fluvial fan architecture, facies, and interaction with lake: lessons...

166
FLUVIAL FAN ARCHITECTURE, FACIES, AND INTERACTION WITH LAKE: LESSONS LEARNED FROM THE SUNNYSIDE DELTA INTERVAL OF THE GREEN RIVER FORMATION, UINTA BASIN, UTAH by Jianqiao Wang

Upload: others

Post on 03-Feb-2020

3 views

Category:

Documents


0 download

TRANSCRIPT

FLUVIAL FAN ARCHITECTURE, FACIES, AND INTERACTION WITH LAKE:

LESSONS LEARNED FROM THE SUNNYSIDE DELTA INTERVAL OF THE

GREEN RIVER FORMATION, UINTA BASIN, UTAH

by

Jianqiao Wang

c© Copyright by Jianqiao Wang, 2018

All Rights Reserved

A thesis submitted to the Faculty and the Board of Trustees of the Colorado School

of Mines in partial fulfillment of the requirements for the degree of Doctor of Philosophy

(Geology).

Golden, Colorado

Date

Signed:Jianqiao Wang

Signed:Dr. Piret Plink-Bjorklund

Thesis Advisor

Golden, Colorado

Date

Signed:Dr. M. Stephen Enders

Professor and HeadDepartment of Geology and Geological Engineering

ii

ABSTRACT

The Early Eocene Sunnyside Delta interval of the middle Green River Formation in the

Uinta Basin changes from fluvial channel and floodplain deposits to interbedded fluvial,

deltaic, and lacustrine deposits across 20 km in the Nine Mile Canyon. Outcrop measured

sections and photomosaics with GPS survey of excellent cliff face exposures are integrated

with areal mapping of channel dimensions, channel to floodplain ratio, and sedimentary

facies variability. This study identifies the Sunnyside Delta interval as a fluvial fan system,

composed of variable discharge river and floodplain deposits that basinward transitions into

deltaic, mixed and lake facies. Variable discharge signatures include the abundant Froude

supercritical flow and high deposition rate sedimentary structures, in-channel mudstones,

in-channel bioturbation and desiccation, and the low abundance of cross stratification. This

work identifies upstream and downstream bar-scale (macroform) accretion styles that were

not previously recognized, as variable discharge river macroforms are commonly referred to

as poorly developed or even non-existent. This work identifies low angle downstream dipping

accretion, steep upstream accretion, and vertical aggradation as some of the characteristic

accretion styles. The fluvial fan stratigraphy is characterized by multiple scales of upward

sandying and thickening successions with upward increasing channel to floodplain ratio,

channel size, the degree of channel amalgamation, and the proportion of floodplain splay

sandstones. The smallest scale upward sandying and thickening successions is recognized as

avulsion packages that form the building blocks of the stratigraphy. Larger scale successions

are likely to indicate lobe and fan progradation. The common avulsions also determines how

the fan interacts with Lake Uinta, as this work documents lateral transitions between channel

and floodplain, deltaic and lake facies across just a few hundred meters to a few kilometers.

Furthermore, some mouth bar deposits consist of alternating carbonate grainstones and

siliciclastic sandstones, and some abandoned channels are filled with dolomitic mudstones.

iii

All these transitions indicates a highly irregular shoreline, where fluvial and deltaic deposits

build out locally at the active channel locations, laminated dolomitic mudstones accumulate

in protected embayments or abandoned channels, and lime grainstones where lake’s wave

and current energy is high. We interpret these fluvial-lacustrine interactions in Sunnyside

Delta interval as a result of river avulsions and contemporaneous carbonate productions.

iv

TABLE OF CONTENTS

ABSTRACT .. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .iii

LIST OF FIGURES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ix

LIST OF TABLES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xii

ACKNOWLEDGEMENTS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xiii

DEDICATION .. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xv

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

1.1 Motivation. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1

1.2 Dissertation Organization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3

1.3 Reference Cited . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4

CHAPTER 2 VARIABLE DISCHARGE RIVER MACROFORMS IN SUNNYSIDEDELTA INTERVAL OF THE GREEN RIVER FORMATION,UINTA BASIN, UTAH .. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6

2.1 Abstract . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6

2.2 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7

2.3 Geological Setting and Background . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8

2.4 Dataset and Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9

2.5 Facies Groups . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10

2.5.1 Facies Group 1: Sandstones and Conglomerates withLong-wavelength, Low-angle Laminations with Foresets andBacksets, and Planar Laminations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10

2.5.2 Facies Group 2: Sandstones with Steep Foresets . . . . . . . . . . . . . . . . . . . . . . . . . . . 14

2.6 Channel Marcoforms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15

v

2.6.1 Large Vertically and Laterally Amalgamated Channel Lithosomes . . . . . . . 15

2.6.2 Large Channel Lithosomes with Upstream Migrating Macroforms . . . . . . . 17

2.6.3 Channel Lithosomes with Low Angle Downstream Accretion Sets. . . . . . . . 19

2.6.4 Amalgamated Channel Lithosomes with High Angle Accretion Sets . . . . . 20

2.6.5 Heterolithic Aggradational Channel Lithosomes . . . . . . . . . . . . . . . . . . . . . . . . . . . 21

2.6.6 Small Lenticular Channel Lithosomes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23

2.7 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24

2.7.1 Lateral Accretion Sets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24

2.7.2 Downstream Accretion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25

2.7.3 Upstream Accretion and the Hierarchy of Froude Supercritical FlowDeposits . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26

2.8 Conclusions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27

2.9 Reference Cited . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44

CHAPTER 3 VERTICAL AND LATERAL FACIES VARIATIONS IN FLUVIALFANS: EARLY EOCENE GREEN RIVER FORMATION, UINTABASIN . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55

3.1 Abstract . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55

3.2 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56

3.3 Geological setting and background . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57

3.4 Methods and dataset . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58

3.5 Facies Associations in Fluvial Channel-floodplain Lithosomes . . . . . . . . . . . . . . . . . . . . 59

3.5.1 Facies Association 1: Large Vertically and Laterally AmalgamatedSandy Channel and Floodplain Lithosomes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59

3.5.2 Facies Association 2: Laterally Amalgamated Sandy to HeterolithicChannel Accretion Set and Floodplain Lithosomes . . . . . . . . . . . . . . . . . . . . . . . 61

vi

3.5.3 Facies Association 3: Small Isolated Channel and FloodplainLithosomes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64

3.6 Vertical Trends . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65

3.7 Lateral Trends . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66

3.8 Discussion on Vertical and Lateral Trends . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68

3.9 Controls on Fluvial Fan Formation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69

3.10 Modern Analogue . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70

3.11 Comparison to Fluvial Fan Models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71

3.12 Conclusion. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72

3.13 Reference Cited . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94

CHAPTER 4 FLUVIAL-LAKE INTERACTIONS IN SUNNYSIDE DELTAINTERVAL, EOCENE GREEN RIVER FORMATION, UINTABASIN, UTAH .. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102

4.1 Abstract . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102

4.2 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103

4.3 Geological Setting . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 104

4.4 Stratigraphic Framework . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105

4.5 Dataset and Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105

4.6 Fluvial and Lacustrine Facies Associations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106

4.6.1 Fluvial Facies Association. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106

4.6.1.1 Facies Association 1.1: Sandstone to Heterolithic LenticularLithosomes with Accretion Sets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106

4.6.1.2 Facies Association 1.2: Interbedded Mudstones andSandstones with Variegated Color . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 108

4.6.2 Lacustrine Facies Associations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109

vii

4.6.2.1 Facies Association 2.1: Lime Grainstone . . . . . . . . . . . . . . . . . . . . . . . 109

4.6.2.2 Facies Association 2.2: Stromatolites . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109

4.6.2.3 Facies Association 2.3: Gastropod Wackestone . . . . . . . . . . . . . . . . . 110

4.6.3 Deltaic Facies Association . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 110

4.6.3.1 Facies Association 3.1: Sharp-based Tabular Sandstone . . . . . . . 110

4.6.3.2 Facies Association 3.2: Interbedded Siliciclastic andCarbonate Mudstone and Sandstone . . . . . . . . . . . . . . . . . . . . . . . . . . . . 111

4.6.4 Fluvial-lacustrine Mixed Facies Associations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 112

4.6.4.1 Facies Association 4.1: Mixed Siliciclastic-CarbonateAccretion Sets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 112

4.6.4.2 Facies Association 4.2: Sigmoidal Sandy GrainstoneMacroforms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 113

4.6.4.3 Facies Association 4.3: Mixed Siliciclastic-carbonateLenticular Mudstone . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114

4.6.4.4 Facies Association 4.4: Laminated Dolomitic Mudstone . . . . . . . 115

4.7 Lateral and Vertical Trends . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 115

4.8 Discussion on Fluvial-lake Interaction. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 116

4.9 Conclusions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 117

4.10 Reference Cited . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 142

CHAPTER 5 CONCLUSIONS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 150

viii

LIST OF FIGURES

Figure 2.1 Stratigraphic column and base map of Uinta Basin and Nine Mile canyon. 28

Figure 2.2 Representative examples of measured sections organized by channellithesome types. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29

Figure 2.3 Example outcrop photos and measured sections of Facies group 1:sandstones and conglomerates interpreted as Froude supercritical deposits. 30

Figure 2.4 Example outcrop photos and measured sections of Facies group 2:sandstones with steep foresets interpreted as Froude subcritical flowdeposits.. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31

Figure 2.5 Example photos of in-channel mudstones and bioturbation. . . . . . . . . . . . . . . . . . 32

Figure 2.6 Large vertically and laterally amalgamated channel lithosomes withlow-angle accretion sets. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33

Figure 2.7 Large channel lithosomes with upstream migrating macroforms. . . . . . . . . . . . . 34

Figure 2.8 Channel lithosomes with low-angle accretion sets. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35

Figure 2.9 Channel lithosomes with high-angle (up to 20◦) accretion sets withpreserved mudstone between sand bodies. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36

Figure 2.10 Heterolithic aggradational channel lithosomes.. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37

Figure 2.11 Small lenticular channel lithosomes. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38

Figure 2.12 Summary diagram of the 6 channel lithesome types in flow parallel (tothe left, arrow points downstream) and perpendicular views (right). . . . . . . . 39

Figure 3.1 Stratigraphic column and base map of Uinta Basin and Nine Mile canyon. 73

Figure 3.2 Outcrop examples of facies association 1. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74

Figure 3.3 Photos of common sedimentary structures in facies association 1. . . . . . . . . . . 75

Figure 3.4 Outcrop examples of facies association 2. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76

ix

Figure 3.5 Photos of common facies in facies associations 2 and 3. . . . . . . . . . . . . . . . . . . . . . . 77

Figure 3.6 Outcrop examples of facies association 3. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78

Figure 3.7 Photos of common facies in facies association 3. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79

Figure 3.8 Examples of characteristic vertical trends. Photos (A, B) and ameasured section (C) of upward thickening and sandying trends infloodplain deposits with capping channel lithosomes. . . . . . . . . . . . . . . . . . . . . . . . . 80

Figure 3.9 Vertical changes in channel (yellow) vs. floodplain (grey) proportions ateach examined location (by mile marker). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82

Figure 3.10 Photos showing thin floodplain-rich fluvial intervals along Road 191outcrops. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83

Figure 3.11 Example photomosaics that illustrate lateral changes from mostproximal to most distal locations, respectively (A through G). See Fig.3.1 for mile marker locations. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84

Figure 3.12 Google Earth image showing Ili River that originates from theTianshan Mountains and builds a large fluvial fan on the southernmargin of the Balkhash Lake. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88

Figure 4.1 Stratigraphic column and base map of Uinta Basin and Nine Milecanyon. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 118

Figure 4.2 Example outcrop photos of fluvial facies association 1. . . . . . . . . . . . . . . . . . . . . . 119

Figure 4.3 Outcrop photos of lime grainstone (FA2.1). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 122

Figure 4.4 Outcrop photos of stromatolite (FA2.2) and gastropod wackestone(FA2.3) facies associations. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 123

Figure 4.5 Outcrop photos of the deltaic facies association 3. . . . . . . . . . . . . . . . . . . . . . . . . . . 124

Figure 4.6 Vertical trends of facies associations. Two styles are presented in A andB with a typical measured section and an outcrop photo. . . . . . . . . . . . . . . . . . . 127

Figure 4.7 Outcrop photos of the mixed siliciclastic-carbonate accretion sets(FA4.1). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 128

Figure 4.8 Outcrop photos of sedimentary structures in oolitic sandstone (F26) inthe mixed siliciclastic-carbonate accretion sets (FA4.1). . . . . . . . . . . . . . . . . . . . . 129

x

Figure 4.9 Outcrop photos of sigmoidal sandy grainstone macroform (FA4.2). . . . . . . . 130

Figure 4.10 Outcrop photos of mixed siliciclastic-carbonate lenticular mudstone(FA4.3). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 131

Figure 4.11 Outcrop photos of laminated dolomitic mudstone (FA4.4). . . . . . . . . . . . . . . . . . 132

Figure 4.12 Depositional trends across the Nine Mile Canyon. . . . . . . . . . . . . . . . . . . . . . . . . . . . 133

Figure 4.13 Outcrop examples showing lateral and vertical transitions. . . . . . . . . . . . . . . . . . 134

Figure 4.14 Examples of lake-fluvial interaction in Ile River, Lake Balkhash. . . . . . . . . . . . 135

xi

LIST OF TABLES

Table 2.1 Description and Interpretation of Sedimentary Facies . . . . . . . . . . . . . . . . . . . . . . . . . . 40

Table 3.1 Description and Interpretation of Sedimentary Facies . . . . . . . . . . . . . . . . . . . . . . . . . . 89

Table 3.2 Description, Interpretation, and Comparison of Facies Associations. . . . . . . . . . . 93

Table 4.1 Description and Interpretation of Sedimentary Facies . . . . . . . . . . . . . . . . . . . . . . . . . 136

Table 4.2 Description and Interpretation of Sedimentary Facies Association . . . . . . . . . . . 142

xii

ACKNOWLEDGEMENTS

First and foremost, I would like to thank my advisor Piret Plink-Bjorklund, who has

taught me observation-based geology, critical thinking, and scientific writing. Her passion

and philosophy in science has inspired me in the field of sedimentology and stratigraphy. I

am grateful how accessible and helpful she is when I had difficulties with writing, formulating

ideas, rearranging papers. I won’t forget the countless hours she spent and how much work

she contributed in my thesis writing. Piret, thank you!

I would also like to thank my committee members: Donna Anderson, Rick Sarg, Yu-shu

Wu. Thank you for the support and confidence in this project. Thank you for the flexibility

in scheduling thesis proposal, comprehensive exams, committee meetings, and thesis defense!

Special thanks go to Donna Anderson and Kristine Zellman, who provided constructive

reviews for two manuscripts, which have significantly improved their quality.

I would like to thank Rick Sarg for encouraging me to explore the unfamiliar carbonate

world in my study area. The insights provided by Rick helped me understand the complexity

of the system. The world of carbonates humbled me as a geologist.

I very much appreciate all the thoughtful discussions with professors, colleagues, and in-

dustry professionals. They are Zane Jobe, Mary Carr, Lauren Schumaker, Theresa Schwartz,

Fabien Laugier, Jeremiah Moody, Kenya Ono, Evan Jones, Kristine Zellman, Haipeng Li,

Luke Pettinga, Mark Hansford, Marta Hodan, Sonia Campos Soto, and many other CoRE

and Dynamic Stratigraphy Group fellows.

This work won’t be done without the help from my field assistants. Thanks go to Jessica

Don, Chayawan Jaikla, Allison Ramsay, Evan Jones, Theresa Schwartz, Lauren Shumaker,

Zane Jobe, and Chiara White. Thank you for asking me questions and challenging my

interpretations in the field. I learned a lot from you.

xiii

Landowners of Nine Mile Ranch are very friendly. Thank you for the hospitality when we

camped in the ranch and the friendly warning about snakes when we were checking rocks. I

also thank Debora Cockburn, Dawn Umpleby, Dorie Chen, Cheryl Medford and Mary Carr

at Colorado School of Mines for providing field work logistics support.

This project is generously supported by Chevron ETC, BP scholarship, AAPG Grants-

in-aid (John and Erika Lockridge Named Grant), SEPM-RMS Fluvial Sedimentology Award,

and REPSEA.

I would like to express my gratitude to my parents for unconditionally supporting me

through all these years of study in the United States. I also thank my friends for the cheering,

shouting, crying, comforting, and all those unforgettable moments.

Lastly, I thank my incredible husband, Xiaopeng Li, for supporting me to overcome

difficulties, as well as enjoying the laughters and tears with me during the entire process. I

can’t imagine how all these can be accomplished without you.

xiv

This dissertation is dedicated to my parents, who taught me hard work and unconditionally

support me to pursue my interest in Geology. This work is also dedicated to my incredible

husband, Xiaopeng Li, for giving the strength and encouraging me to believe myself.

xv

CHAPTER 1

INTRODUCTION

1.1 Motivation

Large fluvial fans are significant unconfined depositional systems that are in need of

comprehensive facies and stratigraphic models and a better understanding of the system

dynamics, as well as a predictive reservoir model. There are currently multiple fluvial fan

models among which only the fluvial megafan model (e.g. Singh et al., 1993; Shukla et al.,

2001; Chakraborty and Gosh, 2010) consider avulsions rather than bifurcations as the key

mechanism for fan formation. All other fan models, such as the distributive fluvial systems

(DFS) (Weissmann et al., 2010, 2013), terminal fans (Kelly and Olsen, 1993), and fluvial

distributary systems (Nichols and Fisher, 2007). Interestingly, the facies and stratigraphic

model recently proposed for the DFS model (Davidson et al., 2013) is the same as the

fluvial megafans model proposed already by Singh et al. (1993). Furthermore, occurrence

of fluvial megafans has been linked to variable discharge rivers, and the control of variable

precipitation, such as today in the monsoon zone and subtropics (Leier et al., 2005). In

contrast DFS have been interpreted to occur in any climate types (Hartley et al., 2010).

The Sunnyside Delta interval of the middle Green River Formation in the Uinta Basin

was chosen for this study due to its laterally continuous exposure across ca 20 km in the

Nine Mile Canyon. The interval has been identified as a possible fluvial fan systems, and

deposits of variable discharge rivers, and linked to global warming during the Early Eocene

Climate Optimum (EECO) (Plink-Bjorklund and Birgenheier, 2012, 2013; Plink-Bjorklund

et al., 2014; Plink-Bjorklund, 2015).

Outcrops of the Sunnyside Delta Interval has been previously interpreted as a terminal

fan (Pusca, 2003) or deltaic deposits (Keighley et al., 2003; Schomacker et al., 2010; Moore

et al., 2012; Burton et al., 2014). This discrepancy elevates our interests to study what

1

causes the differences in interpretations.

The fluvial megafan model was firstly well illustrated by Shukla et al. (2001) using Ganga

Megafan. A systematic channel style change from braided, to anastomosing, to meandering

channels due to a decreasing discharge from proximal to distal part of the fan were observed

(Singh et al., 1993; Shukla et al., 2001). It is the most well developed model and has

numerous modern and ancient examples. Numerous studies in early Eocene fluvial outcrops

(Schmitz and Pujalte, 2007; Storey et al., 2009; Foreman et al., 2012; Plink-Bjorklund and

Birgenheier, 2012) propose that Paleocene Eocene Thermal Maximum (PETM) promotes

a dramatic change in water and sediment discharge and promotes formation of fuvial fans.

Fluvial megafans in Himalayan and central Andean foreland basins began to develop as the

Asian monsoon and seasonal precipitation started respectively (Leier et al., 2005). Highly

seasonal discharge triggers frequent avulsion and channel migration at times, thus the lateral

instability promotes fluvial fan formation (Leier et al., 2005).

This study shows the differences between channel and floodplain deposits and deltaic

mouth-bar deposits in the Sunnyside Delta Interval, and suggests that the depositional

intervals were in places incorrectly identified due to the unfamiliar combinations of sedi-

mentary structures and depositional architecture, not represented in current fluvial facies

models. The study confirms the Sunnyside Interval as a fluvial fans deposit, and further

documents the facies as well as bar-scale architecture of variable discharge rivers. The study

also documents that a considerable lateral variability in fluvial and lacustrine deposits occurs

where the fluvial fan enters the lake, as a result of frequent river avulsions characteristic for

fluvial fans. Considerable lateral variability also occurs in siliciclastic, mixed and carbonate

facies, as the river avulsions and the resultant fluvial sediment input to the lake interact

with contemporaneous carbonate production.

2

1.2 Dissertation Organization

The dissertation is organized into five chapters: introduction to the dissertation, three

papers to be submitted/submitted to peer reviewed journals, and general conclusions. Each

paper provides introduction, geological background, methodology and data, results, and

conclusions on its own. 1st paper (Chapter 2) has been submitted to Sedimentology, 2nd

paper (Chapter 3) has been submitted to Basin Research, and 3rd paper (Chapter 4) will be

submitted to Journal of Sedimentary Research in 2018. The figures, tables, and references

cited in each paper are included in the end part of each paper. A brief summary of three

papers and their contributions is provided below.

Chapter 2 (1st paper) recognizes variable discharge signatures in fluvial channel litho-

somes and identifies 6 types of macroforms varying from low angle dipping downstream

accretion to steep obliquely lateral or upstream accretion, and vertical aggradation. This

paper presents a first systematic description of such macroforms that have hitherto been re-

ferred to as poorly developed (Plink-Bjorklund, 2015) or even missing (Fielding et al., 2018)

in variable discharge rivers.

Chapter 3 (2nd paper) identifies a fluvial megafan system characterized by an avulsion

dominated stratigraphy by documenting the lateral extent, internal architecture, and lat-

eral and vertical facies transitions of channel-floodplain deposits. This paper compares the

results with a modern analogue and with the existing fluvial fan models such as megafans,

distributive fluvial systems, fluvial distributary systems, and terminal fan fluvial systems.

Chapter 4 (3rd paper) investigates the fluvial-lacustrine interaction and the carbonate-

siliciclastic mixing process by documenting the complex fluvial-lacustrine facies transition

styles and lateral changes in mixed siliciclastic-carbonate facies. These transitions are inter-

preted as a result of river avulsions and contemporaneous carbonate productions at a highly

irregular lake shoreline. The results caution the interpretation of transgressive-regressive cy-

cles that link vertical transitions between fluvial and lacustrine deposits to lake-level changes.

3

1.3 Reference Cited

Burton, D., Woolf, K., and Sullivan, B., 2014, Lacustrine depositional environments in theGreen River Formation, Uinta Basin: Expression in outcrop and wireline logs: AAPGBulletin, v. 98, p. 1699-1715.

Davidson, S.K., Hartley, A.J., Weissmann, G.S., Nichols, G.J., and Scuderi, L.A., 2013,Geomorphic elements on modern distributive fluvial systems: Geomorphology, v. 180-181, p. 82-95.

Foreman, B.Z., Heller, P.L., and Clementz, M.T., 2012, Fluvial response to abrupt globalwarming at the Palaeocene/Eocene boundary: Nature, v. 491, p. 92-95.

Hartley, A.J., Weissmann, G.S., Nichols, G.J., and Warwick, G.L., 2010, Large Distribu-tive Fluvial Systems: Characteristics, Distribution, and Controls on Development:Journal of Sedimentary Research, v. 80, p. 167-183.

Keighley, D., Flint, S., Howell, J., and Moscariello, A., 2003, Sequence stratigraphy inlacustrine basins: a model for part of the Green River Formation (Eocene), SouthwestUinta Basin, Utah, U.S.A.: Journal of Sedimentary Research, v. 73, p. 987-1006.

Kelly, S.B., and Olsen, H., 1993, Terminal fans-a review with reference to Devonian examples:Sedimentary Geology, v. 85, p. 339-374.

Leier, A.L., DeCelles, P.G., and Pelletier, J.D., 2005, Mountains, monsoons, and megafans:Geology, v. 33, p. 289-292.

Littler, K., Rohl, U., Westerhold, T., and Zachos, J.C., 2014, A high-resolution benthicstable-isotope record for the South Atlantic: Implications for orbital-scale changesin Late Paleocene-Early Eocene climate and carbon cycling: Earth and PlanetaryScience Letters, v. 401, p. 18-30.

Moore, J., Taylor, A., Johnson, C., Ritts, B.D., and Archer, R., 2012, Facies Analysis, Reser-voir Characterization, and LIDAR Modeling of an Eocene Lacustrine Delta, GreenRiver Formation, Southwest Uinta Basin, Utah, in Baganz, O.W., Bartov, Y., Bo-hacs, K.M., and Nummedal, D. eds., acustrine Sandstone Reservoirs and HydrocarbonSystems, American Association of Petroleum Geologists, Memoir 95, p. 183-208.

Nichols, G.J., and Fisher, J.A., 2007, Processes, facies and architecture of fluvial distributarysystem deposits: Sedimentary Geology, v. 195, p. 75-90.

Plink-Bjorklund, P., 2015, Morphodynamics of rivers strongly affected by monsoon precip-itation: Review of depositional style and forcing factors: Sedimentary Geology, p.110-147.

4

Plink-Bjorklund, P., and Birgenheier, L., 2012, Extreme Seasonality During Early EoceneHyperthermals, in American Geophysical Union, Fall Meeting 2012, abstract#GC21D-1005, http://adsabs.harvard.edu/abs/2012AGUFMGC21D1005P (accessedOctober 2017).

Pusca, V.A., 2003, Wet/Dry, Terminal fan-dominated sequence architecture: A new,outcrop-based model for the lower Green River Formation, Utah: University ofWyoming.

Schmitz, B., and Pujalte, V., 2007, Abrupt increase in seasonal extreme precipitation at thePaleocene-Eocene boundary: Geology, v. 35, p. 215.

Schomacker, E.R., Kjemperud, A.V., Nystuen, J.P., and Jahren, J.S., 2010, Recognition andsignificance of sharp-based mouth-bar deposits in the Eocene Green River Formation,Uinta Basin, Utah: Sedimentology, v. 57, p. 1069-1087.

Shukla, U.K., Singh, I.B., Sharma, M., and Sharma, S., 2001, A model of alluvial megafansedimentation: Ganga Megafan: Sedimentary Geology, v. 144, p. 243-262.

Singh, H., Parkash, B., and Gohain, K., 1993, Facies analysis of the Kosi megafan deposits:Sedimentary Geology, v. 85, p. 87-113.

Storey, M., Condon, D., Stecher, O., and Hald, N., 2009, On the PETM and ETM2global warming events: New evidence for a tectonic-magmatic trigger mechanism (In-vited), in American Geophysical Union, Fall Meeting 2009, Abstract ID. NH33B-1144,http://adsabs.harvard.edu/abs/2009AGUFMNH33B1144S (accessed May 2018).

Weissmann, G.S., Hartley, A.J., Nichols, G.J., Scuderi, L.A., Olson, M., Buehler, H., andBanteah, R., 2010, Fluvial form in modern continental sedimentary basins: Distribu-tive fluvial systems: Geology, v. 38, p. 39-42.

Weissmann, G.S., Hartley, A.J., Scuderi, L.A., Nichols, G.J., Davidson, S.K., Owen, A.,Atchley, S.C., Bhattacharyya, P., Chakraborty, T., Ghosh, P., Nordt, L.C., Michel,L., and Tabor, N.J., 2013, Prograding Distributive Fluvial Systems—GeomorphicModels and Ancient Exampless: SEPM Special Publication No. 104, p. 131-147.

5

CHAPTER 2

VARIABLE DISCHARGE RIVER MACROFORMS IN SUNNYSIDE DELTA INTERVAL

OF THE GREEN RIVER FORMATION, UINTA BASIN, UTAH

A paper submitted to Sedimentology

Jianqiao Wang*,1 and Piret Plink-Bjorklund1

2.1 Abstract

The channel lithosomes of the Sunnyside Delta Interval of the Green River Formation

in the Uinta Basin, Utah, USA display an abundance of Froude supercritical flow sedi-

mentary structures and high deposition rate sedimentary structures, and a low abundance

of dune cross strata, in-channel mud layers, and in-channel bioturbation and desiccation.

Such abundance or even dominance of Froude supercritical flow as well as high deposition

rate sedimentary structures has been recognized as a signature of variable discharge rivers,

where sediment transport and deposition is related to high magnitude floods, and base flow

discharge is characteristically low and commonly below the sediment movement threshold.

While this link between the abundance of Froude supercritical flow sedimentary structures

and variable river discharge is relatively well established, the bar-scale channel architecture

is essentially unknown, and is commonly described as poorly developed, or even absent.

Using excellent cliff face exposures, this study identifies a variety of bar-scale (macroform)

accretion styles that vary from low-angle downstream accretion to steep upstream accretion,

and vertical aggradation in six different types of channel lithosomes. The channel lithosomes

∗Primary and corresponding author. Direct correspondence to [email protected] Stratigraphy Group, Department of Geology and Geological Engineering, Colorado School ofMines, 1500 Illinois Street, Golden, Colorado 80401, USA.

6

vary in size and lithology and range from sandstones to interbedded sandstones and mud-

stones. The macroforms are linked with sedimentary facies that vary from abundant Froude

supercritical flow deposits to climbing ripple laminated deposits and in-channel mudstones.

2.2 Introduction

River discharge variability, as linked to a combination of seasonal and inter-annual vari-

ability, is suggested to be a first-order control on river morphodynamics (e.g., Fielding et al.,

2009, 2017; Plink-Bjorklund, 2015; Nicholas et al., 2016). As a result, variable discharge river

deposits differ from established fluvial facies models in terms of their meso- and macro-scale

bedforms (see e.g., Fielding et al., 2009; 2017; Plink-Bjorklund, 2015; Nicholas et al., 2016).

On the mesoform scale, a variable discharge river facies model is emerging that consists of

an abundance of Froude supercritical and transcritical flow sedimentary structures and a

reduced proportion of cross strata (e.g. Fielding et al., 2009, 2017; Plink-Bjorklund, 2015

and references therein). Considerably less data is available regarding the macroform (bar)

scale, and bars are considered to be significantly modified (Nicholas et al., 2016; Ghinassi

et al., 2017), poorly developed (see Plink-Bjorklund, 2015 and references therein) or even

absent (Fielding et al., 2018).

This paper documents outcrop examples of barforms of variable discharge river deposits

of the Eocene Green River Formation in Uinta Basin, Utah, US. These deposits are rec-

ognized as variable discharge rivers based on the abundance of Froude supercritical flow

and high deposition rate sedimentary structures, in-channel mud deposits, flood event beds,

in-channel bioturbation, and mud-clast conglomerates (see Plink-Bjorklund, 2015 and refer-

ences therein). A variety of barforms are documented that are distinct from the models for

braid bars (Ashley, 1990; Lunt et al., 2004; Lunt and Bridge, 2004; Sambrook Smith et al.,

2009) and point bars (Allen, 1965, 1970; Ethridge and Schumm, 1978; Tyler and Ethridge,

1983; Miall, 1985, 1988; Labrecque et al., 2011; Durkin et al., 2015), and provide docu-

mentation on their facies. These barforms range from low-angle tabular sets of flood-event

7

beds, to steep obliquely accreting upward coarsening sets, to large-scale backsets, to vertical

aggradation sets. This paper thus show that the macroforms are not absent, but rather differ

from the established facies models, and have thus not been readily recognized.

