), Stratigraphy ofthe Pennsylvanian formations of north-central Texas, Texas University Bulletin 2132, 273 pages.

Potter, P. E., Maynard, J. B., and Pryor, W. A., 1980, Sedimentology of Shale, Smdy Guide and Reference Source, Springer-Veriag, New York.

Potter, P. E., Shimp, N. F., and Witter, J., 1963, Trace elements in marine and fresh water argillaceous sediments, Geochimica et Cosmochimica Acta, v. 27, pp. 669-694.

Prince, A. L., 1955, Methods in soil analysis, in Bear, F. E., editor. Chemistry ofthe Soil, Reinhold Publishing Corporation.

Quinby-Hunt, M. S., Wilde, P., Berry, W. B. N., and Orth, C. J., 1988, The redox-related facies of black shales, Geological Society of America, v. 20, p. A193.

Raiswell, R., and Bemer, R. A., 1985, Pyrite formation in euxinic and semi-euxinic sediments, American Joumal of Science, v. 285, pp. 710-724.

Raiswell, R., Buckley, F., Bemer, R. A., Anderson, T. F., 1988, Degree of pyritization of iron as a paleoenvironmental indicator of bottom-water oxygenation, Joumal of Sedimentary Petrology, v. 58, pp. 812-819.

Rankama, K., and Sahama, Th. G., 1950, Geochemistry, University of Chicago Press.

Salomons, W., and Forstner, U., 1984, Metals in the Hydrocycle, Springer-Veriag.

S.A.S.G.S. guidebook, Southwestem Association of Student Geological Societies, 1963, Field Conference on the Geology of West Central Texas, Hardin-Simmons University Geological Society, 84 pages.

Schutter and Heckel, 1985, Missourian (eariy Late Pennsylvanian) climate in Midcontinent North America, Intemational Joumal of Coal Geology, v. 5, pp. 111-140.

Schultz, R. B., 1987, Mineral determination ofthe Heebner Shale Member (Shawnee Group, Upper Pennsylvania) in southeastem Kansas using X-ray diffraction techniques, in Merriam, D. F., editor. The Compass, Sigma Gamma Epsilon, University of Kansas, v. 64, pp. 149-159.

Schultz, R. B., 1989, Geochemical characteristics and inferred depositional environments of upper Pennsylvanian (Virgilian) black shales in eastern Kansas, in Merriam, D. F., editor. The Compass, Sigma Gamma Epsilon, University of Kansas, v. 67, pp. 47-54.

120

Seward, R. W., editor, 1986, NBS Standard Reference Materials Catalog 1986-1987, NBS Special Publication 260, National Bureau of Standards, United States Department of Commerce.

Shaw, T. J., Gieskes, J. M., and Jahnke, R. A., 1990, Eariy diagenesis in differing depositional environments: The response of transition metals in pore water, Geochimica et Cosmochimica Acta, v. 54, pp. 1233-1246.

Shaw, W. H. R., 1960, Sttidies in biogeochemistry-I. A biogeochemical periodic table. The data. Geochimica et Cosmochimica Acta, v. 19, pp. 196-207.

Shelton, J. H., 1958, Strawn-Canyon (Pennsylvanian) Boundary in North-Central Texas, unpublished Master's thesis. University of Illinois, Campaign-Urban, 120 pages.

Siy, S. E., 1988, Geochemical and petrographic study of phosphate nodules ofthe Woodford shale (Upper Devonian-Lower Mississippian) of southem Oklahoma, unpublished Master's thesis. Department of Geosciences, Texas Tech University, 172 pages.

Swaine, D. J., 1983, Geological aspects of trace elements in coal, in Augustitis, S. S., The Significance of Trace Elements in Solving Petrogenetic Problems and Controversies, Theophrastus Publishing and Proprietary Company, S. A., Athens, Greece.

Tourtelot, H. A., 1979, Black shale - its deposition and diagenesis, Clays and Clay Minerals, v. 27, pp. 313-321.

Toyoda, K., and Masuda, A., 1990, Sedimentary environments and chemical composition of Pacific pelagic sediments. Chemical Geology, v. 88, pp. 127-141.

Trabelsi, A. M. S., 1990, Sedimentology, petrography, and diagenesis of selected Paleozoic source rocks from the Permian basin, unpublished Ph. D. dissertation. Department of Geosciences, Texas Tech University, 222 pages.

Twenhofel, W. H., 1939, Environments of origin of black shales, American Association of Petroleum Geologists Bulletin, v. 23, pp. 1178-1198.

Vine, J. D., and Tourtelot, E. B., 1970, Geochemistry of black shale deposits - A summary report. Economic Geology, v. 65, pp. 253-272.

Von Amd, P., (1968), Ein neues Verfahren zur Bestimmung von Eisen (II) oxid in Mineralen und Gesteinen, Neues Jahrbuch Fiir Mineralogie, Monatshefte, pp. 119-125.

Weaver, C. E., 1967, Potassium, illite and the ocean, Geochimica et Cosmochimica Acta, v. 31, pp. 2181-2196.

Wenger, L. M., and Baker, D. R., 1986, Variations in organic geochemistry of anoxic-oxic black shale-carbonate sequences in the Pennsylvanian ofthe midcontinent, USA, Organic Geochemistry, v. 10, pp. 85-92.

121

Wilde, P., Quinby-Hunt, M. S., Berry, W. B. N., and Orth, C. J., 1989, Paleo-oceanography and biogeography in the Tremadoc (Ordovician) lapetus Ocean and the origin ofthe chemostratigraphy o{ Dictyonema flabelliforme black shales. Geological Magazine, v. 126, pp. 19-27.

Yancey, T. E., and McLerran, R. D., 1988, Cyclic stratigraphy ofthe Late Pennsylvanian of north-central Texas, in Cunningham, B. K., editor, Permian and Pennsylvanian Stratigraphy, Midland Basin, West Texas, Sttidies to aid Hydrocarbon Exploration, Society of Economic Paleontologists and Mineralogists, Permian Basin Section, Research Seminar 1, pp. 65-77.

Yin, J. H., Kajiwara, Y., and Fujii, T., 1989, Distribution of transition elements in surface sediments ofthe southwestem margin of Japan Sea, Geochemical Joumal, V. 23, pp. 161-180.

Zangerl, R., and Richardson, E. S., 1963, The paleoecologic history of two Pennsylvanian black shales, Chicago Natural History Museum, Fieldana Geology Memoir, v. 4, 352 pages.

122

APPENDED A

PETROGRAPHY OF SELECTED SAMPLES

sample silty quartz mica (micron) flakes

dark brown clay/ organics/oxides

biogenic features grayish pyrite clay aggregates

EM-97

CCB-103

CCB-109

CCB-121

3027-15

3027-27

3027-38

UPS-11

UPS-23

UPS-35

DB-103

DB-113

P180-7

P180-9

TRR-25

TRR-33

TRR-40

LBK-10

LBK-15

LBK-20

slightly

slightly

yes

yes

slightly

slightly

slightly

very

very

yes

yes

no

yes

very

slightly

yes

yes

very

very

yes

20

25

30

20

30

30

20

40-50

40

20-50

30-50

30

20-40

10-50

10

10

10

50

10

15

few

few

few

few

few

few

few

few

few

few

few

few

many

many

sparse

many

many

many

many

some

many

many

many

many

many

many

many

many

many

many

many

many

few calcitic fossils

few fossils

some calcitic fossils

many calcitic fossils

many fossiIs

many calcitic fossils

many fossils

few fossils

some calcitic fossils

many, big calcitic fossi

burrows

many calcitic fossils, rare phosphatic fossils

many calcitic fossils

few fossils

many fossils

present

present

present

present

present

present

present

present

Is

present

present

present

present

present

present

present

present

present

present

present

present

123

APPENDDC B

GEOCHEMICAL DATA

Legend for Appendix B

LOI = Loss on Ignition at 550°C based on dried samples.

Fe203 = Total Iron as Fe203.

UNIT = Stratigraphic Unit.

Munsell = color code based on Munsell color chart.

Tit-TOC = Total Organic Carbon by titration.

LOI-TOC = Total Oiganic Carbon by Loss on Ignition.

LECO-TOC = Total Organic Carbon by LECO analyzer.

FeO/Fe203 = Ratio of FeO to total Fe203.

X = X-ray diffraction done on sample.

IntStd mean = The mean ofthe intemal standard.

IntStd S.D. = The standard deviation ofthe intemal standard.

