<<

UNLV Retrospective Theses & Dissertations

1-1-2002

Fault segmentation, fault linkage, and hazards along the Sevier fault, southwestern

Ilsa M Schiefelbein University of Nevada, Las Vegas

Follow this and additional works at: https://digitalscholarship.unlv.edu/rtds

Repository Citation Schiefelbein, Ilsa M, "Fault segmentation, fault linkage, and hazards along the Sevier fault, southwestern Utah" (2002). UNLV Retrospective Theses & Dissertations. 1393. http://dx.doi.org/10.25669/pu4m-oa41

This Thesis is protected by copyright and/or related rights. It has been brought to you by Digital Scholarship@UNLV with permission from the rights-holder(s). You are free to use this Thesis in any way that is permitted by the copyright and related rights legislation that applies to your use. For other uses you need to obtain permission from the rights-holder(s) directly, unless additional rights are indicated by a Creative Commons license in the record and/ or on the work itself.

This Thesis has been accepted for inclusion in UNLV Retrospective Theses & Dissertations by an authorized administrator of Digital Scholarship@UNLV. For more information, please contact [email protected]. INFORMATION TO USERS

This manuscript has been reproduced from the microfilm master. UMi films the text directly from the original or copy submitted. Thus, some tfiesis and dissertation copies are in typewriter foce, while others may be from any type of computer printer.

The quality of this reproduction is dependent upon the quality of the copy sulMnitted. Broken or indistinct print, colored or poor quality illustrations and photographs, print bleedthrough, substandard margins, and improper alignment can adversely affect reproduction.

In the unlikely event that the author did not send UMI a complete manuscript and there are missing pages, these will be noted. Also, if unautlTorized copyright material had to be removed, a note will indicate the deletion.

Oversize materials (e.g., maps, drawings, charts) are reproduced by sectioning the original, beginning at the upper left-hand comer and continuing from left to right in equal sections with small overlaps.

ProQuest Information and Learning 300 North Zeeb Road, Ann Arbor, Ml 48106-1346 USA 800-521-0600 UMI

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. NOTE TO USERS

Page(s) not included in the original manuscript are unavailable from the author or university. The manuscript was microfilmed as received.

62

This reproduction is the best copy available.

u m t

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. FAULT SEGMENTATION, FAULT LINKAGE, AND HAZARDS ALONG

THE SEVIER FAULT, SOUTHWESTERN UTAH

by

Usa M. Schiefelbein

Bachelor of Science University of Wisconsin - Milwaukee 1998

A thesis submitted in partial fulfillment of the requirements for the

Master of Science Degree Department of Geoscience College of Sciences

G raduate College University of Nevada, Las Vegas August 2002

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. UMI Number: 1411211

UMI*

UMI Microform 1411211 Copyright 2003 by ProQuest Information and Learning Company. All rights reserved. This microform edition is protected against unauthorized copying under Title 17, United States Code.

ProQuest Information and Leaming Company 300 North Zeeb Road P.O. Box 1346 Ann Arbor, Ml 48106-1346

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. Thesis Approval IJNIV The Graduate College University of Nevada, Las Vegas

August 1 ^2002

The Thesis prepared by

Ilsa M. Schiefelbein

Entitled

Fault Segmentation. Fault Linkage, and Hazards along the Sevier

Fault. Southwestern Utah______

is approved in partial fulfillment of the requirements for the degree of

Master of Science

Examination Committee

Dean of the Crnduate Cotte^e

Examination Coniniitte^Member

Examination Cfimmittec Member

Graduate College faculty Representative

I’K i m - IX)

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. ABSTRACT

Fault Segmentation, Fault Linkage, and Hazards along the Sevier Fault, Southwestern Utah

by

Ilsa M. Schiefelbein

Dr. Wanda J. Taylor, Examination Committee Chair Associate Professor of Geology University of Nevada, Las Vegas

Theoretically, short normal faults link to form long faults and salients develop in

the linkage zones. This project provides new data and interpretations on (1) normal fault

linkage zones; (2) earthquake and landslide hazards; and (3) fold formation along the

active Sevier fault, Utah. Geologic tmq>ping, geometric analyses, geochronology,

landslide evaluation, and stream history led to five conclusions. (1) Six principal faults

linked to form the Orderville salient and connect the Mt. Carmel and Spencer Bench

segments. Four relay ramps formed between overlapping faults. This multipartite

linkage zone implies that simple models only imperfectly represent natural examples. (2)

Two relay ramps contain rarely analyzed fault-parallel folds. These folds acconunodate a

downward decrease in space between faults. (3) Fluvial deposits indicate three

downcutting stages of a south-flowing stream. Some slip along the Sevier fault occurred

after the first two stages and tilted the deposits. (4) 570 ka basalt is offset ~3 m. Thus,

the young slip rate is 0.018 mm/yr. (5) Mechanical weathering processes or seismicity

induced 14 landslides and similar future slope failures are likely.

Ill

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. TABLE OF CONTENTS

ABSTRACT...... Hi

LIST OF FIGURES...... v

ACKNOWLEDGEMENTS...... vi

CHAPTER 1 INTRODUCTION...... 1

CHAPTER 2 CONCEPTUAL MODELS...... 3 Introduction ...... 3 Fault Linkage Models ...... 3 Displacement vs. Distance Diagrams ...... 9 Relay Ramps...... 11 Segment Boundaries ...... 14

CHAPTER 3 REGIONAL TECTONIC BACKGROUND...... 17 Introduction ...... 17 Evolution of the Transform Boundary ...... 18 Regional Tectonics of Southwestern U tah ...... 19 Mesozoic Sevier Orogeny ...... 19 Laramide Orogeny ...... 20 Cenozoic Volcanism ...... 20 Late Cenozoic Basin and Range Style Extension ...... 21 Definition of Provinces in Southern U tah ...... 23 Basin and Range Province ...... 23 ...... 23 Utah Transition Zone ...... 24 High Plateaus subprovince of the Colorado Plateau ...... 25

CHAPTER 4 STRATIGRAPHY...... 26 Mesozoic Stratigraphy ...... 26 Cenozoic Stratigraphy ...... 27

CHAPTER 5 STRUCTURAL DESCRIPTIONS...... 29 Introduction ...... 29 Sevier Fault ...... 29 Southern Domain - Orderville Relay Ramp ...... 30 Central Domain - Stewart Canyon Overlap Zone ...... 31 Spencer Bench Domain ...... 35

IV

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. Folding within the Study Area ...... 38

CHAPTER 6 VIRGIN RIVER DEPOSITS ...... 39 Introduction ...... 39 Modem Fluvial Deposits ...... 40 Young Fluvial Conglomerate ...... 41 Older Fluvial Conglomerate ...... 42 Interpretations ...... 43 Discussion of the ^rgin River Deposits ...... 44 Faulting within the Quaternary Fluvial Deposits ...... 44

CHAPTER 7 STRUCTURAL INTERPRETATIONS ...... 47 Linkage ...... 47 Examples of Fault Linkage ...... 47 Fault Capture in the Orderville Area ...... 48 Overlapping Faults in the Stewart Canyon Area ...... 49 Eastern Stewart Canyon Overlap Zone ...... 49 Central Stewart Canyon Zone ...... 50 Western Stewart Canyon Zone ...... 51 Folds within Relay Ramps...... 52 Summary...... 53 Stages of Relay Ramp Development in the Stewart Canyon Overlap Zone ...... 53 Stages of Linkage Along the Central Sevier Fault ...... 54 Segments and Segment Boundaries ...... 56 Age Constraints for the Sevier Fault ...... 57 Slip R ates...... 57 Fault Scarps...... 59 Folding along the central Sevier fault ...... 60

CHAPTER 8 EARTHQUAKE AND LANDSLIDE HAZARDS...... 63 Earthquakes...... 63 Landslides ...... 65 Identifying and Categorizing Landslides ...... 65 Regional Slope Failures ...... 67 Triggering Mechanisms for Long Valley Landslides ...... 68 Classification of the Dry Wash Landslide ...... 70 Ethical Assessment of Hazards ...... 72

CHAPTER 9 REGIONAL TECTONIC INTERPRETATIONS...... 73 Introduction ...... 73 Northern Termination of the Sevier Fault ...... 73 Other Long Normal Fault Relationships ...... 74 Sevier, Laramide, or Cenozoic Regional Folding within the Study Area ...... 76

CHAPTER 10 PETROLEUM POTENTIAL...... 77 Normal Fault Linkage and Hydrocarbon Production ...... 77

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. Hydrocarbon Production near the Study Area ...... 78

CHAPTER 11 CONCLUSIONS...... 80

APPENDIX A METHODS...... 83 Field Mapping Techniques ...... 83 Cross Section Construction Techniques ...... 83 Stereographic Analyses...... 84 "Ar/^’Ar Methods ...... 85

APPENDIX B ARGON DATA TABLE...... 86

FIGURES...... 87

REFERENCES...... 120

VITA...... 132

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. LIST OF FIGURES

Figure 1 Location Map ...... 87 Figure 2 Models of Types of Linkage ...... 88 Figure 3 Stress Field Diagram ...... 90 Figure 4 Displacement vs. Distance Models ...... 91 Figure 5 Relay Ramp Models ...... 92 Figure 6 Segment Boundary Models ...... 93 Figure 7 Tectonic Map ...... 94 Figure 8 Lower Stratigraphie Column ...... 95 Figure 9 Upper Stratigraphie Column ...... 96 Figure 10 Simplified Geologic and Location M ap ...... 97 Figure 11 Basalt Isochrons ...... 98 Figure 12 Fault Trace Map ...... 99 Figure 13 Orderville Relay Ramp Fence Diagram ...... 101 Figure 14 Orderville and Glendale Relay Ramp Stereonets ...... 102 Figure 15 Geologic Map of Quaternary Units ...... 103 Figure 16 Stewart Canyon Overlap Zone Fence Diagram ...... 104 Figure 17 Photos of Different Aged Conglomerate ...... 105 Figure 18 Photos of Channels and Sandbars within the Conglomerate U nit ...... 106 Figure 19 Displacement vs. Distance Diagrams for the Central Sevier Fault ...... 107 Figure 20 Displacement vs. Distance Diagrams for the Overlap Zones ...... 108 Figure 21 Space Accommodation Diagram ...... 109 Figure 22 Stages of Linkage along the Central Sevier Fault ...... 110 Figure 23 Fault Trace Maps of the Stages of Linkage within the Stewart Canyon Overlap Zone ...... I l l Figure 24 Types of Landslides ...... 112 Figure 25 Landslide Map of the Area ...... 113 Figure 26 Photo (A) Glendale Landslide (B) Dry Wash Landslide ...... 114 Figure 27 Photos of Vegetation on the Dry Wash Landslide ...... 115 Figure 28 Photos of Vegetation within and Surrounding the Dry Wash Landslide... 116 Figure 29 Photos of the (A) Toe and (B) Head Scarp of the Dry Wash Landslide.... 117 Figure 30 Photo of Toe of Dry Wash Landslide ...... 118 Figure 31 Relay Ramp Models for Petroleum Accumulation ...... 119 Figure 32 Location Map of Oil Fields in Southwestern Utah ...... 120

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. ACKNOWLEDEMENTS

First and foremost, I would like to thank my parents, Rick and Kathy

Schiefelbein, for their endless love, support, and encouragement and my brother for all

the humorous phone conversations and just making me laugh I would also like to thank

my advisor and friend. Dr. Wanda J. Taylor, for all her support, patience, confidence

boosting, understanding, thoughtfulness, and motivation. I could not have done it

without her. Thanks goes to my committee. Dr. Gene Smith, Dr. Terry Spell, and Dr.

Barbara Luke, for all their helpful reviews, encouragement, and meeting important

deadlines. Additionally, thanks to Dr. Gene Smith for assisting me with sample

collection, motivation in the field, and dealing with all my traumas throughout my final

semester.

I would like to acknowledge the following sources of funding which helped make

this project possible; American Association of Petroleum Geologists Grants-in-Aid,

Geological Society of America Research Grant, Sigma Xi Grants-in-Aid of Research,

Arizona/Nevada Academy of Science research grant, UNLV Graduate College Summer

Session Scholarship, Bemada E. French Scholarship, UNLV Graduate Student

Association Grant-in-Aid, Geological Society of America Travel Grants, Roy Shelmon

Meeting Awards, and a UNLV PIA grant to Dr. Wanda J. Taylor and Dr. Margaret Rees.

I would like to thank my field assistant. Treasure Bailey, for helping me keep my

sanity in the field for twelve weeks, input on geologically difficult areas, collecting

VI

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. samples, friendship, and companionship. Thanks to all of the landowners, especially, the

Spencer family, Barry and Judith Ford, the Sorensen family, the Lamb family, the

Chamberlain family, the Heaton family, the BrinkerhofT family, the Cox family, and

Norman Carroll for access to and help navigating their property. I would also like to

thank Ellen Lamb, her family, and staff the Glendale KOA for their friendship, allowing

us to camp at a very reasonable rate in an enjoyable camping environment, allowing us to

store equipment, use the phone, and watching over us. Thanks for everything Ellen.

You, your family, and staff made summer fieldwork logistics easier and very enjoyable.

I acknowledge all the help and advice on land ownership from the staff at the Kanab

Bureau of Land Management office.

Thanks to each faculty and staff member in the UNLV Geoscience Department

for all the contributions and creating an enjoyable environment to work in. I would

particularly like to thank Dr. Timothy Lawton at New Mexico State University for his

helpful contributions and discussions of the geology of the area. Thanks to Kathy Zanetti

for all the help with understanding the process of dating, dating my samples, and reading

the results in the Nevada Isotope Geochronology Lab

I would like to thank my friends and colleagues for creating an enjoyable leaming

environment. I particularly would like to thank my friends Robyn Howley, Melissa

Hicks, Amy Brock, John VanHoesen, and Jason Smith for all their support both in and

out of the office, helpful reviews and comments, good times, and friendship.

Last, but certainly not least, I would like to thank my loving boyfriend Rob

Kerscher, for all his love, support, patience, encouragement, and being the scale in my

field photos. I could have never done it without him. Thanks again.

VII

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. CHAPTER 1

INTRODUCTION

The purpose of this project is two-fold. First, conduct a detailed study of

multipartite normal fault (several faults that link and form on long fault) linkage zones

using the central Sevier fault, southwestern Utah as a case study (Fig. 1). Second,

evaluate the earthquake and landslide hazards in the region.

Most published data and interpretations focus on simple linkage zones (e.g..

Crone and Haller, 1991; Koukouvelas and Doutsos, 1996; Pavlides et al., 1998).

Although multipartite linkage zones have been mapped in the past (e.g.. Bowers, 1991;

Billingsley, 1993,1994; Barton et al., 1998; Zampieri, 2000; Ferrill and Morris, 2001),

these sites have not been analyzed in detail. Linkage zones contain information

important to understanding how long (lO's-lOO’s of km) faults form.

Linkage zones are also important to hydrocarbon exploration because active

exploration is common within them (e.g., Gulf of Mexico). Understanding multipartite

linkage zones will help determine where migration pathways and potential petroleum

traps exist.

The Sevier fault is an ideal place to study fault linkage because (1) it is located in

a relatively geologically simple region, (2) exposure is excellent, and (3) detailed work

has not been done along the central Sevier fault. The new data document a multipartite

linkage zone between Glendale and Orderville Utah, which separates two geometric

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 2 segments (Mt. Cannel and Spencer Bench segments). The linkage zone forms the

Orderville salient (bend). This salient contains the linkage sites of several faults and four

relay ramps (strain transfer zones). This number of faults and folds is greater than

predicted in simple linkage models. Two of the relay ramps contain fault-parallel folds in

one linkage area in addition to the ramp monoclines (Plates 1 and 2). Understanding the

structure within relay ramps is necessary to fully describe linkage zones. However, fault-

parallel folds within linkage zones are not commonly discussed because these folds are

typically suttle and not recognized.

The second problem addressed in this project is the earthquake and landslide

hazards for the region. Evaluating earthquake and landslide hazards is important because

the area is popular with summer tourists and is populated by small (<250 people) towns

and agricultural communities. However, little detailed work on seismic activity or

landslides has been done along the central Sevier fault. The data presented here show

that the central Sevier fault cuts -570 ka basalt and tilts Quaternary stream terraces.

Historic earthquakes have been felt and reported in the vicinity of the central Sevier fault

trace, which indicates the central Sevier fault may have been active in the Holocene. In

addition, many of the rock formations are weak and are prone to landslides. The

landslides may be either gravity or seismically induced and may pose a risk to the

residents.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. CHAPTER 2

STRUCTURAL CONCEPTUAL MODELS

Introduction

Long (lO’s-lOO’s km) normal faults cannot rupture as a whole during a single

earthquake because earthquakes can not produce enough slip to rupture lengths greater

than a few lO's of kilometers. This relationship suggests that long faults may have

formed by linkage of several shorter faults. Fault linkage forms geometric segments

(bends) that may or may not correlate to rupture segments (discussed later). Models for

fault linkage and structures formed within the linkage zones are described below.

Fault Linkage Models

Fault segment linkage is the mechanical interaction between two separate fault

segments (e.g.. Peacock and Sanderson, 1991; Trudgill and Cartwright, 1994; Cartwright

et al., 1995; Crider and Pollard, 1998; Crider, 2001; Ferrill and Morris, 2001). Linkage

affects the slip distribution on each of the faults because as the faults interact, slip is

transferred from one fault to the other fault (Crider and Pollard, 1998). The bounding

faults in linkage zones have approximately parallel strikes and display opposite

displacement gradients (Ferrill and Morris, 2001).

Two mechanically different styles of linkage are called “soft” and “hard” linkage.

Soft linkage occurs when the fault geometry changes only because of the stress field

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 4 interactions (Gupta and Scholz, 2000; Taylor et al., 2001). During soft linkage, strain is

transferred between faults through the rock. In contrast, during hard linkage the faults

physically connect (Gupta and Scholz, 2000; Taylor et al., 2001), and faults cut relay

ramps (strain transfer zones) or blocks between the faults resulting in a single fault with

multiple bends (Young et al., 2001). To elucidate fault linkage and the structures formed

within linkage zones, a definition and explanation of radial propagation, linkage, and the

stress fields at fault tips is given below.

Geometrically simple faults tend to grow by radial propagation. Radial

propagation typically occurs when an individual fault extends in length and width, where

the fault has zero slip at the fault tips and the maximum slip is located in the middle 1/3

of the fault (Fig. 2) (Cowie and Shipton, 1998). This type of growth is accompanied by

increasing displacement (or displacement gradient) and along strike propagation of fault

tips (Fig. 2) (Wu and Bruhn, 1994). With increased slip along the fault, stress builds up

at the fault tip. Eventually when the stress at the fault tips exceeds the strength of the

surrounding rock, the surrounding rock will fail and the fault will grow in length (Scholz,

1990; Cartwright et al., 1995). In this type of fault growth, the footwall is not deflected

(Cartwright et al., 1995). Faults that grow by radial propagation follow a single growth

path provided the materials have constant material properties when the fault propagates

(Cartwright et al., 1995).

Radially propagating faults may grow according to a relationship between the

fault length and the maximum displacement, which is perpendicular to the fault trace

(Cartwright et al., 1995; Cladouhos and Marrett, 1996; Cowie and Shipton, 1998). This

relationship requires the maximum displacement to be equal to critical shear strain

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 5 multiplied by the maximum length (Cartwright et al., 1995). This relationship is typically

used for faults with single growth paths because complex fault geometries will cause

scatter in the displacement and length data (Cartwright et al., 1995).

Isolated normal faults typically have a gently curved or "shovel shape" in map

view (Fig. 2). This geometry is caused by a variety of complex mechanisms (cf., Mandl,

2(K)0). When an individual fault propagates along strike or dip, the strike according to

the Mohr-Coulomb theory will be parallel to a? (intermediate principal stress) (Mandl,

2000). The shovel shape is caused by the curving of the Ot trajectories (Mandl, 2000).

The curved 0 2 trajectories may have existed prior to faulting or may have been induced

by the initial propagation of the fault and the interaction of local stress fields at the fault

tips (Mandl, 2000).

In contrast, long faults with salients (bends) tend to grow by radial propagation

followed by segment linkage (Fig. 2). The initial faults may be en echelon or non-en

echelon. Fault growth by segment linkage consists of propagation, local stress field

interaction, and linkage of segments with or without deflection of the footwall

(Cartwright et al., 1995; Ferrill and Morris, 2001).

Active normal faults have local stress fields located near the fault tips that are

different from the regional stress field. These local stress fields exist because of a

mechanical interaction between the fault tip and the surrounding rock. The local stress

field along a single dip-slip normal fault has the greatest shear strength near the fault tip.

The normal shear stresses increase beyond the local stress field and decreases near the

center of the fault. The fault is more likely to rupture in the increased shear stress field

than the decreased shear stress field (Fig. 3A) (Cowie and Shipton, 1998).