2.3 Geological Setting and Background

The Uinta basin is an intra-foreland basin formed to the east of the Sevier thrust front

(Fig. 2.1) in response to the Laramide uplift of the Uinta Mountains (Dickinson et al.,

1988; DeCelles, 2004). The basin is asymmetrical with a gentle southern margin and a

steep northern margin. The early to middle Eocene Green River Formation consists of

interbedded fluvial, deltaic and lacustrine deposits (Ryder et al., 1976) that are 0.6 to 1.8

km thick (Morgan and Bereskin, 2003). A large fluvial-deltaic system on the southeastern

basin margin was fed by the California Paleoriver, a ca 750 km long river that drained from

the Mojave region, rather than from local Laramide uplifts (Davis et al., 2010; Dickinson

et al., 2012). Flowing northward, the far-traveled California paleoriver is interpreted to

form a large fluvial fan system, including the Sunnyside Delta Interval (Plink-Bjorklund and

Birgenheier, 2013). Previously documented paleo-flow directions in Colton/Wasatch and

Green River Formations indicate a general northward direction (Fouch et al., 1976; Dickinson

et al., 1986; Remy, 1992; Schomacker et al., 2010; Ford et al., 2016; Gall et al., 2017; Jones,

2017). Two carbonate successions, the Uteland Butte Limestone and the Carbonate Marker

Unit are the first wide-spread lacustrine carbonates of the Green River Formation (Ryder

et al., 1976; Pitman et al., 1982; Burton et al., 2014) that are stratigraphically interbedded

with the fluvial deposits. The overlying Sunnyside Delta Interval of this study (Fig. 2.1) is a

ca 150 meters thick fluvial-deltaic and lacustrine deposit stratigraphically bound by younger

wide-spread carbonate intervals, the D Marker below and the C Marker above (Remy, 1992;

Keighley et al., 2003; Morgan, 2003). The age of the Sunnyside Delta Interval is constrained

by an absolute age date from the C Marker of 49.6 Ma (Smith et al., 2015), as well as the

49 Ma and 48.4 Ma dates below and above the Mahogany interval higher in stratigraphy

8

(Smith et al., 2010) (Fig. 2.1). The Sunnyside Delta Interval is below loosely constrained

by an older 54 Ma age in the Carbonate Marker Unit (Remy, 1992), and Paleocene fauna in

the lower part of the Colton/Wasatch Formation (Fouch et al., 1987) (Fig. 2.1).

The Sunnyside Delta Interval thus coincides with part of the Early Eocene Climatic Opti-

mum (EECO) (e.g. Littler et al., 2014). Changes in precipitation patterns in response to the

EECO global warming have been interpreted to result in highly seasonal and flashy discharge

(Plink-Bjorklund and Birgenheier, 2013; Plink-Bjorklund et al., 2014; Plink-Bjorklund, 2015;

Rosenberg et al., 2015; Gall et al., 2017).

Previous work (Jacob, 1969; Remy, 1992; Morgan, 2003) has documented the volumetric

abundance of fluvial deposits in the Sunnyside Delta Interval in the Nine Mile Canyon.

Pusca (2003) interpreted the extensive fluvial deposits as a terminal fan system (sensu Kelly

and Olsen, 1993; Legarreta and Uliana, 1998) and linked the fluvial deposits to wet-dry

climate cycles in a semi-arid conditions. Pusca (2003) was also the first to observe that

many of the sedimentary structures in Sunnyside Delta Interval indicate high deposition

rates. Recent work has mostly focused on the effects of variable discharge (Plink-Bjorklund

and Birgenheier, 2013; Plink-Bjorklund et al., 2014; Plink-Bjorklund, 2015; Rosenberg et

al., 2015; Gall et al., 2017), and on deltaic deposits (Schomacker et al., 2010; Moore et al.,

2012).

2.4 Dataset and Methods

In this outcrop study, about 550 m of stratigraphic sections at 5-10 cm resolution were

measured at 8 locations in Nine Mile Canyon (Fig. 2.1). Proportions of sandstone sedimen-

tary facies were calculated based on the measured-section dataset. A total of 111 paleocurrent

directions measurements were gained from parting lineations and the dip directions of cross

strata and laminae, or projected by the dip directions measured from the steepest surfaces of

scours related to cross strata (Fig. 2.1). Since the Sunnyside Delta Interval has an abundance

of Froude supercritical flow sedimentary structures that may dip down- or upstream and are

9

commonly gentle, the measurements are few, therefore we also use paleocurrent directions

measured throughout the Wasatch/Colton and Green River Formations and documented

in previous papers (Fouch et al., 1976; Dickinson et al., 1986; Remy, 1992; Schomacker et

al., 2010; Ford et al., 2016; Gall et al., 2017; Jones, 2017). These combined measurements

indicate a consistently northward direction with variability from northwest to northeast

(Fig. 2.1B). Dip directions of accretion sets were measured on exposed bedding surfaces

and plotted on stereonets. Sixteen photomosaics were used to address lateral sedimentary

facies variations and internal architecture in 2D and 3D exposures. A laser range finder

(TRUPULSETM 360B) was used for channel thickness and width measurements. Channel

widths were measured in outcrops as close to perpendicular to paleocurrent directions as

possible. This paper focuses on channel deposits only, and associated floodplain and distal

deltaic deposits are not documented here.

2.5 Facies Groups

Twelve sedimentary facies were defined using field observations on grain size, textures,

sedimentary structures, bar architectures and trace fossils (Table 2.1, Fig. 2.2). The facies

were then divided into two facies groups based on the sedimentary structures.

2.5.1 Facies Group 1: Sandstones and Conglomerates with Long-wavelength,

Low-angle Laminations with Foresets and Backsets, and Planar Lamina-

tions

This facies group is the most abundant, consisting of about 70% of facies in sandstone

and conglomerates of channel lithosomes. The facies consists of conglomerates (<2%, F

1.1) and mostly very fine to lower medium grained sandstones with scour and fill structures

(30%, F 1.2), ripple-filled scour and fill structures (<1%, F 1.3), convex-up low angle laminae

(8%, F 1.4), gradational planar laminae (14%, F 1.5), distinct planar laminae (20%, F 1.6),

structureless (1%, F1.7), and soft sediment deformation (2%, F1.8) (Table 2.1, Figs. 2.2 and

10

2.3). All sedimentary structures in group 1 are of long wavelength (>1m-10s of m) and have

gently dipping laminae (mostly <15◦). The thickness of lamina sets varies from a few dm to

a few m. The internal laminations are on mm to cm scale.

Facies 1.1 consists of erosionally bound conglomerates with sandy matrix (Figs. 2.3A

and B). Conglomerates contain mostly pebbles, but up to cobble sized limestone clasts and

bioclasts, and/or mudstone clasts occur. The clasts are subrounded (Fig. 2.3B) to angular

(Fig. 2.3A) and poorly sorted. Some clasts are elongate in shape and mostly a few cm to 2

dm wide. The clast composition is, in most cases, the same as the underlying siliciclastic or

carbonate deposits. The limestone intraclasts are mostly unorganized rather than stratified

and occur as layers a few cm thick in the upper portion of sandstone bodies. The mixed

limestone and mudstone clasts form a few cm thick conglomerates at the base of sandstone

bodies. In places, they are preserved as scour conformable laminae that flatten upward and

transition into planar or convex-up planar laminae with a wavelength of several meters, which

is essentially a hummock-like configuration. Clast imbrication is also common, and detrital

plant fragments are common in places at the base of basal scour surfaces. The conglomerate

facies occur throughout the Sunnyside Delta Interval.

Facies 1.2 consists of upper-very-fine to lower-medium grained, moderately to poorly

sorted sandstone with scour and fill structures with upward flattening laminae (Fig. 2.3C).

The laminae fill the scours asymmetrically in most cases with laminae dipping down- and

upstream. Some scours are filled symmetrically. Dip angle of the laminae varies from 5-25◦

but is in most cases gentler than 15◦. This facies occurs at the bases of sandstone bodies

above basal erosion surfaces, or within the sandstone bodies alternating with other facies.

Scour and fill structures have a wavelength from a few cm to 10s of m and thickness on cm to

m scale. In places, sets of scour and fill structures occur throughout the beds with thicknesses

from a few dm to a few m. Thicker, meter scale beds with coarser grained sandstones tend

to have larger scour and fill structures. In places, detrital organic matter is distributed along

the laminations.

11

Facies 1.3 consists of upper-very-fine to lower-fine grained moderately sorted sandstone

with scour and fill structures similar to Facies 1.2, but asymmetric ripple sets fill the upward

flattening laminae, or in places occur only at the lowest parts of the scour and fill structures

(Fig. 2.3D). In places, entire beds that are a few dm to a couple of m thick consist of this

facies only (Fig. 2.2), though not as commonly as Facies 1.2.

Facies 1.4 consists of sandstone with convex-up low angle laminations bounded by low-

angle scours (Fig. 2.3E). Laminae dip very gently (< ca 10◦) in down- and upstream direc-

tions. This facies is a few dm to 1m thick and the low-angle laminae have a wavelength of

a few meters. These structures are commonly of longer wavelength and lower amplitude as

compared to Facies 1.2 and 1.3. This facies commonly occurs together with scour and fill

structures (F1.2; F1.3) in lower parts of channel lithosomes.

Facies 1.5 consists of very-fine to fine grained, poorly to moderately sorted sandstone

with gradational planar laminations (Fig. 2.3F). In places this facies contains more than

50% carbonate grains, such as ooids and ostracods. Internally, each lamina is normally or

reversely graded with gradational boundaries. Lamina thickness varies from 2 mm to 2 cm

in a few dm to a few m thick sets. In places, the laminations become undulating laterally

across a few meters. The carbonate grains are evenly distributed or occur as layers. On

lamina scale, both normal and reverse grading trends are common consisting of larger sized

carbonate grains and smaller siliciclastic grains. On a bed scale, in places coarsening upward

trends occur, where gradational planar laminations grade upward into a carbonate grainstone

composed of 100% carbonate grains.

Facies 1.6 contains very-fine to lower-medium grained moderately sorted sandstone with

distinct planar laminations (Fig. 2.3G). Boundaries between laminae are distinct rather than

gradational. Laminations are on mm scale and bed thickness is on a few dm to a few m

scale. Parting lineations occur in some places.

Facies 1.7 is fine to very coarse grained, poorly to moderately sorted structureless sand-

stone with no visible sedimentary structures. Thickness varies from a few cm to a few dm.

12

This facies is only associated with facies group 1.

Facies 1.8 is soft sediment deformed sandstone occurs in facies group 1 as convolute

bedding with mostly vertical limbs and overturned limbs in some cases (Fig. 2.3H). Deformed

zones are a few dm thick and a couple of m wide.

In many places group 1 facies transition vertically (measured sections in Figs. 2.2 and

2.3) or laterally into each other. For instance, sandstone with scour and fills (F1.2; Figs.

2.3C and D) transition laterally and vertically into convex-up low angle lamina (F1.4; Fig.

2.3E), and planar lamina (F1.5 and 1.6; Figs. 2.3F and G). Planar laminations (F1.5 and

1.6) become convex-up low angle laterally in many places. Vertically these facies also grade

into group 2 facies. Contact between different facies can be erosive to gradual. In some

cases, gradational planar laminations (F1.5) are interbedded with climbing ripple laminations

(F2.3). Sandstone with distinct planar lamina (F1.6) is commonly overlain by sandstone with

climbing ripple laminations (F2.3) with a gradual contact.

Interpretation: By comparison to sedimentary structures produced by experiments, the

facies group 1 sedimentary structures are the product of Froude supercritical flow (see

Alexander et al., 2001; Cartigny et al., 2014). Supercritical flow bedforms are mutually tran-

sitional as linked to increasing flow velocities, but primarily to increase in maximum Froude

number (Cartigny et al., 2014). Low-angle concave- and convex-up laminations (F1.1 to

F1.4) are likely to be produced by antidunes and steeper scour and fill structures by chute-

and-pool formation or cyclic steps (Alexander et al., 2001; Cartigny et al., 2014). Cyclic

steps are a more likely interpretation where steep scour and fill structures are amalgamated,

rather than transition into other facies laterally (Cartigny et al., 2014). Stable antidunes

form due to passage of internal waves from entirely in supercritical flow conditions, whereas

unstable antidunes, chutes and pools and cyclic steps form at hydraulic jumps where flow re-

verses to subcritical conditions (Alexander et al., 2001; Cartigny et al., 2014). Foresets occur

where bedforms migrate downstream, whereas backsets indicate upstream migration (Car-

tigny et al., 2014). Planar laminations (F1.5 and 1.6) that laterally transition to low-angle

13

lamination (F1.1 to F1.4) have been produced by experiments lateral to stable antidunes or

where antidune aggradation rates are low (Cartigny et al., 2014). Planar laminations (F1.5

and 1.6) that do not transition laterally to low-angle laminations may form in subcritical,

transcritical or supercritical flow conditions (e.g. Paola, 1989; Cheel, 1990). The transition

from planar lamination to low-angle lamination is especially likely at high deposition rates

(Leclair and Arnott, 2005; Duller et al., 2008). Gradational laminations (F1.5) indicate high

deposition rates (Duller et al., 2008). Experiments further show that high aggradation rates

commonly occur in supercritical flow conditions, especially at hydraulic jumps where flow

abruptly transitions from supercritical to subcritical with a large energy loss (Alexander et

al., 2001; Cartigny et al., 2014), promoting deposition of structureless sandstones (F1.7).

Convolute bedding (F1.8) indicates rapid porewater escape and may be related to rapid

loading by overlain sediments, or to porewater expulsion during rapid lowering of water

levels, such as during waning phase of floods (e.g. Williams, 1971; Stear, 1985; Singh and

Bhardwaj, 1991). The accumulation and preservation of these facies requires high deposition

rates during peak discharge and rapid flow deceleration prevents sediments from reworking

by subcritical flow (Alexander and Fielding, 1997).

2.5.2 Facies Group 2: Sandstones with Steep Foresets

Description: This facies group contains very fine to lower medium grained, moderately

sorted sandstone with cross stratification (F2.1), climbing cross stratification (F2.2), ripple

lamination (F2.3), and climbing ripple lamination (F2.4). The latter two facies have finer

grained sandstone than the former (Table 2.1; Fig. 2.4). Facies of this group form ca 30%

of the sandstone volume. Climbing ripple laminated sandstone (F2.4) is the most common

facies in this group. This facies group is characterized by steep foresets (ca 30◦). Set thickness

is 6 cm to 20 cm in cross strata and 8 cm to dm scale in sandstone with climbing cross strata

(Figs. 2.4A and B). Ripple lamina set thickness is on cm scale and less than 5 cm (Figs.

2.4C and D). Set boundaries in climbing cross strata (F2.2) and climbing ripple laminae

14

(F2.4) dip upstream and opposite to the cross strata and laminae dip direction. Dip angle

of set boundaries is mostly less than 15◦ (Figs. 2.4B and D).

Interpretation: Cross strata form by migration of dunes and cross laminae form by mi-

gration of ripples (e.g. Simons et al., 1965). These are characteristic bedforms in subcritical

flow conditions, where dunes occur at higher flow velocities and in some cases in coarser

grain sizes than ripples (e.g. Ashley, 1990). Subcritical flow bedforms can only migrate

downstream and thus preserve only foresets as bedload sediment is eroded from the stoss

side and deposited on the lee side (Simons et al., 1965; Ashley, 1990; Allen, 1986; Miall,

2014). Climbing bedforms suggest high sediment accumulation rates (Allen, 1970), with

sediment fallout rate exceeding the bedform migration rate.

2.6 Channel Marcoforms

Sunnyside Delta Interval channel lithosomes refers to lithosomes that are recognized as

deposits of active channels and bars within a channel belt. Channel deposits contain 70%

Froude super- and transcritical, and 30% subcritical flow sandstone sedimentary facies, of

which ca 80% indicate high deposition rates. The channel lithosomes also contain some

in-channel mudstone layers (Figs. 2.5A to C), which vary in color from purple to brown to

greenish grey (F3.1 to 3.4 in Table 2.1). In places vertical or horizontal bioturbations occurs

(Figs. 2.5E to G). We document six different channel lithosome styles based on channel

dimensions, macroform geometries, sedimentary facies and the degree of lateral and vertical

amalgamation.

2.6.1 Large Vertically and Laterally Amalgamated Channel Lithosomes

Description: The large vertically and laterally amalgamated channel lithosomes are 10-

35 m thick, and 100s of m to km wide in a flow oblique view. Some such lithosomes are

massive in appearance and form low relief and sharp-based “sandstone walls” along Nine Mile

Canyon (Fig. 2.6A). Others are lenticular in shape in flow perpendicular view, have a smaller

15

width with composite erosional base and a flat top, commonly with a few meters of local

truncation at the base (Fig. 2.6B). They commonly display multiple internal erosion surfaces

in various degrees of continuity outlining lenticular and tabular sandstones with different

shapes (Figs. 2.2A and 2.6). Some internal erosion surfaces cut through the entire lithosome

(Fig. 2.6B), whereas others bound low-angle downstream accretion sets (Fig. 2.6A). Some

channel lithosomes are bounded by a flat top and a concave up basal surface (Fig. 2.6B),

in others both bounding surfaces are concave up or irregular (Fig. 2.6). Erosion surfaces

are commonly aligned with mudstone clasts (Figs. 2.2 and 2.3A), and in a few places wood

fragments exist along the basal erosion surface. The lithosomes are dominated by moderately

to poorly sorted, fine grained sandstones (Figs. 2.2A and 2.3) with Froude supercritical flow

sedimentary structures (85%) and/or high deposition rate sedimentary structures (78%), and

characteristically lack in-channel mudstone layers. Typical sedimentary structures observed

in this type of channel lithesome (Figs. 2.2A and 2.3) include scour and fill structures (38%,

F1.2), gradational planar laminations (21%, F1.5), convex-up low-angle laminations (8%,

F1.4), and soft sediment deformation (4%, F1.8). There is a minor proportion of Froude

subcritical flow sedimentary structures (9%), such as cross strata (2%, F2.1) and climbing

ripple laminae (7%, F2.4). Some lithosomes have a few dm thick conglomerate at the base

(Figs. 2.2A, 2.3A and B), which rarely stratified. No in-channel vegetation or bioturbation

occurs.

Interpretation: Accretion sets abound by erosion surfaces are interpreted as bar clino-

forms (sensu Mohrig et al., 2000), and the erosionally bound lenticular sandstones as individ-

ual channels. Laterally extensive sandstones with internal erosion surfaces indicate channel

amalgamation. The sedimentary structures indicate dominant deposition from Froude super-

and trans-critical flow conditions and high deposition rates (see Alexander et al., 2001; Car-

tigny et al., 2014). Abundance or dominance of Froude super- and trans-critical flow sedi-

mentary structures is characteristic for variable discharge rivers where deposition dominantly

occurs during high magnitude floods (Fielding, 2006; Fielding et al., 2009; Plink-Bjorklund,

16

2015). Erosionally-bound, low-angle downstream accretion sets have been recognized as flood

event beds (sensu flood units of Plink-Bjorklund, 2015) deposited from individual floods or

flood waves (Tunbridge, 1981; Sneh, 1983; Turner, 1986; Abdullatif, 1989; Deluca and Eriks-

son, 1989; Nichols and Hirst, 1998; Shukla et al., 2001; Hinds et al., 2004; Hampton and

Horton, 2007; Chakraborty and Ghosh, 2010; Chakraborty et al., 2010; Mader and Red-

fern, 2011; Donselaar et al., 2013; Plink-Bjorklund, 2015). This channel lithosome type is

similar to that in the proposed facies model of Plink-Bjorklund (2015) for extremely flashy

rivers where most deposition occurs during high-magnitude floods as a result of short intense

rainfall events.

2.6.2 Large Channel Lithosomes with Upstream Migrating Macroforms

Description: The large channel lithosomes with upstream macroform migration are 10-15

m thick, a few 100s m wide in a flow parallel view, and close to a 100 m wide in flow perpen-

dicular view. They form a complex geometry with flat or undulating tops and undulating

erosional bases (Fig. 2.7). The local erosional relief on the compound basal surface is a

few m with a lateral spacing of 10s of m, whereas the accretion-set bounding erosion sur-

faces are steep (5-20◦) and 10-15 m deep, and upstream-facing (towards south) (Fig. 2.7A).

These erosion surfaces bound lenticular sandstones that are individually 10-15 m thick and

at least 150 m wide in flow parallel view. The erosionally-bound sandstone lenses display

steep, upstream facing geometries along the erosion surfaces that flatten laterally upstream

towards the next erosion surface (Fig. 2.7A). The sandstones are dominantly fine grained

with no apparent vertical grain size trends, other than along basal scours, where grain size

is up to very coarse grained sand and mixed with gravel grains. These coarse basal layers

are structureless (F1.7). Wood pieces and mud clasts occur above the basal scours in places.

These sandstones consist of 40% Froude supercritical flow and/or 100% high deposition rate

sedimentary structures. Soft sediment deformed structures are common (40%) and occur in

the lower and upstream part of the lenticular sets (Fig. 2.8B). Soft sediment deformation (F

17

1.8) occurs as convoluted beds with overturned limbs. They gradually change upward and

laterally into climbing cross strata (20%, F2.2), gradational planar laminations (10%, F1.5),

and scour and fill structures (30%, F1.3). The cross-strata foresets (F2.2) dip northeast as

compared with the southward dip of the accretion surfaces. Scour and fill structures (F1.3)

consist of upward flattening laminae that fill meter wide scours, and dip in both in north and

south directions at approximately 5-15◦. They commonly pass into planar and convex-up

low angle laminations (F1.6 and 1.4) laterally and vertically.

Interpretation: The upstream-dipping accretion sets are backsets (sensus Kostic et al.,

2010; Cartigny et al., 2014) characteristic for Froude supercritical flow deposition in chutes

and pools or cyclic steps. The systematic backsets with regularly spaced scours also suggest

deposition in cyclic steps (Kostic et al., 2010; Cartigny et al., 2014). Cyclic steps migrate

upstream as erosion occurs in Froude supercritical flow conditions on the upstream side of a

hydraulic jump, where flow returns to subcritical conditions and causes rapid deposition at

the hydraulic jump (Kostic and Parker, 2006; Cartigny et al., 2014) (Fig. 2.7C). The strata

tend to be flatter downstream of the hydraulic jump, where the current starts accelerating

and gradually reaches supercritical conditions in the upstream part of the next hydraulic

jump. Highest deposition rates occur at the hydraulic jump where there is a large energy

loss (Figs. 2.8B and C). High deposition rate deposits are suggested to be common in

hydraulic jumps (Postma et al., 2009; Postma, 2014; Postma and Cartigny, 2014), such as the

structureless (F1.7) and soft-sediment deformed (F1.8) deposits in this dataset (Fig. 2.7B).

Upstream migration of cyclic steps results in erosionally bounded asymmetrical lenses that

have characteristic vertical trends, with rapidly deposited structureless or graded hydraulic

jump deposits above the basal erosion surface, overlain by sub- to supercritical flow deposits

with internal sedimentary structures (Postma, 2014). Channel-scale or bar-scale cyclic steps

have hitherto not been documented in river deposits, but cyclic steps on similar scales have

been documented in deltas (Massari, 1996; Ventra et al., 2015) and in slope channels (e.g.

Ono and Plink-Bjorklund, 2017).

18

2.6.3 Channel Lithosomes with Low Angle Downstream Accretion Sets

Description: The channel lithosomes with low-angle accretion sets are 10-15 m thick

and at least 100s of m wide in a strike view. These channel lithosomes consist of tabular

or wedge shaped accretion sets, 0.5-2 m thick that dip at 4-10◦ NNE (downstream) (Fig.

2.8). Accretion set bounding surfaces are sharp and draped by mm to cm thick mud layers.

The sandstones are moderately to well sorted, and very fine to fine grained. There are

no obvious vertical grain size trends (Fig. 2.2C). These channel lithosomes have a high

proportion of Froude supercritical flow sedimentary structures (73%) and/or high deposition

rate sedimentary structures (80%). The sedimentary structures are ripple-filled scour and

fill structures (28%, F1.3), distinct or gradational planar lamina (45%, F1.5 or F1.6) and

climbing ripple laminations (27%, F2.4). Some lithosomes consist completely of climbing

ripples (Fig. 2.2C). Sets of climbing ripples with different climb angles are bounded by

erosion surfaces (Fig. 2.4E). A few dm thick couplets with planar lamina (F1.5 or F1.6) and

climbing ripples (F2.4) are common as well (Fig. 2.4F). The transition from planar lamina

(F1.5 or F1.6) to climbing ripples (F2.4) is mostly distinct (Fig. 2.2) and in places erosional.

Foresets of the climbing ripples dip NNE similar to the low-angle accretion sets.

Interpretation: The low angle downstream dipping accretion sets are interpreted as down-

stream accreting macroforms. Dip angle of the accretion sets is much lower than bar slip-

face cross strata commonly described in sandy channel lithosomes (e.g. Bridge and Lunt,

2006), and have been suggested as common macroform geometries in variable discharge rivers

(Plink-Bjorklund, 2015; see also Tunbridge, 1981; Stear, 1985; Olsen, 1989; North and Tay-

lor, 1996; Zaleha, 1997a; Thomas et al., 2002; Kumar et al., 2003; Allen et al., 2013, 2014;

Plink-Bjorklund and Birgenheier, 2013).

Transitions between Froude supercritical and subcritical flow sedimentary structures

(climbing ripples) are suggested to be linked to a rapid flow deceleration, common in high

magnitude floods, where deposition rate remains high during the falling flood limb (e.g.,

Chakraborty and Ghosh, 2010). In-channel mud deposits, and especially lack of ripples be-

19

tween the transitions from supercritical flow structures to overlying mud deposits indicates

extremely rapid decline in flow strength, with no low-stage reworking of the flood deposit

(e.g., Jones, 1977; Tunbridge, 1981). The sandstone and mudstone couplets form flood event

deposits, interpreted as deposits of single floods or flood waves (Tunbridge, 1981; Sneh, 1983;

Turner, 1986; Abdullatif, 1989; Deluca and Eriksson, 1989; Nichols and Hirst, 1998; Shukla

et al., 2001; Hinds et al., 2004; Hampton and Horton, 2007; Chakraborty and Ghosh, 2010;

Chakraborty et al., 2010; Mader and Redfern, 2011; Donselaar et al., 2013; Plink-Bjorklund,

2015). Such newly formed mud layers have been documented after floods in modern variable

discharge rivers (e.g. Stear, 1985; Abdullatif, 1989; Singh et al., 1993; Billi, 2007).

2.6.4 Amalgamated Channel Lithosomes with High Angle Accretion Sets

Description: The channel lithosomes with high-angle (most 10-15◦, but up to 20◦) accre-

tion sets consist of interbedded sandstones and mudstones and are up to 14 m thick, and

10s of m wide in a flow perpendicular view (Fig. 2.9). The accretion sets are a few dm to

a meter thick, and some thin towards the base (Fig. 2.9A and some in Fig. 2.9B), whereas

others thin towards the top of the sets (some in Figs. 2.9B and C). Mudstone layers are

cm to dm thick (Fig. 2.2D) and in places thicken towards the base of the accretion sets

(Fig. 2.9A). The interbedded sandstone and mudstone accretion sets dip at a high angle

(up to 20◦, stereonet plots in Fig. 2.9), and in places display a sigmoidal shape in a flow

parallel view and a tabular or wedge shape in a flow perpendicular view (Figs. 2.9C and

D). Dip direction of accretion set varies from obliquely upstream to lateral (Fig. 2.9A), to

obliquely downstream facing (Figs. 2.9B and C). In places, cosets in accretion sets are bound

by erosion surfaces (Fig. 2.9B). Commonly they lack grain size trends, but in places mud

proportion increases towards the toes of the accretion sets, forming a general upward coars-

ening trend (Figs. 2.2D and 2.9A). The accretion sets consist of 35% Froude supercritical

flow sedimentary structures (scour and fill, F1.2; distinct planar lamina, F1.6; gradational

planar lamina, F1.5), 35% high deposition rate sedimentary structures, and mostly ripple

20

cross laminations (F2.4). Desiccation cracks (Fig. 2.5D), plant material (Figs. 2.5E and

2.F) and trace fossils (Fig. 2.5G) are commonly present at accretion set boundaries (Fig.

2.2D). Isolated trace fossils are mm to a dm long.

Interpretation: The accretion direction indicates obliquely upstream to lateral and down-

stream accretion directions. This is in contrast to the established facies models where steep

accretion sets are assigned exclusively to lateral accretion (Miall, 2014). Upward coarsening

trends in accretion sets have been linked to deposition during floods (Ghinassi et al., 2016,

2017), when helical flow pattern is destroyed by the shift of the zone of maximum boundary

shear stress from channel thalweg to bar top (Hooke, 1975; Dietrich et al., 1979; Dietrich

and Smith, 1983). The erosionally bound downstream accretion sets have geometries similar

to those described from deltas (Massari, 1996; Ventra et al., 2015) and in slope channels

(e.g. Ono and Plink-Bjorklund, 2017) and are linked to aggradational infilling of hydraulic

jump scours.

This channel lithosome type has a lower proportion of Froude supercritical flow deposits

and a higher proportion of climbing ripples. The latter are interpreted to be linked to the

rapid deceleration of flood waters with high sediment concentration (e.g. Mckee et al., 1967;

Picard and High, 1973; Croke et al., 1998). The preservation of mud is unique relative

to the previous lithosomes. Thick mud layers further suggest a rapid deceleration with

accompanying mud deposition during waning stages of floods (e.g., Jones, 1977; Tunbridge,

1981). The in-channel desiccation cracks indicate seasonally dry channels, common in arid

to sub-humid climates (Sneh, 1983; Nichols and Fisher, 2007). However, the water table

level was relatively high as the vertical burrows are less than 1 dm long (Hasiotis, 2004).

2.6.5 Heterolithic Aggradational Channel Lithosomes

Description: Heteroltithic aggradational channel lithosomes consist of interbedded dm to

m thick sandstones and dm thick mudstones that fill scours up to 10 m deep and 100 m wide.

The sandstone layers commonly thin laterally (Fig. 2.10). In places the aggradational sets

21

are conformable with the basal erosion surface and flatten upward (Fig. 2.10A). In other

places the sets are flat lying and terminate laterally onto the scour surface (Fig. 2.10B). Some

set boundary contacts are erosional, especially towards the top of the channel lithosomes,

and overlain by sandstones or mudstones (Fig. 2.10A). Some channel lithosomes extend

outside the visible channel-bounding erosion surface towards the top and form “wings” or

fill other scours laterally (Fig. 2.10). In places they are erosionally truncated above (Fig.

2.10B). The sandstones are moderately sorted, very fine grained and mostly moderately to

highly bioturbated. A considerably smaller amount of Froude supercritical flow sedimentary

structures (20%, planar laminations, F1.5 and 1.6) and/or high deposition rate sedimentary

structures (50%, climbing ripple laminations, F2.4; gradational planar laminations, F1.5)

are present. Most common sedimentary structures are climbing ripple cross laminations

(40%, F2.4). In some places, sandstones are heavily bioturbated throughout or just at the

boundaries of the accretion sets (Fig. 2.2E). Roots occur mostly in growth position on top

of the accretion sets. Mudstone layers in sand-to-mud couplets are composed of greenish

grey or purple mudstone (F3.2 or 3.1; Figs. 2.2E, 2.5A to C) with some plant fragments and

brownish oxidization bands.

Interpretation: The sandstone-mudstone couplets are interpreted as deposits of individ-

ual floods or flood event beds (sensu Plink-Bjorklund, 2015). Similar to channel type 4

a large proportion of sandstones contain climbing ripples (F2.4) and mudstone layers are

relatively thick, indicating dominant deposition from late-stage waning floods (Chakraborty

and Ghosh, 2010). Trace fossils and pedogenic modification at the accretion set boundaries

suggest that channels were dry between the episodic flood events (Nichols and Hirst, 1998;

Shukla et al., 2001; Donselaar et al., 2013; Plink-Bjorklund, 2015). The lateral onlapping

relationship and the upward flattening geometry suggest aggradational channel lithosomes.