124

Sample

EM-127

EM-125

EM-123

EM-121

EM-119

EM-117

EM-113

EM-109

EM-105

EM-103

EM-101

EM-99

EM-98

EM-97

EM-96

EM-95

CCB-127

CCB-126

CCB-125

CCB-124

CCB-123

CCB-122

CCB-121

CCB-119

CCB-117

CCB-115

CCB-113

CCB-111

CCB-109

CCB-107

CCB-105

CCB-103

CCB-102

CCB-101

3027-38

3027-35

3027-31

3027-27

3027-23

3027-19

3027-15

3027-13

3027-11

UNIT

5 5 4 4 4 4 3 3 3 3 3

7 6 6 6 6 5 5 5 5 5 5 4 4 4 4 3 3 3

4 4 4 4 4 3 3 3 2

Munsell

N5 N4 N3 N2.5

N2.5

N2.5

N2.5

N2.5

N2.5

N3 N4 N5 N5 N5 N5 N5

2.5Y 6/2

7.5YR 4/2

7.5YR 4/2

7.SYR 4/2

7.5YR 4/2

5YR 6/3

5YR 6/3

5YR 6/4

5YR 6/4

5YR 6/4

N6 N6 N6 N6 N6 N6 N6 N6

N4 N4 N5 N5 N5 N5 N6 N6 N6

Si 02

(X)

59.76

56.73

54.80

58.54

58.35

58.37

57.65

58.59

58.86

56.70

60.72

58.61

57.35

62.33

59.19

59.26

60.27

58.93

61.11

59.72

65.42

56.04

55.29

52.88

54.69

54.54

65.94

59.49

59.23

55.02

55.65

54.14

59.48

52.23

57.02

46.24

56.07

57.05

58.77

56.45

54.97

56.26

62.80

Ti02

(X)

0.85

0.80

0.77

0.82

0.81

0.80

0.79

0.83

0.82

0.80

0.83

0.84

0.81

0.87

0.84

0.84

0.96

0.77

0.79

0.73

0.80

0.82

0.80

0.76

0.78

0.78

0.84

0.85

0.86

0.83

0.84

0.81

0.87

0.76

0.84

0.72

0.85

0.87

0.89

0.87

0.86

0.86

0.72

A1203

(X)

18.20

18.35

18.30

18.83

18.67

18.24

18.03

17.27

17.96

16.85

16.58

17.31

17.01

17.62

17.86

17.86

20.33

15.61

14.05

14.14

14.89

18.31

15.58

15.01

15.76

14.93

15.30

17.72

17.82

17.59

18.36

17.93

19.80

17.29

18.76

15.76

18.84

19.26

20.18

19.69

19.74

20.59

15.72

Fe203

(X)

6.00

6.41

6.34

7.15

6.42

6.60

6.20

6.01

5.75

5.67

5.81

5.17

5.80

3.18

5.24

4.14

3.56

3.88

4.10

1.50

2.42

6.42

5.48

5.67

5.57

6.48

6.02

6.10

5.83

5.61

5.96

5.94

5.56

4.98

4.96

4.49

4.78

5.19

4.80

7.17

7.67

7.68

3.04

MnO (X)

0.069

0.037

0.038

0.044

0.032

0.039

0.042

0.036

0.038

0.042

0.052

0.023

0.015

0.012

0.013

0.014

0.016

0.020

0.008

0.009

0.009

0.048

0.099

0.078

0.063

0.070

0.032

0.034

0.031

0.025

0.028

0.023

0.020

0.040

0.030

0.565

0.140

0.025

0.025

0.035

0.033

0.021

0.127

MgO (X)

1.76

1.74

1.80

1.95

1.78

1.71

1.61

1.95

1.93

2.44

1.66

1.16

0.99

1.01

1.01

1.11

1.03

0.81

0.64

0.61

0.60

1.32

1.39

1.54

1.62

1.57

1.44

1.76

1.68

1.75

1.77

1.69

2.51

1.62

1.71

5.13

2.34

1.63

1.83

1.77

1.82

2.53

1.34

CaO (X)

0.39

0.36

0.37

0.50

0.35

0.62

0.33

0.66

0.84

3.67

1.98

0.72

0.11

0.09

0.10

0.17

0.37

1.33

1.03

1.26

0.89

0.64

6.60

9.18

5.12

6.47

1.37

1.95

1.40

1.72

0.87

0.86

0.44

5.67

0.57

6.99

1.60

0.30

0.45

0.42

0.38

0.41

2.71

Na20

(X)

0.51

0.43

0.42

0.43

0.42

0.41

0.39

0.47

0.46

0.60

0.53

0.68

0.52

0.52

0.47

0.50

0.19

0.20

0.22

0.22

0.16

0.15

0.13

0.14

0.14

0.19

0.23

0.15

0.15

0.12

0.15

0.18

0.23

0.20

0.11

0.11

0.13

0.13

0.13

0.12

0.11

0.15

0.10

125

Sample

UPS-35

UPS-31

UPS-27

UPS-23

UPS-19

UPS-15

UPS-11

DB-113

DB-111

DB-109

DB-107

DB-105

DB-103

DB-101

P180-10

P180-9

P180-8

P180-7

P180-6

P180-5

P180-4

TRR-40

TRR-39

TRR-37

TRR-35

TRR-33

TRR-29

TRR-25

LBK-20

LBK-19

LBK-18

LBK-17

LBK-16

LBK-15

LBK-13

LBK-11

LBK-10

LBK-8

UNIT

3 2 2 2 2 2 2

4 4 4 4 3 3 3

4 4 3 3 3 2 2

4 4 4 4 4 4 4

4 3 3 3 3 3 3 3 2 1

IntStd mean

IntStd S .D.

Munsell

N5 N4 N3 N2.5

N2.5

N3 N5

N5 N5 N5 N5 N5 N5

5Y 5/2

N2.5

N3 N5 N6 N5 N3 N3

N2 N2 N2.5

N2 N2.5

N3 N3

5Y 4/2

NI NI NI N2 N2 N2 N2 N2 N2

Si 02

(X)

54.19

52.90

54.00

50.24

53.29

55.44

56.65

48.82

57.41

56.98

57.61

58.30

58.28

52.11

48.40

44.00

52.08

53.51

52.70

44.12

40.77

47.90

46.69

57.50

57.37

56.97

57.81

55.76

56.36

51.55

45.37

39.63

44.52

43.42

47.51

56.70

62.65

52.60

60.37

1.47

Ti02

(X)

0.79

0.79

0.80

0.74

0.80

0.82

0.83

0.68

0.80

0.82

0.82

0.81

0.81

0.73

0.70

0.68

0.74

0.80

0.80

0.69

0.65

0.65

0.59

0.76

0.75

0.82

0.78

0.76

0.73

0.78

0.613

0.59

0.62

0.63

0.69

0.75

0.78

0.64

0.88

0.02

A1203

(X)

18.08

17.97

18.24

16.97

17.40

17.57

17.61

14.41

17.46

17.35

17.49

17.71

17.58

16.16

14.39

13.87

15.16

18.47

18.35

15.73

14.26

14.13

14.39

16.72

17.36

17.34

17.02

16.96

17.52

16.35

14.74

12.73

14.56

13.23

15.23

17.13

16.92

13.80

18.66

0.40

Fe203

<X)

6.51

6.96

6.44

6.72

7.45

7.20

6.72

5.59

6.09

6.37

6.50

6.16

5.74

7.56

5.63

5.93

4.14

6.48

6.45

6.66

5.95

7.95

5.13

6.37

5.62

5.54

5.66

5.83

9.41

7.29

7.85

8.38

7.20

7.75

7.04

7.07

7.74

1.81

6.01

0.16

MnO (X)

0.022

0.035

0.034

0.090

0.051

0.050

0.087

0.085

0.051

0.033

0.049

0.037

0.030

0.143

0.065

0.134

0.045

0.043

0.033

0.190

0.074

0.021

0.033

0.061

0.036

0.025

0.059

0.053

0.066

0.066

0.059

0.053

0.035

0.053

0.037

0.059

0.026

0.023

0.029

0.004

MgO (X)

1.49

1.64

1.75

1.78

1.81

1.85

2.11

1.75

1.74

1.75

1.70

1.69

1.72

1.71

1.36

1.30

1.32

1.70

1.55

1.70

1.86

1.32

2.57

2.12

2.47

1.90

2.05

2.06

1.88

2.30

3.85

2.15

2.19

2.18

1.87

1.92

1.32

0.12

1.70

0.09

CaO (X)

1.05

0.43

0.54

4.19

0.83

1.29

2.36

11.10

1.41

0.92

0.97

0.49

0.78

5.40

8.77

6.92

0.19

0.37

0.60

7.48

11.50

1.59

2.16

1.58

0.46

0.17

3.42

3.02

0.55

1.11

2.04

3.34

1.69

3.17

2.28

2.15

0.79

1.07

1.50

0.06

Na20

<X)