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 6 Local stress fields interact when two or more faults propagate and the tips lie

close to each other. When the zones of relatively high shear stresses near two fault tips

touch, the local stress fields begin to curve toward each other and interact (Fig. 3B).

Crider and Pollard (1998) suggest that the local stress field at the fault tip is located in the

footwall of the faults when the increased shear stresses interact (Fig. 3B-D). With

continued slip, these stress fields will curve towards each other causing the faults to

curve. When the local stress fields interact soft linkage exists between the two faults.

The shear stresses also will increase in the future strain transfer zone (relay ramp) (Crider

and Pollard, 1998). With continued slip, the local stress fields and faults will ultimately

curve and physically connect (link), called hard linkage. The hard linkage is not pictured

in Fig. 3 because this type of linkage has not been numerically modeled.

Two major types of segment linkage occur where fault tips: (1) underlap or (2)

overlap. The overlapping-fault linkage type, contains two subtypes: (1) fault capture and

(2) breakthrough or cross faults. These types and subtypes are discussed below.

Underlapping faults or fault tips are two or more individual faults, either en

echelon or non-en echelon in map view, with the tips spatially separated such that they do

not lie across strike from each other (Fig. 2A). The greatest amount of slip is typically in

the center 1/3 of the fault and by definition the fault tips have zero slip. Toward the tip,

the size of the local stress field will decrease rapidly because faulted and unfaulted rocks

are adjacent to each other (Cowie and Shipton, 1998). This difference develops the local

stress field at the fault tips. With continued slip along the faults, soft linkage will occur

(Fig. 2A). This geometry can be viewed in the field as two or more subparallel faults

curved towards each other with underlapping fault tips. However, after physical hard

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 7 linkage occurs, only one fault with one or more salients will be apparent in natural

settings (Fig. 2A).

Overlapping faults are two or more individual subparallel faults with the tip of

one fault across strike from the other fault(s) (Fig. 2). In the basic type of overlapping

fault linkage, faults with a similar amount of slip propagate past each other and

interaction between the local stress fields near the fault tips requires the faults to curve

toward each other (Crider and Pollard, 1998; Ferrill et al., 1999; Crider, 2(X)1; Ferrill and

Morris, 2(X)1 ; Taylor et al., 2(X)1 ). A relay ramp, which accommodates strain transfer

between the faults, forms in the fault overlap zone (discussed in a following section)

(e.g., Larsen, 1988; Morley et al., 1990; Peacock and Sanderson, 1994; Trudgill and

Cartwright, 1994; Crider and Pollard, 1998; Ferrill et al., 1999; Moore and Schultz, 1999;

Walsh et al., 1999; Peacock et al., 2000; Zampieri, 2000; Crider, 2001; Ferrill and

Morris, 2001; Reber et al., 2001). As the two faults continue to slip, strain is transferred

between the faults and the faults may ultimately link into one long fault (Fig. 2B).

Fault capture is a subtype of linkage in which fault tips overlap. As the faults

propagate in the direction of strike, one fault has more or a greater rate of slip than the

other fault (Ferrill et al., 1999; Reber et al., 2001; Taylor et al., 2001). More slip along a

fault may be related to one fault having more frequent and/or larger earthquakes than the

other faults. Ultimately, the stress field developed around the tip of the more active fault

interacts with that of the other fault. Then, only the larger slip or faster moving fault

curves toward the other fault and the faults link. With additional slip, the faults

ultimately transfer strain with or without the formation of a relay ramp (Fig. 2D). In the

field, faults linked by fault capture are recognized by one curved fault that connects with

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 8 another fault with a comparatively straight trace. These faults, when linked, form a

single fault. The end of the captured fault that lies across strike from the capturing fault

(fault with more or more rapid slip) commonly is isolated in the hanging wall (or possibly

the footwall) and no further slip occurs along it. The captured end may also be buried by

the hanging wall-basin sediments. If a relay ramp formed in the overlap zone, it too may

be down dropped in the hanging wall or isolated in the footwall after linkage (discussed

in a later section).

A second subtype of overlapping fault linkage is breakthrough along connecting

faults. Breakthrough along connecting faults occurs when the tips of two or more

individual faults propagate past each other, overlap and new faults with a different strike

form in the overlap zone (Cartwright et al., 1996; Ferrill et al., 1999; Ferrill and Morris,

2001). Generally, the new breakthrough faults form parallel to each other and at

approximately 30° to the bounding faults. Strain is transferred from one fault strand to

the other fault strand along this newly formed fault set (Fig. 2D). The breakthrough fault

set forms because of an increased displacement gradient in the overlap zone, which

causes extension to increase (Ferrill et al., 1999). This extension produces additional

faults (called breakthrough or cross faults) that allow slip transfer between the

overlapping faults, and occurs where fault throw is disproportional to the fault tip

propagation (Ferrill et al., 1999; Ferrill et al., 2001). If breakthrough faults propagate

from both faults, then fault bounded blocks are formed within an anastomosing fault zone

(Ferrill et al., 1999). In the field, linkage via breakthrough faults (or cross faults) is

identified by two nearby faults with similar strikes, and between the two faults are one or

more faults that strike at a moderate to high angle (-30°) to the bounding faults. Upon

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 9 linkage, these faults become one fault. The breakthrough faults are unlikely to cross the

longer bounding faults, but relict fault tips within the overlap zone are likely to be seen

(Ferrill et al., 1999).

Spacing or the distance between faults is important in determining whether fault

linkage has occurred or will occur. If the faults are spaced too far apart, local stress fields

will not interact. Hence, fault linkage will not occur. Fault spacing at the surface and at

depth may differ because of the dips of the faults. For example, the steeper the fault dip,

the closer in space they need to be in order to link at shallow depths. Faults with

shallower dips can be spaced farther apart and still link at a shallow depth.

Displacement vs. Distance Diagrams

Displacement vs. distance diagrams are used to show the total amount of offset at

various positions along strike of a fault. Here, displacement vs. distance diagrams are

used to emphasize displacement variations within linkage zones (cf., Reber et al., 2001;

Taylor et al., 2001). Focusing on the linkage zones is important, because displacement

anomalies in these areas help diagram and indicate which type of linkage occurred. This

use is different from other studies (e.g. Cowie and Scholz, 1992; Cartwright et al., 1995;

Bohnenstiehl and Kleinrock, 2000) that use displacement vs. distance diagrams to model

the entire fault.

In the vicinity of salients or geometric bends that are convex toward the hanging

wall, on displacement vs. distance diagram can be used to evaluate whether linkage

occurred and the type of linkage providing two assumptions are valid: (I) The total

amount of slip on each individual fault before linkage was zero at the fault tip. (2) Since

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 10 the time of linkage, the slip rate was not markedly greater in the linkage zone than in the

middle 1/3 of each original fault. Thus, in a linkage zone, the displacement should

markedly decrease near the former fault tips. Variations in displacement vs. distance

were modeled along idealized faults to show patterns associated with underlapping faults,

overlapping faults, and fault capture types (Fig. 4) (Taylor et al., 2001). New models for

breakthrough hard linkage and soft linkage were created for this study and are described

below (Fig. 4). These models are used in this study to interpret fault linkage zones along

the central Sevier fault.

Along linked underlapping faults, the slip magnitude decreases sharply in the

center of the fault, which is where the slip or displacement would be expected to be the

greatest on unlinked faults (Fig. 4A). This decrease occurs at the linkage zone. After the

faults linked, additional slip may occur along the entire fault, along zones defined by the

geometry or along distinct earthquake rupture segments (discussed in segment boundaries

section).

On a displacement vs. distance diagram, a basic overlapping type linkage zone

shows less displacement across a broad bench at the site of linkage (Fig. 4B). The bench

forms because two faults propagated past each other, so the total amount of displacement,

added across both faults along strike in the linkage zone, is similar or the same. The

bench length is the same as the length of the overlap zone (Taylor et al., 2001).

However, during the soft linkage of overlapping faults, instead of a bench developing in

the zone of linkage, the displacement tapers to zero (Fig. 4E).

Fault capture subtype linkage zones produce a series of plateaus with decreasing

slip on a displacement vs. distance diagram (Fig. 4C). This decreasing series of plateaus

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. il occur because the fault with the greatest amount or rate of slip intercepts the fault with

the lesser amount of slip at a site with significant displacement. However, displacement

at that site may be less than the maximum displacement on the intercepting fault. The

fault with the lesser amount of slip is cut off and becomes a relict fault.

The breakthrough subtype linkage zone produces a pattern similar to the

overlapping fault linkage zone on a displacement vs. distance diagram but with small

amplitude perturbations within the plateau (Fig. 4D). If the displacement data are

sampled with a small spacing, then "steps" occur where the cross faults are located. No

cross faults exist at the tips of the bounding faults hence the total displacement abruptly

increases in slip outside the linkage zone, following the displacement gradient.

Where a relay ramp forms in the linkage zone, the faults that bound the relay

ramp have tapering slip as they enter the overlap zone (Crider and Pollard, 1998). These

geometries can be related to the displacement gradients at the fault tips (Peacock and

Sanderson, 1991).

Relay Ramps

Linkage of overlapping faults typically exhibit an interaction phase during which

a relay ramp forms. A relay ramp is an inclined panel of rocks between two overlapping

faults with smoothly curving tiplines (Fig. 5) (Larsen, 1988; Peacock and Sanderson,

1991 ; Crider and Pollard, 1998). In the relay ramp, the strike of bedding changes from

parallel to the bounding faults outside the ramp to nearly perpendicular to the bounding

faults within the ramp (Fig. 5) (e.g., Trudgill and Cartwright, 1994; Faulds and Varga,

1998; Ferrill et al., 1999; Moore and Schultz, 1999; Peacock et al., 2000: Ferrill and

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 12 Morris, 2001; Reber et al., 2001). Relay ramps can range in area from square meters to

lO's of square kilometers.

Relay ramps and the structures formed within the ramps are important to aid in

the understanding of (1) how segmented long normal faults link, (2) how faults grow, and

(3) how strain is transferred from one fault strand to another fault strand. Relay ramps

also are important because they allow fluid (hydrocarbons or water) migration from the

footwall to the hanging wall, which may form a fluid trap (Morley et al., 1990; Crider

and Pollard, 1998; Dawers and Underhill, 20(X)).

Relay ramps initiate and begin to form in soft linkage zones between two fault

tips. Relay ramps typically form along overlapping faults; including the subtypes of

breakthrough faults and fault capture (Fig. 5). The relay ramp models below will be used

to describe and interpret linkage zone structures exposed along the central Sevier fault.

Relay ramps connect the hanging wall of one fault strand to the footwall of

another fault strand, effectively transfer strain between the two strands, and ultimately

link two normal faults (Larsen, 1988; Morley et al., 1990; Peacock and Sanderson, 1991;

Peacock and Sanderson, 1994; Trudgill and Cartwright, 1994; Crider and Pollard, 1998;

Faulds and Varga, 1998; Ferrill et al., 1999; Moore and Schultz, 1999; Walsh et al., 1999;

Zampieri, 2000; Peacock et al., 2000; Ferrill and Morris, 2001; Reber et al., 2001). The

model or ideal relay ramp transfers strain from the hanging wall of one fault to the

footwall of the other fault by (1) tilting of the bedding within the ramp, (2) vertical axis

rotation, and (3) cutoff-parallel elongation without additional deformation to the relay

ramp or the surrounding rocks (Ferrill and Morris, 2001).

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 13 A relay ramp accommodates displacement gradients along the bounding faults.

These bounding faults typically have opposite displacement gradients (Ferrill and Morris,

2(X)1). Peacock and Sanderson (1991) suggest that a relationship exists between the

geometry of the relay ramp and the displacement variations of the bounding faults; a

more complex displacement gradient leads to a more geometrically complex relay ramp.

In addition, greater displacement gradients on the bounding faults form larger (vertical

vs. horizontal dimension) ramps (Peacock and Sanderson, 1991; Peacock et al., 2000).

As slip continues after the initial formation, relay ramps commonly expand or propagate

parallel to strike of the bounding faults (Fig. 5). Hence, the relay ramp must continue to

deform to keep the hanging wall and footwall connected (Ferrill and Morris, 2001).

Breached relay ramps are completely bound by two hard linked faults and may

contain cross faults or fractures (Peacock and Sanderson, 1994; Ferrill and Morris, 2001).

These ramps can be further subdivided into breached top and base ramps (Crider, 2001).

Breached top and base ramps occur when a bounding fault cuts across the ramp (hard

linkage site) at the structurally higher or lower parts of the ramp, respectively (Fig. 5)

(Crider, 2001). Cross faults in breached relay ramps may form because of tip interaction

prior to fault overlap (Ferrill and Morris, 2001). This may explain why the breakthrough

fault orientation is at an oblique angle to the relay ramp (Ferrill and Morris, 2001).

Relay ramps may contain a variety of structures. By definition, a relay ramp

contains a fold and is essentially a monocline and trends approximately normal to the

bounding faults (Fig. 5) (Larsen, 1988; Peacock and Sanderson, 1991; Peacock and

Sanderson, 1994; Faulds and Varga, 1998). In addition, other contractile structures may

develop within the relay ramp, including anticlines and synclines depending on the

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 14 geometry of the bounding faults. These fold axes are approximately parallel to the

bounding faults and form in addition to the already existing ramp monocline. Structures

such as these have been rarely studied in the past.

Segment Boundaries

Fault segments and the boundaries between them can be subdivided into three

types: (1) structural, (2) geometric, and (3) earthquake (dePolo et al., 1991). It is

possible, therefore, to define segment boundaries based on surface observations of fault

geometry, scarps, footwall structures, kinematic indicators, and/or earthquake epicenters

(McCalpin, 1996; Stewart and Taylor, 1996). However, most segment boundaries are not

a single point, but rather a broad complexly faulted zone (Janecke, 1993).

Geometric segments are recognized by changes in fault zone morphology, fault

trace orientations (bends, step-overs or en echelon faults, separations), and a change in

fault separation or a gap in a fault zone (Fig. 6A) (dePolo et al., 1991; Keller and Pinter,

1996). Geometric segment boundaries typically exhibit a dramatic change in strike that

shape a salient. These changes in strike usually occur in zones of linkage. Geometric

segments may be consistently and objectively identified on aerial photographs or by

detailed geologic mapping.

Structural segment boundaries end at a structural discontinuity (Fig. SB) (dePolo

et al., 1991). Structural discontinuities can be older faults or folds that strike at a high

angle to the segmented fault. A change in the geologic material (e.g., changing from

coherent rock to fault gouge back to coherent rock) crossed by the fault may be a

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 15 characteristic of a structural segment boundary (Fig. 6B) (Janecke, 1993; Keller and

Pinter, 1996).

Earthquake segments are portions of a fault zone that rupture as a unit during an

earthquake (Fig. 6) (dePolo et al., 1991; McCalpin, 1996). McCalpin (1996) further

defines earthquake segments as "discrete portions of faults that have demonstrably

ruptured to the surface two or more times". However, in large earthquakes more than one

segment may rupture. Earthquake segments are separated by earthquake rupture

terminations that are a zone that typically contains several fault splays and high amounts

of fractured, crushed, and faulted rocks (Janecke, 1993). An earthquake segment

boundary or earthquake discontinuity, is therefore defined as "a portion of a fault where

at least two rupture zones have ends" and can potentially arrest earthquake ruptures

(Wheeler, 1989; dePolo and Slemmons, 1990). The most reliable method of

documenting earthquake segments is paleoseismological evaluation and fault behavioral

data from historic earthquakes (Zhang et al., 1991). Determining earthquake segments is

the most important concept for evaluating earthquakes and their associated hazards.

However, earthquake segment boundaries do not always stop all propagating ruptures

(Crone and Haller, 1991). For example, a large (M 7+) earthquake can rupture more than

one earthquake segment boundary. Hence, earthquake hazard evaluation needs to be

done beyond a determined earthquake segment boundary.

A segment boundary can be any one type, or a combination of geometric,

structural, and earthquake boundaries (Fig. 6). For example, a structural boundary where

an older fault strikes at a high angle to the segmented fault can also serve as an

earthquake rupture boundary. An example would be the 1914 Pleasant Valley, central

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 16 Nevada earthquake. This earthquake was terminated by a structural discontinuity, and

thus, is also an earthquake segment boundary (Zhang et al., 1999). Another example is a

dramatic change in strike, a geometric segment boundary, that may also serve as an

earthquake rupture boundary. An example of this type is the Lost River and Lemhi faults

in east central Idaho. These faults have geometric bends that are also earthquake segment

boundaries (Janecke, 1993). Hence, segment boundaries may have the physical

characteristics of one or more type of boundary.

Some studies in seismology and fracture mechanics indicate that fault geometry

can be important in the generation of earthquakes and rupture patterns (Bruhn et al.,

1987; Menges, 1990). Indeed, studies along the document that some

recent earthquake rupture breaks are limited to geometric segments (Stenner et al., 1999a,

1999b). In contrast, dePolo et al. (1991) suggest that geometric segments in map view

may not have a significant effect on earthquake ruptures with a normal sense of

displacement during large (7.0+) magnitude earthquakes. However, smaller (e.g., M 3.0)

earthquakes typically only rupture a single segment and the segment boundary will arrest

the propagating earthquake. The size of the segment boundary that arrests an earthquake

appears to scale with the length of the ruptured fault and the amount of displacement

during the rupture (Zhang et al., 1999).

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. CHAPTER 3

REGIONAL TECTONIC BACKGROUND

Fault linkage models are motivated by and tested using natural examples (for this

study, the central Sevier fault in southwestern Utah). To understand fault linkage models

and the development of fault segments, the regional tectonics must be understood. Also,

the regional deformation represents the regional stress field within which the faults and

local stress field at the fault tips formed.

Introduction

Throughout the history of the western United States plate interactions along the

western plate margin have influenced tectonism. Through the Paleozoic and into the

Tertiary, subduction dominated. From the Jurassic to at least the late Cretaceous, the slab

dip was steep (60°) (Atwater, 1970; Sveringhaus and Atwater, 1990; Atwater and Stock,

1998). However, from late Cretaceous through the Cenozoic, the slab dip changed to flat

or shallow (<30°). As a result, two orogenies occurred near the western edge of the

Colorado Plateau between latest Jurassic and Eocene: the Sevier and Laramide orogenies.

These orogenies occurred because of shallow slab subduction. Folds and thrust faults of

these orogenies can occur throughout the Utah Transition Zone to the High Plateaus

subprovince of the Colorado Plateau. At -23 Ma, a transform boundary began to form

along the western edge of North American (Sveringhaus and Atwater, 1990). This event

17

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 18 is recognized by the replacement of the time-transgressive sequence of arc magmatism by

the San Andreas transform fault (Atwater, 1970; Sveringhaus and Atwater, 1990). The

length of the transform boundary has increased to nearly 1,125 km (700 mi) since its

inception (Sveringhaus and Atwater, 1990).

Evolution of the Transform Boundary

At -23 Ma the San Andreas transform boundary was initiated when the East

Pacific Rise made contact with the North America Plate (Atwater, 1970). This change

from subduction along the plate margin to a transform boundary had a great effect on the

tectonic evolution of western North America. Dickinson and Snyder (1979) suggested

that the formation of a transform boundary created a slab-window beneath part of western

North America. However, Sveringhaus and Atwater (1990) and Atwater and Stock

(1998) suggest that this slab-window is more of a slab-gap, meaning a completely

slabless area exists underneath North America. The slabless area is a sub-rectangular

shape rather than a "triangle" under North America. The formation of the slab-free zone

relates to several geologic events including regional uplift of the Colorado Plateau,

central Basin and Range extension, formation of the San Andreas fault system, and the

end of arc magmatism.

Today most of the central Basin and Range, Utah Transition Zone, and Colorado

Plateau (Fig. 1) overlie this slab-free zone. Basin and Range style normal faults have cut

into the western margin of the Colorado Plateau in southwestern Utah. These faults

probably initiated in the Miocene.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 19 Regional Tectonics of Southwestern Utah

During the Phanerozoic, southwestern Utah was a site of (1) deposition of a thick

cratonal Mesozoic and Cenozoic sedimentary sequence; (2) the Sevier Orogeny, mainly

during the Mesozoic; (3) the Laramide orogeny mainly during the early Tertiary, (4)

Cenozoic volcanism and sedimentation; and (5) Cenozoic extension. The following

section provides a simplified tectonic overview of southwestern Utah including the

southeastern Basin and Range Province, Utah Transition Zone, High Plateaus

subprovince of the Colorado Plateau, and the western edge of the Colorado Plateau.