The dominant climbing ripples and gradational planar laminations indicate high deposition

rates during the floods and lack of reworking and thus efficient base flow (Fisher et al., 2008;

Chakraborty et al., 2010; Plink-Bjorklund, 2015). During the dry season, soil moisture in

22

spots can be high enough for terrestrial plants to grow and creatures to live within the chan-

nels, which is responsible for the preserved root traces in growth position at accretion set

boundaries (Fielding et al., 2009; Allen et al., 2014; Bashforth et al., 2014). Aggradation or

vertical accretion has also been suggested as a common characteristic for variable discharge

rivers and is formed due to vertical channel bed aggradation (e.g. Tunbridge, 1981; Stear,

1985; Olsen, 1989; Zaleha, 1997; Thomas et al., 2002; Kumar et al., 2003; Allen et al., 2013,

2014; Plink-Bjorklund, 2015).

2.6.6 Small Lenticular Channel Lithosomes

Description: The small lenticular channel lithosomes have lenticular shape with flat tops

and convex-up erosional bases (Fig. 2.11). They are commonly encased in purple and

greenish mudstones. The channel lithosomes are ca 2 m thick, and 10-10s of m wide in a flow-

perpendicular view (Fig. 2.11). Some are laterally truncated and partially amalgamated with

some thin mudstones in between (Figs. 2.11A and C). Others thin in both directions (Fig.

2.11B). Sandstones are very fine grained and contain climbing ripple cross laminations (80%,

F2.4), climbing cross strata (10%, F2.2), and soft sediment deformation (10%, F1.8), which

sums up to 10% Froude supercritical flow and/or 100% high deposition rate sedimentary

structures. Some sandstones completely consist of climbing ripple laminae (Fig. 2.2, F2.4).

There are no obvious grain size trends.

Interpretation: The lower proportion of Froude supercritical flow sedimentary structures

and a higher proportion of climbing ripples and mud layers indicates relatively low flow

velocity compared to the other types of channel types (Mckee et al., 1967; Croke et al.,

1998; Chakraborty and Ghosh, 2010), even though the deposition rate is high as seen by the

100% high deposition rate sedimentary structures. This channel lithosome type is somewhat

similar to type 5, but much smaller in dimension, and less amalgamated.

23

2.7 Discussion

The channel deposits in Sunnyside Delta Interval are significantly different from those in

the widely used fluvial facies models in that (1) Froude super- and trans-critical flow bed-

forms are abundant; (2) abundant downstream accretion, upstream accretion, and vertical

aggradation exist with only a minor and oblique lateral accretion, in places with upward-

coarsening trends; (3) heterolithic channel lithosomes with in-channel muds, in-channel bio-

turbations, and mud-clast conglomerates are common; (4) erosionally bound flood units are

abundant. Some of these characteristics have been linked to variable discharge river deposits

(see reviews in Fielding, 2006; Fielding et al., 2009; Plink-Bjorklund, 2015). Although some

of these characteristics are now well established, relatively little is known about depositional

dynamics of such rivers. We follow with a comparison to the existing understanding of chan-

nels and bars both in persistent and variable discharge rivers with the aim to improve our

understanding of variable discharge river dynamics (Fig. 2.12).

2.7.1 Lateral Accretion Sets

Lateral accretion sets occur as minor components in channel types 3 and 4, where they

are low and high angle accretions sets respectively (Figs. 2.7 and 2.8). The latter accretion

sets can be considered morphologically similar to point-bar accretions sets (Allen, 1965,

1970; Miall, 1985). The differences are the strong oblique downstream (channel type 3)

or upstream migration component (channel type 4), and the low accretion angle in type 3

channels, and the upward coarsening trend in channel type 4. Furthermore, a mud plug or

counter point bar deposits (Smith et al., 2011), rather than the distinct sandstone-mudstone

couplets are expected to be associated with traditional point bars as opposed to the ones in

this study.

Point bars are formed by lateral accretion in a direction that is approximately perpendic-

ular to the flow (Bathurst et al., 1977; Dietrich and Smith, 1983; Miall, 1994; Labrecque et

al., 2011). Point bars are the result of helical flow dynamics, as the flow gradually loses mo-

24

mentum from channel base up to bar surface (Bathurst et al., 1977; Ethridge and Schumm,

1978). A fining upward succession is typically developed with a basal erosion surface, coarse

lag, followed by cross stratified sandstone that grades into ripple laminated sandstone, and

is capped by planar laminated mudstone or siltstone (Tyler and Ethridge, 1983; Miall, 1988;

Donselaar and Overeem, 2008; Labrecque et al., 2011). In cross section, point bars occur

in characteristic steep accretion sets in flow perpendicular view (Miall, 1994), and relatively

flat to convex-upward lobate shape in flow parallel view (Durkin et al., 2015).

In summary, the key differences of the documented lateral accretion sets are their partial

downstream or upstream accretion, lower-than-expected angle and in places upward coars-

ening, together with the sand-mud couplets or flood event beds and the dominance of either

Froude supercritical flow deposits or climbing ripples rather than cross strata.

2.7.2 Downstream Accretion

Downstream accretion occurs in channel types 1, 3, and 4, where the former two types

are amalgamated sandstones and the latter consist of sandstone-mudstone couplets (Figs.

2.12A to H). In addition to being suggested to be common in variable discharge rivers (Plink-

Bjorklund, 2015 and references therein), downstream accretion has been documented in some

braid bars, as well as in mouth bars.

Braid bars are fundamentally compound bars built by unit bars superimposed by dunes

(Lunt and Bridge, 2004; Bridge, 2006). Transverse bars are those attached to river bank and

grow in direction perpendicular to the flow. Longitudinal bars are normally developed in the

middle of channel, have a bar head and bar tail geometry, and mainly accrete downstream

at the angle of repose or less (Sambrook Smith et al., 2006, 2009; Bridge and Lunt, 2006).

Internally braid bar deposits consist of sets of cross stratified sandstones on different scales

(Ashley, 1990; Lunt et al., 2004). No obvious vertical grain size trend occurs in braid

bar deposits, though a downstream decreasing grain size is common in both bar scale and

river reach scale (Rice and Church, 2010). In a cross section, complex undulated 3rd order

25

bounding surfaces separate the stacked compound bars (Miall, 1985). The here documented

downstream accretion sets (channel types 1, 3, and 4) differ from mid-channel bars in that

the accretion sets are of low angle or concave-up (erosional) where steep, and in lack of

downstream migrating cross strata and presence of scour and fill structures and climbing

ripple laminae.

Due to the common downstream accretion features, the barforms in the Sunnyside Delta

Interval has been previously interpreted as mouth bar deposits (Schomacker et al., 2010).

Mouth bars occur as sharp or gradationally based lithosomes associated with distributary

channels and delta-front deposits, and typically show coarsening and thickening upward

trends as they form by progradation of deltaic deposits (Gilbert, 1885; Scruton, 1960; Bhat-

tacharya, 2006). Mouth bars may contain both sediment gravity flow and traction flow

deposits (Plink-Bjorklund and Steel, 2005). Cross strata, ripple lamination, climbing rip-

ple laminations are common sedimentary structures documented in some ancient examples

of mouth bars (Olariu and Bhattacharya, 2006), whereas others display gradational planar

laminations (Plink-Bjorklund and Steel, 2004).

The here documented downstream accretion sets (channel types 1, 3, and 4) differ from

mouth bars in that they lack the link to delta-front or prodelta deposits, and that the

accretion set boundaries are commonly concave up (Fig. 2.9).

2.7.3 Upstream Accretion and the Hierarchy of Froude Supercritical Flow De-

posits

The channel lithosomes with upstream accretion sets (channel types 2 and 4; Figs. 2.12I

to K) have a similar geometry to the cyclic steps produced by experiments (Winterwerp et

al., 1992; Parker, 1996; Taki and Parker, 2005; Spinewine et al., 2009), as they display the

characteristic asymmetrical geometry of hydraulic jumps, with a steep upstream and gentle

downstream margin (Cartigny et al., 2014) and a fill that is structureless and deformed

in the lower part and stratified towards the top (Postma et al., 2009, 2014) . The cyclic

26

steps deposits in channel type 2 are 10-15 m thick, but consist internally of dm to m scale

Froude transcritical and supercritical flow sedimentary structures. A similar occurrence is

described in experiments, where smaller dm to cm scale cyclic steps and antidunes occurred

within a large m-scale hydraulic jump (Spinewine et al., 2009). Thus, at least two levels

of hierarchy potentially occurred in cyclic step dynamics. On the modern seafloor, various

scales of scours (100s to 1000s meter long and 10s to 100s meter deep) and sediment waves

(100s or 1000s of meter wavelength and a few meters or 100s meter wave heights) have been

recently identified as formed by cyclic steps (Symons et al., 2016; Carvajal et al., 2017).

Outcrop work further indicates that the thickness and wavelength of Froude supercritical

flow deposits are extremely variable from centimeter to hundreds of meters scale (Postma et

al., 2009; Postma and Cartigny, 2014; Ono and Plink-Bjorklund, 2017).

2.8 Conclusions

Documentation of the Sunnyside Delta Interval fluvial deposits reveals deposition in vari-

able discharge rivers by the abundance of Froude super- and transcritical flow sedimentary

structures, in-channel mud deposits, trace fossils, and desiccation cracks. This paper doc-

uments six macroform geometries, and shows that variable discharge river macroforms are

considerably different from the established point bars or braid bars. This paper presents

a first systematic description of such macroforms that have hitherto been referred to as

poorly developed (Plink-Bjorklund, 2015) or even missing (Fielding et al., 2018) in variable

discharge rivers. This paper document down- and upstream migrating and obliquely lat-

eral accretion sets, and a variety of accretion angles. In many cases the accretion sets are

erosionally bound and resemble geometries produced by hydraulic jumps in experiments.

27

Figure 2.1 Stratigraphic column and base map of Uinta Basin and Nine Mile canyon. (A)The study area (Nine Mile Canyon) in the Uinta Basin (Utah, U.S.A.). The extent of Colton/ Wasatch and Green River Formations are in maroon and orange respectively. Locationof Fig. 1B is highlighted in a small box with dashed lines. Modified from Dickinson et al.(2012), Schomacker et al. (2010), Sato and Chan (2015), and Jones (2017). (B) Measuredstratigraphic sections (M1.1, 1.2; 2; 3; 4; 5; 6.1, 6.2 in Fig. 2.2) and figure numbers aremarked in dark grey star shapes. Grey numbers are mile marks from 37 to 49 miles alongNine Mile Canyon Road. Paleocurrent directions are compiled from several measured sectionsand plotted in rose diagrams to show the variation at different locations from west to east.Black color in rose diagrams represent all paleocurrent measurements, and the grey ones aretaken from true cross stratifications. (C) Stratigraphic framework in this study. SunnysideDelta Interval (outlined in dashed lines) is the focus of this paper, located in the middle partof Green River Formation, and bounded by D and C markers. Modified from Fouch et al.(1987), Remy (1992), and Smith et al. (2010, 2015).

28

Figure 2.2 Representative examples of measured sections organized by channel lithesome types.

29

Figure 2.3 Example outcrop photos and measured sections of Facies group 1: sandstones andconglomerates interpreted as Froude supercritical deposits. (A) Stratified and imbricatedconglomerates (F1.1). (B) Disorganized conglomerates (F1.1). (C) Sandstone with scourand fill structure (F1.2). (D) Sandstone with ripple filled scour and fill structure (F1.3).(E) Sandstone with convex-up low angle lamina (F1.4). (F) Sandstone with gradationalplanar lamina (F1.5). (G) Sandstone with distinct planar lamina (F1.6). (H) Soft sedimentdeformed sandstone (F1.8). Stars mark the position of respective facies in the measuredsections.

30

Figure 2.4 Example outcrop photos and measured sections of Facies group 2: sandstones withsteep foresets interpreted as Froude subcritical flow deposits. (A) Cross stratified sandstone(F2.1). (B) Sandstone with climbing cross strata (F2.2). (C) Rippled sandstone (F2.3). (D)Sandstone with climbing ripple lamina (F2.4). Stars mark the position of respective facies inthe measured sections. (E) Vertical transition from climbing ripples with high angle climbangles to low climb angles. (F) Couplets of planar lamina and climbing ripples.

31

Figure 2.5 Example photos of in-channel mudstones and bioturbation. (A) Purple mudstonewith bioturbation (F3.1). (B) Purple greenish grey mixed color siltstone (F3.5). (C) Finelaminated purple to greenish grey siltstone (F3.1; F3.2). (D) Desiccation cracks preservedat accretion set boundaries. Stars mark the position of respective facies in the measuresections. (E-F) Root traces in structureless sandstones. (G) Ancorichnus in ripple laminatedsandstones.

32

Figure 2.6 Large vertically and laterally amalgamated channel lithosomes with low-angle accretion sets. (A) Example of anamalgamated channel lithesome with low relief concave top and bottom surfaces. (B) Examples of a lenticular amalgamatedchannel lithosomes. Bar clinoforms dipping to the left indicate downstream accretion.

33

Figure 2.7 Large channel lithosomes with upstream migrating macroforms. (A) Channel macroforms consist of upstream dippingaccretion sets, bound by erosion surfaces. (B) is Part of (A). Example of a single erosionally bound set with characteristicdistribution of facies, where structureless or soft-sediment deformed sandstones occur close to the basal scour and more laminatedfacies towards the top in downstream direction. (C) Sketch of a hydraulic jump illustrating the supercritical upstream edge,the hydraulic jump where deposition rates are highest, and the gentler downstream side with subcritical to transcritical flowconditions. (D) A flow perpendicular view of the same channel lithesome. (E) The paleocurrent data plotted in rose diagramto the left was collected from the cross stratified sandstone layer in Cottonwood Canyon. The rose diagram to the right is thedip measurements from macroforms dipping upstream.

34

Figure 2.8 Channel lithosomes with low-angle accretion sets. (A, C) downstream accreting accretion sets or bar clinoforms areclearly visible in flow -parallel view. (B) Obliquely perpendicular view shows the same channel lithosomes with an obliquelylateral component. (D) Plots the paleocurrents and dip angle of the accretion sets in rose diagram and stereonet. White arrowsshow average paleoflow direction in each outcrop photos (A to C) and the plan view map (D).

35

Figure 2.9 Channel lithosomes with high-angle (up to 20◦) accretion sets with preservedmudstone between sand bodies. Paleocurrent directions are from ripple cross lamination.Together with the dip angle of accretion sets are plotted to the right. (A) Obliquely upstreamand lateral accretion sets with sandstone layers thin and mud layers that thicken towards thebase, forming a coarsening upward trend. (B) Amalgamated erosionally bound downstreamaccretion sets. (C) Obliquely downstream and lateral accretion sets. (D) A more flowperpendicular view of the same channel lithesome as in C. White arrows are average paleoflowdirection in each example.

36

Figure 2.10 Heterolithic aggradational channel lithosomes. (A) Upward flattening aggradational sandstone layers interbeddedwith thin mud layers, truncated on top by an erosion surface. (B) Interbedded sandstone and thick mudstone layers onlap ontothe erosive scour surface.

37

Figure 2.11 Small lenticular channel lithosomes. (A) Laterally amalgamated lenticular sand-stones. (B) Vertically amalgamated lenticular sandstones. (C) Laterally and vertically amal-gamated lenticular sandstones.

38

Figure 2.12 Summary diagram of the 6 channel lithesome types in flow parallel (to the left,arrow points downstream) and perpendicular views (right). Red lines are erosion surfaces.Black lines are bedding surfaces and sedimentary structures. Yellow and green/purple rep-resent sandstone and mudstone respectively. Examples of A, C, E, F, and G are downstreamaccretion styles. B, D, and H are the corresponding styles in flow perpendicular view. Up-stream accretion styles are shown in I to K. Examples L and M look aggradational in bothviews.

39

Table 2.1: Description and Interpretation of Sedimentary Facies

Facies % Textures Structures Width/Thickness Interpretation1.1 Stratified and

imbricatedconglomerates

<1% Conglomerates (coarse sandto pebble sized carbonateclast/bioclast and/or mudclast, subrounded to angular,poorly sorted, varied inshape) in sandy matrix, clastor matrix supported

Imbricated,planarstratified, lowanglestratified,scour and fillstructures,

Thickness: a few cmper conglomeratelayer; Total thickness:a few dm - m

Intrabasinal clasts offloodplain mudstonesor lake carbonatesdeposited with sandmatrix by tractionflow (subcritical tosupercritical)

1.1 Disorganizedconglomeratesand diamictites

<1% Conglomerates (coarse sandto pebble sized carbonateclast/bioclast and/or mudclast, rounded to angular,poorly sorted, varied in shape)in sandy or muddy matrix, inplaces with terrestrial organicmatter (woody fragments)

disorganized Thickness: a few dm -1m

Intrabasinal clasts offloodplain mudstonesor lake carbonatedeposited by collapseand fluidization ofchannel banks duringfalling stage of flow

1.2 Sandstone withscour and fillstructures

30% Upper very fine to lowermedium grained sandstone,moderately to poorly sorted,in places mixed withcarbonate grains

Scours filledwith upwardflatteninglamina

Thickness: a few dm -a few m; Width: afew dm - 10s of m

Supercritical flow,high deposition rates,suspensiondeposition, largeantidune or chute andpool formation

1.3 Sandstone withripple-filledscour and fillstructures

<1% Upper very fine to lower finegrained sandstone, moderatelysorted

Ripples fillthe toe partof upwardflatteningscour and filllamina

Thickness: a few dm -1m; Width: a few dm- a few m

Supercritical tosubcritical flow, highdeposition rates,suspension andbedload deposition

40

Table 2.1: Continued.

Facies % Textures Structures Width/Thickness Interpretation1.4 Sandstone with

convex-up lowangle lamina

8% Very fine to fine grainedsandstone, moderately topoorly sorted, in places mixedwith carbonate grains

Low angleconvex-uplaminations,very lowanglehummocky-like crossstrata withupwardincreasingsteepness

Thickness: a few dm -1m; Width: >1m

Supercritical flow,high deposition rates,suspensiondeposition, antiduneformation

1.5 Sandstone withgradationalplanar lamina

14% Very fine to fine grainedsandstone, moderately topoorly sorted

Diffused,gradationalplanarlaminations,each laminais internallyreversely tonormallygraded

Thickness (lamina):2mm - 2cm;Thickness (bed): afew dm - a few m

Subcritical tosupercritical flow,high deposition rates,suspensiondeposition, plane bedformation

1.6 Sandstone withdistinct planarlamina

20% Very fine to lower mediumgrained sandstone, moderatelysorted

Planarlaminationswith distinctlaminaboundaries

Thickness (lamina):mm scale; Thickness(bed): a few dm - afew m

Supercritical ortranscritical flow,normal depositionrates, bedloaddeposition, plane bedformation

1.7 Structurelesssandstone

1% fine to very coarse grainedsandstone, moderately topoorly sorted

No apparentsedimentarystructures

Thickness: a few cm -a few dm

high deposition rates,suspension deposition

41

Table 2.1: Continued.

Facies % Textures Structures Width/Thickness Interpretation1.8 Soft sediment

deformedsandstone

2% Very fine to lower mediumgrained sandstone, moderatelyto poorly sorted

Flamestructures,overturnedfolds, dishstructures,ball andpillowstructures.

Thickness: a few dm Water escape or localcollapse

2.1 Cross stratifiedsandstone

2% Fine to lower medium grainedsandstone, moderately sorted

Planar andtrough crossstratified

Thickness (cross set):5cm -20cm; Thickness(co-set): a few dm

Subcritical flow,normal depositionrates, bedloaddeposition, dunemigration

2.2 Sandstone withclimbing crossstrata

<1% Fine to lower medium grainedsandstone, moderately topoorly sorted

Planar andtrough crossstrata withclimbing setboundaries

Thickness (cross set):8cm - dm scale;Thickness (co-set): afew dm

Subcritical flow, highdeposition rates,suspension andbedload deposition,dune migration andaggradation

2.3 Cross laminatedsandstone

10% Very fine to lower fine grainedsandstone, moderately sorted

Asymmetricalcross-lamination

Thickness (cross set):cm scale < 5cm;Thickness (co-set): afew dm

Subcritical flow,normal depositionrates, bedloaddeposition, ripplemigration

42

Table 2.1: Continued.

Facies % Textures Structures Width/Thickness Interpretation2.4 Sandstone with

climbing ripplelamina

13% Very fine to fine grainedsandstone, moderately sorted

Asymmetricalcrosslaminationwith climbingsetboundaries

Thickness (cross set):cm scale <5cm;Thickness (co-set): afew dm - a few m

Subcritical flow, highdeposition rates,suspension andbedload deposition,ripple migration andaggradation

2.5 Bioturbatedsandstone

- Very fine to lower mediumgrained sandstone, moderatelyto poorly sorted

Vertical,horizontaland 3D tracefossils

Thickness: a few dm -a few m

Trace fossils formedby insects, dwelling,resting, crawlingtraces

3.1 Purpled redmudstone

- Clay to silt Laminated orblocky orcrumbly, inplaces rooted,bioturbated,or deformed

Thickness: a few cm -a few dm (blocky); afew mm (each lamina)

Oxidized, welldrained in-channelmudstone

3.2 Greenish greymudstone

- Clay to silt Laminated orblocky orcrumbly

Thickness: a few cm -a few dm (blocky); afew mm (each lamina)

Reduced to poorlydrained in-channelmudstone

3.3 Brown to darkgrey mudstone

- Clay to silt laminated Thickness: a few cm -a few dm

Oxidized poorlydrained in-channelmudstone

3.4 Mixed colormudstone

- Clay to silt, Blockygreenish greyto purple

Thickness: a few dm Variable Red-oxin-channel mudstone

43

2.9 Reference Cited

Abdullatif, O.M., 1989, Channel-fill and sheet-flood facies sequences in the ephemeral ter-minal River Gash, Kassala, Sudan: Sedimentary Geology, v. 63, p. 171–184.

Alexander, J., Bridge, J.S., Cheel, R.J., and Leclair, S.F., 2001, Bedforms and associatedsedimentary structures formed under supercritical water flows over aggrading sandbeds: Sedimentology, v. 48, p. 133–152.

Alexander, J., and Fielding, C., 1997, Gravel antidunes in the tropical Burdekin River,Queensland, Australia: Sedimentology, v. 44, p. 327–337.

Allen, J.R.L., 1965, A Review of the Origin and Characteristics of Recent Alluvial Sediments:Sedimentology, v. 5, p. 89–191.

Allen, J.R.L., 1986, Principles of physical sedimentology: Macmillan Magazines LTD, PortersSouth, 4 Crinan ST, London, England N1 9XW, v. 319.

Allen, J.R.L., 1970, Studies in fluviatile sedimentation: a comparison of fining- upwardscyclothems, with special reference to coarse-member composition and interpretation1: v. 1, p. 298–323.

Allen, J.P., Fielding, C.R., Gibling, M.R., and Rygel, M.C., 2011, Fluvial response to paleo-equatorial climate fluctuations during the late Paleozoic ice age: Bulletin of the Ge-ological Society of America, v. 123, p. 1524–1538.

Allen, J.P., Fielding, C.R., Gibling, M.R., and Rygel, M.C., 2014, Recognizing productsof palaeoclimate fluctuation in the fluvial stratigraphic record: An example from thePennsylvanian to Lower Permian of Cape Breton Island, Nova Scotia: Sedimentology,v. 61, p. 1332–1381.

Allen, J.P., Fielding, C.R., Rygel, M.C., and Gibling, M.R., 2013, Deconvolving Signals ofTectonic and Climatic Controls From Continental Basins: An Example From the LatePaleozoic Cumberland Basin, Atlantic Canada: Journal of Sedimentary Research, v.83, p. 847–872.

Ashley, G.M., 1990, Classification of Large-Scale Subaqueous Bedforms: A New Look atan Old Problem-SEPM Bedforms and Bedding Structures: Journal of SedimentaryResearch, v. 60, p. 160–172.

Bathurst, J.C., Thorne, C.R., and Hey, R.D., 1977, Direct measurements of secondary cur-rents in river bends: Nature, v. 269, p. 504–506.

Bhattacharya, J.P., 2006, Deltas, in Posamentier, H.W. and Walker, R.G. eds., Facies ModelsRevisited, SEPM Special Publication 84, p. 237–292.

44

Billi, P., 2007, Morphology and sediment dynamics of ephemeral stream terminal distributarysystems in the Kobo Basin (northern Welo, Ethiopia): Geomorphology, v. 85, p.98–113.

Bridge, J.S., 2006, Fluvial facies models: recent developments (H. W. Posamentier & R. G.Walker, Eds.): SEPM Special Publication 84.

Bridge, J.S., and Lunt, I.A., 2006, Depositional Models of Braided Rivers, in SambrookSmith, G.H., Best, J.L., Bristow, C.S., and Petts, G.E. eds., Braided Rivers: Pro-cess, Deposits, Ecology and Management, International Association of SedimentologySpecial Publication, p. 11–50.

Burton, D., Woolf, K., and Sullivan, B., 2014, Lacustrine depositional environments in theGreen River Formation, Uinta Basin: Expression in outcrop and wireline logs: AAPGBulletin, v. 98, p. 1699–1715.

Cartigny, M.J.B., Ventra, D., Postma, G., and van Den Berg, J.H., 2014, Morphodynam-ics and sedimentary structures of bedforms under supercritical-flow conditions: Newinsights from flume experiments: Sedimentology, v. 61, p. 712–748.

Carvajal, C., Paull, C.K., Caress, D.W., Fildani, A., Lundsten, E.V.E., Anderson, K.,Maier, K.L., Gann, M.M.C., Gwiazda, R., Herguera, J.C., Bay, M., and Landing,M., 2017, Unraveling the Channel – Lobe Transition Zone With High-Resolution AuvBathymetry: Navy Fan , Offshore Baja California , Mexico: , p. 1049–1059.

Chakraborty, T., and Ghosh, P., 2010, The geomorphology and sedimentology of the Tistamegafan, Darjeeling Himalaya: Implications for megafan building processes: Geomor-phology, v. 115, p. 252–266.

Chakraborty, T., Kar, R., Ghosh, P., and Basu, S., 2010, Kosi megafan: Historical records,geomorphology and the recent avulsion of the Kosi River: Quaternary International,v. 227, p. 143–160.

Cheel, R.J., 1990, Horizontal lamination and the sequence of bed phases and stratificationunder upper-flow-regime conditions: Sedimentology, v. 37, p. 517–529.

Croke, J.C., Magee, J.M., and Price, D.M., 1998, Stratigraphy and sedimentology of thelower Neales River, West Lake Eyre, Central Australia: From Palaeocene to Holocene:Palaeogeography, Palaeoclimatology, Palaeoecology, v. 144, p. 331–350.

Davis, S.J., Dickinson, W.R., Gehrels, G.E., Spencer, J.E., Lawton, T.F., and Carroll, A.R.,2010, The Paleogene California River: Evidence of Mojave-Uinta paleodrainage fromU-Pb ages of detrital zircons: Geology, v. 38, p. 931–934.

45

DeCelles, P.G., 2004, Late Jurassic to Eocene evolution of the Cordilleran thrust belt andforeland basin system, western USA: American Journal of Science, v. 304, p. 105–168.

Deluca, J.L., and Eriksson, K.A., 1989, Controls on synchronous ephemeral- and perennial-river sedimentation in the middle sandstone member of the Triassic Chinle Formation,northeastern New Mexico, U.S.A.: Sedimentary Geology, v. 61, p. 155–175.

Dickinson, W.R., Klute, M.A., Hayes, M.J., Janecke, S.U., Lundin, E.R., Mckittrick, M.A.,and Olivares, M.D., 1988, Paleogeographic and paleotectonic setting of Laramide sed-imentary basins in the central Rocky Mountain region: Geological Society of AmericaBulletin, v. 100, p. 1023–1039.

Dickinson, W.R., Lawton, T.F., and Inman, K.F., 1986, Sandstone Detrital Modes, Cen-tral Utah Foreland Region: Stratigraphic Record of Cretaceous -Paleocene TectonicEvolution: Journal of Sedimentary Petrology, v. 56, p. 276–293.

Dickinson, W.R., Lawton, T.F., Pecha, M., Davis, S.J., Gehrels, G.E., and Young, R.A.,2012, Provenance of the Paleogene Colton Formation (Uinta Basin) And Cretaceous-Paleogene provenance evolution in the Utah foreland: Evidence from U-Pb ages ofdetrital zircons, paleocurrent trends, and sandstone petrofacies: Geosphere, v. 8, p.854–880.

Dietrich, W.E., and Smith, J.D., 1983, Influence of the point bar on flow through curvedchannels: Water Resources Research, v. 19, p. 1173–1192.

Dietrich, W.E., Smith, J.D., and Dunne, T., 1979, Flow and Sediment Transport in a SandBedded Meander: The Journal of Geology, v. 87, p. 305–315.

Donselaar, M.E., Cuevas Gozalo, M.C., and Moyano, S., 2013, Avulsion processes at theterminus of low-gradient semi-arid fluvial systems: Lessons from the Rıo Colorado,Altiplano endorheic basin, Bolivia: Sedimentary Geology, v. 283, p. 1–14.

Donselaar, M.E., and Overeem, I., 2008, Connectivity of fluvial point-bar deposits: Anexample from the Miocene Huesca fluvial fan, Ebro Basin, Spain: AAPG Bulletin, v.92, p. 1109–1129.

Duller, R.A., Mountney, N.P., Russell, A.J., and Cassidy, N.C., 2008, Architectural analysisof a volcaniclastic jokulhlaup deposit, southern Iceland: Sedimentary evidence forsupercritical flow: Sedimentology, v. 55, p. 939–964.

Durkin, P.R., Hubbard, S.M., Boyd, R.L., and Leckie, D.A., 2015, Stratigraphic expressionof intra-point-bar erosion and rotation: Journal of Sedimentary Research, v. 85, p.1238–1257.

46

Ethridge, F., and Schumm, S.A., 1978, Reconstructing Paleochannel Morphologic and FlowCharacteristics: Methodology, Limitations, and Assessment: Fluvial Sedimentology,v. 5, p. 705–721.

Fielding, C.R., Alexander, J., and Allen, J.P., 2018, The role of discharge variability in theformation and preservation of alluvial sediment bodies: Sedimentary Geology, v. 365,p. 1–20.

Fielding, C.R., Allen, J.P., Alexander, J., and Gibling, M.G., 2009, Facies model for fluvialsystems in the seasonal tropics and subtropics: Geology, v. 37, p. 623–626.

Ford, G.L., Dechesne, M., and Pyles, D.R., 2016, Stratigraphic architecture of a fluvial-lacustrine basin-fill succession at Desolation Canyon, Uinta Basin, Utah: Reference toWalthers’ Law and implications for the petroleum industry: The Mountain Geologist,v. 53.

Fouch, T., Cashion, W., Ryder, R., and Campbell, J., 1976, Field guide to lacustrine andrelated nonmarine depositional environments in Tertiary rocks, Uinta Basin, Utah.

Fouch, T.D., Hanley, J.H., Forester, R.M., Keighin, C.W., Pitman, J.K., and Nichols, D.J.,1987, Chart showing lithology, mineralogy, and paleontology of the nonmarine NorthHorn Formation and Flagstaff Member of the Green River Formation, Price Canyon,central Utah a principal reference section.