0.25

0.27

0.27

0.26

0.38

0.39

0.38

0.40

0.51

0.46

0.49

0.50

0.52

0.46

0.37

0.35

0.47

0.38

0.37

0.32

0.29

0.49

0.55

0.57

0.71

0.64

0.55

0.51

0.85

0.61

0.632

0.43

0.62

0.53

0.62

0.61

0.48

0.09

0.14

0.01

126

Sample

EM-127

EM-125

EM-123

EM-121

EM-119

EM-117

EM-113

EM-109

EM-105

EM-103

EM-101

EM-99

EM-98

EM-97

EM-96

EM-95

CCB-127

CCB-126

CCB-125

CCB-124

CCB-123

CCB-122

CCB-121

CCB-119

CCB-117

CCB-115

CCB-113

CCB-111

CCB-109

CCB-107

CCB-105

CCB-103

CCB-102

CCB-101

3027-38

3027-35

3027-31

3027-27

3027-23

3027-19

3027-15

3027-13

3027-11

UNIT

5 5 4 4 4 4 3 3 3 3 3

7 6 6 6 6 5 5 5 5 5 5 4 4 4 4 3 3 3

4 4 4 4 4 3 3 3 2

K20 <X)

3.23

3.32

3.40

3.46

3.49

3.41

3.40

3.23

3.15

3.04

2.95

3.17

3.22

3.29

3.38

3.27

3.07

2.25

2.21

2.05

2.25

2.83

2.60

2.61

2.70

2.62

2.70

3.02

2.97

2.93

3.03

2.96

3.53

2.77

3.19

2.74

3.24

3.23

3.38

3.35

3.37

3.97

2.78

P205

(X)

0.115

0.089

0.088

0.105

0.081

0.111

0.077

0.129

0.181

0.200

0.061

0.044

0.050

0.041

0.042

0.064

0.063

0.030

0.034

0.032

0.057

0.063

0.225

0.130

0.110

0.157

0.101

0.138

0.148

0.104

0.095

0.090

0.103

0.091

0.159

0.224

0.163

0.100

0.115

0.166

0.101

0.102

0.558

LOI (X)

4.4 4.6 4.8 4.8 4.8 5.2 4.3 5.6 6.4 5.1 4.7 4.5 6.0 4.0 4.8 4.7

4.8 9.9 8.2 10.0

7.6 5.0 4.1 4.0 4.3 4.6 3.8 4.2 4.6 4.5 4.4 4.9 4.5 4.3

4.2 4.0 4.7 4.6 4.7 4.5 4.4 4.4 3.4

H20 (X)

2.0 2.7 3.3 2.6 3.4 3.1 4.0 2.9 2.8 2.5 2.3 3.6 4.7 4.6 4.3 5.4

3.9 4.0 4.1 4.3 2.8 4.6 2.0 2.1 2.5 2.6 2.5 2.9 3.7 3.6 4.0 3.7 3.1 3.2

3.0 2.5 3.1 3.2 2.7 3.8 2.6 2.9 2.2

Total

(X)

97.28

95.57

94.43

99.23

98.61

98.60

96.82

97.68

99.19

97.61

98.17

95.82

96.57

97.56

97.25

97.32

98.56

97.73

96.48

94.58

97.89

96.24

94.29

94.09

93.35

95.00

100.27

98.31

98.42

93.80

95.15

93.23

100.14

93.15

94.55

89.46

95.95

95.57

97.96

98.34

96.05

99.87

95.49

Sc

(Pprn)

17.4

17.7

17.7

16.1

15.9

16.2

15.9

15.8

15.1

13.8

14.6

15.0

16.5

15.8

16.3

16.1

17.2

14.8

12.9

13.3

15.2

15.8

14.2

13.9

15.1

13.8

13.6

15.8

16.4

16.7

17.1

17.2

15.9

14.1

16.2

14.0

16.7

16.7

18.0

17.6

17.4

15.7

13.9

V

(PPn»)

144 161 176 166 172 169 171 143 143 146 152 127 132 127 130 131

134 115 90 94 113 147 133 119 138 126 125 142 148 151 158 164 176 157

170 126 160 172 171 177 185 183 162

Cr

(PP«n)

93 95 99 94 97 109 107 121 135 123 105 84 87 81 87 83

98 76 63 80 94 86 87 83 89 85 80 90 94 99 101 107 116 114

112 82 111 112 133 159 170 154 146

Co

(PPm)

13.1

13.2

12.5

12.8

11.3

12.0

17.1

11.3

11.7

16.7

16.7

10.3

7.5 7.9 7.4 8.2

22.2

25.8

13.3

10.7

8.2 13.5

11.2

8.6 13.2

12.8

12.2

9.0 12.2

10.6

15.6

12.1

15.1

9.9

14.8

14.4

15.5

18.2

14.6

19.7

22.0

23.5

12.2

127

Sample

UPS-35

UPS-31

UPS-27

UPS-23

UPS-19

UPS-15

UPS-11

DB-113

DB-111

DB-109

DB-107

DB-105

DB-103

DB-101

P180-10

P180-9

P180-8

P180-7

P180-6

P180-5

P180-4

TRR-40

TRR-39

TRR-37

TRR-35

TRR-33

TRR-29

TRR-25

LBK-20

LBK-19

LBK-18

LBK-17

LBK-16

LBK-15

LBK-13

LBK-11

LBK-10

LBK-8

UNIT

3 2 2 2 2 2 2

4 4 4 4 3 3 3

4 4 3 3 3 2 2

4 4 4 4 4 4 4

4 3 3 3 3 3 3 3 2 1

IntStd mean

IntStd S .D.

K20 (X)

3.23

3.16

3.19

3.04

3.21

3.16

3.09

2.67

3.11

3.02

2.98

3.03

3.04

2.76

2.27

2.16

2.42

2.94

2.97

2.49

2.34

2.89

2.85

3.32

3.72

3.40

3.13

3.18

3.23

3.23

3.08

2.59

2.72

2.68

2.99

3.29

3.20

0.25

3.11

0.16

P205

(X)

0.083

0.121

0.099

0.828

0.316

0.608

0.726

0.087

0.108

0.101

0.101

0.090

0.070

0.243

0.838

1.310

0.087

0.154

0.159

0.801

1.000

0.597

0.108

0.125

0.166

0.117

0.097

0.099

0.368

0.122

0.138

0.043

0.342

0.066

0.679

0.126

0.317

0.466

0.124

0.011

LOI (X)

4.5 4.5 4.7 4.9 4.8 5.0 4.6

3.2 3.7 3.7 3.8 3.8 3.9 3.7

5.3 4.9 5.9 6.7 6.7 4.5 4.1

15.6

15.7

6.2 7.9 9.1 4.7 5.8

3.8 13.4

18.1

16.8

14.8

14.7

12.9

7.4 5.2

41.3

H20 (X)

3.4 4.0 3.6 3.3 6.7 3.4 2.6

2.1 2.6 4.2 2.6 3.0 2.8 2.3

3.0 3.6 3.3 3.6 3.7 3.8 4.7

2.8 1.9 2.2 2.3 2.1 2.3 1.9

2.7 1.4 1.4 1.2 1.4 1.0 1.6 2.0 2.3 0.1

Total

(X)

93.59

92.77

93.66

93.06

97.04

96.78

97.76

90.89

94.99

95.69

95.10

95.62

95.27

93.27

91.10

85.15

85.85

95.14

94.38

88.48

87.48

95.94

92.67

97.52

98.85

98.12

97.57

95.94

97.47

98.21

97.88

87.93

90.69

89.41

93.44

99.21

101.73

112.26

92.54

Sc (ppm)

16.2

16.4

16.0

15.5

15.1

15.9

15.2

12.5

15.4

16.0

16.0

15.7

15.6

13.3

12.9

13.5

14.4

16.2

15.2

14.2

12.1

10.8

11.6

15.2

14.8

16.4

14.5

15.4

16.6

17.7

14.4

12.8

12.4

13.1

14.7

14.6

14.1

2.0

16.4

0.7

V

(PPn»)

159 152 157 144 152 153 142

114 153 160 157 151 154 138

100 98 109 118 115 96 88

568 578 167 171 165 148 172

167 397 582 1534

1232

1299

921 243 116 45

150

5

Cr

(PPn)

110 110 109 175 182 295 331

70 85 87 86 85 87 98

50 49 60 65 65 49 43

1520

1371

151 201 498 123 152

94 579 1127

979 567 712 332 143 93 16

95

8

Co

(ppm)