Mesozoic Sevier Orogeny

During the Jurassic into the Eocene, the Sevier Orogeny formed a contractile belt,

the frontal part of which extends from southeastern California, through Utah, and into

northern Canada (Armstrong, 1968). The onset of the Sevier Orogeny was probably

caused by the subduction of the Farallon plate with accelerated plate convergence

(Armstrong and Ward, 1991). During the last decade, most workers have agreed that the

majority of deformation associated with the Sevier orogeny in southwestern Utah is

bracketed between late Albian time and late Campanian (DeCelles, 1995; Lawton et al.,

1997). Sevier thrusting resulted in tectonic thickening, development of an extensive

foreland basin east of the thrust front, and the development of broad folds and uplifts

(Armstrong, 1968; Axen et al., 1990; Burchfiel et al., 1992; Wannamaker et al., 2001).

The foreland basin lies within the vicinity of and extends to the east of the present

location of the Sevier fault. Surface breaking thrusts are thought to have only extended

as far east as Cedar City (Fig. 1) (Hintze, 1988; Axen et al., 1990).

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 20

Laramide Orogeny

Shortly after the Sevier orogeny, the Laramide orogeny occurred between 75 and

35 Ma (Dickinson et al., 1988; Bird, 1998). During the early portion of the Laramide

orogeny (75-65 Ma), the Kula Plate subducted under North America (Bird, 1998). After

65 Ma, the Farallon Plate began to subduct (Bird, 1998). At -50 Ma, the subducted

Kula/Farallon Plate transform boundary passed under the present western edge of the

Colorado Plateau (Bird, 1998). Because of the plate configurations, this orogeny had

driving mechanisms different from the Sevier orogeny. The Laramide orogeny was

driven by basal traction of flat slab subduction, not by motions at the plate margin (Bird,

1998). Evidence of this orogeny (gentle folds) is visible to the east and south of the

Paunsaugunt fault (Fig. 1) (Davis, 1999; Wannamaker et al., 2001). The Laramide

orogeny formed gentle folds such as the monoclines in the eastern High Plateaus east of

the study area (i.e., the Kaiparowits and Waterpocket folds). These folds are thought to

overlie high-angle Precambrian normal faults, which were reactivated as reverse faults

(Davis, 1978; Dumitru et al., 1994).

Cenozoic Volcanism

Volcanism in southern Utah began at -30 Ma (Rowley, 1975; Best et al., 1980)

and continued into the Holocene. Most of the basalt flows in southwestern Utah range

from 10.8 to 0.3 Ma in age (Best et al., 1980). The basalt flows are generally younger to

the east (Best et al., 1980; Nelson and Tingey, 1997). Basalt flows along the Sevier fault

are -0.5 Ma (Best et al., 1980; this study).

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 21 Along the latitude of the southern boundary of Utah (37“N), the effects of

subduction of the Farallon Plate ended at -10 Ma indicating that the post 10 Ma volcanic

rocks in southwestern Utah are not related to active plate subduction (Best et al., 1980;

Severinghaus and Atwater, 1990; Atwater and Stock, 1998). However, the volcanism in

southern Utah may be related to Basin and Range style extension (Best et al., 1980;

Nelson and Tingey, 1997). Cenozoic volcanic units crop out (1) near the Hurricane fault,

(2) near Navajo Lake, (3) in the Marysvale volcanic field, and (4) along and near the

central Sevier fault (Figs. 6 and 7) (Rowley, 1975; Best et al., 1980; Smith et al., 1999).

Late Cenozoic Extension in Southern Utah

Near St. George, the Hurricane fault defines the structural boundary between the

Transition Zone and the High Plateaus. The Hurricane fault is an active, segmented long

normal fault zone (Stewart and Taylor, 1996; Stewart et al., 1997; Taylor et al., 2(X)1).

Movement along the Hurricane fault probably initiated during the Miocene or Pliocene

and continues today (Stenner et al., 1999; Davis, 1999; Taylor et al., 2001). Several

earthquakes and Holocene fault scarps have been located along the Hurricane fault

(Arabasz and inlander, 1986; Christenson and Nava, 1992; Stewart et al., 1997; Lund et

al., 2002). A M 5.8 earthquake is thought to have occurred along the Hurricane fault in

1992 (Christenson and Nava, 1992).

The Sevier-Toroweap fault, a segmented long normal fault, lies -65 km (-40 mi)

to the east of the Hurricane fault. I will use the name Sevier fault when referring to the

Utah portion, Toroweap fault when referring to the portion, and Sevier-

Toroweap fault when referring to the entire fault. The Sevier-Toroweap fault can be

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 22 traced -195 km (-120 mi) from the in Arizona north into the Miocene

Marysvale volcanic field in Utah (Davis, 1999). The Sevier fault probably initiated

during the Miocene and may be active today (Smith and Arabasz, 1991; Hecker, 1993;

Davis, 1999). The central Sevier fault has not been observed cutting Holocene

sediments. However, Jackson (1990) and Hecker (1993) suggest that evidence for recent

activity may be lacking or obscured because of active colluvial slopes and thick

vegetation. The Sevier fault cuts Quaternary basalt and sediments suggesting at least

Quaternary activity (Christenson and Nava, 1992; this study). A few earthquakes have

been recorded in the vicinity (<5 km) of the Sevier fault (Arabasz and Inlander, 1986)

suggesting that this fault may have been active in the Holocene.

The Paunsaugunt fault is the easternmost long normal fault in the High Plateaus.

The fault trace extends -65 km (-40 mi) (Davis, 1999). The Paunsaugunt fault is thought

to be the oldest of the three faults because the hanging wall is topographically higher than

the footwall. This topographic reversal may be caused by a more resistant lithologie unit

capping the hanging wall than the footwall, thus allowing the footwall to erode faster

than the hanging wall. Such topographic reversal takes a significant time to form,

supporting the suggestion that the Paunsaugunt fault has not ruptured recently

(Schiefelbein and Taylor, 2000). The balanced rocks (hoodoos) of the Tertiary Claron

Formation in Bryce Canyon National Park (Fig. 1) also may suggest a lack of large recent

seismic events (Bruhn and Wu, 1993; Brune, 1996, Davis, 1999).

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 23 Definitions of Provinces in Southern Utah

In southern Utah, the Transition Zone is generally further defined by three

different criteria: (1) structural geology, (2) physiography, and (3) geophysics. The

structural definition is based on the structural geology of the area. The physiographic

definition includes geographically defined boundaries. The geophysical definition is

based on crustal thicknesses of the region.

Basin and Range Province

The Basin and Range province is characterized by a thin crust, high heat flow,

magmatism, and north-south trending valleys and mountain ranges (Parsons, 1995;

Wannamaker et al., 2001). These valleys and ranges are bounded by north-south striking

long normal faults. The region is seismically active. Prior to Cenozoic extension, rocks

in the Basin and Range province underwent several episodes of deformation (contraction

and extension).

The region also had significant volcanism since its first initiation -35 Ma

(initiation time varies with latitude) and is seismically active. At the latitude of the

Utah/Arizona border (37"N), the eastern boundary is defined as the Hurricane fault

(Stewart et al., 1997).

Colorado Plateau

In southern Utah, the Colorado Plateau is defined as a physiographic and tectonic

province that lies to the east of the Paunsaugunt fault. The Colorado Plateau is

topographically high and contains generally gently dipping strata, broad regional folds,

and lacks major thrust and normal faults. This province is underlain by as much as 50 km

(31 mi) of crust and is considered relatively tectonically stable (e.g., Beghoul and

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 24 Barazangi, 1989; Nelson and Harris, 2001; Wannamaker et al., 2001). It has undergone

little deformation since the Proterozoic (Wannamaker et al., 2001). However,

contraction, volcanism, and extension have been repeated through geologic time along

the flanks.

Utah Transition Zone

The Utah Transition Zone lies between the Basin and Range province and the

Colorado Plateau. The Transition Zone changes in width from north to south. This

transitional region is suggested to be an uplifted rift shoulder in the north along the

Wasatch Front near Salt Lake City, Utah and its southern extension (Wannamaker et al.,

2001). In southwestern Utah, the transition occurs across an -62 mi (-100 km) wide

zone (Wannamaker et al., 2001).

The initiation and boundaries of the Transition 2kne are still not completely

understood. The Transition Zone has characteristics of both the Basin and Range

province and the Colorado Plateau. It contains a gradual change in geologic

characteristics from Basin and Range style deformation to the less deformed tectonic

style of the Colorado Plateau. It is thought that the Transition Zone is not related to

Pacific and North American plate boundary interactions because the plate boundary

stresses could not propagate this far into the continental interior (Wannamaker et al.,

2001). Wannamaker et al. (2001) suggest that the Transition Zone has been an evolving

structure since 25-30 Ma. This timing is based on the onset of extensional events, which

create the present day Transition Zone (Wannamaker et al., 2001).

The physiographic and structural distinctions place the western boundary of the

Transition Zone at the mouth of the Virgin River Gorge in Arizona and the eastern

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 25 boundary at the Hurricane fault. Structurally, the Transition Zone has Basin and Range-

type normal faults. Geophysically, the transition zone is defined by crustal thickness and

generally lies between the Hurricane and Paunsaugunt faults. The Transition Zone crust

is generally between 19-22 mi (30-35 km) thick (Wannamaker et al., 2001).

High Plateaus Subprovince of the Colorado Plateau

The High Plateaus subprovince of the Colorado Plateau is located between the

Transition Zone and the Colorado Plateau. This subprovince generally lies between the

Hurricane and the Paunsaugunt faults (Fig. 1) and incorporates the . The

High Plateaus are divided by generally north-south oriented normal faults: the Hurricane,

Sevier-Toroweap, and Paunsaugunt faults, from west to east (Fig. 1). This region has

characteristics of the Colorado Plateau as well as the Transition Zone. The crustal

thickness in the High Plateaus is generally 22-25 mi (35-40 km), which is between the

thickness of the Transition Zone and the Colorado Plateau (Wannamaker et al., 2(X)1).

The high-angle normal faults resemble those in the Transition Zone. Hence, the High

Plateaus are variously grouped with either the Transition Zone or the Colorado Plateau.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. CHAPTER 4

STRATIGRAPHY

Stratigraphie units with many well-defined thin members are exposed in the High

Plateaus subprovince and along the Sevier fault (Gregory, 1951; Hintze, 1973; Sargent

and Philpott, 1987; Hintze, 1988; Stokes, 1988; Doelling, 1989). These thin members are

ideal for documenting small stratigraphie separations and allow relatively small

uncertainties in the calculations o f stratigraphie separation along the Sevier fault and

related strands and splays.

The stratigraphy exposed along the central Sevier fault consists of a thick

succession of late Triassic to Cretaceous carbonate and siliciclastic rocks, which are

imconformably overlain by Cenozoic volcanic and sedimentary units (Fig. 8). No

metamorphic or intrusive units crop out in the study area. A detailed description of the

lithologie units is presented on Plate 3. Units and thicknesses that are Triassic and older

on the cross sections, but not exposed in the map area are from Hintze (1988) and

Doelling et al. (1989) (Fig. 8). The units are dominantly limestone, dolostone, and

sandstone (Fig. 8).

Mesozoic Stratigraphy

The exposed Mesozoic section is approximately 2,300 m (7,546 ft) thick, gently

folded, and generally dips to the northwest and northeast (Fig. 9) (Gregory, 1951; Hintze,

26

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 27 1973, 1988; Sargent and Philpott, 1987; Stokes, 1988; Doelling et al., 1989; Nelson and

Tingey, 1997). The Late Triassic through Cretaceous units in the study area dominantly

consist of sandstone and shale with a minor amount of limestone of the Jurassic Carmel

Formation (Fig. 9). Three unconformities exist in the Mesozoic section: the J-1, J-2, and

K-1 (Hintze, 1973; Hintze, 1988; Marzolf, 1988; Stokes, 1988; Doelling et al., 1989)

(Fig. 9).

Cenozoic Stratigr^hy

Tertiary and Quaternary volcanic and sedimentary units unconformably overlie

the Mesozoic succession (Fig. 9). The Cenozoic section is approximately 400 m (1,312

ft) thick (Fig. 9) (Gregory, 1951; Hintze, 1973,1988; Sargent and Philpott, 1987; Stokes,

1988; Doelling et al., 1989; Nelson and Tingey, 1997). The Tertiary succession consists

of a fresh water unit with mostly fluvial and lacustrine conglomerate and sandstone

(Rowley et al., 1975; Stokes, 1988; Doelling et al., 1989; Taylor, 1993: Goldstrand,

1994). The Tertiary rocks generally dip to the northeast.

Late Cenozoic basalt is common in the region (Nelson and Tingy, 1997; Smith et

al., 1999, Downing et al., 2001) and a flow crops out in the mapped area. This basalt is

olivine rich with olivine phenocrysts up to 3 mm (0.11 in) in diameter. The ground mass

is fine, crystalline and dark gray. Basalt flows near Spencer Bench have been dated by

K-Ar techniques at 0.56 Ma +/- 0.07 Ma (Fig. 10 and Plate 1) (Best et al., 1980). Best et

al. (1980) interpreted this flow to be a hawaiite. Using ‘*°Ar/^^Ar dating techniques

(Appendices A and B), two samples (BRC-1 and BRM-2) from the northern portion of

Spencer Bench and Black Mountain were dated (Figs. 10 and 11). One sample was

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 28 collected from the hanging wall at Black Rock Canyon, and the other sample was

collected from the footwall of the Sevier fault at Black Mountain (Fig. 11 and Plate 1).

The sample from Black Rock Canyon yielded a ^Ar/^^Ar isochron age o f0.564 +/- 0.02

Ma (Fig. 11). The sample from Black Mountain yielded an ‘'^Ar/^^Ar isochron age of

0.58 +/- 0.05 Ma (Figs. 10 and 11). These ages are similar to the ages determined by

Best et al. (1980) for volcanic rocks in Black Rock Canyon (Fig. 10).

Late Cenozoic sedimentary units include spring deposits, stream terraces, slope

failures, colluvium, sinter-type spring deposits, and alluvium. The older part of the

Quaternary succession contains (now consolidated) fluvial terrace deposits that are

overlain by basalt. The most recent deposits are unconsolidated fluvial alluvial and

surficial debris (Hintze, 1973; Hintze, 1988; Doelling et al., 1989). These late Cenozoic

units unconformably overlie units ranging from Jurassic through the early Tertiary.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. CHAPTERS

STRUCTURE DESCRIPTIONS

Introduction

The Sevier fault brittlely deforms the nearly flat-laying to gently folded strata of

the High Plateaus. Fault splays, strands, and separate normal faults were identified by

fault breccia, gouge, and offset units. The sense of separation on each fault was

determined by ages (inferred by lithology and fossils ages) of juxtaposed strata. The fault

with greatest stratigraphie separation is referred to as the main strand. Because of the

lack of kinematic indicators, the exact net-slip direction was not determined for the faults.

New data from this study show that four relay ramps and a geometric bend, or salient,

occur along the central Sevier fault. This section describes the newly collected data and

the geometries of structures exposed in the study area are described below.

Sevier Fault

The Sevier/Toroweap fault generally strikes ~N30“E and dips 70-85“W along a

trace length of ~3S0 km (220 mi) between the northern tip line (northern most point) in

Miocene volcanic rocks near Richfield and the southern tip south of the Grand Canyon in

Paleozoic rocks (Fig. 7) (Gregory, 1951; Doelling et a l , 1989; Davis, 1999). The Sevier

fault from the Arizona/Utah border to the northern termination is 260 km (158 mi) in

length.

29

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 30 For the purpose of this study, the central Sevier fault will be subdivided into three,

separate domains: the southern domain, central domain, and Spencer Bench domain (Fig

12A). Each of these domains contains different fault strands and splays. The strike of

these strands varies between north-south and east-west, a 90° variation about the general

~N30°E strike. The "main strand" of the Sevier fault is defined here as the fault with the

greatest stratigraphie separation. The fault that is the main strand changes along strike.

The strikes vary from N5°E to N80°E.

Southern Domain - Orderville Relay Ramp

The southern domain extends from the southern boundary of the map area to ~1

km (-0.6 mi) north of the community of Orderville (37°15'50"N to 37°16'54"N) (Fig.

12A). This domain contains the Orderville relay ramp on which Davis (1999) made

preliminary structural and stratigraphie observations. The rocks exposed in this domain

range from late Triassic to (Quaternary in age (Plate 2). Rocks as young as Cretaceous in

age are faulted, but no Quaternary units are offset in this domain.

Two strands of the Sevier fault bound the Orderville relay ramp: A on the west

and B on the east (Figs. 12A and 12B). Fault A, the main strand o f the Sevier fault has a

curved fault trace, but generally strikes N10°E and dips 76°W. The fault strike changes

from approximately north to south in the south to northeast in the north (Figs. 12A and

12B). Fault A juxtaposes the Cretaceous Tropic Shale against Jurassic Navajo Sandstone

for -760 m (-2,500 ft) of stratigraphie separation. Fault B generally strikes N40“E and

dips 79“W (Plate 4). From south to north, the strike of fault B curves from northeast to

north, back to northeast and finally to north-south. This fault juxtaposes the Jurassic

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 31 Navajo Sandstone and the Jurassic Co-op Creek Member of the Carmel Formation for

-140 m (-450 ft) of stratigraphie separation (Plate 4).

Between faults A and B within the Orderville relay ramp (Fig. 13), map patterns,

cross sections, and stereo plots show that the stratigraphie units form a plunging syncline

oriented 17°, N7°W (Fig. 14A and Plate 4). The fold within the Orderville relay ramp is a

gentle, subhorizonal, upright fold (cf., Ramsey, 1967). However, to the west of fault A

and to the east of fault B, the strata are generally flat lying to gently west dipping.

Faults A and B connect at the surface at the southern end of the relay ramp (Figs.

12A and 12B and Plate 2). Cross section construction and retrodeformation indicate the

interpretation that these faults also connect in the subsurface (Fig. 13 and Plate 4).

In the subsurface, the faults in this region are interpreted to have different

geometries. Fault A has a relatively simple and planar geometry (Plate 4). Fault B is

planar at the surface and becomes curved by a depth of-1,220 m (-4,000 ft). This fault

is required to be curved in order to restore cross sections (Fig. 13 and Plate 4).

The depth to the base of the relay ramp is interpreted to increase from south to

north. Retrodeformation cross sections suggest that these faults coruiect in the southern

portion of this region at a depth o f670 m (2,200 ft). However, in the northern portion of

the relay ramp, the ramp probably initiates at a depth of -2,135 m (7,000 ft) (Fig. 13 and

Plate 4).

Central Domain - Stewart Canyon Overlap Zone

The central domain extends from -10 km (6 mi) northeast of the community of

Orderville to -2 km (-1.2 mi) south of the community of Glendale (37°16'54"N to

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 32 37°19^9”N) (Figs. 12A and 12B). The southern part of this domain is here named the

Stewart Canyon overlap zone (Figs. I2A and I2B). The Stewart Canyon overlap zone

exposes rocks that range in age from Jurassic to (Quaternary (Plate 1). The total

stratigraphie separation in this domain ranges from 770 to 1,045 m (2,525 to 3,425 ft).

Rocks as young as Cretaceous are faulted. However, no (Quaternary offset is documented

in this section.

The main strand of the Sevier fault within the Stewart Canyon overltq) zone

(Figs. 12A and 12B), has strikes that vary from N35°E to N65°E, and dips from 76-84°W.

In this region a portion of the main strand of the Sevier fault is buried by Quaternary

alluvium (Figs. 12A and 12B and Plate 1). Retrodeformable cross section construction

geometrically requires the main strand of the fault to be under the alluvium, near the

modem East Fork of the Virgin River (cross sections F-F' and E-E' on Plate 4) (Fig. 15).

To the south, near cross section F-F', the main strand of the Sevier fault has 435 m (1,425

ft) of offset and places the Jurassic Windsor Member of the Carmel Formation next to

Jurassic Navajo Sandstone (Fig. 16 and Plate 4). However, to the north near cross section

E-E’ the main strand of the Sevier fault has 390 m (1,275 ft) of stratigraphie separation

(Fig. 16 and Plate 4), where it places Jurassic Navajo Sandstone against the Jurassic

Windsor.

The Stewart Canyon overlap zone contains eleven fault strands with three

different fault set orientations; striking northwest, northeast, northeast-east and dips of

69°-84° to the north and west (Figs. 12A and 12B). Two of these faults (faults D) strike

~N55°E but bend northward to strike nearly east-west (Figs. 12A and 12B). These

strands are interpreted to connect at depth (cross section F-F' on Plate 4). At the surface.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 33 the entire group o f faults cuts the Jurassic Navajo Sandstone through the Cretaceous

Dakota Sandstone in exposures. At cross section F - F', the Sevier fault zone has 770 m

(2,525 ft) of stratigraphie separation (Fig. 16 and Plate 4). In cross section F-F’ the

westerly splay is interpreted to connect to the main strand at a depth of 1,280 m (4,200 ft)

(Fig. 16 and Plate 4). The center fault (fault C) is interpreted to connect at depth with the

western strand (fault B) at a depth o f960 m (3,150 ft) (Figs. 12A, 12B, and 16 and cross

section F-F on Plate 4).