Gall, R.D., Birgenheier, L.P., and Vanden Berg, M.D., 2017, Highly Seasonal and Peren-nial Fluvial Facies: Implications For Climatic Control On the Douglas Creek andParachute Creek Members, Green River Formation, Southeastern Uinta Basin, Utah,U.S.A.: Journal of Sedimentary Research, v. 87, p. 1019–1047.

Ghinassi, M., Ielpi, A., Aldinucci, M., and Fustic, M., 2016, Downstream-migrating fluvialpoint bars in the rock record: Sedimentary Geology, v. 334, p. 66–96.

Ghinassi, M., Moody, J., Burton, B., and Martin, D., 2017, The effects of extreme versusannual floods on fluvial pointbar architecture: Inferences from the Powder River,Montana, USA., in International Conference on Fluvial Sedimentology, Calgary.

Gilbert, G., 1885, The topographic features of lake shores: U.S. Geological Survey AnnualReport No. 5.

Hampton, B.A., and Horton, B.K., 2007, Sheetflow fluvial processes in a rapidly subsidingbasin, Altiplano plateau, Bolivia: Sedimentology, v. 54, p. 1121–1147.

Hasiotis, S.T., 2004, Reconnaissance of Upper Jurassic Morrison Formation ichnofossils,Rocky Mountain Region, USA: Paleoenvironmental, stratigraphic, and paleoclimaticsignificance of terrestrial and freshwater ichnocoenoses: Sedimentary Geology, v. 167,

47

p. 177–268.

Hinds, D.J., Aliyeva, E., Allen, M.B., Davies, C.E., Kroonenberg, S.B., Simmons, M.D., andVincent, S.J., 2004, Sedimentation in a discharge dominated fluvial-lacustrine system:The Neogene Productive Series of the South Caspian Basin, Azerbaijan: Marine andPetroleum Geology, v. 21, p. 613–638.

Hooke, R.L.B., 1975, Distribution of Sediment Transport and Shear Stress in a MeanderBend: The Journal of Geology, v. 83, p. 543–565.

Jacob, F., 1969, Delta facies of the Green River Formation (Eocene): Carbon and Duchesne:University of Colorado, 364 p.

Jones, C.M., 1977, Effects of varying discharge regimes on bed-form sedimentary structuresin modern rivers: Geology, v. 5, p. 567.

Jones, E.R., 2017, Probabilistic Source-To-Sink Analysis of the Provenance of the Califor-nia Paleoriver: Implications for the Early Eocene Paleogeography of Western NorthAmerica: Colorado School of Mines, 198 p.

Keighley, D., Flint, S., Howell, J., and Moscariello, A., 2003, Sequence stratigraphy inlacustrine basins: a model for part of the Green River Formation (Eocene), SouthwestUinta Basin, Utah, U.S.A.: Journal of Sedimentary Research, v. 73, p. 987–1006.

Kelly, S.B., and Olsen, H., 1993, Terminal fans-a review with reference to Devonian examples:Sedimentary Geology, v. 85, p. 339–374.

Kostic, S., and Parker, G., 2006, The response of turbidity currents to a canyon–fan tran-sition: internal hydraulic jumps and depositional signatures: Journal of HydraulicResearch, v. 44, p. 631–653.

Kostic, S., Sequeiros, O., Spinewine, B., and Parker, G., 2010, Cyclic steps: A phenomenonof supercritical shallow flow from the high mountains to the bottom of the ocean:Journal of Hydro-Environment Research, v. 3, p. 167–172.

Kumar, R., Ghosh, S.K., Mazari, R.K., and Sangode, S.J., 2003, Tectonic impact on thefluvial deposits of Plio-Pleistocene Himalayan foreland basin, India: SedimentaryGeology, v. 158, p. 209–234.

Labrecque, P.A., Jensen, J.L., Hubbard, S.M., and Nielsen, H., 2011, Sedimentology andstratigraphic architecture of a point bar deposit, Lower Cretaceous McMurray Forma-tion, Alberta, Canada: Bulletin of Canadian Petroleum Geology, v. 59, p. 147–171.

Leclair, S.F., and Arnott, R.W.C., 2005, Parallel Lamination Formed by High-Density Tur-bidity Currents: Journal of Sedimentary Research, v. 75, p. 1–5.

48

Legarreta, L., and Uliana, M.A., 1998, Anatomy of Hinterland Depositional Sequences:Upper Cretaceous Fluvial Strata, Neuquen Basin, West-Central Argentina, in RelativeRole of Eustasy, Climate, and Tectonism in Continental Rocks, Special Publicationsof SEPM No. 59.

Littler, K., Rohl, U., Westerhold, T., and Zachos, J.C., 2014, A high-resolution benthicstable-isotope record for the South Atlantic: Implications for orbital-scale changesin Late Paleocene–Early Eocene climate and carbon cycling: Earth and PlanetaryScience Letters, v. 401, p. 18–30.

Lunt, I.A., and Bridge, J.S., 2004, Evolution and deposits of a gravelly braid bar, Saga-vanirktok River, Alaska: Sedimentology, v. 51, p. 415–432.

Lunt, I.A., Bridge, J.S., and Tye, R.S., 2004, A quantitative, three-dimensional depositionlmodel of gravelly braided rivers: Sedimentology, v. 51, p. 377–414.

Mader, N.K., and Redfern, J., 2011, A sedimentological model for the continental UpperTriassic Tadrart Ouadou Sandstone Member: recording an interplay of climate andtectonics (Argana Valley; South-west Morocco): Sedimentology, v. 58, p. 1247–1282.

Massari, F., 1996, Upper-Flow-Regime Stratification Types on Steep-Face, Coarse-Grained,Gilbert-Type Progradational Wedges (Pleistocene, Southern Italy): Journal of Sedi-mentary Research, v. 66, p. 364–375.

Mckee, E.D., Crosby, E.J., and Berryhill, H.L., 1967, Flood Deposits, Bijou Creek, Colorado,June 1965: Journal of Sedimentary Petrology, v. 37, p. 829–851.

Miall, A.D., 1985, Architectural-Element Analysis: A New Method of Facies Analysis Ap-plied to Fluvial Deposits: Earth-Science Reviews Elsevier Science Publishers B.V, v.22, p. 261–308.

Miall, A.D., 1994, Reconstructing Fluvial Macroform Architecture from Two-DimensionalOutcrops: Examples from the Castlegate Sandstone, Book Cliffs, Utah: Journal ofSedimentary Research, v. 64B, p. 146–158.

Miall, A.D., 1988, Reservoir Heterogeneities in Fluvial Sandstones: Lessons from OutcropStudies: AAPG Bulletin, v. 72.

Miall, A., 2014, The Facies and Architecture of Fluvial Systems, in FLuvial DepositionalSystems, Springer Geology.

Morgan, C.D., 2003, Geologic guide and road logs of the willow creek, indian, soldier creek,nine mile, gate, and desolation canyons, uinta basin, utah (OFR-407).

Morgan, C.D., and Bereskin, S.R., 2003, Characterization of petroleum reservoirs in theEocene Green River Formation, Central Uinta Basin, Utah: The Mountain Geologist,

49

v. 39.

Nicholas, A.P., Sambrook Smith, G.H., Amsler, M.L., Ashworth, P.J., Best, J.L., Hardy,R.J., Lane, S.N., Orfeo, O., Parsons, D.R., Reesink, A.J.H., Sandbach, S.D., Simp-son, C.J., and Szupiany, R.N., 2016, The role of discharge variability in determiningalluvial stratigraphy: Geology, v. 44, p. 3–6.

Nichols, G.J., and Fisher, J.A., 2007, Processes, facies and architecture of fluvial distributarysystem deposits: Sedimentary Geology, v. 195, p. 75–90.

Nichols, G.J., and Hirst, J.P., 1998, Alluvial fans and fluvial distributary systems, Oligo-Miocene, northern Spain; contrasting processes and products: Journal of SedimentaryResearch, v. 68, p. 879–889.

North, C.P., and Taylor, K.S., 1996, Ephemeral-fluvial deposits: Integrated outcrop andsimulation studies reveal complexity: AAPG Bulletin, v. 80, p. 811–830.

Olariu, C., and Bhattacharya, J.P., 2006, Terminal Distributary Channels and Delta FrontArchitecture of River-Dominated Delta Systems: Journal of Sedimentary Research,v. 76, p. 212–233.

Olsen, H., 1989, Sandstone-body structures and ephemeral stream processes in the DinosaurCanyon Member, Moenave Formation (Lower Jurassic), Utah, U.S.A.: SedimentaryGeology, v. 61, p. 207–221.

Ono, K., and Plink-Bjorklund, P., 2017, Froude supercritical flow bedforms in deepwaterslope channels? Field examples in conglomerates, sandstones and fine-grained de-posits: Sedimentology, v. 65, p. 639–669.

Paola, C., 1989, A simple basin-filling model for coarse-grained alluvial systems, in Cross, T.ed., Quantitative Dynamic Stratigraphy, Prentice Hall, Englewood Clffs, p. 363–374.

Parker, 1996, Some speculations on the relation between channel morphology and channel-scale flow structures., in Coherent Flow in Open Channels, p. 429–432.

Picard, M.D., and High, L.R., 1973, Sedimentary structures of ephemeral streams: Elsevier,v. 17.

Pitman, J., Fouch, T., Bulletin, M.G.-A., and 1982, U., 1982, Depositional setting anddiagenetic evolution of some Tertiary unconventional reservoir rocks, Uinta Basin,Utah: AAPG Bulletin, v. 66, p. 1581–1596.

Plink-Bjorklund, P., 2015, Morphodynamics of rivers strongly affected by monsoon precip-itation: Review of depositional style and forcing factors: Sedimentary Geology, p.110–147.

50

Plink-Bjorklund, P., Birgeneier, L., and Jones, E., 2014, Extremely bad early Eoceneweather: evidence for extreme precipitation from river deposits, in CBEP Climaticand Biotic Events of the Paleogene 2014, Ferrara, Italy, p. 175–176.

Plink-Bjorklund, P., and Birgenheier, L., 2013, Effects of Extreme Monsoon Precipitationon River Systems Form And Function, an Early Eocene Perspective, in AmericanGeophysical Union, Fall Meeting 2013, San Francisco, California.

Plink-Bjorklund, P., and Steel, R., 2005, Deltas On Falling-Stage and Lowstand Shelf Mar-gins, The Eocene Central Basin of Spitsbergen: Importance of Sediment Supply:Special Publication of SEPM, p. 179–206.

Plink-Bjorklund, P., and Steel, R.J., 2004, Initiation of turbidity currents: outcrop evidencefor Eocene hyperpycnal flow turbidites: Sedimentary Geology, v. 165, p. 29–52.

Postma, G., 2014, Generic autogenic behaviour in fluvial systems: Lessons from experimentalstudies: International Association of Sedimentology Special Publication, v. 46, p.1–18.

Postma, G., and Cartigny, M.J.B., 2014, Supercritical and subcritical turbidity currents andtheir deposits— A synthesis: Geology, v. 42, p. 987–990.

Postma, G., Cartigny, M., and Kleverlaan, K., 2009, Structureless, coarse-tail graded BoumaTa formed by internal hydraulic jump of the turbidity current? Sedimentary Geology,v. 219, p. 1–6.

Postma, G., Kleverlaan, K., and Cartigny, M.J.B., 2014, Recognition of cyclic steps insandy and gravelly turbidite sequences, and consequences for the Bouma facies model:Sedimentology, v. 61, p. 2268–2290.

Remy, R.R., 1992, Stratigraphy of the Eocene Part of the Green River Formation in theSouth-Central Part of the Uinta Basin, Utah: US Geological Survey Bulletin 1787-BB, p. 79.

Rice, S.P., and Church, M., 2010, Grain-size sorting within river bars in relation to down-stream fining along a wandering channel: Sedimentology, v. 57, p. 232–251.

Rosenberg, M.J., Birgenheier, L.P., and Vanden Berg, M.D., 2015, Facies, StratigraphicArchitecture, and Lake Evolution of the Oil Shale Bearing Green River Formation,Eastern Uinta Basin, Utah, in Smith, M.E. and Carroll, A.R. eds., Stratigraphy andPaleolimnology of the Green River Formation, Western USA, Syntheses in Limnogeol-ogy, Dordrecht, Springer Netherlands, Syntheses in Limnogeology, v. 1, p. 211–249.

Ryder, R.T., Fouch, T.D., and Elison, J.H., 1976, Early Tertiary sedimentation in the westernUinta Basin, Utah: Geological Society of America Bulletin, v. 87, p. 496.

51

Sambrook Smith, G.H., Ashworth, P.J., Best, J.L., Lunt, I.A., Orfeo, O., and Parsons, D.R.,2009, The Sedimentology and Alluvial Architecture of a Large Braid Bar, Rio Parana,Argentina: Journal of Sedimentary Research, v. 79, p. 629–642.

Sambrook Smith, G.H., ASHWORTH, P.J., BEST, J.L., WOODWARD, J., and Simpson,C.J., 2006, The sedimentology and alluvial architecture of the sandy braided SouthSaskatchewan River, Canada: Sedimentology, v. 53, p. 413–434.

Schomacker, E.R., Kjemperud, A.V., Nystuen, J.P., and Jahren, J.S., 2010, Recognition andsignificance of sharp-based mouth-bar deposits in the Eocene Green River Formation,Uinta Basin, Utah: Sedimentology, v. 57, p. 1069–1087.

Scruton, P.C., 1960, Delta Building and the Deltaic Sequence, in Shepard, F.P., Phleger,F.B., and Van Andel, T.H. eds., Recent Sediments Northwest Gulf of Mexico, Tulsa,Oklahoma, American Association of Petroleum Geolgosits, v. 143, p. 82–102.

Shukla, U.K., Singh, I.B., Sharma, M., and Sharma, S., 2001, A model of alluvial megafansedimentation: Ganga Megafan: Sedimentary Geology, v. 144, p. 243–262.

Simons, D.B., Richardson, E. V., and Nordin, C.F., 1965, Sedimentary structures generatedby flow in alluvial channels, in Middleton, G. V. ed., Primary Sedimentary Structuresand their Hydrodynamic Interpretation, The Society of Economic Paleotologists andMineralogists (SEPM), p. 34–52.

Singh, A., and Bhardwaj, B.D., 1991, Fluvial facies model of the Ganga River sediments,India: Sedimentary Geology, v. 72, p. 135–146.

Singh, H., Parkash, B., and Gohain, K., 1993, Facies analysis of the Kosi megafan deposits:Sedimentary Geology, v. 85, p. 87–113.

Smith, M.E., Carroll, A.R., and Scott, J.J., 2015, Stratigraphic Expression of Climate, Tec-tonism, and Geomorphic Forcing in an Underfilled Lake Basin: Wilkins Peak Memberof the Green River Formation, in Smith M., C.A. ed., Stratigraphy and Paleolimnologyof the Green River Formation, Western USA. Syntheses in Limnogeology, Springer,Dordrecht, p. 61–102.

Smith, M.E., Chamberlain, K.R., Singer, B.S., and Carroll, A.R., 2010, Eocene clocks agree:Coeval 40Ar/39Ar, U-Pb, and astronomical ages from the Green River Formation:Geology, v. 38, p. 527–530.

Smith, D.G., Hubbard, S.M., Lavigne, J.R., Leckie, D.A., and Fustic, M., 2011, Stratigraphyof counter-point bar and eddy-accretion deposits in low energy meander belts of thePeace–Athabasca delta, northeast Alberta, Canada, in Davidson, S.K., Leleu, S., andNorth, C.P. eds., From River to Rock Record: The Preservation of Fluvial Sedimentsand Their Subsequent Interpretation, SEPM special publication, 97, p. 143–152.

52

Sneh, A., 1983, Desert stream sequences in the Sinai Peninsula: Journal of SedimentaryResearch, v. 53, p. 1271–1279.

Spinewine, B., Sequeiros, O.E., Garcia, M.H., Beaubouef, R.T., Sun, T., Savoye, B., andParker, G., 2009, Experiments on Wedge-Shaped Deep Sea Sedimentary Depositsin Minibasins and/or on Channel Levees Emplaced by Turbidity Currents. Part II.Morphodynamic Evolution of the Wedge and of the Associated Bedforms: Journal ofSedimentary Research, v. 79, p. 608–628.

Stear, W.M., 1985, Comparison of the Bedform Distribution and Dynamics of Modern andAncient Sandy Ephemeral Flood Deposits in the Southwestern Karoo Region, SouthAfrica: Sedimentary Geology, v. 45, p. 209–230.

Symons, W.O., Sumner, E.J., Talling, P.J., Cartigny, M.J.B., and Clare, M.A., 2016, Large-scale sediment waves and scours on the modern seafloor and their implications for theprevalence of supercritical flows: Marine Geology, v. 371, p. 130–148.

Taki, K., and Parker, G., 2005, Transportational cyclic steps created by flow over an erodiblebed. Part 1. Experiments: Journal of Hydraulic Research, v. 43, p. 488–501.

Thomas, J. V., Parkash, B., and Mohindra, R., 2002, Lithofacies and palaeosol analysisof the Middle and Upper Siwalik Groups (Plio-Pleistocene), Haripur-Kolar section,Himachal Pradesh, India: Sedimentary Geology, v. 150, p. 343–366.

Tunbridge, I.P., 1981, Sandy high-energy flood sedimentation - Some criteria for recognition,with an example from the devonian of S.W. England: Sedimentary Geology, v. 28, p.79–95.

Turner, B.R., 1986, Tectonic and climatic controls on continental depositional facies inthe Karoo Basin of Northern Natal, South Africa: Sedimentary Geology, v. 46,p. 231–257.

Tyler, N., and Ethridge, F.G., 1983, Depositional Setting of the Salt Wash Member of theMorrison Formation, Southwest Colorado: SEPM Journal of Sedimentary Research,v. Vol. 53.

Ventra, D., Cartigny, M.J.B., Bijkerk, J.F., and Acikalin, S., 2015, Supercritical-flow struc-tures on a Late Carboniferous delta front: Sedimentologic and paleoclimatic signifi-cance: Geology, v. 43, p. 731–734.

Williams, G.E., 1971, Flood Deposits of the Sand-bed Ephemeral Streams of Central Aus-tralia: Sedimentology, v. 17, p. 1–40.

Winterwerp, J.C., Bakker, W.T., Mastbergen, D.R., and van Rossum, H., 1992, Hyper-concentrated Sand-Water Mixture Flows over Erodible Bed: Journal of Hydraulic

53

Engineering, v. 118, p. 1508–1525.

Zaleha, M.J., 1997, Fluvial and lacustrine palaeoenvironments of the Miocene Siwalik Group,Khaur area, northern Pakistan: Sedimentology, v. 44, p. 349–368.

54

CHAPTER 3

VERTICAL AND LATERAL FACIES VARIATIONS IN FLUVIAL FANS: EARLY

EOCENE GREEN RIVER FORMATION, UINTA BASIN

A paper submitted to Basin Research

Jianqiao Wang*,1 and Piret Plink-Bjorklund1

3.1 Abstract

In this study, outcrop measured sections and photomosaics with a GPS survey are inte-

grated with areal mapping of channel dimensions, channel to floodplain ratio, and sedimen-

tary facies variability to study the channel and floodplain deposits in the Sunnyside Delta

Interval of the Early Eocene Green River Formation in the Uinta Basin. The documented

lateral extent, internal architecture, and lateral and vertical facies transitions indicate a flu-

vial (mega)fan. Sandying and thickening upward successions occur as an increase in channel

to floodplain ratio, channel size, and the degree of channel amalgamation. Similar trends are

also observed laterally as channel fill facies become more heterolithic, channel amalgamation

degree decreases and the proportion of floodplain deposits increases downstream. There are

multiple scales of upward sandying and thickening packages. The smallest scale packages

are avulsion packages that form the building blocks of the stratigraphy. The larger–scale

upward thickening and sandying successions are interpreted as whole fan and lobe progra-

dation. High avulsion rates and channel return frequency are interpreted to control the high

degree of channel amalgamation in the proximal fans.

The avulsion stratigraphy indicates that the succession is comparable to fluvial megafans

rather than to distributive fluvial systems (DFS), fluvial distributary systems or terminal

∗Primary and corresponding author. Direct correspondence to [email protected] Stratigraphy Group, Department of Geology and Geological Engineering, Colorado School ofMines, 1500 Illinois Street, Golden, Colorado 80401, USA.

55

fans that all invoke bifurcations rather than avulsions. Furthermore, the abundant Froude

supercritical flow and high deposition rate facies, in-channel mud deposits and in-channel

bioturbation and desiccation indicate variable discharge rivers, similar to modern fluvial

megafans.

3.2 Introduction

Large fluvial fan systems or fluvial megafans can cover areas of 1,000 to 100,000 km2 (Leier

et al., 2005), and have a significant hydrocarbon reservoir potential. While previous studies

on Pleistocene to modern fluvial fans (e.g. Shukla et al., 2001; Assine, 2005; Wilkinson et

al., 2006; Chakraborty et al., 2010; Latrubesse et al., 2012; Rossetti et al., 2012; Assine et

al., 2014; Pupim et al., 2017) have assessed their geomorphology, self-organization processes,

as well as their link to foreland basins and seasonal climate (Leier et al., 2005; Fontana et

al., 2008, 2014; Assine et al., 2014), documentation on how fluvial facies vary laterally is

rare (Singh et al., 1993; Shukla et al., 2001; Chakraborty et al., 2010). Ancient datasets

provide data on facies and their variability, but some systems are interpreted as fluvial

megafans (e.g. DeCelles and Cavazza, 1999; Horton and Decelles, 2001; Saez et al., 2007;

Trendell et al., 2013), others as distributive fluvial systems (DFS) (Leleu and Hartley, 2010;

Leleu et al., 2010; Weismann et al., 2014; Owen et al., 2015, 2017), fluvial distributary

systems (Nichols and Hirst, 1998; Nichols and Fisher, 2007; Fisher and Nichols, 2013), or

terminal fans (Kelly and Olsen, 1993; Pusca, 2003). In contrast to the widely used sequence

stratigraphy models that link amalgamated and isolated channels zones to lake or sea-level

cycles (e.g. Keighley et al., 2003), degree of channel amalgamation has been shown to vary

from proximal to distal, and thus progradation and backstepping of fluvial fans builds a

predictable stratigraphy (Singh et al., 1993; Shukla et al., 2001; Chakraborty and Ghosh,

2010; Weissmann et al., 2014; Owen et al., 2017). Such proximal-distal relationships in

facies trends (Trendell et al., 2013; Weissmann et al., 2014; Owen, et al., 2015, 2017), and

paleosol moisture variability (Bhattacharyya et al., 2011; Hartley et al., 2013) have also been

56

recognized in ancient systems.

This paper focuses on documentation of vertical and lateral facies relationships that goes

beyond the simple proximal-distal and vertical trends and links channel lithesome types and

floodplain sand content to their position in the fan. We compare to modern fan systems, and

discuss the role of self-organization in building the fan stratigraphy, and how it relates to

the different fan models. The specifics of channel architecture and the interaction of fluvial

and lacustrine deposits are not discussed in this paper, but addressed in companion studies.

3.3 Geological setting and background

The 25,000 km2 Uinta Basin was formed to the east of the Sevier Thrust Front in response

to the Laramide uplift of the Uinta Mountains to the north (Cashion, 1967; Dickinson

et al., 1988; DeCelles, 2004). The basin is approximately 210 km from west to east and

275 km from north to south (Cashion, 1967). Fluvial facies in the Green River Fm are

extensively distributed across the southern margin of the Uinta Basin (Ryder et al., 1976).

The Green River Fm is subdivided into three informal units: lower, middle, and upper, based

on lacustrine pulses including the Uteland Butte Member, the Carbonate Marker Unit, and

the Mahogany Oil Shale Member (Fig. 3.1) (Fouch, 1975; Pitman et al., 1982; Fouch et al.,

1987). The three major siliciclastic intervals are the Colton Tongue, the Renegade Tongue,

and the Sunnyside Delta Interval. D and C Marker beds mark the bottom and top of the

Sunnyside Delta Interval (Remy, 1992; Morgan, 2003). The C marker is equivalent to the C1

marker of Jacob (1969). The timing of the Sunnyside Delta Interval deposition (Fig. 3.1) is

constrained by an absolute age date from the C marker of 49.6 Ma (Smith et al., 2015) and

54 Ma from the Carbonate Marker Unit (Remy, 1992).

Previous work has mainly focused on the fluvial-deltaic deposition and sequence stratig-

raphy of the Sunnyside Delta Interval (Remy, 1992; Keighley et al., 2003; Taylor and Ritts,

2004; Schomacker et al., 2010; Moore et al., 2012), but some recent studies have intperpreted

that the Sunnyside Delta Interval is one of the fluvial fans (Plink-Bjorklund and Birgenheier,

57

2013; Plink-Bjorklund et al., 2014) developed by the California paleoriver that originated

from the Mojave segment of the Cordilleran magmatic arc (Davis et al., 2010; Dickinson et

al., 2012). The Sunnyside Delta Interval has also been previously interpreted as a terminal

fan deposit (Pusca, 2003). The Wasatch/Colton age fan apex is estimated to have prograded

from about 250 km south of the Uinta Basin into the southern part of the basin (Jones, 2017).

Over a thousand paleocurrent measurements from Colton/Wasatch Formation were collected

along the south basin margin (Jones, 2017) to reconstruct the apex location. Paleocurrent

measurements from Jones (2017) and other previous studies indicate consistent northward

(ranging from northwest to northeast) fluvial sediment transport in the southern Uinta Basin

margin in Colton/Wasatch as well as Green River Formations (Fouch et al., 1976; Dickinson

et al., 1986; Remy, 1992; Schomacker et al., 2010; Ford et al., 2016; Gall et al., 2017; Jones,

2017).

3.4 Methods and dataset

Field data from the Sunnyside Delta Interval (Fig. 3.1) was collected from the northwest-

to-southeast oriented Nine Mile Canyon and several side canyons in north-to-south directions

(Fig. 3.1). Observations cover an area from the 32- to 45-mile marker across ca 20 km in

the Nine Mile Canyon, and continue to the mile marker 49 to the south in the Cottonwood

Canyon. A total of about 550 m of stratigraphic sections were measured at 10 cm resolution.

Percentage of sedimentary facies was calculated using measured sections. Lateral relation-

ships of channel-floodplain lithosomes were examined on cliff-face exposures and mapped by

walking out the stratigraphic intervals in continuous outcrops or using the bounding carbon-

ate marker units (D and C markers) as correlation guides in discontinuous outcrops. Where

possible, channel lithosome bounding surfaces were physically followed to where they pinch

out. Photomosaics were used to assist with analyzing channel architecture, and calculating

channel-floodplain proportions. Architectural dimensions (e.g. channel width and thickness)

were obtained by using laser range finder (TRUPULSETM 360B) in field.

58

3.5 Facies Associations in Fluvial Channel-floodplain Lithosomes

Fluvial channel-floodplain lithosomes are organized into facies associations according to

their relative geographic location and lateral relationships, and defined by lithology, channel

size, channel lithesome architecture, degree of channel amalgamation, channel-floodplain

deposit ratio, and trace fossils. Facies are summarized in Table 3.1.

3.5.1 Facies Association 1: Large Vertically and Laterally Amalgamated Sandy

Channel and Floodplain Lithosomes

Facies association 1 consists of amalgamated sandstones 10-35 m thick and 100s of m

to km wide, and interbedded mudstones and sandstones with variegated colors (Fig. 3.2).

Basal surfaces are either concave-up or undulating. These sandstones have the appearance

of a sandstone wall in outcrops, as they form 85% of the facies association. They commonly

display multiple internal irregular concave-up erosion surfaces (Figs. 3.2B and C) with

mudstone clasts along some of these surfaces. The sandstones consist of erosionally bound,

downstream- (Fig. 3.2B) and upstream-oriented accretion sets (Fig. 3.2C). Local erosional

relief on the compound basal surface is a few meters, whereas the accretion set boundaries

that are also erosional are 10-15 m tall (Fig. 3.2B). The thick sandstones of facies association

1 are characterized by a high proportion of climbing and aggradational bedforms (78-100%,

F1.2, 1.4, 1.5, and 1.8 in Table 3.1) low-angle convex- and concave-up sedimentary structures

(40-85%, F1.2, 1.4, 1.5, and 1.6 in Table 3.1). The dominant sedimentary structures are scour

and fill structures (F1.2, Figs. 3.3A and B, Table 3.1), convex-up low-angle laminations

(F1.4, Fig. 3.3C, Table 3.1), gradational and distinct planar laminations (F1.5 and 1.6,

Figs. 3.3D and E, Table 3.1), and soft sediment deformation (F1.8, Fig. 3.3F, Table 3.1). In

the upstream migrating bedforms, a typical vertical transition from soft-sediment deformed

sandstone above the basal scour surfaces to sandstones with gradational planar lamina,

climbing cross stratification, and scour and fill structures is common (Fig. 3.2C).

59

Interbedded mudstones and sandstones form less than 15% of facies association 1 (Fig.

3.2, Table 3.2). In places they only occur as lenses below the irregular erosion surfaces of

the thick sandstones (Fig. 3.2D). The sandstone beds consist of up to meter-thick very-fine

grained sandstone with planar laminations (F1.6, Table 3.1, Fig 3.3E) and ripple cross lamina

(F2.3, Table 3.1, Figs. 3.2D and 3.3G). Sandstones form 70% of the interbedded sandstone

and mudstone component of facies association 1 (Table 3.2). Mudstones vary in color from

purple to greenish grey, and are commonly soft sediment deformed (Figs. 3.3G and H, Table

3.1). Some mudstones are purple (F3.1) or brown (F3.3) throughout and others transition

from purple to greenish grey (F3.4), or have irregular brownish or grey patches (Table 3.1).

In places, the thick sandstones directly overlay mudstones without the sandstone interbeds

(Figs. 3.3G and H). The mudstone facies (Table 3.1) are the same as in facies associations

2 and 3, where they occur in greater thickness and lateral extent, and are better preserved.

Interpretation: The sandstone wall outcrop appearance is due to the high amalgamation

degree of sandstones. The accretion sets are interpreted as bar clinoforms (sensu Mohrig et

al., 2000), and the thick sandstones are interpreted as amalgamated channels. The dominant

sedimentary structures are indicative of Froude supercritical flow, and high deposition rates

(Alexander et al., 2001; Cartigny et al., 2014), which are characteristic for variable discharge

rivers where deposition occurs from high magnitude floods (Fielding, 2006; Fielding et al.,

2009; Plink-Bjorklund, 2015). The internal erosion surfaces outline individual flood event

beds (sensu flood units in Plink-Bjorklund, 2015) (Figs. 3.2B to D). The upstream migrating

macroforms are interpreted as deposition in cyclic steps because the systematic backsets are

similar to cyclic step deposits produced by experiments, where they occur as bounded by

the upstream facing erosion surfaces and separated by regularly spaced scours (Cartigny

and Postma, 2010; Kostic et al., 2010; Cartigny et al., 2014). The resultant vertical trend

in sedimentary structures in the backsets is characteristic for upstream migration of cyclic

steps, indicating rapidly deposited hydraulic jump deposits above the basal erosion surface,

overlain by sub- to super-critical flow deposits deposited further downstream (Postma et al.,

60

2014). The interbedded mudstones and sandstones are interpreted as floodplain deposits

with interbedded floodplain mudstones and crevasse splay or sheetflow deposits. Mottled

coloring is common in paleosols (Retallack, 1988; Leckie et al., 1989; Wright, 1992), and is a

result of local redox changes due to fluctuating drainage conditions (Wright, 1992). Mottled

color and soft sediment deformation are characteristic for paleosol with variable drainage

conditions (Retallack, 1988; Leckie et al., 1989; Wright, 1992). The direct juxtaposition of

channel lithesome and floodplain mudstone indicates abrupt avulsions (Jones and Hajek,

2007). The high degree of vertical and lateral channel amalgamation indicates high avulsion

rates and a high channel return frequency (Hajek and Wolinsky, 2012; Hajek and Edmonds,

2014), and is the reason for low preservation of floodplain facies.