9.9 11.5

9.7 15.2

13.3

19.0

28.0

9.3 14.1

10.1

14.0

12.4

10.1

13.2

13.0

21.7

13.3

12.2

12.6

25.4

11.4

10.9

12.5

20.3

20.2

15.3

16.9

15.9

25.4

29.3

26.5

20.2

15.7

15.5

16.9

16.2

24.6

6.2

14.5

1.5

128

Sample

EM-127

EM-125

EM-123

EM-121

EM-119

EM-117

EM-113

EM-109

EM-105

EM-103

EM-101

EM-99

EM-98

EM-97

EM-96

EM-95

CCB-127

CCB-126

CCB-125

CCB-124

CCB-123

CCB-122

CCB-121

CCB-119

CCB-117

CCB-115

CCB-113

CCB-111

CCB-109

CCB-107

CCB-105

CCB-103

CCB-102

CCB-101

3027-38

3027-35

3027-31

3027-27

3027-23

3027-19

3027-15

3027-13

3027-11

UNIT

5 5 4 4 4 4 3 3 3 3 3

7 6 6 6 6 5 5 5 5 5 5 4 4 4 4 3 3 3

4 4 4 4 4 3 3 3 2

Ni (ppm)

73.5

60.4

61.3

59.3

79.3

91.3

70.3

112.0

138.0

90.4

85.8

24.5

25.0

20.3

27.8

28.2

28.6

85.5

37.9

58.5

145.0

50.3

51.2

48.7

62.1

52.8

46.5

46.1

50.3

56.9

67.4

68.4

70.7

68.5

69.1

60.1

74.7

71.8

74.6

151.0

102.0

106.0

76.2

Cu (ppm)

26.3

24.8

26.7

18.8

22.4

23.8

25.3

27.3

34.4

41.2

24.9

14.5

36.4

15.5

27.1

28.4

16.0

27.9

25.0

23.7

35.6

15.5

13.4

14.6

16.2

16.8

14.6

21.2

23.3

27.0

28.3

28.3

34.7

19.9

338.0

15.6

24.6

64.1

140.0

13.2

10.1

56.0

53.2

Zn

(ppm)

118 140 165 199 230 237 285 269 277 292 283 80 109 72 68 84

161 336 117 96 86 328 186 231 168 157 159 124 110 132 140 171 150 159

143 118 133 137 158 189 228 269 189

Be (ppm)

2.84

2.93

3.00

2.85

2.81

2.88

2.81

2.64

2.56

2.67

2.53

2.24

2.28

2.28

2.31

2.30

2.45

3.46

1.96

2.46

2.09

2.76

2.17

2.18

2.42

2.26

2.31

2.50

2.60

2.68

2.69

2.74

3.00

2.44

2.72

2.00

2.52

2.63

2.90

3.02

3.08

3.24

2.22

Sr

(ppm)

146 148 135 143 146 153 148 135 148 190 153 279 221 157 213 382

161 174 168 166 157 129 194 224 191 221 149 172 165 142 137 135 159 178

136 229 158 138 144 143 143 150 157

Ba

(ppm)

418 398 389 397 392 375 371 351 342 339 335 426 546 415 411 410

438 325 309 279 318 381 374 521 358 347 373 408 410 385 399 389 430 352

377 310 385 397 404 402 391 403 374

Zr

(ppm)

175 156 147 160 165 163 167 183 186 186 219 173 169 178 160 162

179 225 225 187 234 147 152 142 146 162 229 160 166 158 146 142 164 131

163 141 164 165 173 171 162 168 165

Y

(ppm)

27.2

24.0

23.4

27.1

25.6

29.0

22.2

29.4

35.6

44.9

25.2

19.8

22.0

22.6

20.5

24.2

29.6

104.0

23.8

25.3

24.3

19.7

34.3

27.4

24.9

29.1

27.6

25.9

25.5

25.1

23.4

22.9

29.6

19.6

26.4

36.3

27.9

22.0

24.7

27.9

22.9

27.6

39.2

Rb

(ppm)

162 175 184 188 162 180 182 163 156 141 143 162 152 161 157 156

157 116 104 102 112 164 143 134 143 133 133 163 159 159 169 165 177 153

183 144 165 186 194 191 196 194 138

129

Sample

UPS-35

UPS-31

UPS-27

UPS-23

UPS-19

UPS-15

UPS-11

DB-113

DB-111

DB-109

DB-107

DB-105

DB-103

DB-101

P180-10

P180-9

P180-8

P180-7

P180-6

P180-5

P180-4

TRR-40

TRR-39

TRR-37

TRR-35

TRR-33

TRR-29

TRR-25

LBK-20

LBK-19

LBK-18

LBK-17

LBK-16

LBK-15

LBK-13

LBK-11

LBK-10

LBK-8

UNIT

3 2 2 2 2 2 2

4 4 4 4 3 3 3

4 4 3 3 3 2 2

4 4 4 4 4 4 4

4 3 3 3 3 3 3 3 2 1

IntStd mean

IntStd S. .D.

Ni

(ppm)

81.2

146.0

106.0

147.0

165.0

195.0

175.0

95.8

75.0

70.9

74.7

69.2

63.8

109.0

51.9

93.3

47.4

47.2

43.9

76.5

46.8

138.0

335.0

139.0

149.0

109.0

90.5

99.7

86.2

278.0

587 304.0

309.0

252.0

242.0

126.0

86.1

7.0

63.6

6.2

Cu

(ppm)

20.1

26.6

25.5

39.8

37.0

42.8

46.6

14.9

16.7

28.7

24.8

18.2

18.7

18.3

22.3

25.2

21.0

19.4

18.5

21.9

24.1

53.6

84.3

37.7

47.3

38.6

27.6

32.1

32.0

45.7

94.9

124.0

95.4

102.0

71.5

51.7

30.5

15.5

26.8

3.0

Zn

(ppm)

187 234 233 281 245 282 278

158 138 130 140 142 140 190

53 63 66 76 74 85 55

138 454 95 160 220 87 143

158 287 1155

1516

1316

1575

984 238 124 35

115

8

Be

(ppm)

2.50

2.53

2.52

2.32

2.56

2.64

2.48

2.25

2.43

2.60

2.53

2.45

2.49

2.17

1.93

2.02

2.34

2.49

2.38

2.27

1.84

2.14

2.41

2.81

3.24

3.38

2.64

2.81

2.54

3.24

2.83

2.59

2.72

2.15

2.71

2.54

2.03

0.86

2.66

0.12

Sr

(ppm)

147 124 122 169 125 131 151

344 156 144 142 141 138 216

197 190 67 105 102 187 219

142 100 139 116 120 165 161

188 154 120 167 162 165 229 221 321 114

171

11

Ba

(ppm)

381 379 388 348 376 379 371

482 373 374 382 370 406 734

289 297 308 361 386 343 317

320 289 355 353 354 362 378

499 398 301 328 374 364 386 383 364 114

408

22

Zr

(ppm)

151 153 157 150 155 168 177

134 154 158 162 157 158 141

171 158 165 153 160 150 142

130 126 148 147 154 148 140

135 153 113 116 125 131 139 144 200 15

178

4

Y

(ppm)

23.1

24.4

22.1

59.5

24.8

38.1

45.8

21.8

21.1

22.4

20.8

20.2

19.7

23.5

30.1

43.2

19.7

18.0

19.4

40.6

30.7

27.9

24.6

24.7

35.0

19.7

22.4

19.6

27.9

21.6

21.9

12.6

31.8

14.9

40.7

18.8

32.2

33.5

29.2

3.6

Rb

(ppm)

157 173 168 161 164 161 153

135 172 166 172 166 167 151

138 135 150 162 164 139 137

142 133 169 178 174 151 167

155 143 141 117 131 119 137 155 149 10

179

3

130

Sample

EM-127

EM-125

EM-123

EM-121

EM-119

EM-117

EM-113

EM-109

EM-105

EM-103

EM-101

EM-99

EM-98

EM-97

EM-96

EM-95

CCB-127

CCB-126

CCB-125

CCB-124

CCB-123

CCB-122

CCB-121

CCB-119

CCB-117

CCB-115

CCB-113

CCB-111

CCB-109

CCB-107

CCB-105

CCB-103

CCB-102

CCB-101

3027-38

3027-35

3027-31

3027-27

3027-23

3027-19

3027-15

3027-13

3027-11

UNIT

5 5 4 4 4 4 3 3 3 3 3

7 6 6 6 6 5 5 5 5 5 5 4 4 4 4 3 3 3

4 4 4 4 4 3 3 3 2

Tit-TOC

(X)

0.6 0.7 0.9 0.9 1.0 0.8 0.4 1.6 1.9 1.6 0.9 0.4 1.2 0.7 0.5 0.6

0.3 3.6 2.7 3.7 2.3 0.2 0.2 0.3 0.3 0.4 0.3 0.3 0.4 0.4 0.3 0.3 0.3 0.2

0.4 0.6 0.6 0.4 0.4 0.3 0.4 0.2 0.1

LOI-TOC LECO

(X)

2.1 2.2 2.3 2.3 2.3 2.4 2.0 2.6 3.0 2.4 2.2 2.1 2.8 1.9 2.3 2.2

2.3 4.6 3.8 4.7 3.6 2.3 1.9 1.9 2.0 2.2 1.8 2.0 2.2 2.1 2.1 2.3 2.1 2.0

2.0 1.9 2.2 2.2 2.2 2.1 2.1 2.1 1.6

-TOC

(X)

1 1 1 0

0 0 0

.9

.8

.7

.9

.4

.5

.3 0.4 0

0 0 0 0

.4

.5

.3

.2

.2

S (X)

0 0 0 0

0 0 0 0 0

0. 0. 0. 0.