Three relay ramps occur within the Stewart Canyon overlap zone; the Glendale,

Stewart Canyon, and Elkheart Cliffs relay ramps (Figs. 12A, 12B, and 16). The ramps

have steeply dipping bounding faults that either connect or appear to connect at depth.

The differences among these three relay ramps are described below.

The Glendale relay ramp (Figs. 12A, 12B, and 16) is bounded by two faults that

and interpreted to connect at a depth of 2,135 m (7,000 ft) to the north and a depth of

-2,745 m (9,000 ft) to the south (Figs. 12 A, 12B, and 16 and Plate 4). The bounding

faults are steeply dipping and planar. These bounding faults connect at the surface at the

structurally lowest point of the ramp suggesting the ramp is base breached. A cross fault

cuts the relay ramp and is bounded by the ramp bounding faults (Fig. 16 and Plate 4).

This straight and planar cross fault is interpreted to connect with the easterly bounding

fault at the surface as well as at depth. The connection is interpreted to occur at a depth

o f930 m (3,050 ft) along cross section L-L’ (Fig. 16 and Plate 4).

Within the Glendale relay ramp, map patterns and cross sections show a north

plunging anticline oriented 3°, N59°E (Figs. 14B and 16 and Plate 4). This fold is

classified as a gentle, subhorizonal, upright fold. However, to the east and west of the

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 34 Stewart Canyon overlap zone, the stratigraphie units are gently west dipping (3-7°)

suggesting that the anticline is restricted to the ramp. Ramp-restricted folds, such as this

one, have rarely been described because detailed studies of multipartite linkage zones

with relay ramps has not be done.

The Stewart Canyon relay ramp is a breached ramp with both a base and a top

breach. A top breach means that the bounding faults of the relay ramp coimect at the

structurally high part of the relay ramp. The two bounding faults appear to connect at the

surface on both the north and south (Figs. 12A and 12B). However, the western

bounding fault is buried by Quaternary alluvium. The faults are interpreted to be planar

and appear to connect at depth (Fig. 16 and Plate 4). Two cross faults also cut the ramp

(Figs. 12A and 12B). These faults terminate at the bounding faults. No fold in addition

to the ramp monocline was observed within this relay ramp, but the entire relay ramp is

not exposed.

The easternmost portion of the Stewart Canyon overlap zone contains three faults

(Figs. 12A and 12B). The two eastern and subparallel faults have attitudes of

approximately N40°E, ~85°NW (faults U’ and T’ on Figs. 12A and 12B). These faults

juxtapose Jurassic Navajo Sandstone and the Jurassic Co-op Creek Member of the

Carmel Formation at the surface for a total of-100 m (-330 ft) of stratigraphie separation

(Plate 2). The easternmost faults connect in map view and are interpreted to connect in

cross section Y-Y’ at a depth o f 762 m (2,500 ft) (Fig. 16 and Plate 4). The third fault is

orientated at a high angle to the main strand of the Sevier fault; it strikes -N30°W and

dips -75°NE. This fault places the Jurassic Crystal Creek Member of the Carmel

Formation next to the Jurassic Winsor Member with 183 m (600 ft) of stratigraphie

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 35 separation (cross section E-E' on Plate 4). This fault is interpreted to connect with the

Sevier main strand at a depth of 335 m (1,100 A) on cross section E-E’ (Fig. 16 and Plate

4). The total amount of stratigraphie separation across the Sevier fault zone near cross

section E to E' is 1,044 m (3,425 ft) (Plate 4).

Faults U’ and T’ overly in map view with the western fault (V’) of the Glendale

relay ramp (Figs. 12A and 12B). Between these faults is a relay ramp here named the

Elkheart Cliffs relay ramp. The bounding faults do not link in map view, however, they

may link at depth (Figs. 12A, 12B, and 16). No fold besides the ramp monocline was

observed within this ramp.

Map patterns indicate that the faults in the central domain connect at the surface

(Plate 2 and Figs. 12A and 12B). Cross section construction and restoration suggests that

faults within the Stewart Canyon overlap zone are interpreted to connect in the

subsurface (Fig. 16 and Plate 4). However, the eastern two faults do not connect in map

view with the western faults. Cross section analysis reveals that these faults are

interpreted to appear to connect at depth (Fig. 16 and Plate 4).

Spencer Bench Domain

The Spencer Bench domain is located between the conununity of Glendale and

the northern boundary of the mapped area (37°19'36"N to 37°23'29"N) (Fig. 12A). This

domain exposes rocks that range in age from Cretaceous to (Quaternary (Plate I). An

offset (Quaternary (-570 ka) basalt flow is exposed on and near Black Mountain (Plate 1

and Fig. 12A). However, no (Quaternary sediments or sedimentary rocks are offset (Plate

1).

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 36 In contrast to the other domains, in the Spencer Bench domain the Sevier fault is

generally a single strand fault with hanging wall faults, not a complex splay zone. The

main strand of the Sevier fault, labeled fault E on Figure 12 A, strikes -N40°E and dips

76-81°W. The total amount of stratigraphie offset across the four cross sections varies

between 472 and 869 m (1,550 and 2,850 ft) within the Spencer Bench domain (cross

sections A-A* through D-D' on Plate 4). The greatest amount of total stratigr^hic

separation across the Sevier fault zone is within the splay zone near cross section C-C'

(Fig. 12A and Plate 4). Here the central Sevier fault zone has 518 to 640 m (1,700 to

2,100 ft) of stratigr^hic separation. The stratigraphie separation decreases to the north

and south from C-C’.

Eight faults, fault group C, are exposed to the northeast of Glendale (Fig. 12 A).

The faults within group C can be divided into a north-northwest striking fault set and a

north-northeast fault set with one cross fault. The dips are generally 58-81 °W. The total

amount of stratigraphie offset across these faults is 543 m (1,780 ft) near cross section D-

D’ (Plate 4). The western splay connects with the main strand of the Sevier fault at the

surface (cross section D-D' on Plate 4). However, the westernmost fault in the group,

called fault C , dips east. Fault C strikes N12°W and dips 73°E. It is interpreted to

connect at depth with a west-dipping fault (cross section D-D' on Plate 4). All the faults

are generally planar. Cross cutting fault relationships can be seen on the fault trace map

(Fig. 12A). Fault E, the main strand, is cut by fault R’ (Fig. 12A).

Fault set D lies just to the west of a bend in the main strand of the Sevier fault and

comprises five hanging wall faults: three west-dipping, one east to southeast dipping, and

one non-planar fault (inferred to restore cross sections) with northwest and north-dipping

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 37 sections (O') (Fig. 12 A). The east-dipping fault and north-dipping fault section are buried

by Quaternary alluvium (Fig. 12A and Plate 1), but are geometrically required to explain

outcrop patterns. The west-dipping faults generally strike N2S**E and dip between 79**

and 8S**W. The total stratignq)hic separation along these faults is 175 m (575 ft) (cross

section C-C on Plate 4). The three western faults are interpreted to connect at depth

(cross section C-C' on Plate 4). The western faults also appear to connect with fault D’ at

the surface (Fig. 12A and cross section C-C' on Plate 4). In map view, the generally

north-south striking faults qxpear to terminate at the generally east-west striking fault

(Fig. 12A). Thus, this east-west striking fault may be a transfer fault. The fourth fault

from the west. O’, is an east-dipping fault that is buried, but has an inferred attitude of

N25**W, 80**E. It has a stratigraphie separation of 130 m (425 ft). This fault is interpreted

to connect at a depth with the west-dipping strand to the east (cross section C-C on Plate

4). Set D faults offset units as young as the Cretaceous Kaiparowits/Wahweap at the

surface (cross section C-C on Plate 4). The total stratigraphie separation for fault group

D is 503 m (1,650 ft).

The main strand of the Sevier fault is the easternmost fault near fault set D, fault

E (Fig. 12A and cross section C-C on Plate 4). Between cross section D-D’ and C-C’,

the Sevier fault forms a large bend with strikes, which vary from south to north; N70°E,

N20**E, and N5‘*E. The stratigraphie separation is 549 m (1,800 ft). The main strand of

the Sevier fault is interpreted to connect at depth with both the west-dipping and east-

dipping faults of fault set D at (cross section C-C on Plate 4). The total stratigraphie

separation o f fault set D and the main strand is 869 m (2,850 ft). North of fault set D, the

Sevier fault strikes ~N35‘*E and dips 76®W and continues this pattern north to fault set F.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 38 In this domain, the Sevier fault cuts Quaternary (-570 ka) basalt Four fault

strands, including the main strand, cut the Black Moimtain basalt flow (Plate 1 and Fig.

12A). Three of these faults, called fault set F, strike N44®E. One dips 82®SE and the

other two dip -80® NW. The fourth fault is a cross fault with an attitude of N82®E, 80®N

and -1 meter (-3 feet) of stratigraphie separation. The total amount of post basalt

stratigr%q)hic offset across set F faults is -3 m (-10 feet).

Folding within the Study Area

Along the central Sevier fault, the pre-Claron Formation units have a general

north-northwest dip (1-19®) except in fault blocks or close to faults. Cross section

construction and restoration requires that both the hanging wall and footwall to have

undergone some normal fault drag. The cross sections (A-A', B-B', and C-C) show an

anticline in the hanging wall of the main strand of the Sevier fault. In addition to this

anticline, several regional folds exist near the study area (Fig. 7). In map view, the

Claron Formation generally dips between 3 and 8® NE. However, this unit also has

undergone some fault drag (cross section A-A'). The (Quaternary conglomerate units all

have dips o f <10®E to NE.

The other folds are located within relay ramps, which are interpreted to be fault

controlled. The following interpretations are based on combining the structure contour

and map data from Doelling et al. (1989) and Davis (1999) and field data collected for

this study.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. CHAPTER 6

VIRGIN RIVER DEPOSITS

Introduction

The objective of this section is to provide data and interpretations of the fluvial

deposits and how they relate to the central Sevier fault. These fluvial deposits provide

information to help (1) understand the evolution of the East Fork of the Virgin River, and

most importantly, (2) place age controls on faulting or tectonic deformation in the area.

From here on, I will refer to the East Fork of the Virgin River as the Virgin River.

Three Quaternary fluvial deposits of different ages crop out along the central

Sevier fault (Plates 1 and 2). The relative ages were based on the stratigraphie position of

the deposits in the landscape. Assuming the stream has downcut through time, older

deposits are exposed at higher elevations 1,939 - 1,963 m (6,360 - 6,440 ft). The oldest

deposits are conglomerates that crop out 8 km (5 mi) north of Glendale (Qcgl). This

conglomerate is older than the -570 ka basalt, which unconformably overlies the

conglomerate. The intermediate-aged Quaternary fluvial unit (Qcg) is conglomerate that

crops out only south of the Hidden Lake spring system (Plates I and 2). This unit is

located topographically lower, 1,707 - 1,865 m (5,600 - 6,120 ft), than the Qcgl, but

higher than the modem alluvium and stream channel (Fig. 15). The Qcg is also

unconformably overlain by a basalt flow southwest of Glendale (Fig. 15 and Plates 1 and

2). The modem channel is subparallel to the Sevier fault and lies along the fault trace

39

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 40 south of Glendale and -0.5 km (-1,500 ft) west of the fault, north of Glendale. The

Quaternary alluvium and modem channel are younger than the basalt because of the lack

of basalt flows on top of any of the deposits. The deposits contain basalt clasts and

grains.

Modem Fluvial Deposits

The mapped modem Virgin River channel deposits have clasts smaller than four

centimeters in diameter and are dominantly sand and gravel with few sandbars. The

channel deposits become finer grained to the south. However, some large (>1 meter)

clasts occur locally in the south. These clasts are probably evidence of flooding. The

Sevier fault zone does not visibly offset the modem fluvial deposits.

The flood plain of the Virgin River is -610 m (-2,000 ft) across in the south near

37“I5’50"N to 37®20’30"N and <152 m (<500 ft) in the north near 37°20'30"N to

37°23'36"N (Fig. 15). This difference in flood plain width may be caused by the change

in the gradient of the stream, a change in bedrock lithology, or an increase in discharge

from the north to the south. The gradient of the Virgin River is steeper in the north than

in the south (Fig. 15). North of Dry Wash, the gradient of the Virgin River is 0.026, and

to the south of Dry Wash the gradient is 0.011. The bedrock lithology changes from the

more resistant Cretaceous Kaiparowits/Wahweap Sandstone to the north to the less

resistant mudstones and shales of the Tropic Shale, Dakota Sandstone, and Carmel

Formation to the south. The amount of discharge is greater in the south than the north

particularly because the Dry Wash channel contributes to the Virgin River north of

Glendale (approximately 37°20'30"N) (Fig. 15 and Plates 1 and 2). Therefore, changes in

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 41 gradient, bedrock, and discharge may all impact the width of the Virgin River flood plain.

Knowing and being able to predict the width of the flood plain is important because

several people live along the Virgin River.

Young Fluvial Conglomerate

The young fluvial conglomerate (Qcg) has a lithologie make-up, which differs

from the modem stream deposits (Fig. 17A). The conglomerate clasts comprise

limestone, dolostone, chert, sandstone, basalt, and petrified wood. Generally, the clasts

are well rounded and range from one centimeter to greater than a meter in diameter and

graded bedding is present locally. However, the petrified wood is generally angular to

subangular. The petrified wood is dark brown to black and appears to be agatized. The

matrix consists of an indurated lithic sand. Discontinuous layers of lithic sandstone also

occur throughout the conglomerate unit (Fig. 18A). The discontinuous layers of

sandstone contain cross beds of the type (simple trough cross-stratification) seen in

sandbars or migrating ripples (Prothero and Schwab, 1996). The cross beds, graded

bedding, and channels (Fig. 18B) suggest that the Quaternary conglomerate has a fluvial

origin.

From north to south, the gently dipping surfaces of the outcrops become

topographically lower. These surfaces could have formed by (1) erosion, (2) as part of a

floodplain, or (3) as stream terraces. (1) These surfaces may or may not have formed by

erosion. If the surfaces were formed by erosion, the surrounding bedrock would have the

same erosional characteristics. The bedrock contains no evidence of this type of erosion.

(2) The surfaces did not form during flooding because floodplains typically are composed

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 42 of fine-grained sediments such as muds and silts. The deposits under these surfaces are

sand to boulder sized. (3) Stream terraces form when the base level of an active stream

drops. As the base level drops the steam begins to down cut leaving the old streambed at

a higher elevation than the present stream elevation. The older terraces become inactive

benches (Keller and Pinter, 1996; Prothero and Schwab, 1996). I interpret these

conglomerate units to be old stream terraces, which may have been eroded. The paleo-

Virgin River must have flowed from the north to the south primarily because the terraces

become topographically lower toward the south (Fig. 15).

Older Fluvial Deposits

The older conglomerate (Qcgl) crops out in one location east of U.S. 89 and in

three locations west of U.S. 89 (Fig. 15 and Plate 1). This may be because the older

channel is farther west of the U.S. 89 or other outcrops have been eroded or not

preserved.

Limestone, dolostone, chert, sandstone, and basalt are the dominant clasts in the

conglomerate. Generally, the clasts are well rounded and range from one centimeter to

greater than a meter in diameter and graded bedding is present locally. The matrix

consists of a sandy siltstone.

The main lithologie difference between the older fluvial deposits and the younger

fluvial deposits is the absence of petrified wood clasts in the older conglomerate. The

absence of the petrified wood may be due to the fact that this unit unconformably overlies

the Kaiparowits/Wahweap undivided Formation (Kkw) (Fig. 9). Therefore, the Virgin

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 43 River at the time of the deposition of the older conglomerate did not erode layers

containing petrified wood within the Kkw.

Three additional differences between the older conglomerate and the other

conglomerates help to distinguish the units. First, the older conglomerate is generally 73

m (240 ft) above the modem stream channel (Fig. 15), whereas the younger conglomerate

(Qcg) is typically found between 12 - 37 m (40 - 120 ft) above the present Virgin River

channel (Fig. 15). Another difference is that Qcgl is not on the same grade as the Qcg

indicating that they are separate deposits. Lastly, the older conglomerate has a finer

grained matrix with 0.6 m (2 ft) clasts generally of basalt (Fig. 17B).

Both the older and the younger conglomerates are capped by basalt flows (Fig. 15

and Plates 1 and 2). The older conglomerate is capped by the -570 ka basalt flow

indicating that the conglomerate is older than 570 ka. In addition, both conglomerates

contain clasts of basalt from other flows.

Interpretations

The bedding within the conglomerate and the original slope was used to

determine if this unit was tectonically tilted. Beds formed by a south flowing stream

generally dip gently to the south. The top surfaces of the exposures dip gently to the

northeast and east (<10°NE and <10°E) indicating that they may be tectonically tilted.

Two possible sources of the petrified wood clasts in the Qcg are the Cretaceous

Kaiparowits/Wahweap formations undivided or the Triassic Chinie Formation. Petrified

wood was found within the Kaiparowits/Wahweap formations in the mapped area. The

flow of the paleo-Virgin River was most likely from north to the south. Exposures of the

Triassic Chinie Formation occur only to the south of the outcrops of Qcg and

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 44 Kaiparowits/Wahweap Formation is exposed near the outcrops and to the north. Also the

petrified wood found in the conglomerate is darker in color and highly agatized. The

petrified wood found in the Chinie Formation is generally a lighter color and not as

agatized. However, petrified wood crops out in the Chinie Formation near Zion National

Park, but is darker in color and highly agatized. I suggest that the petrified wood in the

Qcg is derived from the Kaiparowits/Wahweap formation.

Discussion of the Virgin River Deposits

Modem Virgin River deposits have an overall smaller grain size than Qcg. This

suggests that the present day flow is slower than the flow in the past. However, the grain

size of the Qcgl and the modem fluvial deposits are relatively similar indicating that the

flow of the river that deposited Qcgl and the modem flow rates are similar. Both are

faster flow rates than the flow that deposited Qcg. Additionally, the original surface of

the conglomerates was slightly to the south.

Faulting within the Quatemary Fluvial Deposits

None of the Quatemary conglomerate units within the mapped area are cut by a

fault at the surface. This observation suggests that either surface mpturcs are older than

the conglomerate units along this section of the fault, or these deposits are not located on

the faults. The later scenario is more probable because none of the fault strands in the

area of interest project under the conglomerate. However, the 570 ka basalt is cut by

faults and several of the conglomerate outcrops are tilted. The tilting was probably

caused by faulting.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 45 The active channel and floodplain of the Virgin River are located in the hanging

wall just west of the main strand of the Sevier fault in the southern part of the area (Plate

2). This geometry fits the Leeder and Gawthorpe (1987) model of a continental half-

graben with axial trough drainage. Both the model and the Sevier fault have footwall

uplands with incised drainages. In both the model and the study area these drainages

seem to flow nearly perpendicular to the main channel and the fault trace. Both have

meandering streams or rivers that flow approximately parallel to the fault trace. The

model suggests that as the faults continues to slip, sediments progressively onlap hanging

wall units during extension as the stream migrates closer to the fault trace (Leeder and

Gawthorpe, 1987). The southern portion of the Virgin River and Sevier fault fit this

model. No abandoned meander belts occur in the modem alluvium to suggest

progressive onlap and tilting during fault movement. However, all of the Qcg surfaces

dip approximately 10°NE to 10“E. Because the fault dips west, when it moves, the

sediments in the hanging wall are tilted toward the fault or to the east. These data suggest

possible tilting during fault movement because the Sevier fault zone is generally to the

east of the Virgin River.

In the northem portion of the area, the Virgin River and the Sevier fault lie farther

apart. Near Glendale, the main strand of the Sevier fault bends to the northeast and the

Virgin River continues to flow south and does not follow the trace of the fault. However,

north of Glendale, a large tributary. Dry Wash Canyon (Plates 1 and 2,) to the Virgin

River follows the Sevier fault as far north as Black Mountain (Fig. 1). The Quatemary

alluvium within Dry Wash Canyon covers a possible transfer fault of set D (Fig. 12).

Several of the Qcg outcrops are exposed in Dry Wash Canyon suggesting that this canyon

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 46 was a tributary to the Virgin River during the time of Qcg deposition (Plates 1 and 2 and

Fig. 15). This implies that the Virgin River followed the trace of the Sevier fault in Dry

Wash Canyon in the past, but a new drainage has formed to the west.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. CHAPTER?

STRUCTURAL INTERPRETATIONS

Two different types of linkage formed near the Orderville relay ramp and within

the Stewart Canyon overlap zone. Within these linkage sites, relay ramps with fault

parallel folds formed. Together these linkage sites form the Orderville geometric

boundary, or salient. The following sections will discuss (1) the linkage zones and the

structures formed within them, (2) stages of linkage, and (3) slip rates along the central

Sevier fault.