3.5.2 Facies Association 2: Laterally Amalgamated Sandy to Heterolithic Chan-

nel Accretion Set and Floodplain Lithosomes

Facies association 2 is dominated by 3-15 m thick laterally amalgamated sandstones with

low (4-10◦) or high angle (up to 20◦) downstream and upstream accretion sets (Fig. 3.4).

The degree of sandstone amalgamation is lower than in facies association 1, and accretion

set thickness is 57-70% of the thickness in facies association 1 (Table 3.2). The overall

sandstone and heterolithic accretion set lithosome percentage is 40-70% (Table 3.2). The

low angle accretion sets consist of tabular or wedge-shaped sandstones, 0.5-2 m thick, and

form at least 100s of m wide lithosomes in strike view (Fig. 3.4A). The high angle sets

are 10s of m wide in a flow perpendicular view, and composed of sandstone layers that

are a few dm to one meter thick and thin towards either the base or the top of the sets.

Facies association 2 has a relatively lower proportion of climbing and aggradational bedforms

(35-80%, F1.2, 1.3, 1.5, 1.6, and 2.4 in Table 3.1) and low-angle convex- and concave-up

sedimentary structures (35-73%, F1.2, 1.3, 1.5, and 1.6 in Table 3.1). The most common

facies are scour and fill structures (15%, F1.2), ripple-filled scour and fill structures (28%,

F1.3), gradational and distinct planar laminations (30%, F1.5 and 1.6), and climbing ripple

61

laminations (27-35%, F2.4) (Table 3.1, Fig. 3.5). Vertical transition from low-angle convex-

and concave-up sedimentary structures (F1.2, 1.3, 1.5, and 1.6) to climbing ripples (F2.4) is

common (Fig. 3.5A). Some lithosomes are entirely composed of climbing ripple laminations

(Fig. 3.5B). Some accretion sets fine from sandstones to mm-dm thick mudstones (F3.1 to

3.4) laterally within individual accretion sets. In places flute casts occur below the accretion

sets (Fig. 3.5C). Mm to dm thick in-channel mud layers (Fig. 3.5D), with desiccation cracks

(Fig. 3.5E) and terrestrial trace fossils up to 1 dm long (Figs. 3.5G and H) are common at

accretion set boundaries.

The sandstone and heterolithic accretion set lithosomes are laterally and vertically as-

sociated with interbedded mudstones and sandstones with varigated color. The mudstones

are dm thick and purple (F3.1, Fig. 3.5I) or brown (F3.3) throughout, transition from pur-

ple to greenish grey (F3.4), or have irregular brownish or grey patches (F3.4, Table 3.1).

Sandstones are commonly dm thick, but thicken upward to m-thick beds in up to 10 m

thick intervals that have a lateral extent of 100s of meters (Figs. 3.4A and C). The vertical

transition from interbedded sandstone and mudstone with upward thickening sandstones to

sandstone and heterolithic accretion set lithosomes is common (Figs. 3.4A and C). In a few

places there is an abrupt transition from mudstones to sandstone (Fig. 3.5I) and heterolithic

accretion set lithesomes. The sandstone proportion in the interbedded mudstone and sand-

stone lithosomes decreases from 70% to around 50% compared to facies association 1 (Table

3.2). Bioturbation is common in the purple mudstone. In one area, possible mammal tracks

occur in a 3 dm thick heavily bioturbated purple mudstone (Fig. 3.5H). Calcite filled cracks

and calcite nodules are common in all four mudstone facies.

Interpretation: The accretion sets are interpreted as bar clinoforms (sensu Mohrig et

al., 2000), and the sandstones lithosomes are interpreted as laterally amalgamated chan-

nels. The dominant sedimentary structures are indicative of Froude supercritical flow, and

high deposition rates (Alexander et al., 2001; Cartigny et al., 2014), and are characteristic

for variable discharge rivers where deposition occurs from high magnitude floods (Field-

62

ing, 2006; Fielding et al., 2009; Plink-Bjorklund, 2015). The common transition between

Froude supercritical flow structures and climbing ripples is suggestive of a rapid decline in

flow velocity, a characteristic of flood events (e.g. Mckee et al., 1967; Picard and High,

1973; Croke et al., 1998; Plink- Bjorklund, 2015). In-channel mud layers are interpreted

to be deposited from even more rapidly waning floods (e.g., Jones, 1977; Tunbridge, 1981;

Plink- Bjorklund, 2015). The presence of trace fossils and desiccation cracks in channels

at accretion set boundaries indicate that the channels were dry and the beds exposed for

sustained periods between floods (Sneh, 1983; Nichols and Fisher, 2007; Plink- Bjorklund,

2015). Mottled color, roots, and nodules are common in paleosols (Retallack, 1988; Leckie et

al., 1989; Wright, 1992). Burrowing activity and roots are indicators for subaerial exposure

and surface stability (Retallack, 1988; Leckie et al., 1989). The calcite filled cracks are pos-

sible calcified root structures (Leckie et al., 1989). Mottling is a result of local redox changes

due to fluctuating drainage conditions (Wright, 1992). Red mudstones indicate well drained

floodplain, whereas purple mudstones had higher soil moisture, and brown mudstones indi-

cate sustained soil moisture (Kraus and Aslan, 1996; Alonso-Zarza, 2003). The thickening

upward trends have been interpreted as transitional river avulsion deposits (Kraus and Wells,

1999; Jones and Hajek, 2007) that indicate successively larger amounts of sand transported

to floodplain, followed by a channel avulsion. In contrast, direct transition from floodplain

mudstone to channel lithesome is interpreted to be associated with abrupt avulsion, where

channel avulsion is not preceded by gradual splay buildup (Jones and Hajek, 2007). While

the latter is likely to indicate abrupt avulsions, it is unclear whether the former is a strong

indicator for gradual avulsions, as there is no robust way to test the genetic link between the

channel and the floodplain deposits. Furthermore, the floodplain sandstones may originate

from sheetfloods rather than distinct splays.

63

3.5.3 Facies Association 3: Small Isolated Channel and Floodplain Lithosomes

Facies association 3 consists of isolated or partially amalgamated lenticular lithosomes

with heterolithic aggradational accretion sets (Fig. 3.6B) that form ca 10-30% of the facies

association, and interbedded sandstones and mudstones. The lenticular lithosome thickness

decreases by ca 30-50% compared to the accretion set thickness in facies association 2 (Table

3.2). The heterolithic aggradation sets consist of interbedded dm to m thick sandstones and

dm thick mudstone layers in 2-10 m deep and 10s to a 100 m wide erosion surfaces. The

sandstone layers commonly thin towards channel margins (Figs. 3.6A and B). Some termi-

nate onto the basal erosion surface and others are truncated on top. Only 10-20% low-angle

convex- and concave-up sedimentary structures (F1.5, 1.6, and 1.8 in Table 3.1) including

gradational and distinct planar laminations (F1.5 and 1.6) and soft sediment deformation

(F1.8) occur in these lenticular lithosomes. Climbing ripple laminations are dominant with

40-80% (F2.4, Fig. 3.7A, Table 3.1). Mudstone layers are greenish grey or purple (Figs. 3.5E

and F), and contain some plant fragments and brownish oxidization bands (Fig. 3.7B). Des-

iccation cracks occur at some accretion set boundaries. Sandstones are commonly burrowed

throughout or at the accretion set boundaries (Fig. 3.7D).

The interbedded mudstones and sandstone lithosomes form 70-90% of facies association

3 and consist of dm-thick mudstone (F3.1-3.4) and cm to dm thick ripple laminated (F2.3)

or structureless (F1.7) very-fine grained sandstone. Interbedded mudstones and sandstones

form up to 10s of meter thick lithosomes that are laterally continuous across 100s of m to

kms and show upward thickening sandstone trends (Figs. 3.6B and 3.7E), or are laterally

eroded by lenticular lithosomes. Commonly, the interbedded mudstone and sandstone beds

are extensively rooted or bioturbated and sedimentary structures are not preserved (Fig.

3.7C). In places the mudstones coarsen upward into ripple laminated sandstones on dm scale

(Fig. 3.7F).

Interpretation: The erosionally based lenticular lithosomes are interpreted as isolated

channels encased in floodplain deposits. There is a considerably smaller amount of Froude

64

supercritical flow sedimentary structures, and the abundance of climbing ripples and grada-

tional planar laminations indicates high deposition rates. Desiccation cracks, trace fossils,

and pedogenically modified in-channel mudstones bound flood event beds indicate that chan-

nels were dry for sustained periods. Such desiccated in-channel mudstones, and climbing-

ripple filled channels are shown to occur in variable discharge rivers (Sneh, 1983; Nichols

and Hirst, 1998; Shukla et al., 2001; Nichols and Fisher, 2007; Allen et al., 2011; Donselaar

et al., 2013; Plink-Bjorklund, 2015). The interbedded mudstone and sandstone beds are

interpreted as crevasse splays and floodplain mudstones, due to their lateral extent, presence

of roots, and the internal coarsening upward trend from structureless mudstone to ripple

laminated sandstone (Mohrig et al., 2000; Jones and Hajek, 2007; Sendziak, 2012).

3.6 Vertical Trends

Upward sandying and thickening trends on multiple scales are common in all the facies

associations. The smallest scale of such trends consist of upward thickening floodplain sand-

stones, interbedded with mudstones and capped with channels, mostly 2-5 m thick (Figs.

3.8A-C; see also Figs. 3.2D, 3.4A, C, and 3.6). These upward sandying packages are lat-

erally continuous for at least 100s of m to kms (Figs. 3.4A and 3.6A). Statistical analysis

shows that such packages can be recognized on 87% of the beds in measured sections as bed

thickening as well as grain-size-coarsening trends (Figs. 3.8C-E). A total number of 82 such

packages are documented in the numerical dataset and their thickness ranges from 1.5-20

m with a median of 8 m, and 25th and 75th percentiles at 2-12 m (Fig. 3.8D). The thick

outliers are likely to contain multiple packages that are not recognized by the analyses as

they lack the capping channel deposits.

Intermediate scale upward thickening and sandying packages are most commonly 23-40

m thick and consist of upward transitions from facies association FA3 to FA2, from FA2 to

FA1, or from FA3 to FA1 (Figs. 3.2A, 3.4A, and 3.6A). These larger-scale packages extend

laterally for 10s of kilometers as many of such packages can be followed across the Nine Mile

65

Canyon study area (Fig. 3.9). A total number of 15 such packages are documented in the

numerical dataset and their thickness ranges from 22-55 m with a median of 29 m, and 25th

and 75th percentiles at 23-40 m (Fig. 3.8D).

The largest scale upward thickening and sandying trends are most commonly 59-103 m

thick (Fig. 3.8D) and consist of multiple upward facies association transitions, such as for

example the double stack from FA3 to FA2 in Fig. 3.6A. Lateral extent of these largest

scale packages is at least 20 km as they extend across the Nine-Mile Canyon study area, but

it is yet unclear if they extend across the whole basin. A total number of 4 such packages

are documented in the numerical dataset and their thickness ranges from 56-139 m with a

median of 78 m, and 25th and 75th percentiles at 59-103 m (Fig. 3.8D).

Interpretation: The smallest-scale packages are interpreted as avulsion packages as they

indicate splay or sheetflood buildout, followed by channel avulsions (Jones and Hajek, 2007).

Their lateral extent is dependent on the alluvial ridge width as determined by the distance of

overbank sediment transport during floods. These avulsion packages are the basic building

blocks of the studied fluvial stratigraphy. The origin of the intermediate scale packages is

less obvious and will be discussed later in this paper.

3.7 Lateral Trends

The Sunnyside Delta Interval has been mapped across the more than 150 km wide south-

ern basin margin. This mapping shows considerable variability in the distribution of facies

associations 1-3 within the fluvial succession, as well as in the proportions of the deltaic

and lake facies. In general, the proportion of FA3 increases considerably within the channel-

floodplain deposits, while channel size and the total proportion of channel-floodplain deposits

decrease towards the northwestern basin margin (Road 191 outcrops, Fig. 3.10). . The pro-

portion of FA1 is largest in the central part of the southern basin margin, such as in the

easternmost part of the Nine Mile Canyon and along the Green River (Fig. 3.1).

66

In addition to these basin-scale trends there is considerable lateral facies variability across

the ca 20 km wide continuous outcrops within the Nine Mile Canyon. A general trend is

that the degree of channel amalgamation, proportion of sandstone, channel size, and thus the

proportion of FA1 decrease from the southeasternmost outcrops in the Cottonwood Canyon

towards the northwest along the Nine Mile Canyon. The channel to floodplain ratio decreases

from ca 80% in the easternmost Nine Mile Canyon (mile markers 46-45) to ca 50% across

ca 20 km towards the northwest (mile marker 32) (Fig. 3.11). In addition to these general

trends there is further internal variability. For example, within certain stratigraphic intervals

the proportion of FA1 decreases from 90% to 30% as it transitions into FA3 over only a few

kilometers distance (Fig. 3.9). Whereas in other intervals a similar transition occurs more

gradually through transitioning into FA2 first and then into FA3 across ca 5 km (Fig. 3.9).

Yet in other intervals FA2 persists across the whole study area. At mile marker 46 and 45

(Figs. 3.11A and B), the upper part of the exposed succession is a “sandstone wall” with

amalgamated channels of FA1, and the lower part FA2 with a large heterolithic channel is at

marker 46. At mile 44.5, ca 800 m downstream, the uppermost “sandstone wall” (FA1) has

transitioned into an upward sandying package of FA3 and FA2 (Fig. 3.11C). The lower part

displays a lateral change from FA1 to FA2 and FA3 across only a few 100 m (Fig. 3.11C).

The exposure at mile marker 44 shows a larger proportion of the Sunnyside stratigraphy

(Fig. 3.11D), where the upper upward sandying packages contain FA3 and FA2, but the

stratigraphically lower section consists of FA1. At mile 41.2 (Fig. 3.11E) the upper part of

the stratigraphy is sandstone dominated as the uppermost upward sandying packages have

transitioned back to FA1 across ca 4.5 km. There are also some lake deposits that appear

at some of the upward sandying package boundaries (orange layers in Fig. 3.11E). At mile

marker 37, the whole succession consists of transitions between FA3 and FA2, and majority

of the channel fills are heterolithic (Fig. 3.11F). At mile marker 32 the whole Sunnyside

Interval is exposed, and much of the succession consists of deltaic and lake deposits. The

channel-floodplain intervals are relatively sandy and channel prone and consist of FA1-3 (Fig.

67

3.11G).

3.8 Discussion on Vertical and Lateral Trends

The large extent of the Sunnyside Interval river and floodplain deposits, together with the

downstream decrease of channel size, the degree of channel amalgamation, and increase in

floodplain deposit proportions suggest deposition in a fluvial (mega)fan system (sensu Singh

et al., 1993; Shukla et al., 2001). Fluvial fans are built by avulsions (Shukla et al., 2001; Leier

et al., 2005; Chakraborty et al., 2010; Latrubesse et al., 2012; Sinha et al., 2012) because

only avulsions as opposed to bifurcations (cf. Weissmann et al., 2010; see also discussion in

North and Warwick, 2007) can generate amalgamated channel belts. The degree of channel

amalgamation is highest in the most proximal fan where the area is smallest and the channel

return frequency is the highest (Chakraborty and Ghosh, 2010; see also Hajek and Wolinsky,

2012). The decrease in channel size is linked to water loss to floodplain and infiltration into

the fan deposits (Singh et al., 1993; Shukla et al., 2001). The increase in floodplain deposits

is related to the increase in outward fan surficial area and the decrease in channel size and

degree of amalgamation. Comparison to modern fluvial fans shows that channels are highly

amalgamated, such as in FA1, in proximal fans within 5-25 km of the fan apex in the Kosi

fan (Singh et al., 1993; Chakraborty et al., 2010), and within 20 km in the Ganga fan (Shukla

et al., 2001). The 40-70% channel lithesome proportion of FA2 is similar to the medial fan

zone in the Kosi fan where the floodplain mud volume increases to 45% from medial to

distal fan within 25-130 km from fan apex (Singh et al., 1993). In the medial zone of the

Ganga fan, sandy to heterolithic low-angle downstream and upstream accretion sets, and

vertical aggradation sets are documented in the braided zone (Shukla et al., 2001). Within

the 50-80 km from fan apex in the Ganga fan, shallow and narrow channels are interbedded

with floodplain muds and thin sandstones (Shukla et al., 2001), similar to FA2. The distal

Ganga fan, 80-100 km from fan apex is characterized by meandering channels with broad

floodplains (Shukla et al., 2001), similar to the laterally extensive floodplain deposits in FA3.

68

Upward thickening and coarsening trends have been recognized in modern fans as indica-

tive of fan progradation (Chakraborty and Ghosh, 2010; Sinha et al., 2014). Such fan-scale

progradation trends have also been suggested in ancient systems (DeCelles and Cavazza,

1999; Uba et al., 2005; Weissmann et al., 2013; Owen et al., 2015, 2017). Yet this dataset

shows that not all upward thickening and coarsening trends occur across the entire fan, or

even across 20 km (Fig. 3.11). Furthermore, the here documented multiple scales of upward

thickening and coarsening packages (Fig. 3.8E), and the recognition of individual avulsion

packages (Figs. 3.8A-C) suggest that there are multiple mechanisms for generating such

trends. The avulsion scale packages are likely to have the smallest lateral extent, as deter-

mined by the width of overbank sediment transport. These channel-floodplain processes on

alluvial ridge-scale are likely to determine the shortest scale lateral changes, such as tran-

sitions from FA1 to FA2 and FA3 across merely a few 100 m to kms (Fig. 3.11C), where

we observe a lateral transition from a locally amalgamated channel belt to a floodplain-

dominated succession. While the largest scale vertical upward thickening trends (Fig. 3.8E)

are likely to correspond to the entire fan progradation, it remains unclear what controls the

intermediate scale packages. A possible hypothesis is lobe-scale progradation, as modern

fans are documented to consist of 3-4 lobes (Assine and Silva, 2009; Chakraborty et al.,

2010; Chakraborty and Ghosh, 2010; Assine et al., 2014). Such lobes form because channel

avulsions cluster within lobes (Chakraborty et al., 2010).

3.9 Controls on Fluvial Fan Formation

Fluvial fan formation has been linked to a sufficient aggradation rate and discharge, if a

river has a variable (seasonally fluctuating) discharge (Leier et al., 2005). Channel bed aggra-

dation and the resultant channel superelevation are the main trigger of avulsions (Bryant et

al., 1995; Mohrig et al., 2000; Makaske, 2001; Stouthamer and Berendsen, 2001; Jerolmack

and Mohrig, 2007; Sinha and Sarkar, 2009; Hajek and Edmonds, 2014). The high depo-

sition rate sedimentary structures (F1.1-1.8, F2.2, and F2.4, Table 3.1) indicate that high

69

deposition rates were common in the Sunnyside system. The latter, together with abundant

Froude supercritical flow sedimentary structures (F1.1-1.8, Table 3.1), in-channel mudstones

(F3.1-3.4, Table 3.1), and in-channel desiccation and pedogenic modification indicate depo-

sition in variable discharge rivers (Plink-Bjorklund, 2015). Large discharge is evidenced by

the large bar clinoform height, as well as suggested by the size of the California paleoriver

(Dickinson et al., 2012).

Modern fluvial fans have high avulsion rates. For example, the Kosi river experienced

avulsions on average every 7 years between 1760 and 1960 (Chakraborty et al., 2010). The

Kosi river is currently confined by artificial levees, but still experienced avulsions in 2008

and 2009 (Chakraborty et al., 2010). The 2008 avulsion was an abrupt avulsion where the

Kosi River shifted more than 100 km eastward across the fan (Sinha, 2009). This avulsion

occurred during the annual monsoon flood (Sinha, 2009; Sinha et al., 2013). The avulsed

channel reoccupied one of its former paleochannels and 80–85% of the river flow was diverted

into the new course. The avulsion occurred as a result of channel bed aggradation and

superelevation and a resultant artificial levee break (Sinha, 2009).

3.10 Modern Analogue

The Ili River fluvial fan is deposited on the southeastern margin of Lake Balkash in

Khazakstan, ca 600 km from the Tian Shan mountain front (Fig. 3.12). It is thus similar

to the setting of the Wasatch/Colton and Green River Formation fluvial systems, in terms

of the fan size and shape, the lake size, as well as the fan position away from the mountain

front (see also Jones, 2017). Lake Balkhash is a closed lake (Birkett, 1995) with a variable

surface area dependent on seasonal fluctuations and water input from the fluvial megafan

(Kezer and Matsuyama, 2006). From Google Earth measurements, the fan area is ca 25,

000 km2 with an apex-distal fan distance of 250 km and a fan width of 200 km (Fig. 3.12),

similar to the estimated apex location of the Wasatch fluvial fan, which was 250 km south

of the Uinta Basin during deposition of the distal fan in the Uinta Basin (Jones, 2017). The

70

distance to Ili River fan from the Tian Shan Mountains is also comparable to that of the

California Paleoriver that drained from the Mojave region ca 750 km away from the Uinta

Basin (Davis et al., 2010; Dickinson et al., 2012).

3.11 Comparison to Fluvial Fan Models

The abundance of avulsion packages, and the channel amalgamation clearly demonstrate

that avulsions are the key process of lateral river mobility in the Sunnyside Delta Interval.

This is in agreement with our understanding of fluvial megafans (Shukla et al., 2001; Leier et

al., 2005; Chakraborty et al., 2010; Latrubesse et al., 2012; Sinha et al., 2012), but contadicts

the distributive fluvial systems (DFS) (Weissmann et al., 2010), fluvial distributary systems

(Nichols and Fisher, 2007) and terminal fan (Kelly and Olsen, 1993) models that all invoke

bifurcations rather than avulsions as a means of lateral channel mobility. For this reason all

these latter models have an inherent discrepancy, as their plan-view models show bifurcations,

but the cross-sectional views are based on data that indicate amalgamated to isolated channel

and floodplain deposits.

Pusca (2003) interpreted the Sunnyside Delta Interval as a terminal fan system (sensu

Kelly and Olsen, 1993; Legarreta and Uliana, 1998) and suggested wet-dry climate cycles in

the semi-arid climate zone. The terminal fan is interpreted to prograde and retrograde in

respond to the increasing and decreasing precipitation during wet and dry periods (Pusca,

2003). Pusca (2003) was also the first to observe that many of the sedimentary structures in

Sunnyside Delta Interval indicate high deposition rates. However, the term “terminal” that

implies the fan never reached to the lake (sensu Kelly and Olsen, 1993), is misleading, since

the sediments prograded into the basin and formed mouth bars and deltas as the fan met

the lake (Schomacker et al., 2010).

71

3.12 Conclusion

This outcrop study recognizes the Sunnyside Delta Interval channel and floodplain de-

posits as a fluvial fan system. Detailed study across 20 km in the Nine Mile Canyon links

channel and floodplain lithesome types to their position in the fan, and recognizes that self-

organization by avulsions is the key component of fan stratigraphy. Upward thickening and

sandying trends are recognized on multiple scales, and their origin ranges from avulsions to

lobe and fan progradation. Laterally, channel-floodplain lithosomes show a decreasing de-

gree of channel amalgamation, sand to mud ratio, and channel size from proximal to distal

fan. In addition to these general lateral trends, the degree of cannel amalgamation and the

channel to floodplain proportions vary laterally on multiple scales. We identify that as flu-

vial megafans rather than distributive fluvial system (DFS), fluvial distributary system, or

terminal fan models, as the latter invoke bifurcations rather than avulsions as the mechanism

for lateral channel mobility. A modern analogy of this type system was also provided.

72

Figure 3.1 Stratigraphic column and base map of Uinta Basin and Nine Mile canyon. (A)Geological map of the Wasatch/Colton (maroon) and Green River (orange) Formations inthe Uinta Basin, with study area locations at Nine Mile Canyon, Road 191 and Hay Canyon(modified from Dickinson et al., 2012; Schomacker et al., 2010; Sato and Chan, 2015; Jones,2017). (B) Study area at Nine Mile Canyon (see location in A) with figure locations andmile markers (grey numbers). Rose diagrams show paleocurrent directions measured in thisstudy. (C) Stratigraphic table with available age constraints (modified from Fouch et al.,1987; Remy, 1992; Smith et al., 2010, 2015).

73

Figure 3.2 Outcrop examples of facies association 1. Each has an interpreted pair of pho-tos. (A) Cliff-view at 46-mile marker in Nine Mile Canyon. (B) Vertically and laterallyamalgamated channel lithosomes with no floodplain deposits preserved. Internal erosionsurfaces bound flood event beds in downstream accretion sets. (C) Upstream migratingmacroforms with limited mudstone preserved below local erosional relief. (D) A lens of pre-served floodplain deposits (purple) with thick splay sandstones below a channel lithosome(yellow). Facies association and facies components are marked in figures. Large white arrowspoint in the general paleoflow direction.

74

Figure 3.3 Photos of common sedimentary structures in facies association 1. (A) Sandstonewith scour and fill structure (F1.2). (B) Sandstone with ripple filled scour and fill struc-ture (F1.3). (C) Sandstone with convex-up low angle lamina (F1.4). (D) Sandstone withgradational planar lamina and scour and fill (F1.5). (E) Sandstone with distinct planarlamina (F1.6). (F) Soft sediment deformed sandstone (F1.8). (G) Meter thick soft sedimentdeformed purple mudstone (F3.1) interbedded with bioturbated sandstone (F2.5) below achannel. (H) soft sediment deformed greenish grey mudstone (F3.2) below a channel.

75

Figure 3.4 Outcrop examples of facies association 2. (A) Cliff-view at 39-mile marker in NineMile Canyon. (B) Low angle downstream accretion sets. (C) High angle accretion sets dipobliquely upstream. (D) Downstream accretion sets. (E) Channel (yellow) and floodplain(purple) deposits with dm to m thick splay sandstones. Large white arrows point into thegeneral paleoflow direction.

76

Figure 3.5 Photos of common facies in facies associations 2 and 3. (A) Transition fromclimbing ripples (F2.4) to planar lamination (F1.5). (B) Sandstone with climbing ripplelaminations (F2.4). (C) Flute casts under a channel sandstone. (D) Thin layer of mixed colormudstone (F3.4) preserved between two channel sandstones. (E) Brownish mudstone withcalcite nodules. (F) Desiccation cracks preserved at accretion set boundary. (G) Roots insandstone (F2.5). (H) Vertical burrows in ripple laminated sandstone. (I) Possible mammaltracks preserved at the base of channel sandstone.

77

Figure 3.6 Outcrop examples of facies association 3. Each has an interpreted pair of photos.(A) Cliff-view of interbedded facies associations 2 and 3 in two upward thickening andsandying units. (B) Isolated channel lithosomes (yellow) with in-channel mudstone layers(gray) encased in floodplain deposits (purple) with dm-thick splay deposits. Large whitearrows point in the general paleoflow direction.

78

Figure 3.7 Photos of common facies in facies association 3. (A) Climbing ripple laminatedsandstone (F2.4). (B) Nodules in floodplain mudstone (F3.1). (C) Heavily rooted sandstonesurface. (D) Dm long root/burrow at accretion set boundary. (E) Coarsening upwardtrend from thinly laminated mudstone to ripple laminated sandstone (F2.3) in crevasse splaydeposits. (F) Close up of the ripple laminated sandstone (F2.3) in (E).

79

Figure 3.8 Examples of characteristic vertical trends. Photos (A, B) and a measured section(C) of upward thickening and sandying trends in floodplain deposits with capping channellithosomes. These upward sandying and thickening packages correspond to the smallest,avulsion scale packages. Statistical analysis reveal multiple scales of upward thickeningtrends (D, E). The blue lines in (D) trace the smallest-scale upward thickening trends. Thebox and whisker diagrams in (E) show three different scales in upward thickening trends. (F)Comparison of upward thickening trends to trends in grain size and sedimentary structuresshows that thickening trends are linked to sandying trends, but sedimentary structures donot follow these trends.

80

81

Figure 3.9 Vertical changes in channel (yellow) vs. floodplain (grey) proportions at eachexamined location (by mile marker). Percentages are estimated from combining measuredsection and photomosaics. Note the top interval gradually transitions from FA1 to FA2 thento FA3 from mile marker 46 to 43 (across ca 5km). Some other transitions from FA1 to FA3are more abrupt across over only a few km. Other intervals show FA2 persists across thewhole study area.

82

Figure 3.10 Photos showing thin floodplain-rich fluvial intervals along Road 191 outcrops.(A) Overview of fluvial (channels in yellow and floodplain in red) and lake facies (not colored)proportions. (B) Example of a thin fluvial interval with a heterolithic lenticular channeldeposit (yellow) and floodplain deposit with thin splay sands.

83

Figure 3.11 Example photomosaics that illustrate lateral changes from most proximal tomost distal locations, respectively (A through G). See Fig. 3.1 for mile marker locations.

84

85

86

87

Figure 3.12 Google Earth image showing Ili River that originates from the Tianshan Moun-tains and builds a large fluvial fan on the southern margin of the Balkhash Lake. (A) TheBalkhash Lake. (B) Close-up of the fan. White arrows show river’s path. Note that the fan(B) is not built at the foot of the mountains, but rather, the river transits for a few 100kmbefore it builds the fan.

88

Table 3.1: Description and Interpretation of Sedimentary Facies

Facies % Textures Structures Width/Thickness Interpretation1.1 Stratified and

imbricatedconglomerates

<1% Conglomerates (coarse sandto pebble sized carbonateclast/bioclast and/or mudclast, subrounded to angular,poorly sorted, varied inshape) in sandy matrix, clastor matrix supported

Imbricated,planarstratified, lowanglestratified,scour and fillstructures,

Thickness: a few cmper conglomeratelayer; Total thickness:a few dm - m

Intrabasinal clasts offloodplain mudstonesor lake carbonatesdeposited with sandmatrix by tractionflow (subcritical tosupercritical)

1.1 Disorganizedconglomeratesand diamictites

<1% Conglomerates (coarse sandto pebble sized carbonateclast/bioclast and/or mudclast, rounded to angular,poorly sorted, varied in shape)in sandy or muddy matrix, inplaces with terrestrial organicmatter (woody fragments)

Disorganized Thickness: a few dm -1m

Intrabasinal clasts offloodplain mudstonesor lake carbonatedeposited by collapseand fluidization ofchannel banks duringfalling stage of flow

1.2 Sandstone withscour and fillstructures

30% Upper very fine to lowermedium grained sandstone,moderately to poorly sorted,in places mixed withcarbonate grains

Scours filledwith upwardflatteninglamina

Thickness: a few dm -a few m; Width: afew dm - 10s of m

Supercritical flow,high deposition rates,suspensiondeposition, largeantidune or chute andpool formation

1.3 Sandstone withripple-filledscour and fillstructures

<1% Upper very fine to lower finegrained sandstone, moderatelysorted

Ripples fillthe toe partof upwardflatteningscour and filllamina

Thickness: a few dm -1m; Width: a few dm- a few m

Supercritical tosubcritical flow, highdeposition rates,suspension andbedload deposition

89

Table 3.1: Continued.