.048

.067

.057

.257

.031

.025

.027

.143

.043

036 037 045 063

FeO (X)

3.76

2.83

3.66

3.71

2.56

2.06

0.99

3.66

3.46

3.35

2.46

0.83

1.00

0.50

0.42

0.57

0.42

1.95

2.05

1.82

1.62

0.90

1.37

1.18

1.53

1.16

1.80

1.81

2.26

1.93

1.61

0.84

1.75

2.05

2.05

1.85

1.75

1.25

1.35

1.57

2.48

Fe0/Fe203

(ratio)

0.70

0.49

0.64

0.55

0.44

0.35

0.18

0.68

0.67

0.66

0.47

0.18

0.19

0.17

0.09

0.15

0.13

0.56

0.55

0.84

0.28

0.18

0.27

0.24

0.26

0.21

0.33

0.34

0.45

0.36

0.30

0.19

0.39

0.51

0.48

0.40

0.41

0.19

0.20

0.23

0.91

X-ray

X

X

X

X

X

X

X

X

X

X

X

Chemical

Index of

Alterati

80.8

81.0

80.7

80.9

80.7

80.7

80.7

80.2

81.2

79.8

80.3

79.2

79.7

80.0

80.1

80.4

84.8

84.9

83.7

84.6

84.7

84.7

83.7

83.1

83.4

82.6

82.2

83.5

83.7

83.9

83.9

83.7

82.5

83.9

83.7

83.4

83.5

83.9

83.9

83.7

83.7

81.9

83.2

131

Sample UNIT Tit-TOC LOI-TOC LECO-TOC

(X) (X) (%)

UPS-35 UPS-31 UPS-27 UPS-23 UPS-19 UPS-15 UPS-11

DB-113 DB-111 DB-109 DB-107 DB-105 DB-103 DB-101

P180-10 P180-9

P180-8 P180-7

P180-6 P180-5

P180-4

TRR-40

TRR-39

TRR-37

TRR-35

TRR-33

TRR-29

TRR-25

LBK-20

LBK-19

LBK-18

LBK-17

LBK-16

LBK-15

LBK-13

LBK-11

LBK-10

LBK-8

3 2 2 2 2 2 2

4 4 4 4 3 3 3

4 4 3 3 3 2 2

4 4 4 4 4 4 4

4 3 3 3 3 3 3 3 2 1

IntStd mean

IntStd S.D.

0.3 0.5 0.6 0.8 0.7 1.0 0.5

0.2 0.4 0.3 0.3 0.4 0.3 0.2

2.5 2.0 0.8 0.7 1.0 1.7 1.2

10.9

11.3

3.7 4.6 5.1 2.3 3.1

0.8 11.4 22.3

19.9

17.7

15.7

13.4

5.8 2.7 50.0

0.4

0.1

2.1 2.1 2.2 2.3 2.3 2.3 2.2

1.5 1.7 1.7 1.8 1.8 1.8 1.7

2.5 2.3 2.8 3.1 3.1 2.1 1.9

7.3 7.4 2.9 3.7 4.3 2.2 2.7

1.8 6.3 8.50

7.9 6.9 6.9 6.1 3.5 2.4 19.4

10.4 18.655

16.0

13.5

11.1

4.2

0.961 2.385 4.345

2.980 2.290

2.075

S (X)

FeO (X)

2.07 2.40

2.85 1.97

1.77 3.86 2.38

1.51

3.06 1.93

2.17 1.93

Fe0/Fe203

(ratio)

0.35 0.38 0.49 0.33

0.26 0.60 0.39

0.30

0.56 0.34

0.37 0.35

X-ray

2.03 2.30 2.47

0.55 0.39 0.43

4.78 0.57

X

X

X

X

X X

Chemical Index of A l terat ion

82.2 82.3 82.4 82.0 80.9 81.2 81.6

80.3 80.6 81.2 81.3 81.2 80.9 81.2

82.4 82.6 81.8 82.9 82.7 83.0 82.6

78.2 78.3 78.7 77.0 78.6 79.9 79.8

78.1 78.4 77.1 78.4 78.6 77.8 78.2 78.9 79.9 97.1

1.63

0.14

132

APPENDIX C

STRATIGRAPfflC DISTRIBUTION

133

a

'v * C o

j Q

O

o c o

"o

31 2 3

2 -3 8 1 5

1 0 9 ""OS 101

3 8

3 0 2 7 - 3 8 U P S - 3 5 3 0 2 7 - 1 1

D by Titration

D B - 1 1 3 U P S - 1 1

I P 1 8 0 - 1 0 D B - 1 0 1 P180—••

-t- by Loss on Ignition

£ 3 C

E _3 < C

o o o

o o

3 0 2 7 - 3 8 U P S - 3 5 3027-11

D B - 1 1 3 U P S - 1 1

P I 8 0 - 1 O D B - 1 0 1 P180 —•*

Figure 48. Stratigraphic distribution oftotal organic carbon (TOC) in Upper Salesville (3027 and UPS), Lower Salesville (DB), and P180 sections. Base ofsection is on the right. Top: total organic carbon (weight percent). Bottom: total organic carbon/ aluminum ratio.

134

a

c o

o o

c C7>

o

1 2 6 1 2 4

— I

East Mountain —127

D by Titration

I Colony Creek—127 E M - 9 5

+ by Loss on Ignition

C C B - 1 0 1

E 3 C

E _3 <

o

o o

c C71

"5

l O I

East Mounta in—127 E M - 9 5 C C B - 1 0 1

Figure 49. Stratigr^hic distribution oftotal organic carbon (TOC) in East Mountain (EM) and Colony Creek at Brad (CCB) sections. Base of section is on the right. Top: total oiganic carbon (weight percent). Bottom: total organic carbon/aluminum ratio.

135

a.

*^ c o

SI

o o

c

"o

Tulso RR —40 T R R - 2 5

D by Titration

L B K - 8

by Loss on Ignition

E 3 C

"E _3 <

C O

o o c o

"5

7 -

6 -

5 -

3 -

2 -

1 -

T R R - 2 5 L B K - 8

Figure 50. Stratigraphic distribution oftotal organic carbon (TOC) in Tulsa RR (TRR) and Lost Branch (LBK) sections. Base of section is on the right. Top: total organic carbon (weight percent). Bottom: total organic carbon/aluminum ratio.

136

o

8 o

E _3

E 3

••D C o u

3 0 2 7 - 3 8 U P S - 3 5 3 0 2 7 - 1 1

D B - 1 1 3 U P S - 1 1

P 1 8 0 - 1 0 D B - 1 0 1 P 1 8 0 — 4

0 . 0 4

E 3 C

E _3

E 3

'c O

0 . 0 3 5 -

0 . 0 3

0 . 0 2 5

0 . 0 2

0 . 0 1 5

0 . 0 1

0 . 0 0 5

3 8 31

1 O 8

3 0 2 7 - 3 8 I U P S - 3 5 3 0 2 7 - 1 1

I D B - 1 1 3 U P S - 1 1

| P 1 8 0 - 1 0 D B - 1 0 1 P 1 8 0 — 4

Figure 51. Stratigraphic distribution of scandium and titanium in Upper Salesville (3027 and UPS), Lower Salesville (DB), and PI 80 sections. Base ofsection is or right. Top: scandium. Bottom: titanium.

137

on the

o o o o

E 3 C 'E _3 <

c o u in

1 .4

1 .3

1 .2

1 .1

1

0 .9

0 . 8

0 . 7 h

0 .6

0 .5

0 . 4

0 . 3

0 .2

O.I

O

12V25 1 0 7 1 0 3

Eas t M o u n t a i n — 1 2 7 E M - 9 5

Co lony C reek—127 C C B - 1 0 1

E 3 C

E _3 < E 3 'c a

0 . 0 4

0 . 0 3 5

0 . 0 3 -

0 . 0 2 5

0 . 0 2 -

0 . 0 1 5 -

0 . 0 1 -

0 . 0 0 5 -

East M o u n t a i n — 1 2 7 E M - 9 5 C C B - 1 0 1

Figure 52. Stratigraphic distribution of scandium and titanium in East Mountain (EM) and Colony Creek at Brad (CCB) sections. Base ofsection is on the right. Top: scandium. Bottom: titanium.