Linkage

Examples of Fault Linkage

Originally isolated faults are interpreted to have linked between Orderville and

the Stewart Canyon overlap zone to form the central Sevier fault (Figs. 12A and 12B).

Total displacement vs. distance diagrams were created from where the central Sevier

fault is generally a single strand in the north to the southern boundary of the mapped area

(Fig. 19). The total displacement decreases within the linkage zones (Fig. 4). Within

these linkage zones, two different styles of linkage were identified. Near Orderville, fault

capture is suggested. In the Stewart Canyon overlap zone, linkage of overlapping faults

is suggested. However, several fault tips overlap, and thus this fault zone is multipartite.

47

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 48 more complex than typically described (i.e., Ferrill et al., 1999). The two different styles

are discussed below.

Fault Capture in the Orderville Area

Near Orderville, the central Sevier fault linked by fault capture. On displacement

vs. distance diagrams the total displacement decreases near the Orderville relay ramp

(Fig. 20A). Displacement decreases are expected to occur near linkage sites because

linkage zones are typically located near original fault tips where the displacement is less.

The pattern on the displacement vs. distance diagram for the Orderville area closely

resembles the pattern for the model of fault capture (Fig. 4C). Both diagrams have a

plateau in the area where the faults overlap. In this type of linkage, one fault has more or

more rapid slip than the other fault. The fault with more or more rapid slip is the fault

with the most displacement (Fig. 4C). In addition, the map fault trace pattern (Figs. 12A

and 12B) closely resembles the fault trace pattern for the model of fault capture (Fig. 4C).

Faults A and B (Figs. 12A and 12B), have nearly the same surface fault trace pattern as

the model surface fault trace pattern (Fig. 4C).

Different stages of linkage were identified by the mapped fault trace pattern. For

example, stage one, the least advanced stage of linkage, is indicated by two faults that are

separate and isolate. Stage two is indicated by the linkage of two faults at one linkage

site. Stage three, or the most advanced stage of linkage, occurs where two faults link at

two linkage sites. One linkage site is buried under the Quatemary alluvium to the north

of Orderville (Figs. 12A and 12B). The other linkage site is located where faults A’ and

B’ connect at the surface to the south (Figs. 12A and 12B). The Orderville area is in an

advanced stage of linkage via fault capture because the faults have two linkage sites.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 49 I suggest that the faults in the Orderville area linked by fault capture because their

initial fault spacing and orientations did not allow the fault tip stress fields to interact

until the fault with the greater amount of slip, the western fault, propagated to

approximately the center 1/3 of the eastern fault (Figs. 12A and 12B).

In the fault capture linkage area near Orderville, the Orderville relay ramp formed

(Figs. 12A and 12B) (Davis, 1999; this study). The Orderville relay formed in a fault

capture linkage situation, which is a type of hard linkage. The attitudes within the

Orderville relay ramp reveal the typical ramp monocline (Fig. S) as well as an additional

fold (discussed later). This ramp is also breached at both the top and bottom at the two

linkage sites (base and top breaches) (Figs. 12A and 12B).

Overlapping Faults in the Stewart Canyon Area

The Stewart Canyon overlap zone may be a strain transfer zone, as indicated by

four major linked overlapping faults. The Stewart Canyon overlap zone is more complex

than simple overlapping fault tip linkage. It is a strain transfer zone that involves at least

twelve faults, four sites of linkage, and three relay ramps. The zone is subdivided into

three domains: eastern, central, and western (Figs. 12A and 12B).

Eastern Stewart Canyon Overlap Zone

The eastern most fault (T’) is interpreted to be a splay of the next fault to the west

(U’) (Figs. 12A and 12B). To the north near cross section E-E , the northeast dipping

fault (Q’) is also interpreted to be a splay of fault U’ (Figs. 12A and 12B).

Farther to the north, faults U’ and S’, which connect in map view (Figs. 12A and

12B), form a fault slice. Fault S’ was once the main strand, however, the stress fields

changed and a new fault (U”) propagated, which became the main strand.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 50 I suggest that the eastern portion of the Stewart Canyon overlap zone is a relay

ramp and herein name it the Elkheart Cliffs relay ramp (Figs. 12A and 12B). This relay

ramp formed between overlapping fault tips. On displacement vs. distance diagrams, the

plotted data (Fig. 20D) closely resemble the shape predicted by the model for soft linkage

via overlapping faults (Fig. 4E). Additionally, the bounding faults do not link at the

surface suggesting soft linkage. However they may link at depth suggesting a hard

linkage situation (Plate 4). Thus, both hard and soft linkage occur along these two faults.

Central Stewart Canyon Overlap Zone

In the central portion of the Stewart Canyon overlap zone, the faults that form the

Glendale relay ramp are linked overlapping faults for several reasons (Figs. 12A and

12B). First, the displacement vs. distance profile (Fig. 20B), generally resembles the

overlapping fault tips model (Fig. 4B). However, an anomaly exists within the linkage

zone along the northern part of the linkage site (Fig. 20B). The location of the anomaly

corresponds to the location of fault D, a breakthrough or cross fault. Therefore, the

anomaly is most likely a direct result of that cross fault. Second, the mapped patterns for

the central Stewart Canyon overlap zone closely resemble the model fault trace map

patterns for overlapping faults (Figs. 4B and 20B). Both have overlapping faults that

connect at the surface. The faults in the central Stewart Canyon overlap zone and in the

models in map pattern overlap and connect (Figs. 4B, 12A, 12B, and 12B). In this case,

hard linkage occurred because the ramp is breached on the basal end. However, this area

also contains cross faults, which, require a combination of the overlapping and cross fault

models (Figs. 4B and 20B). The central Stewart Canyon overlap zone mapped fault trace

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 51 patterns are interpreted to have formed by a combination of the overlapping faults

(Taylor et al., 2(X)1) and breakthrough fault linkage models (Fig. 4).

The Glendale relay ramp (Figs. 12A and 12B) formed between overlapping faults

(V’ and W’) within the central portion of the Stewart Canyon overlap zone, with hard

linkage and a base breach (physical connection of the faults at the structurally lowest

point of the relay ramp). The attitudes of bedding within the ramp reveal the typical

ramp monocline as well as an additional fold (discussed later). The Glendale relay ramp

also contains a cross fault that may ultimately break down the relay ramp by continued

slip along the faults.

Western Stewart Canyon Overlap Zone

The western faults linked overlapping faults with hard linkage at two sites, one to

the north and one to the south (Figs. 12A and 12B). The pattern for the western portion

of the Stewart Canyon overlap zone has a lens shape, like in the overlapping fault tips

model. The displacement vs. distance diagram for the pattern of the Stewart Canyon

relay ramp area (Figs. 12A and 12B) most closely matches the model of overlapping fault

tips (Fig. 20B). However, the fault trace map shows cross faults. In addition, an

anomaly or step is present on the displacement vs. distance diagram (Fig. 20B). I suggest

that these faults formed after linkage because of the cross cutting relationships of the

faults (Figs. 12A and 12B). The cross faults terminate sharply at the bounding faults with

a high angle of intersection rather than the 30° typical of cross faults. Therefore, in this

area late cross faults and overlapping faults, combined, formed the linkage site.

I suggest that these faults linked across overlapping fault tips because of fault

spacing and fault tip stress field interaction (Chapter 2). The fault spacing was small

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 52 enough to allow both stress fields to interact with each other at the same time. Linkage

by overlapping faults may also be an issue of propagation rates. For example, both faults

have similar propagation rates and displacement, so when the stress fields interact, the

faults link via overlapping type of linkage rather than fault capture.

Folds Within Relay Ramps

The Orderville relay ramp contains a fault-parallel syncline and the Glendale

relay ramp contains a fault-parallel anticline (Figs. 13 and 14). The rocks outside of the

bounding faults are generally flat lying or gently west dipping. If these folds formed by

drag along the fault, the beds would "roll-over " and dip more steeply towards the fault or

exhibit normal drag and dip steeply away from the fault. If the folds are fault

propagation folds, the beds in the footwall would be deformed. However, neither

geometry is the case. I suggest that the folds formed as a result of a downward decrease

in space between the bounding faults (Figs. 13 and 16).

A downward decrease in space between two relay ramp bounding faults occurs

because the faults merge at depth in the linkage zone (Fig. 21). At the moment when the

ramp-bounding faults link, the rocks within the relay ramp fill the space between the

faults (Fig. 21). However, if fault B or both fault A and B move (Fig. 21), the rocks

within the fault-bounded wedge-shaped block are dropped down into a smaller space.

This decrease requires the rocks to shorten normal to the bounding faults, which can be

accomplished through a fault-parallel fold (Fig. 21). The more the faults slip, the tighter

the fold becomes within the fault block. Therefore, fault-parallel folds within relay

ramps suggest that faults bounding the relay ramp approach each other and/or merge with

depth and the bounding faults have had significant slip after formation of the relay ramp.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 53 The Glendale relay ramp contains a generally fault-parallel anticline. On Figs.

12A and 12B, notice that the axial trace of the fold bends and changes strike. This

change suggests that the east-bounding fault is not planar at depth, but changes strike

and/or dip with depth causing the axial surface of the fold to bend.

Because most of the Stewart Canyon relay ramp is buried under Quatemary

alluvium, it is not know whether or not it contains a fold. Where the relay ramp is

exposed, attitudes were difficult to collect because the Cretaceous Dakota Sandstone is a

poorly exposed dominantly muddy shale. The attitudes that were collected are on coal

seams. Based on mapping and cross section construction analyses, I suggest that the

Stewart Canyon relay ramp is a zone of fault linkage.

Summary

The Orderville, Glendale, Stewart Canyon, and Elkheart Cliffs relay ramps have

formed between overlapping faults. However, the fault capture subtype of overlapping

linkage occurred near the Orderville relay ramp (Fig. 20). Breakthrough cross faults are

important in the Glendale and Stewart Canyon relay ramps. These relay ramps were all

identified by fault patterns and/or the attitudes of bedding between the bounding faults.

The unique pattern of faults and relay ramps in the Stewart Canyon overlap zone may

represent a triple ramp linkage situation where one ramp forms and with continued slip,

another ramp forms while the first formed relay ramp becomes both base and top

breached. Later another ramp forms, while the second ramp becomes base breached.

Stages of Relay Ramp Development in the Stewart Canyon Overlap Zone

The Stewart Canyon relay ramp is interpreted to have formed first, next the

Glendale relay ramp, and finally the Elkheart Cliffs relay ramp. These stages of

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 54 development are based on (1) cross cutting fault trace patterns, (2) whether the relay

ramp is breached or not (if the ramp is breached at how many sites), and (3) the number

of cross faults, which ultimately lead to the break down the ramp. Breaching of the relay

ramp suggests that the ramp is in its final stage of development and will begin to

breakdown (Ferrill et al., 1999).

The Stewart Canyon relay ramp probably formed first because it has both a head

and top breach. No anticline or syncline was found within the ramp suggesting that the

ramp formed prior to the other ramps (i.e., the ramp couldn’t have formed through a pre­

existing ramp with a fault parallel fold). However, there may be a fold, but most of the

ramp is covered by Quaternary alluvium. The Glendale relay ramp probably formed

next. This ramp has a base breach and a fault-parallel anticline. With additional slip

along the bounding faults, the ramp will probably form a top breach and the fault-parallel

anticline will become tighter. I suggest that the Elkheart Cliffs relay ramp formed last.

This is interpreted because (1) the ramp is not breached, (2) there is no evidence of a

fault-parallel fold, and (3) the bounding faults may connect at depth but not at the

surface.

Stages of Linkage along the Central Sevier fault

The new data and analyses provided here allow recognition of five stages of

linkage along the central Sevier fault.

During stage 1, the central Sevier fault inititated as six isolated faults: Spencer

Bench segment (U’), X’, V’, W’, Y’, and Mt. Carmel segment (Z’) (Fig. 22A). A

possible relay ramp developed between faults X’, V’, and +/- W’ (Figs. 22A and 23A).

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 55 This relay ramp is a combination of the Glendale and Stewart Canyon relay ramps. This

stage of development is soft linkage.

During stage 2, near the community of Orderville, fault Y’ continued to propagate

south and fault Z (Mt. Carmel segment) continued to propagate north (Fig. 22B). As

these faults overlapped, they formed the Orderville relay ramp. After this ramp formed,

hard linkage occurred. The relay ramp is double breached with a top and base breach.

The syncline formed just before or after hard linkage. To the north, faults X and V’

continue to propagate ultimately linking by hard linkage to form a breached relay ramp

(Figs. 22B and 23B). This breached ramp is the combined Glendale and Stewart Canyon

relay ramps. While faults X and V’ link, fault W’ propagated and divided the area

between faults X’ and V’ into the Glendale and Stewart Canyon relay ramps (Figs. 22B

and 23B). An anticline also formed within the Glendale relay ramp. Soft linkage

occurred between faults X and W’ as well as W’ and V’ (Figs. 22B and 23B). To the

northeast, fault U’ (Spencer Bench segment) continued to propagate south. Between

faults U’ and V’, the Elkheart Cliffs relay ramp developed soft linkage (Figs. 22B and

23B).

During stage 3, fault W’ propagated to the north and linked by hard linkage with

fault V’ (Figs. 22C and 23C). This hard linkage site is a base breach of the Glendale

relay ramp and isolates the Stewart Canyon relay ramp. In this stage the Stewart Canyon

relay ramp is bound by three faults X’, V’ and W’ (Figs. 22C and 23C). Cross faults

continued to develop in the Stewart Canyon relay ramp (Figs. 22C and 23C). To the east,

fault U’ continued to propagate to the south and the Elkheart Cliffs relay ramp developed

further by soft linkage (Figs. 22C and 23C).

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 56 During stage 4, a cross fault developed and cut the Glendale relay ramp and

anticline (Figs. 22D and 23D). Fault W’ propagated and linked with fault Z to form a

top breach on the Stewart Canyon relay ramp (Figs. 22D and 23D). In addition, a cross

fault cutting the Stewart Canyon relay ramp formed. These cross faults have an atypical

orientation of -90° to faults X and V’ and subparallel (20°) to V’ (Figs. 22D and 23D).

This non-standard orientation probably formed because the Stewart Canyon relay ramp is

has three bounding faults rather than the typical two bounding faults. To the east, fault

U' continues to propagate farther south (Figs. 22D and 23D). Fault U’ may possibly link

at depth with fault V’ to form a hard linkage site (Figs. 22D and 23D). Farther to the

east, fault T’ begins to splay from fault U’ and propagates north (Figs. 22D and 23D).

During the final stage (stage 5), cross and breakthrough faults continued to

develop and fault T’ propagated farther northward (Figs. 22E and 23E). Stage 5 is also

the modem fault trace map. Several of the faults have been buried by Quatemary

alluvium.

Segments and Segment Boundaries

Typically geometric salients and segment boundaries form at linkage sites or

along linkage zones between reentrants. The salient along the central Sevier fault

between Orderville and Glendale is a large geometric bend containing linkage zones

called the Orderville salient (Fig. 22E). This salient contains four linkage sites and the

Orderville, Glendale, Stewart Canyon, and Elkheart Cliffs relay ramps. A salient or

geometric segment boundary is a zone rather than a particular point (Fig. 22E). The

Orderville salient lies between the Mt. Carmel segment to the south and the Spencer

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 57 Bench segment to the north. Both the Mt. Carmel and the Spencer Bench segments are

reentrants to the Orderville salient (Fig. 22E).

Age Constraints and Slip Rates on the Sevier Fault

Different portions of the Sevier fault have different ages of last known surface

rupture. The northem portion is geographically defined from Panguitch south to Black

Mountain and is called the Spencer Bench geometric segment. The central portion is

defined from Black Mountain south to Orderville and is called the Orderville salient. The

southem portion is defined from Orderville south to Yellow Jacket Spring called the Mt.

Carmel geometric segment (Fig. 1).

Slip Rates

Preliminary slip rate studies have been done for the northem portion of the

Spencer Bench geometric segment (Hecker, 1993). Hecker (1993) states that the slip rate

is 0.360 mm/year (0.014 in/year) near Red Canyon (Fig. 1). This slip rate was calculated

for the last 560 ka using the post Red Canyon andésite slip and the age of the andésite

(Best et al., 1980).

No slip calculations have been published for the central and southem portions of

the Sevier fault. This study introduces new data and calculations of slip rates along the

central portion of the Sevier fault.

Pre- and post-basalt slip rates for the central Sevier fault are useful to estimate (I)

the age of initiation, (2) whether the slip rate has been constant through time, (3) possibly

whether the region is at seismic risk in the future. To determine the age of initiation and

pre-basalt slip rate, the post-basalt slip rate must be determined.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 58 Post-basalt slip rates were calculated in the Spencer Bench domain near Black

Mountain. Here, the Sevier fault displaced 570 ka basalt 10 meters (33 ft) (Best et al.,

1980; this study) (Figs. 12A and 12B). Assuming the slip rate was constant since 570 ka,

an average yearly slip rate for the Spencer Bench segment can be calculated from the

equation; total offset of the basalt divided by the age of the basalt. Using this formula the

slip rate for the last 570 ka is 0.0180 mm/year (0.0007 in/year).

After calculating the post-basalt offset the initiation of the Sevier fault can be

determined by using the formula r = d/t where t = time, d = distance or stratigraphie

separation, and r = post-basalt offset slip rate. The d used was the greatest amount of

stratigraphie separation along the Spencer Bench segment of the main strand of the

central Sevier fault. This calculation assumes that the fault has had a constant slip rate

throughout its entire history. By using this formula, central Sevier fault initiated 44 Ma.

However, the 44 Ma age is older than expected because entire Tertiary Claron Formation

is cut by the Sevier fault. Furthermore, the Claron Formation does not contain any

evidence of syn-faulting deposition. Therefore, it is required that the initiation of the

central Sevier fault is post-Claron. The top of the Claron Formation is -30 Ma based on

a date for a tuffaceous sandstone within the upper Claron (Lundin, 1989).

The pre-basalt slip rate was calculated based on cross section construction and

using the formula r = d/t. Where d = stratigraphie separation and t = age of the top of the

Claron Formation (30 Ma). The total offset was calculated on the top of the Jurassic

Navajo Sandstone. This method of determining the total offset assumes that the slip rate

has been constant through pre-basalt time and that slip initiated immediately after the

cessation of deposition of the Tertiary Claron Formation. Thus, the minimum slip rate

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 59 for the pre-basalt offset is 0.0229 mm/year (0.0009 in/year). If slip initiated significantly

later than when deposition of Claron Formation ceased, the rate would be faster.

So, based on the above calculations, two interpretations can be made. (1) The

southem section of the central portion (Spencer Bench segment) of the Sevier fault has

become less active through time. (2) The Red Canyon area is a separate geometric

segment from the Spencer Bench segment with a different slip rate. Either interpretation

agrees with calculations of decreasing slip rates along the Hurricane fault with time

(Lund et al., 2002).

The post-basalt slip rate along the Spencer Bench segment is significantly less

than the Quatemary slip rates for the northem portion of the Sevier fault. This suggests

that the northem portion of the fault has been more active in the last 570 ka. These slip

rates only suggest that the central Sevier fault has been active recently and may produce

seismicity in the future.

Fault Scarps

Several fault scarps occur along the Sevier fault. In the northem portion, near

Red Canyon, Hecker (1993) documented scarps, which based on scarp morphology may

be as young as Holocene in age (Fig. 1 ). However, at the westem entrance to Red

Canyon, the Sevier fault has a total of -900 m (-2,950 ft) of offset (Lundin, 1989). Here

a 560 ka andésite is juxtaposed against the Tertiary Claron Formation with -200 m (-660

ft) of offset (Best et al., 1980; Hecker, 1993). Christenson and Nava (1992) reported

fault scarps in Quatemary alluvial fans to the east of Panguitch (Fig. 1). Along the

central Sevier fault, no faults cut the Quatemary sediments. However, the 570 ka basalt

on Black Mountain is offset (Figs. 12 and 1). Stream terraces of late Tertiary/Quatemary

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 60 age are tectonically tilted. Along the southem portion of the Sevier fault, south of Mt.

Carmel Junction, Doelling et al. (1989) mapped alluvial gravels, eolian sands, and mixed

eolian sands and alluvium cut by strands of the Sevier fault. The above fault scarp data

suggest that seismic activity along the Sevier fault is youngest to the north and oldest in

the central portion.

Both the slip rate and the ages of scarps suggest that the most recently active

portion of the Sevier fault is the northem portion followed by the southem portion then

the central portion. However, further studies and additional mapping will help constrain

these relationships.