Facies % Textures Structures Width/Thickness Interpretation1.4 Sandstone with

convex-up lowangle lamina

8% Very fine to fine grainedsandstone, moderately topoorly sorted, in places mixedwith carbonate grains

Low angleconvex-uplaminations,very lowanglehummocky-like crossstrata withupwardincreasingsteepness

Thickness: a few dm -1m; Width: >1m

Supercritical flow,high deposition rates,suspensiondeposition, antiduneformation

1.5 Sandstone withgradationalplanar lamina

14% Very fine to fine grainedsandstone, moderately topoorly sorted

Diffused,gradationalplanarlaminations,each laminais internallyreversely tonormallygraded

Thickness (lamina):2mm - 2cm;Thickness (bed): afew dm - a few m

Subcritical tosupercritical flow,high deposition rates,suspensiondeposition, plane bedformation

1.6 Sandstone withdistinct planarlamina

20% Very fine to lower mediumgrained sandstone, moderatelysorted

Planarlaminationswith distinctlaminaboundaries

Thickness (lamina):mm scale; Thickness(bed): a few dm - afew m

Supercritical ortranscritical flow,normal depositionrates, bedloaddeposition, plane bedformation

1.7 Structurelesssandstone

1% Very fine to very coarsegrained sandstone, moderatelyto poorly sorted

No apparentsedimentarystructures

Thickness: a few cm -a few dm

high deposition rates,suspension deposition

90

Table 3.1: Continued.

Facies % Textures Structures Width/Thickness Interpretation1.8 Soft sediment

deformedsandstone

2% Very fine to lower mediumgrained sandstone, moderatelyto poorly sorted

Flamestructures,overturnedfolds, dishstructures,ball andpillowstructures.

Thickness: a few dm Water escape or localcollapse

2.1 Cross stratifiedsandstone

2% Fine to lower medium grainedsandstone, moderately sorted

Planar andtrough crossstratified

Thickness (cross set):5cm -20cm; Thickness(co-set): a few dm

Subcritical flow,normal depositionrates, bedloaddeposition, dunemigration

2.2 Sandstone withclimbing crossstrata

<1% Fine to lower medium grainedsandstone, moderately topoorly sorted

Planar andtrough crossstrata withclimbing setboundaries

Thickness (cross set):8cm - dm scale;Thickness (co-set): afew dm

Subcritical flow, highdeposition rates,suspension andbedload deposition,dune migration andaggradation

2.3 Cross laminatedsandstone

10% Very fine to lower fine grainedsandstone, moderately sorted

Asymmetricalcross-lamination

Thickness (cross set):cm scale < 5cm;Thickness (co-set): afew dm

Subcritical flow,normal depositionrates, bedloaddeposition, ripplemigration

91

Table 3.1: Continued.

Facies % Textures Structures Width/Thickness Interpretation2.4 Sandstone with

climbing ripplelamina

13% Very fine to fine grainedsandstone, moderately sorted

Asymmetricalcrosslaminationwith climbingsetboundaries

Thickness (cross set):cm scale <5cm;Thickness (co-set): afew dm - a few m

Subcritical flow, highdeposition rates,suspension andbedload deposition,ripple migration andaggradation

2.5 Bioturbatedsandstone

- Very fine to lower mediumgrained sandstone, moderatelyto poorly sorted

Vertical,horizontaland 3D tracefossils

Thickness: a few dm -a few m

Trace fossils formedby insects, dwelling,resting, crawlingtraces

3.1 Purpled redmudstone

- Clay to silt Laminated orblocky orcrumbly, inplaces rooted,bioturbated,or deformed

Thickness: a few cm -a few dm (blocky); afew mm (each lamina)

Oxidized, welldrained floodplainand in-channelmudstone

3.2 Greenish greymudstone

- Clay to silt Laminated orblocky orcrumbly

Thickness: a few cm -a few dm (blocky); afew mm (each lamina)

Reduced to poorlydrained floodplainand in-channelmudstone

3.3 Brown to darkgrey mudstone

- Clay to silt Laminated Thickness: a few cm -a few dm

Oxidized poorlydrained floodplainand in-channelmudstone

3.4 Mixed colormudstone

- Clay to silt Blockygreenish greyto purple

Thickness: a few dm Red-ox variability onfloodplain andin-channel mudstone

92

Table 3.2: Description, Interpretation, and Comparison of Facies Associations.

Facies Association Environment Channel % Sandstone% ofFloodplain

ChannelThickness

1 Large vertically andlaterally amalgamated

sandy channel-floodplainlithosomes

Proximal fluvial fan 85-100% >70% 10-35m

2 Laterally amalgamatedchannel-floodplain

lithosomes with sandy toheterolithic accretion sets

Medial fluvial fan 40-70% ˜50% 3-15m (57-70%decrease from

FA1)

3 Small isolatedchannel-floodplain

lithosomes

Distal fluvial fan 10-30% ˜30% 2-10 (30%decrease from

FA2)

93

3.13 Reference Cited

Abdullatif, O.M., 1989, Channel-fill and sheet-flood facies sequences in the ephemeral ter-minal River Gash, Kassala, Sudan: Sedimentary Geology, v. 63, p. 171–184.

Alexander, J., Bridge, J.S., Cheel, R.J., and Leclair, S.F., 2001, Bedforms and associatedsedimentary structures formed under supercritical water flows over aggrading sandbeds: Sedimentology, v. 48, p. 133–152.

Allen, J.P., Fielding, C.R., Gibling, M.R., and Rygel, M.C., 2011, Fluvial response to paleo-equatorial climate fluctuations during the late Paleozoic ice age: Bulletin of the Ge-ological Society of America, v. 123, p. 1524–1538.

Alonso-Zarza, A.M., 2003, Palaeoenvironmental significance of palustrine carbonates andcalcretes in the geological record: Earth-Science Reviews, v. 60, p. 261–298.

Assine, M.L., 2005, River avulsions on the Taquari megafan, Pantanal wetland, Brazil:Geomorphology, v. 70, p. 357–371.

Assine, M.L., Corradini, F.A., Pupim, F. do N., and McGlue, M.M., 2014, Channel arrange-ments and depositional styles in the Sao Lourenco fluvial megafan, Brazilian Pantanalwetland: Sedimentary Geology, v. 301, p. 172–184.

Assine, M.L., and Silva, A., 2009, Contrasting fluvial styles of the Paraguay River in thenorthwestern border of the Pantanal wetland, Brazil: Geomorphology, v. 113, p.189–199.

Bhattacharyya, P., Scuderi, L.A., Weissmann, G.S., Hartley, A.J., Davidson, S.K., andNichols, G.J., 2011, Satellite Imagery Evaluation of Soil Moisture Variability in theGanges Basin, India, in American Geophysical Union, Fall Meeting 2011, San Fran-cisco, California.

Billi, P., 2007, Morphology and sediment dynamics of ephemeral stream terminal distributarysystems in the Kobo Basin (northern Welo, Ethiopia): Geomorphology, v. 85, p.98–113.

Birkett, C.M., 1995, The contribution of TOPEX/POSEIDON to the global monitoring ofclimatically sensitive lakes: Journal of Geophysical Research, v. 100, p. 25179.

Bryant, M., Falk, P., and Paola, C., 1995, Experimental study of avulsion frequency andrate of deposition: Geology, v. 23, p. 365–368.

Cartigny, M., and Postma, G., 2010, Experiments on internal hydraulic jumps in stratifiedturbidity currents and their relation to structureless sands., in EGU General AssemblyConference Abstracts, Vienna, Austria, v. 12, p. 4952.

94

Cartigny, M.J.B., Ventra, D., Postma, G., and van Den Berg, J.H., 2014, Morphodynam-ics and sedimentary structures of bedforms under supercritical-flow conditions: Newinsights from flume experiments: Sedimentology, v. 61, p. 712–748.

Cashion, W., 1967, Geology and fuel resources of the Green River Formation, southeasternUinta basin, Utah and Colorado: U.S. Geological Survey Professional Paper 548, p.48.

Chakraborty, T., and Ghosh, P., 2010, The geomorphology and sedimentology of the Tistamegafan, Darjeeling Himalaya: Implications for megafan building processes: Geomor-phology, v. 115, p. 252–266.

Chakraborty, T., Kar, R., Ghosh, P., and Basu, S., 2010, Kosi megafan: Historical records,geomorphology and the recent avulsion of the Kosi River: Quaternary International,v. 227, p. 143–160.

Croke, J.C., Magee, J.M., and Price, D.M., 1998, Stratigraphy and sedimentology of thelower Neales River, West Lake Eyre, Central Australia: From Palaeocene to Holocene:Palaeogeography, Palaeoclimatology, Palaeoecology, v. 144, p. 331–350.

Davis, S.J., Dickinson, W.R., Gehrels, G.E., Spencer, J.E., Lawton, T.F., and Carroll, A.R.,2010, The Paleogene California River: Evidence of Mojave-Uinta paleodrainage fromU-Pb ages of detrital zircons: Geology, v. 38, p. 931–934.

DeCelles, P.G., 2004, Late Jurassic to Eocene evolution of the Cordilleran thrust belt andforeland basin system, western USA: American Journal of Science, v. 304, p. 105–168.

DeCelles, P.G., and Cavazza, W., 1999, A comparison of fluvial megafans in the Cordilleran(Upper Cretaceous) and modern Himalayan foreland basin systems: Bulletin of theGeological Society of America, v. 111, p. 1315–1334.

Dickinson, W.R., Klute, M.A., Hayes, M.J., Janecke, S.U., Lundin, E.R., Mckittrick, M.A.,and Olivares, M.D., 1988, Paleogeographic and paleotectonic setting of Laramide sed-imentary basins in the central Rocky Mountain region: Geological Society of AmericaBulletin, v. 100, p. 1023–1039.

Dickinson, W.R., Lawton, T.F., and Inman, K.F., 1986, Sandstone Detrital Modes, Cen-tral Utah Foreland Region: Stratigraphic Record of Cretaceous -Paleocene TectonicEvolution: Journal of Sedimentary Petrology, v. 56, p. 276–293.

Dickinson, W.R., Lawton, T.F., Pecha, M., Davis, S.J., Gehrels, G.E., and Young, R.A.,2012, Provenance of the Paleogene Colton Formation (Uinta Basin) And Cretaceous-Paleogene provenance evolution in the Utah foreland: Evidence from U-Pb ages ofdetrital zircons, paleocurrent trends, and sandstone petrofacies: Geosphere, v. 8, p.854–880.

95

Donselaar, M.E., Cuevas Gozalo, M.C., and Moyano, S., 2013, Avulsion processes at theterminus of low-gradient semi-arid fluvial systems: Lessons from the Rıo Colorado,Altiplano endorheic basin, Bolivia: Sedimentary Geology, v. 283, p. 1–14.

Fielding, C.R., 2006, Upper flow regime sheets, lenses and scour fills: Extending the rangeof architectural elements for fluvial sediment bodies: Sedimentary Geology, v. 190,p. 227–240.

Fielding, C.R., Allen, J.P., Alexander, J., and Gibling, M.G., 2009, Facies model for fluvialsystems in the seasonal tropics and subtropics: Geology, v. 37, p. 623–626.

Fisher, J. a., and Nichols, G.J., 2013, Interpreting the stratigraphic architecture of fluvialsystems in internally drained basins: Journal of the Geological Society, v. 170, p.57–65.

Ford, G.L., Dechesne, M., and Pyles, D.R., 2016, Stratigraphic architecture of a fluvial-lacustrine basin-fill succession at Desolation Canyon, Uinta Basin, Utah: Reference toWalthers’ Law and implications for the petroleum industry: The Mountain Geologist,v. 53.

Fouch, T., 1975, Lithofacies and related hydrocarbon accumulations in Tertiary strata of thewestern and central Uinta Basin, Utah: Rocky Mountain Association of Geologists,Symposium, p. 163–173.

Fouch, T., Cashion, W., Ryder, R., and Campbell, J., 1976, Field guide to lacustrine andrelated nonmarine depositional environments in Tertiary rocks, Uinta Basin, Utah.

Fouch, T.D., Hanley, J.H., Forester, R.M., Keighin, C.W., Pitman, J.K., and Nichols, D.J.,1987, Chart showing lithology, mineralogy, and paleontology of the nonmarine NorthHorn Formation and Flagstaff Member of the Green River Formation, Price Canyon,central Utah: a principal reference section: U.S. Geological Survey, MiscellaneousInvestigation Series.

Gall, R.D., Birgenheier, L.P., and Vanden Berg, M.D., 2017, Highly Seasonal and Peren-nial Fluvial Facies: Implications For Climatic Control On the Douglas Creek andParachute Creek Members, Green River Formation, Southeastern Uinta Basin, Utah,U.S.A.: Journal of Sedimentary Research, v. 87, p. 1019–1047.

Hajek, E.A., and Edmonds, D.A., 2014, Is river avulsion style controlled by floodplain mor-phodynamics? Geology, v. 42, p. 199–202.

Hajek, E.A., and Wolinsky, M.A., 2012, Simplified process modeling of river avulsion andalluvial architecture: Connecting models and field data: Sedimentary Geology, v.257–260, p. 1–30.

96

Hartley, A., Weissmann, G., Bhattacharayya, P., Nichols, G., and ..., 2013, Soil developmenton modern distributive fluvial systems: preliminary observations with implicationsfor interpretation of paleosols in the rock record: SEPM Special Publication No. 104,p. 149–158.

Horton, B.K., and Decelles, P.G., 2001, Modern and ancient fuvial megafans in the fore-land basin system of the central Andes , southern Bolivia: Implications for drainagenetwork evolution in fold- thrust belts: Basin Research, v. 13, p. 43–63.

Jacob, F., 1969, Delta facies of the Green River Formation (Eocene): Carbon and Duchesne:University of Colorado, 364 p.

Jerolmack, D.J., and Mohrig, D., 2007, Conditions for branching in depositional rives: Ge-ology, v. 35, p. 463–466.

Jones, E.R., 2017, Probabilistic Source-To-Sink Analysis of the Provenance of the Califor-nia Paleoriver: Implications for the Early Eocene Paleogeography of Western NorthAmerica: Colorado School of Mines, 198 p.

Jones, H.L., and Hajek, E.A., 2007, Characterizing avulsion stratigraphy in ancient alluvialdeposits: Sedimentary Geology, v. 202, p. 124–137.

Keighley, D., Flint, S., Howell, J., and Moscariello, A., 2003, Sequence stratigraphy inlacustrine basins: a model for part of the Green River Formation (Eocene), SouthwestUinta Basin, Utah, U.S.A.: Journal of Sedimentary Research, v. 73, p. 987–1006.

Kelly, S.B., and Olsen, H., 1993, Terminal fans-a review with reference to Devonian examples:Sedimentary Geology, v. 85, p. 339–374.

Kezer, K., and Matsuyama, H., 2006, Decrease of river runoff in the Lake Balkhash basin inCentral Asia: Hydrological Processes, v. 20, p. 1407–1423.

Kostic, S., Sequeiros, O., Spinewine, B., and Parker, G., 2010, Cyclic steps: A phenomenonof supercritical shallow flow from the high mountains to the bottom of the ocean:Journal of Hydro-Environment Research, v. 3, p. 167–172.

Kraus, M.J., and Aslan, A., 1996, Eocene Hydromorphic Paleosols: Significance for Inter-preting Ancient Floodplain Processes: Journal of Sedimentary Research, v. 63.

Kraus, M.J., and Wells, T.M., 1999, Recognizing Avulsion Deposits in the Ancient Strati-graphical Record, in Fluvial Sedimentology VI, Oxford, UK, Blackwell PublishingLtd., p. 251–268.

Latrubesse, E.M., Stevaux, J.C., Cremon, E.H., May, J.H., Tatumi, S.H., Hurtado, M.A.,Bezada, M., and Argollo, J.B., 2012, Late Quaternary megafans, fans and fluvio-aeolian interactions in the Bolivian Chaco, Tropical South America: Palaeogeography,

97

Palaeoclimatology, Palaeoecology, v. 356–357, p. 75–88.

Leckie, D., Fox, C., and Tarnocai, C., 1989, Multiple paleosols of the late Albian BoulderCreek Formation, British Columbia, Canada: Sedimentology, v. 36, p. 307–323.

Legarreta, L., and Uliana, M.A., 1998, Anatomy of Hinterland Depositional Sequences:Upper Cretaceous Fluvial Strata, Neuquen Basin, West-Central Argentina, in RelativeRole of Eustasy, Climate, and Tectonism in Continental Rocks, Special Publicationsof SEPM No. 59.

Leier, A.L., DeCelles, P.G., and Pelletier, J.D., 2005, Mountains, monsoons, and megafans:Geology, v. 33, p. 289–292.

Makaske, B., 2001, Anastomosing rivers: a review of their classification, origin and sedimen-tary products: Earth-Science Reviews, v. 53, p. 149–196.

Mckee, E.D., Crosby, E.J., and Berryhill, H.L., 1967, Flood Deposits, Bijou Creek, Colorado,June 1965: Journal of Sedimentary Petrology, v. 37, p. 829–851.

Mohrig, D., Heller, P., Paola, C., and Lyons, W., 2000, Interpreting avulsion processfrom ancient alluvial sequences: Guadalope-Matarranya system (northern Spain) andWasatch Formation (western Colorado): Geological Society of America Bulletin, v.112, p. 1787–1803.

Moore, J., Taylor, A., Johnson, C., Ritts, B.D., and Archer, R., 2012, Facies Analysis, Reser-voir Characterization, and LIDAR Modeling of an Eocene Lacustrine Delta, GreenRiver Formation, Southwest Uinta Basin, Utah, in Baganz, O.W., Bartov, Y., Bo-hacs, K.M., and Nummedal, D. eds., acustrine Sandstone Reservoirs and HydrocarbonSystems, American Association of Petroleum Geologists, Memoir 95, p. 183–208.

Morgan, C.D., 2003, Geologic guide and road logs of the willow creek, indian, soldier creek,nine mile, gate, and desolation canyons, uinta basin, utah (OFR-407).

Nichols, G.J., and Fisher, J.A., 2007, Processes, facies and architecture of fluvial distributarysystem deposits: Sedimentary Geology, v. 195, p. 75–90.

Nichols, G.J., and Hirst, J.P., 1998, Alluvial fans and fluvial distributary systems, Oligo-Miocene, northern Spain; contrasting processes and products: Journal of SedimentaryResearch, v. 68, p. 879–889.

North, C.P., and Warwick, G.L., 2007, Fluvial Fans: Myths, Misconceptions, and the Endof the Terminal-Fan Model: Journal of Sedimentary Research, v. 77, p. 693–701.

Owen, A., Nichols, G.J., Hartley, A.J., and Weissmann, G.S., 2017, Vertical trends within theprograding Salt Wash distributive fluvial system, SW United States: Basin Research,v. 29, p. 64–80.

98

Owen, A., Nichols, G.J., Hartley, A.J., Weissmann, G.S., and Scuderi, L.A., 2015, Quan-tification of a Distributive Fluvial System: The Salt Wash DFS of the MorrisonFormation, SW U.S.A.: Journal of Sedimentary Research, v. 85, p. 544–561.

Picard, M.D., and High, L.R., 1973, Sedimentary structures of ephemeral streams: Elsevier,v. 17.

Pitman, J., Fouch, T., Bulletin, M.G.-A., and 1982, U., 1982, Depositional setting anddiagenetic evolution of some Tertiary unconventional reservoir rocks, Uinta Basin,Utah: AAPG Bulletin, v. 66, p. 1581–1596.

Plink-Bjorklund, P., 2015, Morphodynamics of rivers strongly affected by monsoon precip-itation: Review of depositional style and forcing factors: Sedimentary Geology, p.110–147.

Plink-Bjorklund, P., Birgeneier, L., and Jones, E., 2014, Extremely bad early Eoceneweather: evidence for extreme precipitation from river deposits, in CBEP Climaticand Biotic Events of the Paleogene 2014, Ferrara, Italy, p. 175–176.

Plink-Bjorklund, P., and Birgenheier, L., 2013, Effects of Extreme Monsoon Precipitationon River Systems Form And Function, an Early Eocene Perspective, in AmericanGeophysical Union, Fall Meeting 2013, San Francisco, California.

Postma, G., Kleverlaan, K., and Cartigny, M.J.B., 2014, Recognition of cyclic steps insandy and gravelly turbidite sequences, and consequences for the Bouma facies model:Sedimentology, v. 61, p. 2268–2290.

Pupim, F. do N., Assine, M.L., and Sawakuchi, A.O., 2017, Late Quaternary Cuiabamegafan, Brazilian Pantanal: Channel patterns and paleoenvironmental changes:Quaternary International, v. 438, p. 108–125.

Pusca, V.A., 2003, Wet/Dry, Terminal fan-dominated sequence architecture: A new,outcrop-based model for the lower Green River Formation, Utah: University ofWyoming.

Remy, R.R., 1992, Stratigraphy of the Eocene Part of the Green River Formation in theSouth-Central Part of the Uinta Basin, Utah: US Geological Survey Bulletin 1787-BB, p. 79.

Retallack, G., 1988, Field recognition of paleosols (Juergen Reinhardt & Wayne R. Sigleo,Eds.): Reston, Virginia, Geological Society of America.

Rossetti, D.F., Zani, H., Cohen, M.C.L., and Cremon, E.H., 2012, A Late Pleistocene-Holocene wetland megafan in the Brazilian Amazonia: Sedimentary Geology, v. 282,p. 276–293.

99

Ryder, R.T., Fouch, T.D., and Elison, J.H., 1976, Early Tertiary sedimentation in the westernUinta Basin, Utah: Geological Society of America Bulletin, v. 87, p. 496.

Schomacker, E.R., Kjemperud, A.V., Nystuen, J.P., and Jahren, J.S., 2010, Recognition andsignificance of sharp-based mouth-bar deposits in the Eocene Green River Formation,Uinta Basin, Utah: Sedimentology, v. 57, p. 1069–1087.

Sendziak, K.L., 2012, Stratigraphic architecture and avulsion deposits in a low net-sand-content fluvial succession: Lower Wasatch Formation, Uinta Basin: Colorado Schoolof Mines, 53 p.

Shukla, U.K., Singh, I.B., Sharma, M., and Sharma, S., 2001, A model of alluvial megafansedimentation: Ganga Megafan: Sedimentary Geology, v. 144, p. 243–262.

Singh, H., Parkash, B., and Gohain, K., 1993, Facies analysis of the Kosi megafan deposits:Sedimentary Geology, v. 85, p. 87–113.

Sinha, R., 2009, The Great avulsion of Kosi on 18 August 2008: Current Science, v. 97, p.429–433.

Sinha, R., Ahmad, J., Gaurav, K., and Morin, G., 2014, Shallow subsurface stratigraphyand alluvial architecture of the Kosi and Gandak megafans in the Himalayan forelandbasin, India: Sedimentary Geology, v. 301, p. 133–149.

Sinha, R., Gaurav, K., Chandra, S., and Tandon, S.K., 2013, Exploring the channel connec-tivity structure of the August 2008 avulsion belt of the Kosi River, India: Applicationto fl ood risk assessment: Geology, v. 41, p. 1099–1102.

Sinha, R., Latrubesse, E.M., and Nanson, G.C., 2012, Quaternary fluvial systems of tropics:Major issues and status of research: Palaeogeography, Palaeoclimatology, Palaeoecol-ogy, v. 356–357, p. 1–15.

Sinha, R., and Sarkar, S., 2009, Climate-induced variability in the Late Pleistocene-Holocenefluvial and fluvio-deltaic successions in the Ganga plains, India: A synthesis: Geo-morphology, v. 113, p. 173–188.

Smith, M.E., Carroll, A.R., and Scott, J.J., 2015, Stratigraphic Expression of Climate, Tec-tonism, and Geomorphic Forcing in an Underfilled Lake Basin: Wilkins Peak Memberof the Green River Formation, in Smith M., C.A. ed., Stratigraphy and Paleolimnologyof the Green River Formation, Western USA. Syntheses in Limnogeology, Springer,Dordrecht, p. 61–102.

Sneh, A., 1983, Desert stream sequences in the Sinai Peninsula: Journal of SedimentaryResearch, v. 53, p. 1271–1279.

100

Stear, W.M., 1985, Comparison of the Bedform Distribution and Dynamics of Modern andAncient Sandy Ephemeral Flood Deposits in the Southwestern Karoo Region, SouthAfrica: Sedimentary Geology, v. 45, p. 209–230.

Stouthamer, E., and Berendsen, H.J. a, 2001, Avulsion Frequency, Avulsion Duration,and Interavulsion Period of Holocene Channel Belts in the Rhine-Meuse Delta, TheNetherlands: Journal of Sedimentary Research, v. 71, p. 589–598.

Taylor, A.W., and Ritts, B.D., 2004, Mesoscale heterogeneity of fluvial-lacustrine reservoiranalogues: Examples from the eocene green river and colton formations, Uinta Basin,Utah, USA: Journal of Petroleum Geology, v. 27, p. 3–26.

Trendell, A.M., Atchley, S.C., and Nordt, L.C., 2013, Facies Analysis of A Probable Large-Fluvial-Fan Depositional System: The Upper Triassic Chinle Formation At PetrifiedForest National Park, Arizona, U.S.A: Journal of Sedimentary Research, v. 83, p.873–895.

Uba, C.E., Heubeck, C., and Hulka, C., 2005, Facies analysis and basin architecture of theNeogene Subandean synorogenic wedge, southern Bolivia: Sedimentary Geology, v.180, p. 91–123.

Weissmann, G.S., Hartley, A.J., Nichols, G.J., Scuderi, L.A., Olson, M., Buehler, H., andBanteah, R., 2010, Fluvial form in modern continental sedimentary basins: Distribu-tive fluvial systems: Geology, v. 38, p. 39–42.

Weissmann, G.S., Hartley, A.J., Scuderi, L.A., Nichols, G.J., Davidson, S.K., Owen, A.,Atchley, S.C., Bhattacharyya, P., Chakraborty, T., Ghosh, P., Nordt, L.C., Michel,L., and Tabor, N.J., 2013, Prograding Distributive Fluvial Systems—GeomorphicModels and Ancient Exampless: SEPM Special Publication No. 104, p. 131–147.

Wilkinson, J.M., Marshall, L.G., and Lundberg, J.G., 2006, River behavior on megafans andpotential influences on diversification and distribution of aquatic organisms: Journalof South American Earth Sciences, v. 21, p. 151–172.

Wright, V.P., 1992, Paleosol Recognition: a Guide to Early Diagenesis in Terrestrial Settings,in Developments in Sedimentology, Elsevier, v. 47, p. 591–619.

101

CHAPTER 4

FLUVIAL-LAKE INTERACTIONS IN SUNNYSIDE DELTA INTERVAL, EOCENE

GREEN RIVER FORMATION, UINTA BASIN, UTAH

A paper to be submitted to Journal of Sedimentary Research

Jianqiao Wang*,1 and Piret Plink-Bjorklund1

4.1 Abstract

Eocene-aged fluvial and lake deposits in the Sunnyside Delta Interval in the Nine Mile

Canyon show lateral fluvial-lake interactions. River and floodplain deposits transition into

an interbedded succession of river and floodplain deposits, deltaic deposits, and lake de-

posits across the 20-km wide study area in Nine Mile canyon. However, this transition is not

gradual, but rather, river channel and floodplain facies laterally alternate with carbonate or

mixed carbonate-siliciclastic deposits across distances of only a few hundred meters to a few

kilometers. Furthermore, some mouth bar deposits consist of alternating carbonate grain-

stones and siliciclastic sandstones, and some abandoned channels are filled with dolomitic

mudstones. These numerous transitions between depositional environments indicate a highly

irregular shoreline, where fluvial and deltaic deposits build out locally at the active chan-

nel locations, and laminated dolomitic mudstones accumulate in protected embayments or

abandoned channels, and lime grainstones where lake’s wave and current energy is high.

We interpret these fluvial-lake interactions in Sunnyside Delta interval as a result of river

avulsions and contemporaneous carbonate productions.

∗Primary and corresponding author. Direct correspondence to [email protected] Stratigraphy Group, Department of Geology and Geological Engineering, Colorado School ofMines, 1500 Illinois Street, Golden, Colorado 80401, USA.

102

4.2 Introduction

Vertical transitions between fluvial and lacustrine deposits are commonly interpreted as

transgressive-regressive cycles linked to lake level changes (Astin, 1990; Abdul Aziz et al.,

2003; Keighley et al., 2003; Olsen, 2009; Zhao et al., 2015), where fluvial deposits are linked

to lowstands and lake carbonates to highstands in a reciprocal sedimentation relationship

(Thrana and Talbot, 2006; Ford et al., 2016; Rahiminejad et al., 2018). Such interpretation

is driven by the sequence stratigraphy concept and considers the shut-off or dilution of

carbonate production when or where siliciclastic input is high (Handford and Loucks, 1993).

Although this assumption has been demonstrated in outcrop studies including the well-

known Capitan margin of the Permian Delaware Basin (Harris and Saller, 1999; Kerans and

Tinker, 1999), it does not apply in many modern coastal and shallow marine environments

where siliciclastic and carbonate sediments are coevally mixed, resulting in lateral transitions

between carbonate and fluvial sediments (Chiarella et al., 2017). The carbonate factory in a

closed lake is even less sensitive to siliciclastic input, as chemical precipitation occurs when

the lake water is supersaturated (James and Jones, 2016).

Outcrops of the Eocene Green River Formation Sunnyside Delta Interval in the Uinta

Basin (Utah) display lateral fluvial and lake (carbonate) transitions across only a few hun-

dreds of meters. For example, thinly laminated dolomitic mudstone-filled channels change

laterally into fluvial channel sandstones, and/or interbedded siltstone and sandstones with

paleosols. There are also examples of mouth-bar sandstones interbedded with carbonate

grainstones and mudstones, and sandy carbonate grainstones that laterally pass into mouth

bar sandstones. All of these lateral transitions caution the interpretation that vertical lake

and fluvial transitions necessarily involve lake level changes as they are shown to have Walthe-

rian relationships.

While growing amount of research has been conducted in modern and ancient systems

(Halfar et al., 2004; McNeill et al., 2004; Quesne et al., 2009; Tanavsuu-Milkeviciene et

al., 2009; Zeller et al., 2015), since Mount (1984) first discussed the mixing processes in

103

modern shallow marine environment, not much data are available on the compositional and

stratal mixing processes (Chiarella et al., 2017) in ramp settings of lacustrine basins (Song

et al., 2017). This study aims to investigate the fluvial-lacustrine interaction and the mixing

process by documenting the fluvial-lacustrine facies transition styles and lateral changes in

mixed siliciclastic-carbonate facies. We discuss the fluvial avulsion processes as a significant

control on the mixing processes.

4.3 Geological Setting

Uinta Basin formed in response to the Laramide uplift of the Uinta Mountains (Cashion,

1967; Dickinson et al., 1988; DeCelles, 2004). During the interfingering fluvial-lacustrine

deposition of Green River Formation in early to middle Eocene between ca 53 to 46 Ma,

Lake Uinta has been recognized as an underfilled lake (Picard, 1955; Fouch, 1975; Ryder

et al., 1976; Castle, 1990; Fouch et al., 1994; Smith et al., 2003, 2008; Davis et al., 2010).