138

o

8 o E 3 C

E 3

TJ C D

1 .

1.3 -

1 .2 -

1 .1

1

0 .9

0 . 8

0 . 7

0 . 6

0 . 5

0 . 4 -

0 . 3 -

0 . 2

O.I

O

3 3 3 7 2 5

_ 4 0

—I Tulsa RR —40

— I Lost Branch —20

TRR—25 L B K - 8

E 3 C 'E _3

E 3 'c

0 . 0 4

0 . 0 3 5 -

0 . 0 3

0 . 0 2 5 -

0 . 0 2 -

0 . 0 1 5 -

0 . 0 1 -

0 . 0 0 5 -

Tulsa RR—40 T R R - 2 5 L B K - 8

Figure 53. Stratigraphic distribution of scandium and titanium in Tulsa RR (TRR) and Lost Branch (LBK) sections. Base ofsection is on the right. Top: scandium. Bottom: titanium.

139

E 3 C 'E 3

O

(O

3.5 -

3 -

2.5

2 -

1 .5 -

1 -

0.5 -

3027-38 UPS-35 3027—11

DB-113 UPS—11

P180—10 DB-101 P180—4

10

O O O O

E 3 C

E _3 < E 3

'c O u k.

3027-38 UPS-35 3027-11

DB—113 UPS-11

P180—10 DB-101 P180-4

Figure 54. Stratigraphic distribution of silicon and zirconium in Upper Salesville (3027 and UPS), Lower Salesville (DB), and PI 80 sections. Base ofsection is on the right. Top: silicon. Bottom: zirconium.

140

E 3 C 'E _3

< C

o

(/)

3 . 5 -

3 -

2 .5 -

2 -

1 .5

0 . 5

1 2 7

East M o u n t a i n — 1 2 7 E M - 9 5

Co lony Creek—127

0 3 101

C C B - 1 0 1

1 0

O

8 o E 3 C

'E _3 < E 3 ' c O

9 -

8 -

7 -

6 -

5 -

4 -

3 -

2 -

1 -

East Mounta in—127 E M - 9 5 C C B - 1 0 1

Figure 55. Stratigraphic distribution of silicon and zirconium in East Mountain (EM) and Colony Creek at Brad (CCB) sections. Base ofsection is on the right. Top: silicon. Bottom: zirconium.

141

E 3 C

E _3

o

3 . 5 -

4 0

2 . 5

1 . 5

0 . 5

3 7

2 0

Tulsa RR —40 Lost Branch —20 T R R - 2 5 L B K - 8

1 0

O O O O

E 3

"E _3

E 3

'c O

Tulsa RR—40 TRR—25 L B K - 8

Figure 56. Stratigraphic distribution of silicon and zirconium in Tulsa RR (TRR) and Lost Branch (LBK) sections. Base ofsection is on the right. Top: silicon. Bottom: zirconium.

142

E 3

E _3 <

O CD

UPS-35 3027-11

DB-113 UPS—11

P180-DB-101 P180—4

O O O O

E 3 C

E _3

E _3

!• v

3027-38 UPS—35 3027-11

DB-113 UPS-11

P180-10 DB-101 P180—4

Figure 57. Stratigraphic distribution of barium and beryllium in Upper Salesville (3027 and UPS), Lower Salesville (DB), and PI 80 sections. Base ofsection is on the right. Top: barium. Bottom: beryllium.

143

E 3 C 'E _3 < E 3 O

OQ

19

1 8 -

17 -

16

15

14

1 3

12

1 1

1 0

9

8

7

6

5

4

3

2

1

O

1 19

1 2 7

— I East Mounta in—127

7 1 0 3 - o a CLIOI . « 1 0 1

E M - 9 5 Colony Creek—127

C C B - 1 0 1

O

8 o E 3 C 'E _3

E _3

% CO

1 . 4

1 . 3 -

1 . 2 -

1 . 1

1 I -

0 .9

0 .8

0 .7

0 . 6

0 .5

0 . 4

0 . 3

0 . 2

O.I

O

1 2 6

East Mounta in—127 Colony Creek—127 EM —95 C C B - 1 0 1

Figure 58. Stratigraphic distribution of barium and beryllium in East Mountain (EM) and Colony Creek at Brad (CCB) sections. Base ofsection is on the right. Top: barium. Bottom: beryllium.

144

E 3 C

'E _3 < E

_3 * i _ O

m

Tulsa RR—40 TRR—25 L B K - 8

O O O O

E 3 C

'E _3

E _3

% V

CD

1 .4

1.3 -

1 .2

1 .1

1

0 .9

0 . 8

0 .7

0 . 6

0 .5

0 .4

0 . 3

0 . 2 -

0.1 -

O

- 4 0

1 7 3 3

2 5

Tulsa RR—40 Lost Branch —20 T R R - 2 5 L B K - 8

Figure 59. Stratigraphic distribution of barium and beryllium in Tulsa RR (TRR) and Lost Branch (LBK) sections. Base ofsection is on the right. Top: barium. Bottom: beryllium.

145

E 3 g "E _3

E 3

• ( / ) V)

_o "o Q.

0 . 2 6

0 . 2 4 -

0 . 2 2

0 . 2

0 . 1 8

0 . 1 6

0 . 1 4

0 . 1 2

O.I

0 . 0 8

0 . 0 6

0 . 0 4

0 . 0 2

O

3 8 I H ® 1 0 5 101

3 0 2 7 - 3 8 I UPS-3 0 2 7 - 1 1

• 3 5 I D B - 1 1 3 U P S - 1 1 D B -

| P 1 8 0 -•101

1 0 P 1 8 0 — 4

O

8 o E 3 C

"E _3

E 3

lo 3

_ 3 8

2 -

1 -

2 3 1 5

1 2 7 .1 1

19 i i 3 s ; E 5 e k i 0 5 „ i o i •'°

11

3 0 2 7 - 3 8 I U P S - 3 5 3 0 2 7 - 1 1

I D B - 1 1 3 U P S - 1 1

I P 1 8 0 - 1 0 D B - 1 0 1 P 1 8 0 — 4

Figure 60. Stratigraphic distribution of potassium and rubidium in Upper Salesville (3027 and UPS), Lower Salesville (DB), and PI 80 sections. Base ofsection is on the ri^t. Top: potassium. Bottom: mbidium.

146

E 3 C

'E _3

E 3

O Q.

0 . 2 6

0 . 2 4

0 . 2 2

0 . 2

0 . 1 8

0 . 1 6

0 . 1 4

0 . 1 2 h

0 .1

0 . 0 8

0 . 0 6

0 . 0 4

0 . 0 2

O

- 1 1 2 1 1 1 7 1 0 9 . 9 7

p i

East Mounta in—127 E M - 9 5

Colony Creek—127 C C B - 1 0 1

O O o o

E 3 C 'E _3 < E 3

3

111 , „ - , 1 0 3

2 -

East Mounto in—127 Colony Creek—127 E M - 9 5 C C B - 1 0 1

Figure 61. Stratigraphic distribution of potassium and mbidium in East Mountain (EM) and Colony Creek at Brad (CCB) sections. Base ofsection is on the nght. Top: potassium. Bottom: mbidium.

147

E 3 C

'E 3

E 3

O

- ^ ^ - ^

0 . 2 6

0 . 2 4 I -

0 . 2 2

0 . 2

0 . 1 8

0 . 1 6

0 . 1 4

0 . 1 2 l -

0 .1

0 . 0 8

0 . 0 6

0 . 0 4

0 . 0 2

O

3 5 4 0

3 9 3 7 1 8

1 7

2 5

Tulsa RR—40 Lost Branch —20 TRR—25 L B K - 8

O O O O

E 3 C

'E _3

E 3

• • o

'Xi 3

Tulsa RR —40 T R R - 2 5 L B K - 8

Figure 62. Stratigraphic distribution of potassium and mbidium in Tulsa RR (TRR) and Lost Branch (LBK) sections. Base ofsection is on the right. Top: potassium. Bottom: mbidium.