Folding along the Central Sevier Fault

By using the collected field data, I suggest that the folds in question do affect the

Tertiary Claron Formation and therefore are a result of early Laramide with additional

Cenozoic normal fault folding. The trace of the axial surface of the westem syncline

(Fig. 7) (Doelling et al., 1989) would be approximately along U.S. 89 (Fig. 7). The field

area lies in the east limb of this fold, thus dips in the area should be north to northwest.

The trace of the Harris Mountain anticlinal axial surface to the south (Doelling et al.,

1989) crosses the trace of the Sevier fault near Orderville and continue north in the

footwall (Fig. 7). The structure contours and attitudes of bedding suggest that these fold

pattems are possible.

The Tertiary Claron Formation has an ~10°NE dip and the Quatemary

conglomerate units have <10°E dips. This strata rotation was probably formed by

tectonic tilting or fault drag along the Sevier fault. This warping deformed the already

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 61 existing Laramide folds and the Tertiary Claron Formation. This interpretation helps

explain the north to northwest dipping pre- Tertiary Claron units and the east dips of the

Tertiary Claron and post- Tertiary Claron units.

Cross section restoration suggests that the units near the faults are folded into the

fault. This fold is parallel to the fault and is interpreted to be caused by Cenozoic normal

fault drag along the Sevier fault.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. NOTE TO USERS

Page(s) not included In the original manuscript are unavailable from the author or university. The manuscript was microfilmed as received.

62

This reproduction Is the best copy available.

UMI‘

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. CHAPTER 8

EARTHQUAKE AND LANDSLIDE HAZARDS

Earthquakes

The central Sevier fault is located in the southern part of the Intermountain

Seismic Belt (ISO), a tectonically active region with large (M 7.0+) earthquakes (Fig. 1).

This belt extends from northwestern Montana to southern Nevada and northern Arizona

and is characterized by shallow (<15 km focal depths) earthquakes (Smith and Sbar 1974;

Christenson and Nava, 1992; Mason, 1996). Large prehistoric earthquakes of M 7.0 -

7.5 occurred in southwestern Utah, and thus, may possibly occur in the future

(Christenson and Nava, 1992). Earthquakes that occurred prior to July 1962 are based on

felt reports (shaking felt and reported by humans) whereas earthquakes occurring after

July 1962 are based on instrumental data (Christenson and Nava, 1992). Historical

earthquakes in southwestern Utah have reached magnitudes of 6.0 - 6.5 (Christenson and

Nava, 1992). Some examples are the 1902 M 6.5 Richfield, 1902 M 6.3 Pine Valley, and

1921 M 6.3 Elsinore earthquakes (Christenson and Nava, 1992). Approximately 2,300

earthquakes >M 2.0 occurred in southwestern Utah since 1850 (Christenson and Nava,

1992), and approximately 20 earthquakes greater than M 4.0 occurred during this century

in southwestern Utah and northwestern Arizona (Christenson and Nava, 1992). A M 5.5

- 5.7 eanhquake was recorded in 1959 southeast of Kanab (Doelling et al., 1989;

Christenson and Nava, 1992). Christenson and Nava (1992) document an earthquake of

63

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 64 Mercalli Intensity VII (Richter magnitude of -5.7) near Kanab in 1887 and several M 2.0

- 4.0 earthquakes near the trace of the Sevier fault. However, it is difficult to determine

whether these earthquakes occurred along the Sevier fault zone because earthquake

locations and focal depth resolutions are poor. On January 1, 1924, a Mercalli Intensity

III earthquake was felt in the town of Orderville (Fig. 1) (Doelling et al., 1989). From

January 1, 1924, through November 27, 1927, fourteen aftershocks (Mercalli Intensity of

II to III) were also felt in Orderville (Doelling et al., 1989). Doelling et al. (1989)

suggest the January 1, 1924 earthquake and aftershocks were caused by movement along

the Sevier fault zone. However, no Holocene fault scarps or ground ruptures are

documented along the Sevier fault zone near Orderville, however deep earthquakes may

not cause surface rupture. The proximity of these reported earthquakes to the Sevier

fault, suggests that this fault may have Holocene activity.

The Sevier fault lacks evidence for Holocene surface rupture, but evidence for

late Quaternary surface rupture does exist (Jackson, 1990; Christenson and Nava, 1992;

Hecker, 1993). The Sevier fault offsets -570 ka basalt near Black Mountain (Plate 1)

(Best et al., 1980; this study), indicating surface rupture during the Quaternary. The lack

of documented Holocene surface ruptures suggests recurrence intervals are greater than

10 to 30 ka (Christenson and Nava, 1992). However, Jackson (1990) and Hecker (1993)

state that evidence for recent activity may be lacking or obscured because of active

colluvial slopes and thick vegetation. Also, erosion by the Virgin River may have

removed scarps. Christenson and Nava (1992) suggest relatively high late Quaternary

slip rates of 0.30 - 0.47 mm/year (0.012 - 0.019 in/year) that seem to contradict the lack

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 65 of evidence for Holocene surface rupture. Little information is known about the average

earthquake recurrence intervals along the Sevier fault in southwestern Utah.

Earthquakes of M 4.0 and higher generally cause a variety of earthquake and

associated hazards. Some of the hazards in the region include ground shaking, surface

rupture, liquefaction, flooding, and slope failures (rock falls, landslides, debris flows,

avalanches, etc.) (Christenson and Nava, 1992; Harty, 1992; Hecker, 1993; Lund, 1997;

Stewart et al., 1997a, 1997b; Reber et al., 2001). This study will focus on slope failures.

Landslides

A landslide is the downward movement of earth material along a distinct failure

surface. Active landslides can range in velocity from one meter per day to as fast as 300

km/hour (186 mi/hour) (Rahn, 1996). Landslides that no longer move are considered

inactive or stable (Rahn, 1996). This study provides new information on 14 landslides.

Identifying and Categorizing Landslides

Most of the landslides documented in southern Utah are at elevations greater than

1,829 m (6,000 ft) above sea level (Schroder, 1971) and are derived from the from the

Cretaceous Tropic Shale, Dakota Sandstone, and Jurassic Carmel, Triassic Chinle, and

Moenkopi Formations (Schroder, 1971; Doelling et al., 1989; Christenson and Nava,

1992; Harty, 1992; Lund, 1997; Stewart et al., 1997; Reber et al., 2(X)1). These

formations contain expandable clays such as montmorillonite (Mulvey, 1992). When

wet, these clays can expand up to 2,(X)0 times their original dry volume causing

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 66 landslides to occur within these high clay content collapsible soils and rock units

(Mulvey, 1992).

One natural trigger for many landslides is earthquakes. Earthquake induced

landslides typically occur on steep slopes (>25°) that have a relief greater than 152 m

(500 ft) (Keefer, 1984; Rahn, 1996). Typically, such slopes can be found along streams,

washes, or young faults. Failure occurs when the material is dislodged by earthquake

shaking. After failure, the slopes are then typically undercut by streams or active washes

(Keefer, 1984; Rahn, 1996). The amount of undercutting is dependent on how much and

what type of material fell. Also, the amount of undercutting is dependent on whether or

not the debris fell into the stream.

Naturally occurring landslides can also form by gravitational failure along

bedding planes of weak units (i.e., shales and mudstones) or clay rich soils. The

weakness and potential for failure can be increased by prolonged rain and freeze-thaw

cycles (Harty, 1992; Rahn, 1996).

Human induced landslides are also common. Two ways in which humans induce

landslides are ( 1 ) clear-cutting vegetation and (2) adding weight (i.e., building a

structure) on or above a potential or existing stable slide. Clear-cutting vegetation causes

a loss of roots, which lowers the shear strength of the soil or rock unit allowing it to

potentially fail (Rahn, 1996). Adding weight to the potential slide or a stable pre-existing

slide can accelerate or trigger movement of the slide because the angle of repose has been

changed. Humans also cause landslides by providing a triggering mechanism for slope

failure (Rahn, 1996). An example is cutting the toe of an existing landslide to build a

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 67 roadway or structure. Cutting the toe causes the slide to become unstable and begin to

move.

Landslides are categorized into three types based on the lithology in which the

landslide occurred; (1) bedrock, (2) soils, and (3) unconsolidated materials (Rahn, 1996).

This study focuses on the landslides that occur in bedrock. Bedrock landslides are further

categorized into four types: (1) rockfall, (2) rotational slump, (3) planar block slide, and

(4) rockslide (Rahn, 1996) (Fig. 24). Rockfalls generally occur along steep cliffs and

may occur in M 4.0 and greater earthquakes (Keefer, 1984; Harty, 1992). Along the

central Sevier fault, fourteen rockfalls and/or rotational slumps were documented (Plates

1 and 2). The rockfalls are located near the high cliffs of the Navajo Sandstone and the

Co-op Creek Member of the Carmel Formation. The rotational slumps are derived from

the weak shale and mudstone units (Tropic Shale, Dakota Formation, Carmel Formation)

near Glendale (Plates 2 and 3).

Regional Slope Failures

Landslides are common in southwestern Utah because the region experiences

relatively high precipitation and has high elevation, steep slopes, earthquakes, and

unstable geologic formations (Doelling et al., 1989; Harty, 1992). Many of these

landslides have damaged infrastructure, homes, and businesses. However, these slides

are relatively unstudied. Below are possible triggering mechanisms and classification of

the fourteen landslides along the central Sevier fault.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 68 Triggering Mechanisms for Long Valley Landslides

Landslides are common within the vicinity of the Sevier fault (Schroder, 1971;

Doelling et al., 1989; Harty, 1992). Characteristic features of landslides include: (1)

steep topography, (2) arcuate to linear scarps, (3) benched or hummocky topography, (4)

bulging toes, (5) ponded or re-routed drainages, and (6) immature vegetation (compared

the vegetation around the landslide) on the landslide (McCalpin, 1996). All of the

fourteen landslides in the study area may have been natural slope failures in weak units

triggered by high rainfall and/or the freeze-thaw cycle. Three landslides, labeled A, B,

and C, were identified along the Glendale Bench Road (Plate 2 and Fig. 25) (Harty, 1992;

this study). One of these landslides, slide A, has either been triggered or reactivated by

humans (Fig. 25). The center of the landslide has been cut for road construction (Fig.

26A). Several failure precaution techniques have been applied to this landslide to prevent

future failure. The first technique is to cut back the slope (above the road) of the existing

landslide to a lower angle to prevent the landslide from sliding over the road (Fig. 26A).

The second technique is to reinforce the landslide topographically below the road to

prevent a new landslide from forming within the old slide (Fig. 26A). The other

landslides, slides B and C, along the Glendale Bench Road may or may not have been

human induced. No stabilization techniques have been performed on these slides.

In contrast, the landslide in the eastern portion of Dry Wash, called the Dry Wash

Landslide, may have been triggered by ground shaking from a historical earthquake

(Plate I, Figs. 25 and 26B). To state that a landslide is seismically induced the landslide

must be dated and associated with earthquake shaking (Keefer, 1994; McCalpin, 1996;

Bamhardt and Kayen, 1999; Papadopoulos and Plessa, 2000).

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 69 To demonstrate whether an earthquake triggered a landslide is difficult because

only a few studies have been done on seismically induced landslides. Therefore, the

interpretations vary widely. Rahn (1996), McCalpin (1996), and Bamhardt and Kayen

(1999) summarize several basic criteria to determine whether a landslide has a seismic

origin. These techniques are used in this study. First, is the regional analysis of

landslides. Commonly, in a seismically active zone, a group of landslides will occur

simultaneously (Nikonov, 1988). However, single large landslides can also be triggered

by earthquakes (Nikonov, 1988). Second, the landslide morphology must be analyzed.

The hillslope from which the slide was derived must be at a steep angle with some

topographic relief. In regions where the topographic relief is due to streams downcutting,

the landslide is typically undercut by a stream. These landslides also may contain

characteristics such as liquefaction features. Seismically induced landslides typically

have a more blocky shape whereas slides that form because of intense rainfall typically

have a more spread out and less blocky shape (Perrin and Hancox, 1992; Bamhardt and

Kayen, 1999). Lastly, the historical records of reported earthquakes and landslides must

match, keeping in mind that earthquake-triggered landslides may have a three to five day

delay in movement (Rojstaczer and Wolf, 1992). This delay in movement may be caused

by increased groundwater flow locally or time-dependent manifestations of increased

pore water pressure due to ground shaking during an earthquake (Rojstaczer and Wolf,

1992). Papadopoulos and Plessa (2(XK)) and Keefer (1984) suggest that earthquake

triggered landslides usually occur in weak materials such as soils, weathered, sheared,

heavily fractured, jointed, or saturated rocks.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 70 Classification of the Dry Wash Landslide

By using the above stated techniques, the Dry Wash Landslide can be classified

and its triggering mechanism can be evaluated. I suggest that the Dry Wash Landslide is

a rotational slump (Fig. 24) and may have been triggered by the 1924 Orderville

earthquake and/or associated aftershocks.

Evidence to support the Dry Wash Landslide as a rotation slump include steep

topography, an arcuate to linear scarp, and hummocky topography (Fig. 26A). The Dry

Wash Landslide also contains blocks that were tilted or rotated during failure. Several of

the trees incorporated in the landslide are also rotated (Fig. 27A). A vegetation

difference is notable between the landslide and the surrounding topography. Little to no

new vegetation exists on the deposit and what new growth exists is sparse (Fig. 27A).

The new growth consists of -0.9 m (-3 ft) tall pine trees and immature manzanita, sage

brush, yucca, and tall grass (Fig. 27B). This vegetation is less mature than the same types

of vegetation on surfaces surrounding the landslide (Fig. 28A). However, the dead

vegetation within the deposit is similar in maturity to the live vegetation surrounding the

feature. Most of the vegetation incorporated within the landslide is highly decayed and

lacks foliage (Figs 28B and 29A). This vegetation is rotated and incorporated within the

landslide. However, some of incorporated vegetation survived the landslide and is

growing at an oblique angle to its original growth direction giving the trunks of the trees

a bent or hook shape (Fig. 27A). The original surface as well as vegetation before the

failure is preserved both above and below (in the scarp graben) the head scarp of the slide

(Figs. 24 and 29B). The toe is bulging and has ponded and rerouted drainages (Fig.

30A).

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 71 The age of the Dry Wash Landslide was determined by using historical methods,

vegetation analysis, and geomorphic analysis. Historical methods include reports of

events from local inhabitants. Local inhabitants of the region state that the slide occurred

approximately 75 years ago. However, no written documentation of when the landslide

occurred is available. Geomorphic analyses suggest that the landslide occurred relatively

recently perhaps in the last century. The scarp of the Dry Wash Landslide is very steep

(nearly vertical) with little degradation (Fig. 26B). The slip surface is on a steep (nearly

vertical) slope with 37 - 46 m (120 - 150 ft) of relief. The surface of the landslide has a

blocky to angular and hummocky shape with relatively few (1-2) stream channels cutting

the landslide indicating little erosion has occurred on the slide. However, a stream

undercuts the toe (Fig. 30A). Vegetation analysis also suggests that the landslide

occurred recently, perhaps within the past century because the maturity of the vegetation

incorporated within and living within the landslide is -<100 years old. By the above

analyses, I suggest that the Dry Wash Landslide is young (<100 years old).

Even though the Dry Wash Landslide fits the above stated criteria to be

categorized as a seismically induced earthquake (cf., Rahn, 1996; McCalpin, 1996), I

suggest that the evidence is lacking to be certain whether the Dry Wash Landslide was

seismically induced. First, only an approximate date of when the landslide occurred was

documented. No written historical records were found. Therefore, it is difficult to say

whether the landslide occurred approximately at the same time as the 1924 Orderville

earthquake swarm. Secondly, the vegetation and geomorphic descriptions of the Dry

Wash Landslide also fit the descriptions of landslides that were not triggered by

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 72 earthquakes. Therefore, by using the above criteria and data the Dry Wash Landslide

may or may not be seismically induced.

Ethical Assessment of Hazards

Natural disasters such as earthquakes and landslides are unavoidable and can

threaten life and property. Investigations of landslides and potentially active faults

involve good science as well as ethical scientific practices (Cronin and Sverdrup, 1998).

Thus, great care should be taken when locating new sites for structures. Areas above,

below, as well as on the weak shale and mudstones should be avoided. Potential

earthquake and landslide hazards should be analyzed before new structures are built.

Land and building owners should be made aware of any potential earthquake and

associated hazards. A geoscientist must follow a code of ethics to ensure the reduction of

life and property loss during a natural disaster (Cronin and Sverdrup, 1998). Therefore,

with better earthquake building codes, understanding of landslides and their locations,

ethical scientific practice, and increased public awareness, the possible life and property

loss from such natural hazards can be minimized.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. CHAPTER 9

TECTONIC INTERPRETATIONS

Intrcxluction

Additional data and analysis may help resolve some of the controversial topics

about the evolution of the western edge of the Colorado Plateau. Discussed below are the

northern termination of the Sevier fault, relationships of the Basin and Range style faults

in southwestern Utah, and the age of folding within and near the study area.

Northern Termination of the Sevier Fault

Controversy about where the Sevier fault terminates to the north is found

throughout the literature. Most workers agree that the Sevier fault ends near or within the

Marysvale Volcanic field (Fig. 7). Determining the northern termination is important to

determine the seismic potential of the Sevier fault as well as to determine if the Sevier

fault connects to the zone to the north.

Previous workers (i.e., Eardley and Beutner, 1934) mapped the Sevier fault as a

single strand cutting the Marysvale Volcanic Field and place the northern termination

near Central, Utah (Fig. 7). However, the Utah state geological map shows the Sevier

fault ending in the middle of the Marysvale Volcanic Field. Later mapping (Rowley et

al., 1981) suggested that the Sevier fault has several splays and strands cutting the

Marysvale Volcanic Field up to 5 km (3.1 mi) north of Piute Reservoir and possibly

73

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 74 farther north. Hecker (1993) showed Quaternary faulting along trend of the Sevier fault

to Richfield. However, Hecker (1993) showed no Quaternary scarps between Panguitch

and the southern portion of the Piute Reservoir (Fig. 7). This suggests that either the

Sevier fault does not exist in this region or its last movement is pre-Quatemary. However

D. Simon (personal communication 2002) documented Holocene events near the Piute

Reservoir. Davis (1999) suggested that the Sevier fault lost its characteristics within the

Marysvale Volcanic Field. However, Rowley (personal communication 2002) observed

scarps indicating that the Sevier fault does cut the Marysvale Volcanic Field and

terminates near Richfield. These observations suggest that the Sevier fault does not

connect with the Wasatch fault, and terminates near Central or Richfield (Fig. 7).

Other Long Normal Fault Relationships

The Hurricane, Sevier, and Paunsaugunt faults have the same general structural

characteristics (i.e., down-to-the-west, steeply dipping, segmented, etc.). However, their

relationships to each other may provide an insight to the evolution of the western margin

of the Colorado Plateau.

First, the greatest amount of seismicity is along the Hurricane fault with the

largest recorded earthquake in the last decade (M 5.8) (Arabasz et al., 1992; Pechmann et

al., 1992). Some seismicity occurs near the Sevier fault, but less than along the

Hurricane fault (Davis, 1999). Two earthquakes have been recorded along the trace of

the Sevier fault in the last century (Doelling et al., 1989; Davis 1999). The Paunsaugunt

fault has had no recorded seismic events within the last 100 years (Doelling et al., 1989;

Engdahl and Rinehart, 1991; Davis, 1999).

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 75 Second, offset units become younger to the west. For example, the Hurricane

fault offsets Holocene units and several Quaternary sedimentary deposits (Stewart et al.,

1997). Quaternary basalt is also offset (Stewart and Taylor, 1996; Stewart et al., 1997).

The Sevier fault offsets Quaternary basalts (Best et al., 1980) and Quaternary sediments

(Doelling et al., 1989; Hecker, 1993). The Sevier fault also tectonically tilts stream

terrace deposits near Orderville and Glendale (Fig. 1). However, along the northern

portion, near the Piute Reservoir, as many as four events have occurred in the Holocene

(Fig. 7) (D. Simon personal communication, 2002). Mapping by Bowers (1991) along

the Paunsaugunt faults shows only one location, south of Bryce Canyon National Park,

with fault scarps that cut Quaternary units. The two scarps are <1 meter (<~3 ft) high and

are located in Pliestocene to possibly Holocene pediment surfaces (Fig. 7).

Third, the Paunsaugunt fault has reversed topography (the hanging wall is

topographically higher than the footwall) (Doelling et al., 1989; Bowers, 1991). This

indicates either the units in the hanging wall are more susceptible to erosion than the

footwall units or the fault has not had recent movement. Along the Paunsaugunt fault,

erosion, little to no movement along the fault, or both options are probable. Using the

above previously published data and new data from this study, 1 agree with the previous

suggestion that the faults are more active or younger to the west (i.e., Doelling et al.,

1989; Davis 1999; Schiefelbein and Taylor, 2000).