It is featured a ramp style southern lake margin (Carroll and Bohacs, 1999) with shallow

water depth for long distances. Mixed siliciclastic and carbonate deposits of Green River

Formation developed as Lake Uinta experienced from fresh to transitional, rising to high-

lake, and closing lake stages, defined by organic richness (Cashion and Donnell, 1972, 1974;

Johnson et al., 2010; Tanavsuu-Milkeviciene and Sarg, 2012; Tanavsuu-Milkeviciene and

Sarg, 2017). Recent work on the siliciclastic part of Green River Formation in the Uinta

Basin has mostly focused on the effects of variable discharge on rivers (Plink-Bjorklund and

Birgenheier, 2013; Plink-Bjorklund et al., 2014; Plink-Bjorklund, 2015; Rosenberg et al.,

2015; Gall et al., 2017), as well as on deltaic deposits (Jacob, 1969; Schomacker et al., 2010;

Moore et al., 2012). On the carbonate part, lacustrine profundal and shoal deposits have

been documented (Pitman et al., 1982; Smith et al., 2008; Burton et al., 2014). Mixed

littoral to sublittoral carbonates and siliciclastics were recently documented in the eastern

basin but the main focus was on vertical trend and climate (Gall et al., 2017) as well as lake

level control (Tanavsuu-Milkeviciene and Sarg, 2017).

104

4.4 Stratigraphic Framework

The Eocene Green River Formation is subdivided into three informal units: lower, middle

and upper, based on regionally extensive lacustrine intervals including the Uteland Butte

Member, the Carbonate Marker Unit, and the Mahogany Oil Shale Member (Fig. 4.1)

(Fouch, 1975; Pitman et al., 1982; Fouch et al., 1987). For the Nine Mile canyon area, which

is located in the southwestern Uinta Basin, the three major siliciclastic intervals are the

Colton Tongue, the Renegade Tongue, and the Sunnyside Delta Interval. D and C Marker

beds mark the bottom and top of the Sunnyside Delta interval and can be followed across

the study area in the Nine Mile Canyon (Remy, 1992; Morgan, 2003). D marker is located ca

160 m above the Carbonate Marker Unit (Picard and High, 1973; Ryder et al., 1976; Remy,

1992). C marker is equivalent to the C1 marker of Jacob (1969) and correlated to ‘R4’ zone

of the eastern Uinta Basin, thus, Sunnyside Delta Interval is equivalent to Douglas Creek

Member in the eastern Uinta Basin (Roberts, 1964; Tanavsuu-Milkeviciene and Sarg, 2017).

The age of the Sunnyside Delta interval deposition is constrained by an absolute date range

from the C marker of 49.6 Ma (Smith et al., 2015) and 54 Ma from the Carbonate Marker

Unit (Remy, 1992) (Fig. 4.1).

4.5 Dataset and Methods

Field data from Sunnyside Delta interval (Fig. 4.1) was collected from the west-to-east

oriented Nine Mile Canyon and several north-to-south oriented side canyons (Fig. 4.1).

Observations start from mile marker 32 to 45 in Nine Mile Canyon, and continue to mile

marker 49 to the south in Cottonwood Canyon. A total of about 550 m of measured section

data were collected at 10 cm resolution. Acid testing was used to determine dolomite vs.

limestone in the field. The Dunham classification modified by Embry and Klovan (1971)

is applied in this study to include coarse grained carbonates. Thin sections were made

in the Thin Section Laboratory (Department of Geology and Geological Engineering at

Colorado School of Mine) to further observe mixed siliciclastic and carbonate facies. Lateral

105

relationships of sedimentary facies were examined on cliff-face exposures and mapped by

walking out the stratigraphic intervals in continuous outcrops. Photomosaics were used to

assist with analyzing channel architecture. Architectural dimensions (e.g. channel width

and thickness) were obtained by using laser range finder (TRUPULSETM 360B) in field.

4.6 Fluvial and Lacustrine Facies Associations

4.6.1 Fluvial Facies Association

4.6.1.1 Facies Association 1.1: Sandstone to Heterolithic Lenticular Lithosomes

with Accretion Sets

Facies association 1 includes lenticular and laterally amalgamated sandstone to het-

erolithic erosionally based lithosomes, 10-35 m thick (Figs. 4.2A-C). Thicker lithosomes con-

tain downstream and upstream facing accretion sets 10-15 m thick (Figs. 4.2A-C), whereas

thinner lithosomes display vertical aggradation, a few to 10 m thick (Fig. 4.2D). Some litho-

somes are sandy throughout, whereas others contain up to dm-thick mud layers in sand-mud

couplets (Fig. 4.2D). This facies association consists primarily of very-fine to fine grained,

moderately to poorly sorted sandstone with up to 10% of mudstone and up to 2% con-

glomerate. The latter occurs along the erosional base of channel lithosomes. Sedimentary

structures are dominated by (over 70%) low-angle convex- and concave up sedimentary struc-

tures (F3, 5, 6, and 7; Figs. 4.2E-H; Table 4.1). Cross strata (F10-13) are the minor with

less than 26% and commonly occur in climbing sets (Figs. 4.2I and J). Sandstone with scour

and fill structures (F3; 30%), distinct planar lamina (F7; 20%), gradational planar lamina

(F6; 14%), climbing ripple lamina (F13; 13%), ripple lamina (F12; 10%), and convex-up

low angle lamina (F5; 8%) are the dominant facies (Fig. 4.2; Table 4.1). Most lithosomes

do not display obvious vertical grain size trends, but some upstream accretion sets coarsen

upward in places because mudstone layers in accretion sets thicken downward. Basal and

internal erosion surfaces are concave up and flat or undulating, and in places aligned with

106

mudstone clasts. Accretion sets are commonly erosionally bound (Figs. 4.2A-D). Wood

fragments are associated with erosion surfaces in places. In places bioturbation occurs at

accretion set boundaries, including cm to dm long roots and vertical burrows (Figs. 4.2K

and L). Desiccation cracks occur in places at the mud-to-sand contact within the accretion

sets (Fig. 4.2M). This facies association is commonly associated with interbedded mudstones

and sandstones of FA1.2 (Figs. 4.2A-D), but in places directly overlies micritic or carbonate

cemented siltstones of FA3.2.

Interpretation: The erosionally bound lenticular accretion sets are interpreted as channels

with a variable degree of lateral and vertical amalgamation. Sandstone facies indicate dom-

inant Froude supercritical flow, as seen by comparison of the sedimentary structures with

experimentally produced supercritical flow deposits (e.g. Alexander et al., 2001; Cartigny

et al., 2014). In channel bioturbation and desiccation indicates that channels were season-

ally or inter-annually dry, such as is common in arid to sub-humid climates (Sneh, 1983;

Stear, 1985). Abundance or dominance of Froude supercritical flow sedimentary structures

is characteristic for variable discharge rivers, where deposition dominantly occurs during

high magnitude floods (Fielding, 2006; Fielding et al., 2009; Plink-Bjorklund, 2015). In

channel mud layers suggest a rapid deceleration of floods with mud deposition during wan-

ing stages (e.g., Jones, 1977; Tunbridge, 1981). The erosionally bound sandy accretion sets

and the sandstone-mudstone couplets are interpreted as flood event beds (Tunbridge, 1981;

Sneh, 1983; Turner, 1986; Abdullatif, 1989; Deluca and Eriksson, 1989; Nichols and Hirst,

1998; Shukla et al., 2001; Hinds et al., 2004; Hampton and Horton, 2007; Chakraborty and

Ghosh, 2010; Chakraborty et al., 2010; Mader and Redfern, 2011; Donselaar et al., 2013;

Plink-Bjorklund, 2015). Despite the seasonal or inter-annual desiccation, the water table

was relatively high as the vertical burrows are less than 1 dm long (Hasiotis, 2004).

107

4.6.1.2 Facies Association 1.2: Interbedded Mudstones and Sandstones with

Variegated Color

Facies association 1.2 consists of interbedded mudstones and sandstones and is purple

(F15), greenish grey (F16), brown (F17), or shows color transitions from purple to greenish

grey (F18) on cm to dm scale. In places irregular brownish or grey patches occur. Mud-

stones are dm-thick (F15-18) and interbedded with ripple laminated (F12) or structureless

(F8) very-fine grained sandstones, cm to a m thick. In places mudstones coarsen upward into

sandstones (Fig. 4.2N). The interbedded mudstones and sandstones commonly form upward

sandying and thickening packages, erosionally overlain and laterally truncated by lenticular

or amalgamated accretion sets of FA 1.1 (Figs. 4.2A-D). The lateral extent of interbedded

sandstones and mudstones varies from only 10s of meters where severely erosionally trun-

cated to 100s of m or kms. Bioturbation and soft sediment deformation are common in the

purple mudstone (F15; Fig. 4.2O). Possible mammal tracks occur in a 3 dm thick heavily

bioturbated purple mudstone layer (Fig. 4.2O). Calcite filled cracks (Fig. 4.2P) and calcite

nodules are common in all mudstone facies (F15-18).

Interpretation: The mottled color, burrowing and tracks, and carbonate nodules are

characteristic for paleosols and subaerial exposure (e.g. Retallack, 1988; Leckie et al., 1989;

Wright, 1992). Some calcite filled cracks are possible calcified root structures. Mottling is a

result of local redox changes due to fluctuating drainage conditions and is also reflected in

the purple to brown colors (Wright, 1992). The paleosols and the association with channels

of FA 1.1 indicate that the interbedded mudstones and sandstones are floodplain deposits.

The sandstone interlayers are interpreted as crevasse splays, and the upward coarsening and

thickening trends with capping channel sandstones as avulsion packages (Mohrig et al., 2000;

Jones and Hajek, 2007; Sendziak, 2012). The upward coarsening and sandstone thickening

trends overlain by channels indicate transitional avulsions (Jones and Hajek, 2007). The

splay rich floodplain indicates frequent avulsions (Shukla et al., 2001; Hajek and Wolinsky,

2012).

108

4.6.2 Lacustrine Facies Associations

4.6.2.1 Facies Association 2.1: Lime Grainstone

Ooids and ostracods are the dominant grains (F23) in the lime grainstone and they are

commonly well sorted (Fig. 4.3A). Fragments of bivalve (Fig. 4.3C), gastropod (Fig. 4.3C),

and fish scales (Fig. 4.3D) are present only locally in the lower part of the 10-50 cm thick

grainstone layers (Fig. 4.3H) and form a floatstone or rudstone (F24). Grainstone layers

are laterally consistent and commonly continuous for 10s to 100s of m (Fig. 4.3I), but

vary in allochem content and layer thickness. In places ripple laminations with round and

symmetrical crests occur (Figs. 4.3F and G). In places, cm long desiccation cracks occur at

the top of the beds (Fig. 4.3G). In outcrop, this facies association is white-tan to yellowish

orange in color and forms erosion-resistant layers.

Interpretation: Strong waves or currents roll grains in shallow littoral zone near the lake

shoreline to form ooids (Ryder et al., 1976). Chemical carbonate precipitation allows the

accumulation of coated grains, such as ooids, oncolites and pisolites (Cole and Picard, 1978).

Fragmented bivalve, gastropod, and fish scales indicate that they are redistributed by storm

waves or currents (Milroy and Wright, 2002). The abundance of ooids and ostracods and

lack of carbonate mud suggest an agitated environment where wave energy is high (Platt

and Wright, 1991). The well sorted grains inherently prohibit the formation of sedimentary

structures and explain the poorly developed ripple laminations.

4.6.2.2 Facies Association 2.2: Stromatolites

Stromatolite boundstone (F27) has hemispherical heads, cm to m tall (Figs. 4.4C and D),

with internal mm thick convex-up lamination (Fig. 4.4E). Stromatolites (F27) are commonly

yellowish in color and change laterally into oolite and pisolite/oncolite grainstone (F23; Fig.

4.3B), and stromatolite rudstone (F25). This transition exclusively occurs in the D marker

(Fig. 4.1C) that is mapped across the Nine Mile Canyon. Inverse grading from ooids to

109

pisolites and oncolites is common in dm thick grainstones (F23; Figs. 4.3E and H).

Interpretation: The internally laminated stromatolites are interpreted to form in the lower

littoral to sublittoral zone (Platt and Wright, 1991; Bridge and Demicco, 2008). The yellow-

ish color has been shown to indicate dolomite content (Symcox, 2015). The lateral transition

between stromatolite (F26) and stromatolite rudstone (F25) and oolite and pisolite/oncolite

grainstone (F23) is indicative of high hydrodynamic energy including storm events in littoral

and sublittoral zone of lake (Krylov, 1982; Cohen and Thouin, 1987; Suriamin, 2010; Sarg

et al., 2013). The inverse grading from ooids to pisolites and oncolites indicates a shoaling

upward succession as wave energy increases shoreward (Platt and Wright, 1991).

4.6.2.3 Facies Association 2.3: Gastropod Wackestone

Gastropods-bearing fine grained structureless micrite (F22) occur as up to 2 dm thick,

laterally continuous, sharply bound, dark resistant beds (Figs. 4.4A and B). These wacke-

stones are commonly vertically associated with oolitic or ostracodal grainstone (F23) and

oolitic or ostracodal sandstone (F26) (Fig. 4.4A).

Interpretation: The occurrence of mudstone matrix in gastropod wackestone suggests

deposition in a lower littoral to sublittoral zone or lagoon where the environment is calm

(Renaut and Gierlowski-Kordesch, 2010).

4.6.3 Deltaic Facies Association

4.6.3.1 Facies Association 3.1: Sharp-based Tabular Sandstone

The sharp-based tabular sandstone is 5 to 10 m thick, and 100s of meter wide (Figs. 4.5A

and B). It consists of fine to upper-fine grained moderately sorted sandstone dominated

by gradational planar lamina (F6; 33%), scour and fill structures (F3; 31%), cross strata

(F10; 12%), soft sediment deformation (F9; 9%), convex-up low angle lamina (F5; 8%), and

climbing ripple cross lamina (F13; 7%). Convex- and concave-up laminated sandstones (F3,

5, and 6) form 72% of the succession and climbing and aggradational bedforms occur (F3, 5,

110

6, 9, and 13) in 88% of the deposits. The tabular sandstones do not display grain size trends

(Fig. 4.5), but wood and carbonate fragments occur in places. Some sandstones display

low-angle downstream accretion sets. The characteristic sharp but flat basal surface changes

in places laterally into a concave-up erosion surface that bounds a lenticular sandstone (Figs.

4.5D and E). Multiple yellow layers of mm thick siliciclastic mudstone form dewatering flame

structures. Vertically this facies association overlies interbedded mudstone and sandstone of

FA3.2 (Figs. 4.5A and B) or lacustrine carbonate or mixed facies associations (FA2.1-2.4;

FA4.1-4.3) (Fig. 4.5D). This facies association transitions laterally into fluvial (FA1.1-1.2),

and lacustrine and mixed facies associations (FA2.1-2.4; FA4.1-4.3).

Interpretation: The tabular geometry together with downstream accretion, high depo-

sition rate sedimentary structures, and the association with lacustrine deposits suggests

deposition in mouth bars (Scruton, 1960; Plink-Bjorklund and Steel, 2005; Olariu and Bhat-

tacharya, 2006). The lateral transition of the sharp based tabular sandstone to lenticular

shaped sandstone with concave-up erosion surface suggests mouth bar transition into its

adjacent distributary channel (Olariu and Bhattacharya, 2006). Sedimentary structures in-

dicate high deposition rates and abundant Froude supercritical flow.

4.6.3.2 Facies Association 3.2: Interbedded Siliciclastic and Carbonate Mud-

stone and Sandstone

This facies association consists of cm to 1 m thick very-fine grained sandstones (F12)

and greenish grey or dark grey, and in places limy mudstones (F19) (Figs. 4.5A-C). Individ-

ual beds commonly fine upward from ripple laminated sandstone (F12) to thinly laminated

mudstone (F19) (Fig. 4.5C). Dewatering structures and Scoyenia trace fossil are preserved

at mudstone to sandstone boundaries. The interbedded sandstones and mudstones form

upward coarsening packages where sandstone thickness in successive beds increases upward

(Fig. 4.5C) or the mudstones occur as a dm to m thick layer sharply overlain by FA 3.1,

and in places interbedded with carbonate facies associations (FA2.1 and 2.3) (Figs. 4.5A,

111

B). Vertically, interbedded mudstone and sandstone is overlain by sharp-based tabular sand-

stone (FA3.1), or amalgamated to lenticular lithosomes of FA1.1, or variegated interbedded

mudstone and sandstone of FA1.2 (Fig. 4.5A).

Interpretation: The Scoyenia traces, the gray color, and the lack of pedogenic modi-

fication indicate a subaqueous lake environment (Hasiotis, 2004). The upward coarsening

deposits of interbedded mudstone and sandstone beds are interpreted as delta front deposits.

The fining upward trend of individual beds is characteristic for turbidites that are likely of

hyperpycnal flow origin (Plink-Bjorklund and Steel, 2004; Bhattacharya and MacEachern,

2009) where mudstones are truncated.

4.6.4 Fluvial-lacustrine Mixed Facies Associations

4.6.4.1 Facies Association 4.1: Mixed Siliciclastic-Carbonate Accretion Sets

This facies association appears orange brown in outcrop and occurs as sharp-based ac-

cretion sets, dm to m thick and 10s of meter wide (Figs. 4.7 and 4.8). It is composed of

oolitic sandstones (F26), oolitic and ostracod grainstones (F23), siliciclastic sandstones (F3,

6, 9, and 10), and limy mudstones (F19) (Figs. 4.7A and B). Oolitic sandstones consist

of poorly sorted lower fine to lower-medium grained siliciclastic sands (30-40%) and similar

sized carbonate grains (ooids and/or ostracods) (F26; Fig. 4.8). Sedimentary structures

in oolitic sandstones (F26; Fig. 4.8) include cross stratification (30%), gradational planar

lamina (30%), scour and fill structures (20%), symmetrical ripples with sharp crests (10%),

unidirectional asymmetric ripple laminations (8%), and soft sediment deformation (2%).

Carbonate grains are either evenly distributed through the tabular mixed unit (Fig. 4.7B)

or align with lamina (Fig. 4.8). Siliciclastic sandstones commonly display scour and fill struc-

tures (F3), gradational planar lamina (F6), structureless (F9), and soft sediment deformation

(F10). Dewatering and loading structures are characteristic at the boundary with siliciclastic

sandstones (Figs. 4.7E and 4.8D). The limy mudstone layer (F19) between the siliciclastic

sandstone (F3, 6, 9, and 10) and oolitic sandstone is thinly laminated (F26) (Fig. 4.7). Cross

112

strata foresets strata dip NNE similar to the accretion sets. A typical vertical transition (Fig.

4.7B) starts with a couple of dm thick carbonate grainstone (FA2.1) or stromatolite (FA2.2)

interbedded with calcareous mudstone (FA3.2), then top truncated by sharp-based tabular

sandstone (FA3.1) or mixed siliciclastic-carbonate accretion sets (FA4.1).

Interpretation: The mixed accretion sets are interpreted as a mixed carbonate-siliciclastic

mouth bar complex, based on the occurrence as basinward-dipping accretion sets, the simi-

larity of siliciclastic sandstones to mouth bar sandstones on FA 3.1, and the similar vertical

transitions of this FA to FA 3.1 (Fig. 4.7A). The poor sorting suggests the source of ooids

comes from a contemporaneous carbonate production nearshore and in-situ mixing with sand

(Mount, 1984; McNeill et al., 2004). Fluvial-derived siliciclastic sands may have been mixed

with carbonate grainstones due to lateral mixing, as the allochems were transported along

the lake shoreline by waves or currents. The dewatering and loading features are indicative of

high deposition rate (Lowe and LoPiccolo, 1996). The presence of symmetrical wave ripples

suggests wave reworking (Leckie, 2003).

4.6.4.2 Facies Association 4.2: Sigmoidal Sandy Grainstone Macroforms

Sigmoidal sandy grainstone macroforms consists of light brown moderately sorted upper

fine to lower medium grained siliciclastic sands and slightly larger sized carbonate grains

(ooids and/or ostracods; >85%). The deposit coarsens upward with larger size carbonate

grains on top and more siliciclastic sand at the bottom. The macroform is 2-3m thick and

exhibits tabular shape with flat and sharp lower and upper boundaries (Fig. 4.9). Lateral

extent of the macroform is ca 400 m, and it consists of sigmoidally or erosionally bound

cross strata 1.5-2 m thick with topset, foreset, and bottomset geometries in flow parallel

view (Fig. 4.9). Foresets dip obliquely perpendicular (SW230◦) as compared to fluvial

sediment transport direction at an angle of 15-25◦. In flow perpendicular view, the bedding

looks planar. The macroform is capped by a 20 cm thick grainstone layer (FA2.1), and it lies

on top of a meter thick greenish grey laminated dolomitic carbonate mudstone bed (FA2.3).

113

Interpretation: The sigmoidal sandy grainstone macroform is interpreted as formed by

punctuated mixing of large amount of carbonate grains with detrital sand (Mount, 1984).

The migration direction suggests shore-parallel transport and storm-generated currents,

which may occur during periodic high energy storms (Mount, 1984; Halfar et al., 2004).

4.6.4.3 Facies Association 4.3: Mixed Siliciclastic-carbonate Lenticular Mud-

stone

The lenticular mudstones occur as lenticular erosionally based deposits (Fig. 4.10) that

consists of limy mudstone (F19), argillaceous dolomitic mudstone (F21), and in places also

interbeds of oolitic grainstone (F23), oolitic sandstone (F26), and limy mudstones (F19)

(Fig. 4.10C). The argellaceous dolomitic mudstone has cm-scale laminations and contains

detrital clay as well as dolomite (F21). It is light/dark grey altered to dark brown (Fig.

4.11). The limy and dolomitic mudstones (F19 and 21) fill dm-scale scours or multi-m scale

channels (Fig. 4.10), and display internal scour surfaces. In some places wood fragments

occur at these scour surfaces. This facies association is laterally and vertically associated

with mouth bars of FA 3.1 and oolitic sandstone accretion sets (FA4.1).

Interpretation: The occurrence of dolomitic mudstones, interbedded mudstones, siliciclas-

tic sandstones, and carbonate grainstones in channels, the lateral association with mouth

bars and distributary channel deposits, and the presence of wood fragments suggest deposi-

tion in abandoned channels. The alternation of color and the laminations are associated with

variation of organic matter content, which is possibly associated with algal blooms (Renaut

and Gierlowski-Kordesch, 2010). Such dolomitic mudstones and organic rich marl deposits

(Platt and Wright, 1991) tend to occur in nearshore protected lagoons or embayments (Teller

et al., 2000), and could thus also accumulate in channels abandoned by avulsions, as such

channels would form areas protected from waves. Internal erosion surfaces, woody fragments,

occasional siliciclastic sandstone and argillaceous layers suggest that the abandoned chan-

nels transferred discharge and minor amount of sediment at times. The grainstone interbeds

114

indicate transient current or wave action.

4.6.4.4 Facies Association 4.4: Laminated Dolomitic Mudstone

Dolomitic mudstone forms thinly laminated beds, mm to cm thick (Fig. 4.11). It is

internally black but weathers into bluish grey color (F20; Figs. 4.11A and B). In places,

there are sand-filled burrows on top of the beds. The laminated dolomitic mudstone layer

extends laterally for kilometers.

Interpretation: The black mm thick dolomitic mudstone is interpreted as organic rich

marl deposits (Platt and Wright, 1991). The presence of bioturbation and the km scale

extent suggests deposition in a nearshore protected lagoon or embayment (Teller et al.,

2000).

4.7 Lateral and Vertical Trends

In general, the ca 200 m thick Sunnyside Delta Interval across the 20 km of Nine Mile

Canyon exposures displays a distinct change from southeast to northwest, as the south-

easternmost outcrops only contain fluvial deposits (FA1.1. and 1.2), and the proportion of

lacustrine facies increases gradually towards northwest (Fig. 4.12). First thin lime grain-

stone (FA2.1) and dolomitic mudstone (FA4.4) intervals alternate with the thick packages

of fluvial deposits (Figs. 4.12B and C), but eventually the lacustrine deposits and deltaic

deposits form a significant part of the stratigraphy (Fig. 4.11D).

This transition from river to lake deposits is not a simple gradual change, but rather, this

transition is complex, as lateral transitions between fluvial (FA1), lacustrine (FA2), deltaic

(FA3), and mixed fluvial-lacustrine (FA4) deposits occur in 100s of m to km scale (Figs.

4.12-4.14). For example, at mile marker 45 ca 35 m thick amalgamated sandy erosionally

based fluvial channel deposit (FA1.1) overlying a dm thick gastropod wackestone (FA2.3)

transitions westward into a mouth bar complex with dolomitic mudstone filled scours at

the base (FA4.3), overlain by a mixed siliciclastic-carbonate mouth bar (FA4.1), floodplain

115

deposits (FA1.2), a lens of limey mudstone (FA3.2), and then a siliciclastic mouth bar (FA3.1)

across 300 m (Fig. 4.13). The siliciclastic mouth bar is associated with underlying interbedd

mudstone and sandstone (FA3.2) and mudstone with variegated color (FA1.2). Further

northwest, this whole complex pinches out into a thick floodplain succession across 200 m

(Fig. 4.13). In another example at mile marker 40.8, a mixed siliciclastic-carbonate mouth

bar (FA4.1) that overlies oolitic or ostrocodal grainstone (FA 2.1), transitions southeastward

into floodplain deposits (FA 1.2), where it is truncated by a river channel (FA1.1) towards

northwest across 430 m (Fig. 4.13).

4.8 Discussion on Fluvial-lake Interaction

These numerous transitions between depositional environments indicate a highly irregular

shoreline, where fluvial and deltaic deposits build out locally at the active channel locations,

laminated dolomitic mudstones accumulate in protected embayments or abandoned channels,

and lime grainstones, where lake wave and current energy is high. In a modern example, sim-

ilar lateral transitions occur where Ile River fan enters Lake Balkhash (Fig. 4.14). Ili River

is highly avulsive as it builds a large fluvial fan. Where the active river channel is, multiple

deltaic lobes develop and prograde into the lake (Fig. 4.14A). At paleochannel locations

(Figs. 4.14B and C), abandoned channels and different types of lagoons occur, as well as

wave or current built bars or shoals. Thus the observed lateral variability in the Sunnyside

Delta interval is interpreted to be a function of river avulsions and the contemporaneous

carbonate grain and mud production. This mixing process along an irregular lake shoreline

is also responsible for the mixed siliciclastic-carbonate mouth bars, sandy grainstone bars,

and dolomitic mudstone filled channels.

This lateral variability further has a large impact on the nature of transgressions and

regressions, as river and deltaic successions can locally prograde as seen by the lakeward

perturbations, where the active channels are (Fig. 4.14D), whereas coeval transgression may

occur laterally in areas without fluvial sediment input.

116

4.9 Conclusions

The Eocene Sunnyside Delta Interval transitions from river deposits to interbedded river,

deltaic, and lake deposits across 20 km in the Nine Mile Canyon. This transition is not

gradual but rather highly variable as river and floodplain deposits transition laterally into

lake carbonates or deltaic deposits across a few hundred meters or a few kilometers. Some

mouth bar deposits are mixed siliciclastic-carbonate and indicate coeval carbonate grain

production and fluvial sediment input. Some abandoned channels are filled with dolomitic

mudstones. River and delta lobe avulsions and the contemporaneous carbonate production

are interpreted to control carbonate-siliciclastic mixing along a highly irregular shoreline.

117

Figure 4.1 Stratigraphic column and base map of Uinta Basin and Nine Mile canyon. (A)Geological map of the Wasatch/Colton (maroon) and Green River (orange) Formations inthe Uinta Basin, with study area locations at Nine Mile Canyon, Road 191 and Hay Canyon(modified from Dickinson et al., 2012; Schomacker et al., 2010; Sato and Chan, 2015; Jones,2017). (B) Study area at Nine Mile Canyon (see location in A) with figure locations andmile markers (grey numbers). Rose diagrams show paleocurrent directions measured in thisstudy. (C) stratigraphic table with available age constraints (modified from Fouch et al.,1987; Remy, 1992; Smith et al., 2010, 2015).

118

Figure 4.2 Example outcrop photos of fluvial facies association 1. (A-D) Vertical transitionsbetween channel (FA1.1) and floodplain (FA1.2) facies associations. (E) Sandstone withscour and fill structure (F3). (F) Sandstone with convex-up low angle lamina (F5). (G)Sandstone with gradational planar lamina and scour and fill structures (F6 and F3). (H)Sandstone with distinct planar lamina (F7). (I) Sandstone with ripple cross lamina (F12).(J) Sandstone with climbing ripple cross lamina (F13). (K) Roots in sandstone. (L) Verticalburrows in ripple laminated sandstone. (M) Desiccation cracks at the bottom of sandstone.(N) Upward coarsening floodplain mudstone (F15-18) to sandstone (F12) (FA1.2). (O)Mammal tracks preserved at the base of a sandstone overlying a purple mudstone. (P)Possible calcite filled cracks in purple mudstone (F15).

119

120

121

Figure 4.3 Outcrop photos of lime grainstone (FA2.1). (A) Well sorted oolitic grainstone(F23). (B) Pisoid and oncoid grainstone (F23). (C) Gastropods and bivalve shells andfragments in non-stromatolite bioclastic floatstone (F24). (D) Fish scale fragments in non-stromatolite bioclastic floatstone (F24). (E) Inverse grading from ooids to pisolites andoncolites is common in dm thick grainstone beds (F23). (F-G) Sharp crested symmetricalripple lamina. Note the cm long desiccation cracks at the upper boundary of rippled ooliticgrainstone (F23) in G. (H) dm thick oolitic grainstone (F23) overlain by mixed siliciclastic-carbonate mouth bars (FA4.1).

122

Figure 4.4 Outcrop photos of stromatolite (FA2.2) and gastropod wackestone (FA2.3) faciesassociations. (A) A 20 cm thick gastropod bearing wackestone bed (F22) overlying ostracodalgrainstone (F23). (B) A gastopod in FA2.3 or F22. (C) D marker with laterally linkedstromatolites. (D) A few dm tall stromatolite layer in D marker. (E) Internal laminations instromatolites in D marker. (F) Fragments of mm scale stromatolites in stromatolite rudstone(F25).

123

Figure 4.5 Outcrop photos of the deltaic facies association 3. (A, B) Characteristic verticaltransition from delta front (FA3.2) interbedded sandstones and mudstones to mouth barand distributary complex (FA3.1) amalgamated sandstones. (C) Upward fining succession(marked by white triangle) from sandstone (F12) to mudstone beds (F19) in delta front(FA3.2) . (D-F) Photos and measured sections of sharp based mouth bar and distributarycomplex (FA3.1). Photos in D and E are 30 m apart.

124

125

126

Figure 4.6 Vertical trends of facies associations. Two styles are presented in A and B with a typical measured section and anoutcrop photo. (A) Fluvial channel and floodplain (FA1.1 and 1.2) deposits are on top of deltaic (FA3.1 and 3.2) and lacustrine(FA2.2) deposits. The measured section is obtained from the outcrop in the figure. (B) Floodplain deposits (FA1.2) verticallytransition into lacustrine (FA2.1) and deltaic (FA3.2) deposits, which are overlain by mixed siliciclastic accretion sets (FA4.1).The measured section is not from the outcrop in the figure.

127

Figure 4.7 Outcrop photos of the mixed siliciclastic-carbonate accretion sets (FA4.1). (A) Photo and line drawing of accretionset architecture. (B) Evenly distributed carbonate grains. (C) Greenish grey mudstone (F19) below cross stratified ooliticsandstone (F26). (D) Cross stratified grainstone (F26). (E) Loading structure at the base of a sandstone. (F) Cross stratifiedand soft sediment deformed oolitic sandstone (F26).

128

Figure 4.8 Outcrop photos of sedimentary structures in oolitic sandstone (F26) in the mixed siliciclastic-carbonate accretionsets (FA4.1). (A) Symmetrical ripples with sharp crests. (B) Gradational planar lamination. (C) Scour and fill structures. (D)Soft sediment deformation.