148

E 3 C 'E _3

E _3 '</) 0) c C7> D

0 . 4 5

0 . 4 -

0 . 3 5 -

0 . 3 -

0 . 2 5 -

0 . 2 -

0 . 1 5 -

O . I

0 . 0 5 -

U P S - 3 5 3 0 2 7 - 1 1

D B - 1 1 3 U P S - 1 1

P I 8 0 - 1 O D B - 1 0 1 P 1 8 0 — 4

E 3 C 'E 3

E _3 TJ O

m

0 . 0 9

0 . 0 8 -

0 . 0 7 -

0 . 0 6 -

0 . 0 5 -

0 . 0 4 -

0 . 0 3

0 . 0 2 -

0 . 0 1 -

3 0 2 7 - 3 8 U P S - 3 5 3 0 2 7 - 1 1

D B - 1 1 3 U P S - 1 1

P 1 8 0 - 1 0 D B - 1 0 1 P1B0—4

Figure 63. Stratigraphic distribution of magnesium and sodium in Upper Salesville (3027 and UPS), Lower Salesville (DB), and PI 80 sections. Base ofsection is on the right. Top: magnesium. Bottom: sodium.

149

E 3 C

'E _3 <

V c en o

0 . 4 5

0 . 4 -

0 . 3 5 -

0 . 3

0 . 2 5

0 . 2

0 . 1 5

O . I

0 . 0 5

1 0 3

1 2 7 2 5

— I East M o u n t a i n — 1 2 7

E M - 9 5 Co lony C reek—127

C C B - 1 O l

E 3 C

'E 3

\ E 3

"•D O

(/I

0 . 0 9

0 . 0 8 -

0 . 0 7 -

0 . 0 6

0 . 0 5 -

0 . 0 4 -

0 . 0 3

0 . 0 2

0 . 0 1

East M o u n t a i n — 1 2 7 E M - 9 5 CCB—101

Figure 64. Stratigraphic distribution of magnesium and sodium in East Mountain (EM) and Colony Creek at Brad (CCB) sections. Base ofsection is on the right. Top: magnesium. Bottom: sodium.

150

E 3 C 'E 3

E _3 '<0 V

c u 2

0 . 4 5

0 . 4 -

0 . 3 5 -

0 . 3 -

0 . 2 5 -

0 . 2 -

0 . 1 5 -

O.I -

0 . 0 5 -

Tu lsa RR —40 T R R - 2 5 L B K - 8

E 3 C

'E 3

E _3

O

0 . 0 9

0 . 0 8 -

0 . 0 7 -

0 . 0 6

0 . 0 5 -

0 . 0 4

0 . 0 3 -

0 . 0 2 -

0 . 0 1 -

Tu l sa RR—40 T R R - 2 5 LBK —8

Figure 65. Stratigraphic distribution of magnesium and sodium in Tulsa RR (TRR) and Lost Branch (LBK) sections. Base ofsection is on the right. Top: magnesium. Bottom: sodium.

151

'E _3

E 3

o

0 . 8

0 . 7 -

0 . 6 -

0 . 5

0 . 4 -

0 . 3 -

0 . 2 -

O.I -

U P S - 3 5 3 0 2 7 - 1 1

P 1 8 0 -

O O O O

E 3 C

'E _3

E _3 "c O

(75

U P S - 3 5 3 0 2 7 - 1 1

D B - 1 1 3 U P S - 1 1

P 1 8 0 - 1 0 DB—101 P 1 8 0 — 4

Figure 66. Stratigraphic distribution of calcium and strontium in Upper Salesville (3027 and UPS), Lower Salesville (DB), and P180 sections. Base ofsection is on the right Top: calcium. Bottom: strontium.

152

E 3 C

'E 3

E 3

0 . 8

0 .7 -

0 . 6 -

0 . 5 -

0 . 4 -

0 . 3 -

0 . 2 -

O.I -

East Mounta in—127 E M - 9 5 C C B - 1 0 1

8

E 3 C

E _3

E _3

o

(7)

15

1 4 -

13 -

12

1 1 -

1 0 -

9 -

8

7

6

5

4

3 -

2 -

1

O

9 5

_ 1 2 T 2 5

East Mounta in—127 E M - 9 5

Colony Creek—127

1 0 1

CCB —101

Figure 67. Stratigraphic distribution of calcium and strontium in East Mountain (EM) and Colony Creek at Brad (CCB) sections. Base ofsection is on the right. Top: calcium. Bottom: strontium.

153

E 3 C

'E _3

E 3

0 . 8

0 .7 -

0 . 6 -

0 . 5

0 . 4

0 . 3

0 . 2

O . I

2 9

4 0

2 0

Tu lsa RR —40 Los t B r a n c h —20 T R R - 2 5 L B K - 8

O O O O

E 3 C

E

£ 3 "c O

(75

15

1 4 -

13 -

12

1 1

1 0

9

8

7

6

5

4

3

2 I-

1

O

4 0

T Tu lsa RR—40

2 9 2 5 2 0

1 0

Lost B r a n c h —20 T R R - 2 5 LBK—8

Figure 68. Stratigraphic distribution of calcium and strontium in Tulsa RR (TRR) and Lost Branch (LBK) sections. Base ofsection is on the right. Top: calcium. Bottom: strontium.

154

E 3 C

E _3 <

V c C

0 . 0 1

0 . 0 0 9 -

0 . 0 0 8 -

0 . 0 0 7

0 . 0 0 6

0 . 0 0 5

0 . 0 0 4

0 . 0 0 3

0 . 0 0 2

0 . 0 0 1 *

1 3 0 2 7 - 3 8 I U P S - 3 5

3 0 2 7 - 1 1 I D B - 1 1 3

UPS—11 I P 1 8 0 - 1 0

D B - 1 0 1 P 1 8 0 — 4

E 3 C

'E _3

<

V)

a c o

o

0 . 4 5

0 .4 -

0 . 3 5 -

0 .3 -

0 . 2 5 -

0 . 2 -

0 . 1 5 -

O.I -

0 . 0 5

U P S - 3 5 3 0 2 7 - 1 1 P 1 8 0 — 4

Figure 69. Stratigraphic distribution ofmanganese and total iron in Upper Salesville (3027 and UPS), Lower Salesville (DB), and P180 sections. Base ofsection is on the right. Top: manganese. Bottom: total iron.

155

E 3 C

"E _3

V l/> V c o o> c o

0 .01

0 . 0 0 9 -

0 . 0 0 8 -

0 . 0 0 7 -

0 . 0 0 6 -

0 . 0 0 5

0 . 0 0 4

0 . 0 0 3

0 . 0 0 2

0 . 0 0 1

- 1 2 7

1 0 1

—I East Mounta in—127

E M - 9 5 Colony Creek—127

C C B - 1 0 1

E 3 C

E _3 <

V

V)

o c o

"o

0 . 4 5

0 . 4 -

0 . 3 5

0 . 3 -

0 . 2 5 -

0 . 2 -

0 . 1 5 -

O.I -

0 . 0 5

East Mounta in—127 E M - 9 5 C C B - 1 0 1

Figure 70. Stratigraphic distribution ofmanganese and total iron in East Mountain (EM) and Colony Creek at Brad (CCB) sections. Base ofsection is on the right. Top: manganese. Bottom: total iron.

156

E 3 C

'E <

C

c C7

0.01

0 . 0 0 9 -

0 . 0 0 8 -

0 . 0 0 7 -

0 . 0 0 6 -

0 . 0 0 5 -

0 . 0 0 4 -

0 . 0 0 3

0 . 0 0 2

0 .001

2 5

2 0 19 18 1 7 1 5

4 0

Tu lsa RR—40 Lost B r a n c h — 2 0 T R R - 2 5 L B K - e

E 3 c

'E _3 <

Q>

C o

13 "o

0 . 4 5

0 . 4 -

0 . 3 5 -

0 .3 -

0 . 2 5 -

0 .2 -

0 . 1 5 -

0.1 -

0 . 0 5 -

Tulsa RR—40 T R R - 2 5 L B K - 8

Figure 71. Stratigraphic distribution ofmanganese and total iron in Tulsa RR (TRR) and Lost Branch (LBK) sections. Base ofsection is on the right. Top: manganese. Bottom: total iron.

157

E 3 C

'E 3

V) 3

Q. (O O

D O S

0 . 0 7 -

0 . 0 6 -

0 . 0 5

0 . 0 4 -

0 . 0 3

0 . 0 2

0 . 0 1 1 0 1

1 1 3 1 0 9 1 0 5 B - e - B -

1 3 0 2 7 - 3 8 I U P S - 3 5

3 0 2 7 —1 1 I D B - 1 1 3

U P S - 1 1 I P 1 8 0 - 1 0

D B - 1 0 1 P 1 8 0 — 4

O O O O

E 3 C

'E 3

E 3

4 . 5

3 . 5 -

2 .5 -

1 .5 -

0 . 5 -

3 0 2 7 - 3 8 U P S - 3 5 3 0 2 7 - 1 1

D B - 1 1 3 U P S - 1 1

P I 8 0 — 1O D B - 1 0 1 P 1 8 0 — 4

Figure 72. Stratigraphic distribution of phosphorus and yttrium in Upper Salesville (3027 and UPS), Lower Salesville (DB), and P180 sections. Base ofsection is on the right. Top: phosphorus. Bottom: yttrium.