Sevier, Laramide, or Cenozoic Regional Folding within the Study Area

Broad regional folds can be found from the vicinity of Zion National Park east

through the Colorado Plateau (Doelling et al., 1989; Davis, 1999). From structure

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 76 contour maps Doelling et al. (1989) suggested that the Mt. Harris anticline lies in the

hanging wall of the Sevier fault south of Mt. Carmel Junction (Fig. 7). They also

suggested that the synclines lie in the hanging wall north of Black Mountain and in the

footwall east of Black Mountain (Figs. 7 and 1).

These folds may have been formed during the Sevier orogeny, Laramide orogeny,

or by Cenozoic normal faulting. Davis (1978) and Wannamaker et al. (2001) suggested a

Laramide age for these folds. However, Davis (1999) suggested that several folds

located near the normal fault traces formed by fault drag. Data from this study suggest

that these folds were formed during the Laramide orogeny, Cenozoic normal faulting,

and/or both.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. CHAPTER 10

PETROLEUM POTENTIAL

Normal Fault Linkage and Hydrocarbon Production

The hanging walls and footwalls linked by relay ramps as well as structures

within the relay ramp can form migration pathways and/or traps for hydrocarbons

(Morley et al., 1990; Peacock and Sanderson, 1994; Dawers and Underhill, 2000; this

study). Relay ramps can be important for finding hydrocarbon traps because fault-

parallel folds within the ramps, in addition to the ramp monoclines may be traps (Morley

et al., 1990; Peacock and Sanderson, 1994). The bounding faults of and/or breakthrough

faults within relay ramps can ( 1 ) form barriers to hydrocarbon migration, (2) act as

hydrocarbon flow pathways, or (3) be permeable such that the hydrocarbons can pass

through the fault. Below, relay ramps will be discussed as potential traps and/or

migration pathways for hydrocarbons.

Relay ramps with different internal structures trap hydrocarbons in different

locations. Relay ramps that do not have associated fault parallel folds, but have

impermeable faults typically trap hydrocarbons on the up-dip side of the ramp (Fig. 31 A).

Relay ramps that contain fault parallel anticlines, such as those recognized in this study,

trap hydrocarbons within the structurally highest portion of the anticline (Fig. 3 IB).

Relay ramps with fault parallel synclines should trap hydrocarbons along the limbs of the

fold near the bounding faults, if the faults are impermeable (Fig. 31C). With

77

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 78 breakthrough faults, hydrocarbons may accumulate within fault bounded blocks

depending on whether the faults are permeable (Fig. 3ID). These hydrocarbons can

either migrate from the source rock up along a non-permeable fault or through a

permeable fault and into the topographically highest part of the ramp below the cap rock

(Fig. 31 ). The amount of hydrocarbon accumulation depends upon the source rock.

In addition to relay ramps, physical fault linkage may be important for

hydrocarbon exploration. For example, if two faults with overlapping fault tips link at

one site and a fault-parallel anticline develops within the relay ramp, then four-way-

closure can form. The four closures are; (1) the linkage site, (2) the ramp monocline, and

(3 and 4) the two limbs of the anticline. However, if the faults link via overlapping fault

tips along breakthrough faults, then a relay ramp may or may not form. However, a

structure to trap the hydrocarbons can still form (Ferrill and Morris, 2001).

Hydrocarbon Production Near Study Area

Hydrocarbon producing fields located near the central Sevier fault include the

Virgin oil field. Upper Valley field, and Anderson Junction field (Fig. 32). The source

rocks for these fields include the Pennsylvanian Hermosa Formation, Permian Hermit and

Kaibab Formations as well as the Virgin Limestone and Timpoweap Members of the

Triassic Moenkopi Formation (Fig. 8) (Peterson, 1974; Hintze, 1988; Van Kooten, 1988;

Doelling et a!., 1989; Harris, 1994; S.J. Reber, personal communication 2001). The

structures that have been explored include folds associated with the Laramide and Sevier

orogenies and Cenozoic normal faults. For the Anderson Junction and Virgin oil fields

the Hurricane fault was considered to be a migration pathway (Peterson, 1974; Harris,

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 79 1994). In the Anderson Junction field, the oil was trapped either on an anticlinal closure

of the Kanarra anticline, a fault wedge of the Hurricane fault, or the Virgin anticlinorium

(Peterson, 1974; Harris, 1994). In the Upper Valley field, hydrocarbons were trapped

and pumped out of the Upper Valley anticline. The total number of barrels produced in

southwestern Utah between the first oil production in the state in 1907 (Virgin field) and

1988 is approximately 26 million barrels (26 MMBO) (Peterson, 1974; Van Kooten,

1988; Doelling et al., 1989). However, the total amount of oil produced in the entire state

of Utah is -900 million barrels (900 MMBO).

The High Plateaus subprovince near the Sevier fault may be a region for future

hydrocarbon exploration. Today, the number of oil producing fields are minimal in the

area. However, four exploration holes with shows of oil and/or gas have been drilled in

the footwall near Mt. Carmel, and one hole with no shows of oil or gas was drilled in the

footwall near Orderville (Fig. 32) (Doelling et al., 1989). Four relay ramps have been

identified along the Sevier fault (Davis, 1999; this study). These relay ramps and

associated faults fit the above models for hydrocarbon traps and/or migration pathways.

By understanding relay ramps and the types of linkage zones along the Sevier fault, these

structures may be used as analogs for potential traps and/or migration pathways in other

oil producing extensional regimes (e.g.. North Sea, Gulf of Suez, and the East African

Rift).

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. CHAPTER II

CONCLUSIONS

Recognition of large convex geometric bends, or salients, along faults coincides

with concepts that long faults form by linkage of two or more shorter faults. Linkage

typically creates salients along the trace of the fault that geometrically segment the faults.

Segmented long faults can form by a number of different types of linkage However,

models of segment linkage generally do not address multipartite linkage zones. Such

studies of active linked or linking faults have several applied significances Fault linkage

zones are important to the petroleum industry to understand migration pathways and

where potential traps exist, and important to the public to understand the potential

earthquake and landslide hazards for the region. This study focuses on multipartite

linkage zones and the structures formed within these zones using the central Sevier fault

as a case study

Two main types of fault linkage and several subtypes are currently recognized;

overlapping tips, simple underlapping tips, fault capture, and breakthrough faults.

Within overlapping fault linkage zones, relay ramps typically form. Understanding the

relay ramps and the structures within is necessary to fully describe linkage zones

However, fault-parallel folds within relay ramps have not been discussed at length.

80

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 81 The Sevier fault is a segmented long normal fault that probably initiated in the

Miocene (IS -12 Ma) (Davis, 1999). The Sevier fault is generally a N30°E striking,

steeply (>75°) west dipping, multi-strand normal fault.

Two subtypes of fault linkage were documented along the central Sevier fault.

The first is a fault capture situation in the Orderville relay ramp area (Fig. 12). The

second is a series of overl^ping faults with cross faults in the Stewart Canyon overly

zone (Fig. 12). This overlap zone contains three relay ramps (Glendale, Stewart Canyon,

and Elkheart Clififs) (Fig. 12). Relay ramps are a structure that allows strain transfer

between two faults (Fig. 5). The Stewart Canyon relay ramp is different from the simple

relay ramp model; it is bound by three faults.

Along the central Sevier fault, two structures were identified within the relay

ramps. In the Orderville relay ramp and the Glendale relay ramp, fault parallel folds are

exposed within the relay ramp. In addition, in the Glendale relay ramp cross or

breakthrough faults were identified. These faults cut the fault parallel fold. The

multipartite nature of the region between Orderville and Glendale has a series of linkage

sites that form a large salient across which the Mt. Carmel segment and Spencer Bench

segment linked.

The Sevier fault is possibly an active, at least active in the Quaternary, fault. It

offsets ~570 ka basalt and tilts yotmg (Pleistocene?) river deposits. Earthquakes have

also been felt and recorded near the Sevier fault. The area of interest also has landslide

hazards. Several landslides which maybe seismically induced have been identified in the

area.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 82 In conclusion, this study provided (1) new natural examples of multipartite

normal fault linkage zones, (2) an evaluation of the earthquake and landslide hazards

along the central Sevier fault, and (3) new data and interpretations for the regional

tectonics of southern Utah.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. APPENDIX A

METHODS

The techniques of data collection and analyses include field mapping, cross-

section construction, stereographic analyses, and ^Ar/^^Ar isotopic dating. An overview

of each technique is presented below.

Field Mapping Techniques

Geologic mapping was the primary technique used for data collection.

Approximately 30 km^ (19 mi^) were mapped in Kane County, southwestern Utah during

the summer o f2000. Mapping was done using standard geologic techniques at a scale of

1:12,000. The topographic base was enlarged from the Orderville, Glendale, and Long

Valley 7.5’ U.S.G.S. quadrangles printed from the MapTech topographic map computer

program. Color aerial photographs were used to aid in locating geologic features,

structures, and formations. However, all mapping was based on direct field observation.

Cross Section Construction Techniques

Retrodeformable cross sections were constructed from structural and stratigraphie

data and relationships observed and mapped in the field. All cross sections were drawn

approximately perpendicular to the strike of the faults in order to analyze along strike

variations in splay zones and along single strands o f the fault. All fault attitudes were

83

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 84 calculated by three point problems and/or structure contours (Marshak and Mitra, 1988;

Rowland and Duebendorfer, 1994) because no faults in the study area yielded a

measurable surface. Apparent dips were calculated by using the alignment diagram for

solving iq)paient dip problems (Rowland and Duebendorfer, 1994). The cross sections

were constructed under the assumptions that plane strain occurred, the volume o f rock

remained constant, and the bedding thickness remained constant. The bed lengths

balance and no loss of area occurred. Bed lengths in both restored and deformed cross

sections can be and are within the standard uncertainty of 5-10% (Rowland and

Duebendorfer, 1994). Constant unit thickness is based on (1) the calculated thickness

from the map pattern of exposed units in the study area and (2) the published thickness of

subsurface units (Figs. 8 and 9) (Hintze, 1988; Stokes, 1988; Doelling et al., 1989).

The geometries o f the faults are based on the surface data and the ability to be

retrodeformed along each cross section. Standard cross section retrodeformation

techniques were used for the faults (e.g., Gibbs, 1983; Davison, 1986; Williams and

Vann, 1987; Vendeville, 1991; White and Yielding, 1991; White, 1992; Kerr and White,

1994; Song and Cawood, 2001).

Stereographic Analyses

Several stereographic analyses were done on structures within the study area.

Stereonets were created to show the attitudes of folds within the relay ramps. All

stereographic analyses were done using the computer program GeOrient.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 85 ^Ar^^Ar Methods

Two Quaternary basalt samples from the same flow were collected for ^Ar/^^Ar

isotopic dating as part of this study. BRC-I was collected from the hanging wall and

BRM-2 was collected from the footwall of the Sevier fault. Approximately 2.25 kg (5

lbs) of each sample was collected and trimmed of weathered surfaces in the field. Thin

sections were made and analyzed from each sample to ensure that the samples could be

isotopically dated. The samples were crushed and seived to uniform sizes. The seive

size chosen was the size that yielded the largest possible individual crystals that lacked an

adhering matrix. Olivine and volcanic glass were removed from the samples by mineral

separation methods including heavy liquids and hand picking. All mineral separations

were completed by the author at the UNLV Mineral Separation Lab.

Approximately 0.1 oz (250 mg) of basaltic groundmass firom each sample were

sent to the TRIGA Reactor at the University of Michigan for irradiation. Both samples

were then dated using the ^Ar/^^Ar dating techniques at the Nevada Isotope

Geochronology Lab under the direction of Dr. Terry Spell. Methods of ^Ar/^®Ar dating

are described in McDougall and Harrison (1988).

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. ■o I 8 Q.

■D CD

BRC-1, Hangingwall baaatt J = 0.0009860 +/- 0.5% step 36Ar 37Ar 38Ar 39Ar 40Ar %40Ar‘ % 39Ar Had Ca/K 40Ar*/39ArK Age(ka) lad. 1 1.168 21.043 7.099 46.681 347.471 4.5 3.2 3.02204972 385.3463 581 45 2 0.373 21.399 4.397 65.996 127.538 21.7 4.5 2.17319728 474.9860 693 18 3 0.399 34.899 8 4.783 137.595 157.175 33.7 9.5 1.69969812 437.4950 647 14 ■D 4 0.463 50.267 4.95 228.484 201.768 40.6 15.7 1.47420945 408.8139 611 14 5 0.459 52.0 4.278 247.133 211.537 43.1 17.0 1.40992569 415.1381 619 14 6 0.452 44.595 3.282 199.718 192.101 37.5 13.7 1.49624894 400.9480 601 15 7 0.631 62.979 3.666 3 210.743 243.816 30.9 14.5 2.00282796 403.3032 604 18 CD 8 0.523 48.595 2.555 106.222 183.313 25.0 7.3 3.06702336 493.2572 715 21 9 0.648 59.985 2.684 86.579 206.163 16.7 6.0 4.64703833 450.4271 663 26 10 2.409 3. 561,781 6.719 107.924 612.525 7.7 7.4 35.2344524 533.9035 763 142 3" 11 0.962 53.545 0.949 12.408 274.831 CD 3.4 0.9 29.1574354 974.6847 1215 159 12 0.998 12.953 0.371 3.255 288.36 0.7 S 0.2 26.6692229 790.8199 1040 545 ■o Total gas age = O 644.26 15.01 CQ. steps 4-7 Plateau age = 609.62 15.01 g. BRM>2, Footwall Isasalt J = 0.0009920 +/-0.5% o step 36Ar 37Ar 38Ar 39Ar 40Ar %40Ar* 3 %39Arrlsd Ca/K 40Ar*/39ArK Aae(ka) lad. 1 4.661 ■o 35.768 9.069 73.269 1334.228 1.6 5.5 3.33034434 338.2579 522 123 o 2 1.152 44.009 8.491 102.421 363.437 11.5 7.6 2.93099522 488.6136 713 25 3 1.15 52.307 10.25 187.277 386.931 17.2 14.0 1.90460093 414.9916 622 18 CD 4 1.15 49.131 Q. 9.457 241.597 409.738 21.5 18.0 1.38651649 429.2611 640 19 5 1.133 41.9 6.166 203.626 388.245 17.4 15.2 1.40205111 379.1655 576 21 6 1.01 35.703 3.478 142.427 338.564 15.6 10.6 1.70928902 429.2611 640 29 7 1.223 47.142 3.528 143.132 404.517 14.7 10.7 2.24617964 498.6021 725 26 "D 8 1.038 38.102 2.812 CD 75.708 322.777 9.9 5.6 3.4334775 502.7836 730 52 9 1.437 59.372 3.47 75.865 426.994 5.9 5.7 5.34216699 381.4743 579 49 (/) 10 3.555 488.233 C/) 6.421 82.936 952.688 4.1 6.2 40.610412 597.7710 640 181 11 1.022 68.762 1.164 13.183 307.063 9.9 1.0 35.9321625 7036.4809 3747 283 Total gas age = 680.50 20.70 steps 3-6 Plateau age = 618.51 20.70 Appendix B. ^"Ar/^^Ar step heating data for basalt with olivinie and glass removed. Only nine steps are used in each isochron calculation (Fig. 11). * 87

Panguitch

Red Canyon H urricaneT Sevier Bryce Canyon Fault Fault National Park St. George Grand W ash Glendale Orderville Mt. Carrne YeSi, VRSure*» Paunsauaun G Jacket -pEoraS ARIZONA Sevier Orogenic Belt & Thrust Front /

^ Washin^tonj

/ \ ’■ / Approximate ISB Southern Boundary COLORADO PLATEAU

Kilometers

Figure I. This location map shows the Hurricane fault, the Sevier-Toroweap fault, the Paunsaugunt fault and their locations relative to the Basin and Range province. Transition Zone, and the Colorado Plateau. The ISB (Intermountain Seismic Belt) is located north of the labeled dashed line. The study area is indicated by the dark gray polygon.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 7] ■DCD Q.O C g Q.

■D CD

C/) C/) A. Linkage of Underlapping B. Linkage of Overlapping C Fault Capture Along D. Linkage of Overlapping Type Fault Segments Type Fault Segments Connecting FauHs Subtype witti Breakttvougl) Faults Subtype

8

(O'

3.3" CD

CD T3 O Q. Ca o3 T3 O

CDQ.

Figure 2. Stages of fault propagation and linkage, with stages 1-3 in chronological order. Diagram A is

T3 the three general steps for linkage of underlapping type faults. Diagram B (overlapping type fault tips), C CD (fault capture subtype), and D (breakthrough fault linkage subtype) show the three types of linkage that (/) (/) occurs in overlapping fault tips. Boxes labeled I show faults growing by radial propagation. Boxes labeled 2 show the second stage of linkage. The faults and stress fields have interacted with each other but have not linked. Boxes labeled 3 show the final phase of linkage. The shaded areas and arrows represent the total fault displacement. The heavier lines represent the normal faults. This diagram is modified from Reber et al. (2001) and Taylor et al. (2001). % 7] ■DCD O Q. C 8 Q.

■D CD

C/) C /i

A. Single Propagating B. Stress Field Interaction c . Continued Stress Field D. Stress Field Interaction Fault with Underlapping Interaction witti Under- with Overlapping Fault ■D8 Faults lapping Faults and Relay Ramp Development

CD

3. 3" CD ■DCD i O Q. C a if 3O "O O

CD Q.

Figure 3. These diagrams show the local stress field location and interactions. The + signs represent the locations ■D where the faults may be located aficr the next rupture. The - signs are areas that are not likely to fail during the next CD rupture. The contour lines represent the numerically modeled shear stress with darker gray fill for higher values. The

C/Î C/) enclosed shapes represent the local stress field for each fault. Diagram A is a representation of radial propagation. B is underlapping faults. C is continued slip along the faults in diagram B (underlapping faults). D is overlapping faults. In diagram D, a relay ramp (strain transfer zone) has formed between the overlapping fault tips. Ball and bar are on the hanging wall of fault. Diagram A is modified from Cowie and Shipton (1998). Diagrams B - D are modified from Crider and Pollard (1998) and Crider (2001). * 90

Underlapping Faults LEQEND U = Up-thrown block Map View D = Down-dropped block — ' = Normal fault S /fe o f —» = Magnitude of offset Unlage, Figure 4. This diagram shows model displacement 100 200 Distance Along Strike vs. distance diagrams for two types (A and B) and two subtypes (C and D) o f B Overlapping Faults fault linkage. The distance Map View is measured along the strike of a piece of a fault across the linkage zone. The data density for diagram D was «■gio doubled within the linkage zone to show the "step" 100 200 pattern. The ball and bar Distance Along Strike are on the hanging wall. The shaded area and arrows Fault Capture represent the amount of Map View displacement along the fault. The axes are unitless because the models are Plateau independent o f scale. S l4 0 Diagrams A through C are a 1 20 modified from Taylor et al. (2001). 100 200 300 Distance Along Strike

Breakthrough Faults u Map View

^ -^30

100 200 300 400 Distance Along Strike

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 91

Soft Linkage Map View

100 200 Distance Along Strike

Figure 4 cont. This diagram (E) shows model displacement vs. distance diagrams for soft linkage in an overlapping fault linkage situation.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 92

Footwall

Relay Ramp

Hanging-wall

Upper Fault Fpotwall Top Breach

Relay Ramp

Hanging-wall Lower Fault

Footwall Relay Rami

Hanging-wall

Figure 5. The diagramed faults overlap in A, link by fault capture in B, and link by breakthrough faults in C. A relay ramp formed between the overlapping faults in each diagram. Diagram B is an example of a top breached ramp. Notice the changes in the strikes of the beds in the ramp. The different shades of gray represent different strata. The arrows represent the amount of displacement along the fault. The heavy lines represent the normal faults.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 93

Geometric Segment Boundary

Geometric Segment Boundary Structural Segment Boundary

Geometnc Segment Boundary

Propagation of earttiquake rupture Earthquake Segment Boundary

Figure 6. This figure shows three types of segment boundaries: (A) a geometric segment boundary, (B) a combined structural and geometric segment boundary, and (C) a combined earthquake and geometric segment boundary at which the earthquake rupture terminated. Earthquakes may also initiate at boundaries. The dark gray shaded regions and thin arrows represent the total amount of offset along the fault. The heavier weight lines represent normal faults, and the thicker arrows represent motion along the fault.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 94

Figure 7. This location Richfield _ . map shows the Sevier Central/^* fault, axial surfaces, and the approximate northern Monroe. ' ' termination. Axial surface data are from Doelling et al. (1989) and Bowers (1991). Fault Marysvaf^ trace data are from Hecker(1993). The northern termination of Reseivoir the Sevier fault was Kingston; • i determined using literature. The approximate location fo the Marysvale Volcanic Field is represented by the light gray shaded area. Panguitch Canyon National

Cedar City Cedar Breaks ^National Monument y ^ m r / ,i

Bryce Canyon Navajo Lake Anticline Legend Paunsaugunt Syncline Normal Fault Mt. Carmel Syncline Jet. \//TOrdefville 4 - - f - Anticline — Roadways • Towns ' Kanab Miles Hams Mt. 15 30 Anticline m m l^AH ARIZONA Kilometers 20 40

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 95

Age Unit Thickness Rock (ft) Type Tennv Canvon Tongue Lamb Point Tongue 400 (aventa Fm 300 .L Mostly fluvial and flat coastal plain deposits with aeolian Moenave Fm sandstone and shallow marine limestone layers. Chinie Fm 700 — % Petrified wood occunrs in the Chinie Formation. Moenkopi Fm 1200 ri-fn

Kaibab Ls 370 Toroweao Fm 450 Hermit Fm 650 r .' T A Cedar Mesa (Queantoweap) Ss 620 Dominantly marine limestone, Hermosa Fm 590 shale, and sandstone. Some limestone layers contain chert. Redwall Ls 840

Ouray Ls 150 Ls _LUL "Supramuav" 630 Muav Ls 590 Bright Angel Fm 340 Tapeats Ss 330 pG Grand Canyon Series 820+ Crystalline basement rock.