129

Figure 4.9 Outcrop photos of sigmoidal sandy grainstone macroform (FA4.2). The outcrop is continuous in west-to-eastorientation. (A) is located to the east of (B). Note the composing bedforms and red lines represent their boundaries. Anoverview of the outcrop is provided in (C). This deposits laterally transition into mouth bar deposits across ca 200m tonorthwest.

130

Figure 4.10 Outcrop photos of mixed siliciclastic-carbonate lenticular mudstone (FA4.3). (A,B) Erosionally based lenticular mudstones with internal erosion surfaces. (C, D) Erosionallybased lenticular interbedded mudstones, silisiclastic sandstones and carbonate grainstones.

131

Figure 4.11 Outcrop photos of laminated dolomitic mudstone (FA4.4). (A, B) The characteristic bluish gray color of dolomiticmudstone (F20). (B) Sand-filled horizonal burrows on top of a bed. (C) Dark brown color altered with light grey color inargillaceous dolomitic mudstone (F21).

132

Figure 4.12 Depositional trends across the Nine Mile Canyon. (A) Channel and floodplain deposits. (B, C) Channel andfloodplain deposits with thin lake interbeds. (D) Interbedded and laterally variable lake and fluvial deposits.

133

Figure 4.13 Outcrop examples showing lateral and vertical transitions. (A) Lateral and vertical trends at mile marker 44.5to 45 outcrop oriented in west (left) to east (right). (B) Lateral trends at mile marker 40.8 outcrop oriented in NW330◦ tothe left. A mixed siliciclastic-carbonate mouth bar (FA4.1) that overlies oolitic or ostrocodal grainstone (FA 2.1), transitionssoutheastward into floodplain deposits (FA 1.2) and is truncated by a river channel (FA1.1) towards northwest across 430 m.

134

Figure 4.14 Examples of lake-fluvial interaction in Ile River, Lake Balkhash. (A) Activelyprograding delta lobes with an interlobe lagoon. (B) Multiple styles of lagoons, abandonedchannels, and wave or current reworking of sediment. (C) Abandoned with standing water.Note how river and floodplain deposits are laterally associated with the lake. (D) Deltapertrubations along the Lake Balkash shoreline.

135

Table 4.1: Description and Interpretation of Sedimentary Facies

Facies Textures Structures Width/Thickness Interpretation1 Stratified and

imbricatedconglomerates

Conglomerates (coarsesand to pebble sizedcarbonate clast/bioclastand/or mud clast,subrounded to angular,poorly sorted, varied inshape) in sandy matrix,clast or matrix supported

Imbricated,planar stratified,low anglestratified, scourand fillstructures,

Thickness: afew cm perconglomeratelayer; Totalthickness: a fewdm - m

Intrabasinal clasts offloodplain mudstones orlake carbonates depositedwith sand matrix bytraction flow (subcritical tosupercritical)

2 Disorganizedconglomerates

Conglomerates (coarsesand to pebble sizedcarbonate clast/bioclastand/or mud clast, roundedto angular, poorly sorted,varied in shape) in sandyor muddy matrix, in placeswith terrestrial organicmatter (woody fragments)

Disorganized Thickness: afew dm - 1m

Intrabasinal clasts offloodplain mudstones orlake carbonate depositedby collapse and fluidizationof channel banks duringfalling stage of flow

3 Sandstone withscour and fillstructures

Upper very fine to lowermedium grained sandstone,moderately to poorlysorted, in places mixedwith carbonate grains

Scours filledwith upwardflatteninglamina

Thickness: afew dm - a fewm; Width: a fewdm - 10s of m

Supercritical flow, highdeposition rates,suspension deposition,large antidune or chute andpool formation

4 Sandstone withripple filledscour and fillstructures

Upper very fine to lowerfine-grained sandstone,moderately sorted

Ripples fill thetoe part ofupwardflattening scourand fill lamina

Thickness: Afew dm - 1m;Width: A fewdm - a few m

Supercritical to subcriticalflow, high deposition rates,suspension and bedloaddeposition

136

Table 4.1: Continued.

Facies Textures Structures Width/Thickness Interpretation5 Sandstone with

convex-up lowangle lamina

Very fine to fine grainedsandstone, moderately topoorly sorted, in placesmixed with carbonategrains

Low angleconvex-uplaminations,very low anglehummocky-likecross strata withupwardincreasingsteepness

Thickness: Afew dm - 1m;Width: >1m

Supercritical flow, highdeposition rates,suspension deposition,antidune formation

6 Sandstone withgradationalplanar lamina

Very fine to fine grainedsandstone, moderately topoorly sorted

Diffused,gradationalplanarlaminations,each lamina isinternallyreversely tonormally graded

Thickness(lamina): 2mm -2cm; Thickness(bed): a few dm- a few m

Subcritical to supercriticalflow, high deposition rates,suspension deposition,plane bed formation

7 Sandstone withdistinct planarlamina

Very fine to lower mediumgrained sandstone,moderately sorted

Planarlaminationswith distinctlaminaboundaries

Thickness(lamina): mmscale; Thickness(bed): a few dm- a few m

Supercritical ortranscritical flow, normaldeposition rates, bedloaddeposition, plane bedformation

8 Structurelesssandstone

fine to lower mediumgrained sandstone,moderately to poorlysorted

No apparentsedimentarystructures

Thickness: a fewcm - a few dm

high deposition rates,suspension deposition

137

Table 4.1: Continued.

Facies Textures Structures Width/Thickness Interpretation9 Soft sediment

deformedsandstone

Very fine to lower mediumgrained sandstone,moderately to poorlysorted

Flamestructures,overturnedfolds, dishstructures, balland pillowstructures.

Thickness: afew dm

Water escape or localcollapse

10 Cross stratifiedsandstone

Fine to lower mediumgrained sandstone,moderately sorted

Planar andtrough crossstratified

Thickness (crossset): 5cm-20cm;Thickness(co-set): a fewdm

Subcritical flow, normaldeposition rates, bedloaddeposition, dune migration

11 Sandstone withclimbing crossstrata

Fine to lower mediumgrained sandstone,moderately to poorlysorted

Planar andtrough crossstrata withclimbing setboundaries

Thickness (crossset): 8cm - dmscale; Thickness(co-set): a fewdm

Subcritical flow, highdeposition rates,suspension and bedloaddeposition, dune migrationand aggradation

12 Cross laminatedsandstone

Very fine to lower finegrained sandstone,moderately sorted

Asymmetricalcross-lamination

Thickness (crossset): cm scale <5cm; Thickness(co-set): a fewdm

Subcritical flow, normaldeposition rates, bedloaddeposition, ripple migration

13 Sandstone withclimbing ripplelamina

Very fine to fine grainedsandstone, moderatelysorted

Asymmetricalcross laminationwith climbingset boundaries

Thickness (crossset): cm scale<5cm;Thickness(co-set): a fewdm - a few m

Subcritical flow, highdeposition rates,suspension and bedloaddeposition, ripplemigration and aggradation

138

Table 4.1: Continued.

Facies Textures Structures Width/Thickness Interpretation14 Bioturbated

sandstoneVery fine to lower mediumgrained sandstone,moderately to poorlysorted

Vertical,horizontal and3D trace fossils

Thickness: a fewdm - a few m

Trace fossils formed byinsects, dwelling, resting,crawling traces

15 Purpled redmudstone

Clay to silt Laminated orblocky orcrumbly, inplaces rooted,bioturbated, ordeformed

Thickness: afew cm - a fewdm (blocky); afew mm (eachlamina)

Oxidized, well drainedfloodplain and in-channelmudstone

16 Greenish greymudstone

Clay to silt Laminated orblocky orcrumbly

Thickness: afew cm - a fewdm (blocky); afew mm (eachlamina)

Reduced to poorly drainedfloodplain and in-channelmudstone

17 Brown to darkgrey mudstone

Clay to silt Laminated Oxidized poorly drainedfloodplain and in-channelmudstone

18 Varigated colormudstone

Clay to silt, Blocky greenishgrey to purple

Thickness: afew dm

Red-ox variability onfloodplain

139

Table 4.1: Continued.

Facies Textures Structures Width/Thickness Interpretation19 Limy mudstone Detrital clay to silt, lime

mudStructureless,some faintlamina,commonlyinterbeddedwith ripplelaminatedsandstone(F12), occur intabular,laterallyextensive beds.

Thickness: afew dm - 1m

siliciclastic derived,deposited in lake

20 Oraganic richdolomiticmudstone

Lime mud, organic rich Structureless,thinly laminated

Thickness: afew mm

Nearshore protected lagoonor embayment, low energy

21 Argillaceousdolomiticmudstone

Dolomitic mudstone withdetrital clay

Planarlaminated

Thickness(lamina): a fewcm; Thickness(bed): a few dm

sublittoral

22 Gastropodwackestone

Lime mud, gastropods Structureless Thickness: afew cm - 2dm

Lower littoral to sublittoral

23 Oolitic orostracodalgrainstone

Ooids and ostracodsdominated, pisoids andoncolites also present in Dmarker

Symmetricalripple lamina,structureless,planar lamina,inverse gradingin places,

Thickness:10-50 cm;Width: laterallycontinuous for10s - 100s of m

Shallow littoral, highenergy

140

Table 4.1: Continued.

Facies Textures Structures Width/Thickness Interpretation24 Non-

stromatolitebioclasticfloatstone orrudstone

Ca 30% fragmentedbivalves, gastropods, fishscales, muddy intraclasts,size of the clasts are largerthan 2mm

Structureless Thickness: afew cm - 1dm

Lower littoral tosublittoral, high energy

25 Stromatoliterudstone

Stromatolite fragments,muddy intraclasts,elongate,

structureless Thicknes: dm Restricted lake

26 Oolitic orostracodalsandstone

Poorly sorted, ooids orostracods, upper fine tolower medium grainedsiliciclastic sand

Symmetricalripple lamina,asymmetricalripple lamina,gradationalplanar lamina,scour and fillstructures, softsedimentdeformation

Total thickness:up to 1.5m;Thickness perlayer: ca 50cm

Littoral, high engergy

27 Stromatolite Hemispheroidalmicrobialites

Internallaminated

Thickness: cm -m; Thinkness(lamina): mm

Littoral, biogenic

141

Table 4.2: Description and Interpretation of Sedimentary Facies Association

Environment Facies Association Interpretation

Fluvial1.1 Amalgamated and isolated

channel lithosomesFluvial channel

1.2 Purple, greenish grey,brownish mudstone

Floodplain

Lacustrine2.1 Lime grainstone Littoral, high energy2.2 Stromatolite Lower littoral to sublittoral

(restricted lake)2.3 Gastropod wackestone Sublittoral, lagoon, low

energy

Delta3.1 Amalgamated Distributary

channel-mouth bar complex3.2 Thinly interbedded

mudstone and sandstoneDelta front

Mixed fluvial-lacustrine

4.1 Mixedsiliciclastic-carbonate

accretion sets

Carbonate shoal mixedwith siliciclastic grains;

wave reworked mouth bars4.2 Sigmoidal sandy grainstone

macroformStormed induced carbonate

shoal4.3 Mixed

siliciclastic-carbonatelenticular mudstone

Siliciclastic derivedmudstone deposited incalm lake environment

4.4 Laminated dolomiticmudstone

Nearshore lagoon

4.10 Reference Cited

Abdul Aziz, H., Sanz-Rubio, E., Calvo, J.P., Hilgen, F.J., and Krijgsman, W., 2003,Palaeoenvironmental reconstruction of a middle Miocene alluvial fan to cyclic shallowlacustrine depositional system in the Calatayud Basin (NE Spain): Sedimentology, v.50, p. 211–236.

Abdullatif, O.M., 1989, Channel-fill and sheet-flood facies sequences in the ephemeral ter-minal River Gash, Kassala, Sudan: Sedimentary Geology, v. 63, p. 171–184.

Alexander, J., Bridge, J.S., Cheel, R.J., and Leclair, S.F., 2001, Bedforms and associatedsedimentary structures formed under supercritical water flows over aggrading sandbeds: Sedimentology, v. 48, p. 133–152.

142

Astin, T.R., 1990, The Devonian lacustrine sediments of Orkney, Scotland; implications forclimate cyclicity, basin structure and maturation history: Journal of the GeologicalSociety, v. 147, p. 141–151.

Bhattacharya, J.P., and MacEachern, J.A., 2009, Hyperpycnal Rivers and Prodeltaic Shelvesin the Cretaceous Seaway of North America: Journal of Sedimentary Research, v. 79,p. 184–209.

Billi, P., 2007, Morphology and sediment dynamics of ephemeral stream terminal distributarysystems in the Kobo Basin (northern Welo, Ethiopia): Geomorphology, v. 85, p.98–113.

Bridge, J., and Demicco, R., 2008, Earth Surface Processes, Landforms and Sediment De-posits: New York, Cambridge University Press.

Burton, D., Woolf, K., and Sullivan, B., 2014, Lacustrine depositional environments in theGreen River Formation, Uinta Basin: Expression in outcrop and wireline logs: AAPGBulletin, v. 98, p. 1699–1715.

Carroll, A.R., and Bohacs, K.M., 1999, Stratigraphic classification of ancient lakes: Balanc-ing tectonic and climatic controls: Geology, v. 27, p. 99–102.

Cashion, W., 1967, Geology and fuel resources of the Green River Formation, southeasternUinta basin, Utah and Colorado: U.S. Geological Survey Professional Paper 548, p.48.

Cashion, W., and Donnell, J.R., 1972, Chart showing correlation of selected key units in theorganic-rich sequence of the Green River Formation, Piceance Creek Basin, Colorado,and Uinta Basin, Utah.

Cashion, W.B., and Donnell, J.R., 1974, Revision of nomenclature of the upper part of theGreen River Formation, Piceance Creek Basin, Colorado, and Eastern Uinta Basin,Utah.

Chakraborty, T., and Ghosh, P., 2010, The geomorphology and sedimentology of the Tistamegafan, Darjeeling Himalaya: Implications for megafan building processes: Geomor-phology, v. 115, p. 252–266.

Chiarella, D., Longhitano, S.G., and Tropeano, M., 2017, Types of mixing and hetero-geneities in siliciclastic-carbonate sediments: Marine and Petroleum Geology, v. 88,p. 617–627.

Cohen, A.S., and Thouin, C., 1987, Nearshore carbonate deposits in Lake Tanganyika: Ge-ology, v. 15, p. 414.

143

Cole, R.D., and Picard, M.D., 1978, Comparative mineralogy of nearshore and offshore la-custrine lithofacies, Parachute Creek Member of the Green River Formation, PiceanceCreek Basin, Colorado, and eastern Uinta Basin, Utah: Geological Society of AmericaBulletin, v. 89, p. 1441.

DeCelles, P.G., 2004, Late Jurassic to Eocene evolution of the Cordilleran thrust belt andforeland basin system, western USA: American Journal of Science, v. 304, p. 105–168.

Dickinson, W.R., Klute, M.A., Hayes, M.J., Janecke, S.U., Lundin, E.R., Mckittrick, M.A.,and Olivares, M.D., 1988, Paleogeographic and paleotectonic setting of Laramide sed-imentary basins in the central Rocky Mountain region: Geological Society of AmericaBulletin, v. 100, p. 1023–1039.

Embry, A.F., and Klovan, J.E., 1971, A Late Devonian Reef Tract on Northeastern BanksIsland, N.W.T.: Bulletin of Canadian Petroleum Geology, v. 19, p. 730–781.

Fielding, C.R., 2006, Upper flow regime sheets, lenses and scour fills: Extending the rangeof architectural elements for fluvial sediment bodies: Sedimentary Geology, v. 190,p. 227–240.

Fielding, C.R., Allen, J.P., Alexander, J., and Gibling, M.G., 2009, Facies model for fluvialsystems in the seasonal tropics and subtropics: Geology, v. 37, p. 623–626.

Ford, G.L., Dechesne, M., and Pyles, D.R., 2016, Stratigraphic architecture of a fluvial-lacustrine basin-fill succession at Desolation Canyon, Uinta Basin, Utah: Reference toWalthers’ Law and implications for the petroleum industry: The Mountain Geologist,v. 53.

Fouch, T., 1975, Lithofacies and related hydrocarbon accumulations in Tertiary strata of thewestern and central Uinta Basin, Utah: Rocky Mountain Association of Geologists,Symposium, p. 163–173.

Fouch, T.D., Hanley, J.H., Forester, R.M., Keighin, C.W., Pitman, J.K., and Nichols, D.J.,1987, Chart showing lithology, mineralogy, and paleontology of the nonmarine NorthHorn Formation and Flagstaff Member of the Green River Formation, Price Canyon,central Utah: a principal reference section.

Frederick Sarg, J., Suriamin, J., Tanavsuu-Milkevidene, S.K., and Humphrey, J.D., 2013,Lithofacies, stable isotopic composition, and stratigraphic evolution of microbial andassociated carbonates, Green River Formation (Eocene), Piceance Basin, Colorado:AAPG Bulletin, v. 97, p. 1937–1966.

Gall, R.D., Birgenheier, L.P., and Vanden Berg, M.D., 2017, Highly Seasonal and Peren-nial Fluvial Facies: Implications For Climatic Control On the Douglas Creek andParachute Creek Members, Green River Formation, Southeastern Uinta Basin, Utah,

144

U.S.A.: Journal of Sedimentary Research, v. 87, p. 1019–1047.

Hajek, E.A., and Wolinsky, M.A., 2012, Simplified process modeling of river avulsion andalluvial architecture: Connecting models and field data: Sedimentary Geology, v.257–260, p. 1–30.

Halfar, J., Ingle, J.C., and Godinez-Orta, L., 2004, Modern non-tropical mixed carbonate-siliciclastic sediments and environments of the southwestern Gulf of California, Mex-ico: Sedimentary Geology, v. 165, p. 93–115.

Handford, C.R., and Loucks, R.G., 1993, Carbonate Depositional Sequences and SystemsTracts–Responses of Carbonate Platforms to Relative Sea-Level Changes: Chapter 1,in Loucks, R.G. and Sarg, J.F. eds., M 57: Carbonate Sequence Stratigraphy: RecentDevelopments and Applications, AAPG Special Volumes 57, v. 168, p. 3–41.

Harris, P., and Saller, A., 1999, Subsurface expression of the Capitan depositional systemand implications for hydrocarbon reservoirs, northeastern Delaware Basin: SEPMSpecial Publication No. 65, p. 37–49.

Hasiotis, S.T., 2004, Reconnaissance of Upper Jurassic Morrison Formation ichnofossils,Rocky Mountain Region, USA: Paleoenvironmental, stratigraphic, and paleoclimaticsignificance of terrestrial and freshwater ichnocoenoses: Sedimentary Geology, v. 167,p. 177–268.

Jacob, F., 1969, Delta facies of the Green River Formation (Eocene): Carbon and Duchesne:University of Colorado, 364 p.

James, N.P., and Jones, B. (Geology professor), 2016, Origin of carbonate sedimentary rocks:New Jersey, John Wiley & Sons Ltd.

Johnson, R.C., MERCIER, T.J., BROWNFIELD, M.E., and SELF, J.G., 2010, Assessmentof in- place oil shale resources in the Eocene Green River Formation, Uinta Basin,Utah and Colorado.

Jones, H.L., and Hajek, E.A., 2007, Characterizing avulsion stratigraphy in ancient alluvialdeposits: Sedimentary Geology, v. 202, p. 124–137.

Keighley, D., Flint, S., Howell, J., and Moscariello, A., 2003, Sequence stratigraphy inlacustrine basins: a model for part of the Green River Formation (Eocene), SouthwestUinta Basin, Utah, U.S.A.: Journal of Sedimentary Research, v. 73, p. 987–1006.

Kerans, C., and Tinker, S., 1999, Extrinsic stratigraphic controls on development of theCapitan reef complex: SEPM sepcial publication No. 65, p. 15–36.

Krylov, I.N., 1982, Lacustrine stromatolites from the Kyzylgir formation (Pliocene), Altaimountains, USSR: Sedimentary Geology, v. 32, p. 27–38.

145

Leckie, D.A., 2003, Modern environments of the Canterbury Plains and adjacent offshoreareas, New Zealand - An analog for ancient conglomeratic depositional systems innonmarine and coastal zone settings: Bulletin of Canadian Petroleum Geology, v. 51,p. 389–425.

Leckie, D., Fox, C., and Tarnocai, C., 1989, Multiple paleosols of the late Albian BoulderCreek Formation, British Columbia, Canada: Sedimentology, v. 36, p. 307–323.

Lowe, D.R., and LoPiccolo, R.D., 1996, The Characteristics and Origins of Dish and PillarStructures: Journal of Sedimentary Research, v. 44.

McNeill, D.F., Cunningham, K.J., Guertin, L.A., and Anselmetti, F.S., 2004, DepositionalThemes of Mixed Carbonate-siliciclastics in the South Florida Neogene: Applicationto Ancient Deposits: Integration of outcrop and modern analogs in reservoir modeling:AAPG Memoir 80, p. 23–43.

Milroy, P.G., and Wright, V.P., 2002, Fabrics, facies control and diagenesis of lacustrineooids and associated grains from the Upper Triassic, southwest England: GeologicalJournal, v. 37, p. 35–53.

Mohrig, D., Heller, P., Paola, C., and Lyons, W., 2000, Interpreting avulsion processfrom ancient alluvial sequences: Guadalope-Matarranya system (northern Spain) andWasatch Formation (western Colorado): Geological Society of America Bulletin, v.112, p. 1787–1803.

Morgan, C.D., 2003, Geologic guide and road logs of the willow creek, indian, soldier creek,nine mile, gate, and desolation canyons, uinta basin, utah (OFR-407).

Mount, J.F., 1984, Mixing of siliciclastic and carbonate sediments in shallow shelf environ-ments: Geology, v. 12, p. 432–435.

Olariu, C., and Bhattacharya, J.P., 2006, Terminal Distributary Channels and Delta FrontArchitecture of River-Dominated Delta Systems: Journal of Sedimentary Research,v. 76, p. 212–233.

Olsen, H., 2009, Orbital Forcing on Continental Depositional System-Lacustrine and FluvialCyclicity in the Devonian of East Greenland, in Orbital Forcing and Cyclic Sequences,Oxford, UK, Blackwell Publishing Ltd., p. 429–438.

Picard, M.D., and High, L.R., 1973, Sedimentary structures of ephemeral streams: Elsevier,v. 17.

Pitman, J., Fouch, T., Bulletin, M.G.-A., and 1982, U., 1982, Depositional setting anddiagenetic evolution of some Tertiary unconventional reservoir rocks, Uinta Basin,Utah: AAPG Bulletin, v. 66, p. 1581–1596.

146

Platt, N.H., and Wright, V.P., 1991, Lacustrine carbonates: facies models, facies distri-butions and hydrocarbon aspects: Sepcial Publication International Association ofSedimentologist, v. 13, p. 57–74.

Plink-Bjorklund, P., 2015, Morphodynamics of rivers strongly affected by monsoon precip-itation: Review of depositional style and forcing factors: Sedimentary Geology, p.110–147.

Plink-Bjorklund, P., and Steel, R., 2005, Deltas On Falling-Stage and Lowstand Shelf Mar-gins, The Eocene Central Basin of Spitsbergen: Importance of Sediment Supply:Special Publication of SEPM, p. 179–206.

Plink-Bjorklund, P., and Steel, R.J., 2004, Initiation of turbidity currents: outcrop evidencefor Eocene hyperpycnal flow turbidites: Sedimentary Geology, v. 165, p. 29–52.

Postma, G., and Cartigny, M.J.B., 2014, Supercritical and subcritical turbidity currents andtheir deposits— A synthesis: Geology, v. 42, p. 987–990.

Quesne, D., Buta-Neto, A., Benard, D., and Guiraud, M., 2009, Distribution of Albianclastic deposits in the Benguela basin (Angola): evidence of a Benguela palaeocurrent?Bulletin de la Societe Geologique de France, v. 180, p. 117–129.

Rahiminejad, A.H., Yazdi, M., Ashouri, A.R., and Kangazian, A., 2018, Siliciclastic inputand sedimentation rate as controls on paleoecological distribution of Lower Miocenebenthic carbonate producers in a fan–delta: Zagros foreland basin, Iran: HistoricalBiology, v. 30, p. 608–626.

Remy, R.R., 1992, Stratigraphy of the Eocene Part of the Green River Formation in theSouth-Central Part of the Uinta Basin, Utah: US Geological Survey Bulletin 1787-BB, p. 79.

Renaut, R.W., and Gierlowski-Kordesch, E.H., 2010, Lakes, in James, N.P. and Dalrymple,R.W. eds., Facies Models, Renaut, Geol. Assoc. Can. IV Series, GEOtext, 6, p.541–575.

Retallack, G., 1988, Field recognition of paleosols (Juergen Reinhardt & Wayne R. Sigleo,Eds.): Reston, Virginia, Geological Society of America.

Roberts, P.K., 1964, Stratigraphy of the Green River Formation, Uinta Basin, Utah: UtahUniversity, 212 p.

Ryder, R.T., Fouch, T.D., and Elison, J.H., 1976, Early Tertiary sedimentation in the westernUinta Basin, Utah: Geological Society of America Bulletin, v. 87, p. 496.

Scruton, P.C., 1960, Delta Building and the Deltaic Sequence, in Shepard, F.P., Phleger,F.B., and Van Andel, T.H. eds., Recent Sediments Northwest Gulf of Mexico, Tulsa,

147

Oklahoma, American Association of Petroleum Geolgosits, v. 143, p. 82–102.

Sendziak, K.L., 2012, Stratigraphic architecture and avulsion deposits in a low net-sand-content fluvial succession: Lower Wasatch Formation, Uinta Basin: Colorado Schoolof Mines, 53 p.

Shukla, U.K., Singh, I.B., Sharma, M., and Sharma, S., 2001, A model of alluvial megafansedimentation: Ganga Megafan: Sedimentary Geology, v. 144, p. 243–262.

Singh, A., and Bhardwaj, B.D., 1991, Fluvial facies model of the Ganga River sediments,India: Sedimentary Geology, v. 72, p. 135–146.

Singh, H., Parkash, B., and Gohain, K., 1993, Facies analysis of the Kosi megafan deposits:Sedimentary Geology, v. 85, p. 87–113.

Smith, M.E., Carroll, A.R., and Scott, J.J., 2015, Stratigraphic Expression of Climate, Tec-tonism, and Geomorphic Forcing in an Underfilled Lake Basin: Wilkins Peak Memberof the Green River Formation, in Smith M., C.A. ed., Stratigraphy and Paleolimnologyof the Green River Formation, Western USA. Syntheses in Limnogeology, Springer,Dordrecht, p. 61–102.

Smith, M.E., Carroll, A.R., and Singer, B.S., 2008, Synoptic reconstruction of a majorancient lake system: Eocene Green River Formation, western United States: Bulletinof the Geological Society of America, v. 120, p. 54–84.

Sneh, A., 1983, Desert stream sequences in the Sinai Peninsula: Journal of SedimentaryResearch, v. 53, p. 1271–1279.

Song, L., Guo, Y., Zhang, K., Sinica, L.Z.-A.G., and 2017, U., 2017, Mixed Siliciclastic-Carbonate Sediments in China: Recent Advances and Prospects: Acta Geologic Sinica(English Edition), v. 91, p. 283–284.

Stear, W.M., 1985, Comparison of the Bedform Distribution and Dynamics of Modern andAncient Sandy Ephemeral Flood Deposits in the Southwestern Karoo Region, SouthAfrica: Sedimentary Geology, v. 45, p. 209–230.

Suriamin, H., 2010, Facies, Diagenesis, and Geochemistry of the Eocene Green River For-mation: Colorado School of Mines, 214 p.

Symcox, C., 2015, Facies, stratigraphy, and mineralogy of the carbonate marker and Dmarker units, Lower Green River Formation, Uinta Basin, Utah: Colorado School ofMines, 178 p.

Tanavsuu-Milkeviciene, K., Plink-Bjorklund, P., K, K., and L., A., 2009, Coeval versus recip-rocal mixed carbonate-siliciclastic deposition, Middle Devonian Baltic Basin, EasternEurope: implications from the regional tectonic development: Sedimentology, v. 56,

148

p. 1250–1274.

Tanavsuu-Milkeviciene, K., and Sarg, J.F., 2017, Depositional Cycles and Sequences In AnOrganic-Rich Lake Basin: Eocene Green River Formation, Lake Uinta, Colorado andUtah, U.S.A.: Journal of Sedimentary Research, v. 87, p. 210–229.

Tanavsuu-Milkeviciene, K., and Sarg, F.J., 2012, Evolution of an organic-rich lake basin -stratigraphy, climate and tectonics: Piceance creek basin, eocene green river forma-tion: Sedimentology, v. 59, p. 1735–1768.

Teller, J.T., Risberg, J., Matile, G., and Zoltai, S., 2000, Postglacial history and paleoecologyof Wampum, Manitoba, a former lagoon in the Lake Agassiz basin: Geological Societyof America Bulletin, v. 112, p. 943–958.

Thrana, C., and Talbot, M.R., 2006, High-frequency carbonate-siliciclastic cycles in theMiocene ofthe Lorca Basin (Western Mediterranean, SE Spain): Geologica Acta, v.4, p. 343–354.

Williams, G.E., 1971, Flood Deposits of the Sand-bed Ephemeral Streams of Central Aus-tralia: Sedimentology, v. 17, p. 1–40.

Wright, V.P., 1992, Paleosol Recognition: a Guide to Early Diagenesis in Terrestrial Settings,in Developments in Sedimentology, Elsevier, v. 47, p. 591–619.

Zeller, M., Verwer, K., Eberli, G.P., Massaferro, J.L., Schwarz, E., and Spalletti, L., 2015,Depositional controls on mixed carbonate-siliciclastic cycles and sequences on gentlyinclined shelf profiles (J. Reijmer, Ed.): Sedimentology, v. 62, p. 2009–2037.

Zhao, J., Mountney, N.P., Liu, C., Qu, H., and Lin, J., 2015, Outcrop architecture of afluvio-lacustrine succession: Upper Triassic Yanchang Formation, Ordos Basin, China:Marine and Petroleum Geology, v. 68, p. 394–413.

149

CHAPTER 5

CONCLUSIONS

This dissertation uses high resolution outcrop data and documents channel architecture

in a fluvial fan system, proximal to distal changes within the fan, and the fan transition to

the lake. This field-based study provides systematic descriptions of bar-scale macroforms,

lateral and vertical facies transitions, to challenge the current understanding of fluvial facies

models. The macroforms presented in Chapter 2 (1st paper) raise the awareness that the

well-developed macroforms in variable discharge rivers are different from the ones in the

classical fluvial models. In Chapter 3 (2nd paper), the detailed lateral and vertical transi-

tions documented in a fluvial fan system significantly add to the simple proximal-distal and

vertical trends published so far, and link channel lithesome types and floodplain sandyness

to their position in the fan. This work challenges previous interpretations of the Sunnyside

Delta interval of the middle Green River Formation and links this succession to variable

discharge rivers and their ability to build fluvials fans. The results and the comparison to

modern fans advance our understanding the large fluvial fan systems. Chapter 4 (3rd paper)

documents the effects of avulsions and contemporaneous carbonate production, such as a

highly irregular lake shoreline due to lateral transitions between fluvial progradation and

lacustrine deposition. The findings caution the transgressive-regressive interpretations ap-

plied to vertical transitions between fluvial and lake deposits, and propose that such lateral

variability has a large impact on the nature of transgression and regression.

150