158

E 3 C

3 k_ O .c Q. in O

O 0 8

0 . 0 7 -

0 . 0 6 -

0 . 0 5 -

0 . 0 4 -

0 . 0 3

0 . 0 2

0 . 0 1 1 0 3

East Mounta in—127 E M - 9 5

CLP7 1 0 3 101 O • O •

CCB—101

4 . 5

O O O O

E 3 C

'E 3

E 3

4 -

3 .5 -

2 . 5

2 -

1 . 5

0 . 5 -

1 2 6

1 - • '27 1 2 1

East Mounta in—127 E M - 9 5

Colony Creek—127 C C B - 1 0 1

Figure 73. Stratigraphic distribution of phosphoms and yttrium in East Mountain (EM) and Colony Creek at Brad (CCB) sections. Base ofsection is on the right. Top: phosphorus. Bottom: yttrium.

159

E 3

'E 3

V) 3

Q. (A O

0 . 0 8

0 . 0 7

0 . 0 6 -

0 . 0 5 -

0 . 0 4

0 . 0 3

0 . 0 2

0 . 0 1

4 0

2 0

—I Tu lsa RR —40 Los t B r a n c h —20

T R R - 2 5 L B K - 8

O O o o

E 3

'c E

E 3

4 . 5

3 .5 -

3 -

2 . 5 -

2 -

1 .5

1 -

0 . 5 -

Lost Branch —20 TRR—25 L B K - 8

Figure 74. Stratigraphic distribution of phosphoms and yttrium in Tulsa RR (TRR) and Lost Branch (LBK) sections. Base ofsection is on the right. Top: phosphorus. Bottom: yttrium.

160

8

E 3 C

"E 3 5

3 0 2 7 — 3 8 U P S - 3 5 3 0 2 7 - 1 1 P 1 8 0 -

O o o o

E 3 C

E _3

E 3

•"U O c

3 0 2 7 - 3 8 U P S - 3 5 3 0 2 7 - 1 1

D B - 1 1 3 U P S - 1 1

PT 8 0 - 1 O D B - 1 0 1 P 1 8 0 — 4

Figure 75. Stratigraphic distribution of nickel and vanadium in Upper Salesville (3027 and UPS), Lower Salesville (DB), and PI 80 sections. Base ofsection is on the right. Top: nickel. Bottom: vanadium.

161

o o 8 E 3 C

'E 3

East Mounta in—127 E M - 9 5 C C B - 1 0 1

O O O O

E 3 C

E _3 < E 3

•v o c o >

East Mounta in—127 E M - 9 5 C C B - 1 0 1

Figure 76. Stratigraphic distribution of nickel and vanadium in East Mountain (EM) and Colony Creek at Brad (CCB) sections. Base ofsection is on the right. Top: nickel. Bottom: vanadium.

162

8 8

E 3 C

'E 3

4 0

3 5 -

3 0 -

2 5 -

2 0 -

15 -

10 -

5 -

Tu lsa RR —40 TRR—25 L B K - 8

E 3 C

'E _3 < E 3

o c

1 3 0

1 2 0 -

1 1 0 -

1 0 0

9 0

8 0

7 0

6 0

5 0

4 0

3 0

2 0

1 0

O

1 7

3 5 - B -

3 3 2 9 2 5 - Q B ^

-r Tulsa RR—40 Lost Bronch —20

T R R - 2 5 LBK —8

Figure 77. Stratigraphic distribution of nickel and vanadium in Tulsa RR (TRR) and Lost Branch (LBK) sections. Base ofsection is on the right. Top: nickel. Bottom: vanadium.

163

E 3 C

'E 3

E 3

'E

o o

18

17

16

15

14

13

12

1 1

10

9

8

7

6

5

4

3

2

1

O

101

3027-38 I UPS-35 3027-11

I DB-113 UPS—11

IP180-10 DB-101 P180—4

O o o o

E 3 C 'E _3

C M

3 0 2 7 - 3 8 U P S - 3 5 3027-11

D B — 1 1 3 U P S - 1 1

PI 80 — 10 D B - 1 0 1 P 1 8 0 — 4

Figure 78. Stratigraphic distribution of chromium and zinc in Upper Salesville (3027 and UPS), Lower Salesville (DB), and PI 80 sections. Base ofsection is on the right. Top: chromium. Bottom: zinc.

164

8

3 C 'E 3

E 3 'E o

o

18

17 -

16

I 5

14 -

13 -

12 -

I I -

1 0 -

9 -

8 -

7 -

6 -

5

4

3

2

1

O

1 0 3

- 1 2 V 2 5

East Mounta in—127

'^^J^\r 1 0 3

1 0 1

E M - 9 5 Colony Creek—127

C C B - 1 0 1

O O O O

c E _3

c M

East Mounta in—127 E M - 9 5

Colony Creek—^ 27 C C B - 1 0 1

Figure 79. Stratigraphic distribution of chromium and zinc in East Mountain (EM) and Colony Creek at Brad (CCB) sections. Base ofsection is on the right. Top: chromium. Bottom: zinc.

165

o o o o

E 3 C

'E 3

E 3

E

130

120

1 10

100

90 -

80 -

70 -

60 -

50

40

30

20

10

O

4 0

18 1 7

2 5

Tu lsa RR—40 T R R - 2 5

Los t B r a n c h —20 L B K - 8

O O O O

E 3 C

'E _3 <

1 0 0

9 0

8 0

7 0

6 0

5 0

4 0 -

2 5

Tulso R R - 4 0 Lost Bronch —20 T R R - 2 5 LBK —8

Figure 80. Stratigraphic distribution of chromium and zinc in Tulsa RR (TRR) and Lost Branch (LBK) sections. Base ofsection is on the right. Top: chromium. Bottom: zinc.

166

o o o o

E 3 C 'E _3 < O .o o o

1 .7

1 .6

1 .5

1 .4

1 .3

1 .2

1 .1

1

0 . 9

0 . 8

0 . 7

0 . 6

0 . 5

0 . 4

0 . 3

0 . 2

O . I

o

-

-

JSl^i / l ^cf^A

3 0 2 7 - 3 8

15 „

/ \ / \

J V K ^

1 3 0 2 7

3 1

U P S -- 1 1

- 3 5

1 1

IxF a 'A s r

I "I 1 D B - 1 1 3

U P S - 1 1

t \

i k 101 ° E w \ V '

I P I S O - i o D B - 1 0 1

9 I

S ^ r^

1 P 1 8 0 — 4

o o o o

E 3 C E _3 < t> Q. Q. O L>

8 -

7 -

6 -

5 -

4 -

3 -

1 -

3 0 2 7 - 3 8 U P S - 3 5 3 0 2 7 - 1 1

D B - 1 1 3 U P S - 1 1

P I 8 0 - 1 0 D B - 1 0 1 P 1 8 0 — 4

Figure 81. Stratigraphic distribution of cobalt and copper in Upper Salesville (3027 and UPS), Lower Salesville (DB), and PI 80 sections. Base ofsection is on the right. Top: cobalt. Bottom: copper.

167

E 3 C '£ _3 < O

XI o o

1 .7

1 .6 -

1 .5 -

1 . 4 -

1 .3

1 .2 -

1.1 -

1 -

0 .9 -

0 . 8

0 .7

0 .6

0 .5

0 . 4

0 . 3

0 .2

O.I

O

1 2 6

1 0 3

1 2 V 2 5

—I East Mounta in—127

E M - 9 5 Colony Creek—127

CCB—101

O O O O

E 3 C

E < V Q. O. O O

East Mounta in—127 E M - 9 5 C C B - 1 0 1

Figure 82. Stratigraphic distribution of cobalt and copper in East Mountain (EM) and Colony Creek at Brad (CCB) sections. Base ofsection is on the right. Top: cobalt. Bottom: copper.

168

o

8 o E 3 C

'E _3 < O

.O o o

1 .7

1 .6 -

1 .5 -

1 . 4 -

1 .3 -

1 .2 -

1.1 -

1 -

0 .9

0 . 8

0 . 7

0 . 6

0 . 5

0 . 4

0 . 3

0 . 2

O.I

O

19 18

3 7 3 5

4 0

Tulsa RR —40 —I Lost Branch —20 T R R - 2 5 L B K - 8

O O O O

£ 3 C

E _3 < a. a. o o

Tulso RR—40 Lost Branch —20 TRR—25 L B K - 8

Figure 83. Stratigraphic distribution of cobalt and copper in Tulsa RR (TRR) and Lost Branch (LBK) sections. Base ofsection is on the right. Top: cobalt. Bottom: copper.

169


Recommended