Figure 8. This stratigraphie column represents units in the subsurface of the study area. Standard symbols are used for rock types; wavy bands are used for volcanic flows, black ovals are used for coal, triangles are used for gypsum, open angles with lines represent cross beds, filled circles are used for chert, asterisks represent Precambrian basement rocks. The triangles with the crosses outside the stratigraphie column represent hydrocarbon producing units. Data from Gregory (1951), Hintze (1973, 1988), Stokes (1988), and Doelling et al. (1989).

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 96

Age Unit Thickness Rock (ft)Type Surficial deposits <100 Basalt flows <100

Fresh water unit. Consists a» of mostly fluvial and •I Claron Formation 800 ,9 lacustrine conglomerate, sandstone and limestone.

Kaiparowits Largely terrigenous rocks and Wahweap 1500 derived from the Sevier Formations Orogenic Belt. Deposited undivided along a coastal plane where coal-bearing non-marine ÜI sandstone and Straight Cliffs SS 240 conglomerate interfinger Tropic (Mancos) with marine shale of mid­ 760 Shale continental seaway. Dakota SS 320

Paria River Mbr 240 ICrystal Creek Mbr 170 n r r Co-op Creek Mbr 400 J-1 Represent two distinct paleoenvironments (1 ) Navajo sandy desert - Navajo Sandstone (main body) 2000 Sandstone and (2) shallow marine invasion - Carmel Formation.

Figure 9. This stratigraphie column represents mapped units. Standard symbols are used for rock types; wavy bands are used for volcanic flows, black ovals are used for coal, triangles are used for gypsum, and open angles with lines represent cross beds. Data from Gregory (1951), Cashion (1967), Hintze (1973), Rowley et al. (1975), Geesaman and Voorhees (1980), Hintze (1988), Marzolf (1988), Stokes (1988), Doelling et al. (1989), and Goldstrand (1994).

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 97

112"37'45 12°32'10 BRC-1 Black Rock Canyon

-37°23'35

W

Dry Wash Virgin River Landslide

37®20’30

Dry Wash Canyon Glendale Bench Glendale Road

Legend

□ Quaternary

Orderville HI Tertiary o 0.5 1.0 Cretaceous Kilometers I Jurassic 0 1500 3000 Stewart Feet Canyon Figure 10. The diagram is a simplified geologic map of the study area. Geographic locations and sample collection sites are indicated. The heavy weight lines are faults. The Virgin River is shown by a lighter weight line. 37°15'50 I

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 98

0.0036

BRC-1 BasaK (black) 0.0032 Age = 564 +/- 20 ka «Ar/36Ar = 288.1 +/- 3.9 0.0028 MSWD = 0.48 Steps 1-9 0.0024 BRM-2 Basalt (gray) Age = 580 +/- 50 ka ‘0Ar/36Ar = 291.7+/-3.85 < 0.0020 MSWD = 0.62 Steps 1-9 < 0.0016

0.0012

0.0008

0.0004

^^A r/^A r

Figure 11. The above diagram shows the isochron plots for the Black Rock Canyon basalt and Black Mountain basalt. Nine steps for each sample (shown by ellipses indicating uncertainty) were used to determine the isochron age. The Black Rock Canyon basalt is shown in black and the Black Mountain basalt is shown in gray.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 99

Fault Set F Black Mountain 37°23'30

fl) m Fault E 8

COI

Fault Set D

37°20'30"-

Glendale " Fault Group C

Stewart Canyon. Relay Ramp .g,

Elkheart Cliffs Relay Ramp N

Orderville Stewart Canyon A Fence Diagram Stewart Canyon w / Overlap Zone ^'Glendale Kilometers Relay Ramp 0 1500 3000 Orderville % Relay I Ramp Feet

‘Orderville Fence Diagram 37°15’50 112°37’45” 112°32'10 _L Figure 12A. This diagram represents the fault traces in the Sevier fault zone. The ball and bar symbols are on the hanging wall. The labeled cross sections, fault domains, sets, and strands are discussed in the text. The main strand is colored gray. The light gray dashed lines are where the displacement vs. distance diagrams were constructed.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 100

Glendale #

Spencer Bench Segment

Elkheart Cliffs Stewart Canyon / Relay Relay Ramp ' Ramp

‘ Glendale Relay Ramp Orderville Salient

Orderville Relay Ramp

0.5 Order/ille Kilometers 3000 Mt. Carmel Segment Feet Figure 12B. This diagram represents the fault traces in the Stewart Canyon and Orderville areas. The ball and bar symbols are on the hanging wall. The labeled cross sections, fault domains, sets, and strands are discussed in the text. The main strand is colored gray.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. CD "D O Q. C g Q.

■D CD (/) C/) L

CD ■D8

(O' 3 1 Wash 37°20’20"- ndslide

3"3. CD ■DCD O Quatemary/Tertiary Q. C □Alluvium (Qa) aO Landslide (Qls) 3 □Gong. (Qcg) ■D O □Basait (Qb) □Sinter (Q/Ts) Cretaceous CD Q. □Kaiparowits/Wahweap (Kkw □straight Cliffs (Ks) □Tropic Shale (Kt) Kilometers □Dakota SS (Kd) ■D Jurassic Carmel Fm CD 37“19’20 - □Windsor Mbr. (Jew) □Paria River Mbr. (Jcp) (/) □Crystal Creek Mbr. (Jcc) □Co-og^reek^brJJco)

112°38’0" 112°36’0" Figure 25. The geologic map of the central portion of the study area shows both the Dry Wash Landslide and the landslides along the Glendale Bench Road. Notice that the landslides occur in the Cretaceous Tropic (Kt) and Dakota (Kd) Formations. 101

Figure 13. The fence diagram shows the fault parallel ^ clin e within OnJervMe the Orderville rday Relay Ramp , ramp. Notice that the strata outside the relay ramp are nearly flat lying. Also, notice to the south, the area along N-hT between the faults at the surface and at depth decreases as the m space decreases. The / heavy weight lines represent normal faults. \ The white represents bedding planes within / the syncline with light weight lines representing structure / contours. The stratigraphy and stratigraphie symbols are described on Plate 3. Cross sections on Plate 4 were used to constrain the shape o f the syncline. The unlabeled vertical lines are tie lines. However, several cross sections were removed to show the syncline shape.

No Vertical Exaggeration

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 102 A. Orderville Relay Ramp

22 R s.

B Glendale Relay Ramp N

3®. N59“E

21 R s.

Figure 14. These beta plots show the attitudes o f bedding within two relay ramps and fold axes orientations. A shows beds and the synclinal axis orientation within the Orderville relay ramp B shows the attitude o f beds and anticlinal axis within the Glendale relry ramp.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 103

112°37*46

37°23’36

Black Mountain

virgin River Qcg 37®20‘30 0 1.0 Dry Wash Canyon Kilometers

Glendale 0 1500 3000 Feet

□ Alluvium (Qa) ■ Younger Conglomerate (Qcg) Older Conglomerate (Qcg1) □ Other Quaternary Units Orderville Basalt (Qb) □ Bedrock (brx)

Figure 15. This diagram is a Quaternary geologic map of the central Sevier fault. The faults are in the heavy-weight solid, dashed, and dotted lines indicating a certain location, an approximate location, and a buried fault respectively. The Virgin River is shown with a heavy gray solid line. 37=15'50"

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 104

Figure 16. The fence Elkheart CMS Relay Ramp diagram shows the > V* fault parallel anticline within the Glendale relay ramp. Notice that the strata outside the relay ramp is Glendale Relay Ramp nearly flat lying. However, the strata within the ramp form a fault paialld anticline. Stewart Canyon . The faults appear to Relay Ramp ^ connect at depth hence there is a decrease in space within the relay ramp. Notice within the ramp there are / cross fuilts which cut through the relay / ramp The heavy weight lines represent normal faults. The white represents the bedding planes showing the anticline with the light lines representing structure contours. The stratigraphy and \ stratigraphie ^m bols are described on Plate 3. Cross sections on Plate 4 were used to contrain the shape o f the anticline. The unamed vertical lines represent tie lines. Several cross sections No Vertical Exaggeration were removed to show the anticline shape.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. ■o Û.o c g û.

■o CD

(£(/) 3d

8 -Ov< ë '

3 CD

Cp. 3" CD ■OCD O Q. C Oa 3 ■O O

CD Q.

■D CD

C/) C/) Figure 17. Photo A shows the younger conglomerate (Qcg). Notice the variation in size and composition o f the clasts. In the lower left comer is a sandbar. Photo taken facing north. Rockhammer for scale. The outcrop faces south. Photo B is the older conglomerate (Qcgl). This unit is overall finer grained. Notice the large basalt clasts. Person (Robert Kerscher) for scale is 1.75 meters (S'9”) tall. 106

Figure 18. Photo A is a roadcut along U.S. 89 o f the Cretaceous Dakota Sandstone unconformably overiain by the younger conglomerate. The black unit in the left o f the photo is one o f the coal seams in die Dakota Sandstone. The photo was taken facing west. Photo B is a sandbar within the young conglomerate. This photo was taken facing north.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 7} ■DCD O Q. C g Q.

■D CD

C/) C/î

800

400 ■D8

CD

600 3. 3" CD 400 ■DCD 200 O Q. C 800 a 3o 600 T3 400 O 200

(D Q. Stewart Canyon Orderville c Relay Ramp 1200 Northern Portion of Data Set Overlap Zone 800

T3 (D 400

(/) C/) Distance (km) Figure 19. This diagram shows the displacement vs. distance profiles (A - D) for each major fault. The vertical scale differs among the plots. A is the easternmost fault, is the westernmost fault, and (E) the central Sevier fault as a whole. Displacement was plotted along even increments of distance for each fault. Distance was measured along the fault. 108

Displacement vs. Distance Profiles A 800, Orderville Relay Ramp r 600

5 200 Orderville

0.5 1 1.5 2.5 Distance Along Strike (km)

B N 8 500 Glendale Relay Ramp 1 400 I 300 8 Cross Fault S 200 S' Q 100

0.51 1.5 2 2.5 Distance Along Strike (km)

. Stewart Canyon Relay Ramp ^ 600

1 1.5 2 Distance Along Strike (km) D 400

~ 300 Elkheart Cliffs Relay Ramp 8 200

5 100

0.1 0.2 0.3 0.4 0.5 Distance Along Strike (km) Figure 20. The diagram shows the displacment vs. distance profiles along potential linkage zones in the Sevier fault. A is near the Orderville relay ramp. B is the Glendale relay ramp. C shows the Stewart Canyon relay ramp area. D is the Elkheart Cliffs relay ramp. The relay ramps are located in the center of the diagrams. The displacement is the total stratigraphie offset of the bounding faults.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 109

Relay Ramp Fault B Fault A

Legend Linkage Site Shale

Sandstone Relay Fault A Ramp Fault B Limestone

Fault

Linkage Site

Figure 21. These diagrams show the evolution of a fault bounded panel of rocks with the formation of a fault-parallel syncline within the relay ramp. The faults are the heavy weighted lines and the relay ramp is labeled.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 110

Glendale* Glendale Spencer Bench, f Segment ~-/U' Kilometers 0 IKK) 3000 K -eet

Stewart Orderville. Canyon Overlap Zone

jDrdervllle z / ~ ~ Mt. Carmel Segment

Glendale Glendale

w:

Orderville. Stewart Orderville. Stewart Canyon Canyon Overlap Overlap Zone Zone

E Glendale < Figure 22. This diagram shows the stages Spencer Bench Segment - of linkage for the central Sevier fault. A is Elkheart the oldest and E is the modem fault trace map. Boxes B - D could be in any order. Stewart Canyon' There are no timing constraints of which Relay Ramp ; linkage happened first, second, or third. The black lines represent faults. The gray lines in the last diagram represent the main Orderville , • V Glendale strand. The fold symbols show the V; Relay Ramp approximate timing of folding within the relay ramps. Orderville /Orderville Relay Ramp / Salient />Mt. Carmel Segment

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 7J CD ■ D O Q. C g Q.

■D CD

(/)

Stewart Canyon C. Stewart Canyon, ' Breached j Breached Beginning Stages Relay Ramp / Relay Ramp 8 of Linkage ■D !^ase U' Breach (O'

i X Base Breach Legend ' Glendale 3"3. Relay Ramp CD 2000 Feet Glendale Breached CD W Relay Ramp ■D W 1000 -Top O Breach CQ. Meters a 3O ■D Figure 23. The fault trace map of the O Stewart Canyon, E. Stewart Canyon / Breached Breached J / Stewart Canyon overlap zone shows the Relay Ramp. /Relay Ramp/ '' CD stages of development. Diagrams A - E Q. U’ are in chronological order with A representing the youngest time and E representing the oldest time. The gray ■D CD lines represent the faults with relatively //r unconstrained timing and the black lines (/) VEIkheart indicate faults with moderate to well \ Cliffs /f , v^^Elkheart // Z' /A / Cliffs constrained timing y / // y Relay Ramp Glendale Breached Relay Glendale Breached Relay Ramp With Cross Faults Y Ramp With Cross Faults 112

Rockfall Joint

(0 75

Rotational Slump Planar - Block Slide %

Scarp Graben o Q) Rupture S Surface JO

Extremely Slow to Moderate Moderate Scarp Face Rocksllde % Dip Slope i i Q Very Slow to Extremely Rapid

Figure 24. This diagram shows the different types of landslides. The landslides found in the study are are rockfall and rotational slump types. From Vames (1958) and Rahn (1996).

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 7) "OCD O Q. c 8 Q.

■D CD

C/) C/)

■D8 SfVyash 37°20'20" ndslide

3. 3" CD ■DCD O Û. Quaternary/Tertiary C EX Alluvium (Qa) Oa Landslide (Qls) 3 □Cong. (Qcg) ■o QBasalt (Qb) o □Sinter (Q/Ts) Cretaceous CD Q. □KaiparowitsA/Vahweap (Kkw □straight Cliffs (Ks) □Tropic Shale (Kt) Kilometers □Dakota SS (Kd) ■D 37°19'20- Jurassic Carmel Fm CD □Windsor Mbr. (Jew) (J) □Paria River Mbr. (Jcp) C/Î O □Crystal Creek Mbr. (Jcc) 3 □C o-op Creek Mbr. (Jco) 1 T 112°38'0" 112°36'0" Figure 25. The geologic map of the central portion of the study area shows both the Dry Wash Landslide and the landslides along the Glendale Bench Road. Notice that the landslides occur in the Cretaceous Tropic (Kt) and Dakota (Kd) Formations. u> 114

Figure 26. Photo A shows the Glendale Bench Roadlandslide. Notice remediation techniques used to stabilize the landslide. This photo is facing northeast. Photo B is a view o f the Dry Wash Landslide. This landslide has a steep and arcuate shape scarp. The photo is facing northeast.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 115

Figure 27. A = Young pine tree incorporated into the Dry Wash Landslide. The pine tree survived the landslide and is now growing at an oblique angle to its original growth path. Most o f the landslide has little to no new growth Photo is facing north-northeast B = New growth on the Dry Wash Landslide. The dark green bushes are manzanita. The steep slope in the background is the head scarp o f the landslide. The gray to black layers near the skyline are coal seams. The photo is fiidng southeast. The person (Robert Kerscher) is 1.75 m (5'9") tall.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 116

Figure 28. A = Sparse and immature vegetation on the landslide (in the foreground). Along the skyline the vegetation is denser and more mature. Photo A is facing southeast. B = The dead vegetation incorporated witftin the landslide has no foliage and is highly decayed Photo B is facing south-southeast. The person (Robert Kerscher) is S'9" (1.75 m) tall

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 117

a

Figure 29. A = Wood from a tree incorporated into the landslide. Rock hammer for scale. B = The scarp and scarp graben o f the Dry Wash Landslide. The scarp graben is the topographically lower area with the large trees. These trees were once at the same elevation as the trees on the skyline. The photo faces southeast.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 118

Figure 30. The Dry Wash Landslide is rerouting streams. Person (Robert Kerscher) is standing in one o f the rerouted streams Robert is 1 75 m (S'9") tall The steep slope behind him is the toe of the landslide.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 119

Figure 31. These cross sectional models show hydrocarbon accumulation within relay ramps in extensional linkage zones. A=classic relay ramp. B=fault parallel anticline within the ramp. C=fault parallel syncline within the ramp. D=breakthrough faults within the ramp. The large filled shapes are potential hydrocarbon accumulation B positions within the relay ramp.

Legend \ I Shale or cap rock

Sandstone or reservoir rock

Limestone or source rock

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 120

114“W 113° 112° 38°N Iron Co. Garfield Co. Hurricane Fault'*i Sevier Fault

Upper Valley Field ^ Wastiington Co.i Kane County Carmel Anderson Jet. 'Orderville Field Virgin Oil Field St. George •Kanab UTAH ARIZONA

100

Kilometers

Figure 32. This map of southwestern Utah shows the oil fields in the region and the exploration holes along the Sevier fault. The solid circles represent communities. The open circles represent dry exploration holes. The half filled circles represent exploration holes that have shows of oil and/or gas. The gray filled shapes represent the oil fields in the region. Data are from Doelling et al. (1989).

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 121

REFERENCES

Arabasz, W.J., Pachmann, J.C., and Nava, S.J., 1992, The St. George (Washington County), Utah, earthquake of September 2, 1992: University of Utah Seismograph Stations, Preliminary Earthquake Report, 6 p.

Arabasz, W.J. and Julander, D R., 1986, Geometry of seismically active faults and crustal deformation within the Basin and Range-Colorado Plateau transition in Utah: Geological Society of America Special Paper 208, p. 43-74.

Armstrong, R.L., 1968, Sevier orogenic belt in Nevada and Utah: Geological Society of America Bulletin, v. 79, p. 429-458.

Armstrong, R.L. and Ward, P., 1991, Evolving geographic patterns of Cenozoic magmatism in the North American Cordilleran: the temporal and spatial association of magmatism and metamorphic core complexes: Journal of Geophysical Research, v. 96, p. 13,201-13,224.

Atwater, T.M., 1970, Implications of plate tectonics for the Cenozoic tectonic evolution of western North America: Geological Society of America Bulletin, v. 81, p. 3,581-3,536.

Atwater, T. and Stock, J., 1998, Pacific-North America plate tectonics of the Neogene southwestern United States; an update: International Geology Review, v. 40, p. 375-402.

Axen, G.J., Wernicke, B.P., Skelly, M.J., and Taylor, W.J., 1990, Mesozoic and Cenozoic tectonics of the Sevier thrust belt in the Virgin River Valley area, southern Nevada, in Wernicke, B.P., ed.. Basin and Range extensional tectonics near the latitude of Las Vegas, Nevada: Geological Society of America Memoir 176, p. 123-154.

Bamhardt, W.A. and Kayen, R.E., 1999, Internal structure of earthquake-induced landslides in Anchorage, Alaska: Seismological Research Letters, v. 70, p. 231- 232.

Barton, C M., Evans, D.J., Bristow, C.R., Freshnewy, E C., and Kirby, G.A., 1998, Reactivation of relay ramps and structural evolution of the Mere Fault and

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 122 Wardour Monocline, northern Wessex Basin: Geological Magazine, v. 135, p. 383-395.

Beghoul, N., and Barazangi, M., 1989, Mapping high Pn velocities beneath the Colorado Plateau constrains uplift models: Journal of Geophysical Research, v. 94, p. 7,083-7,104.

Best, M.G. and Grant, S.K., 1987, Stratigraphy of the volcanic Oligocene Needles Range Group in southwestern Utah and eastern Nevada, U.S. Geological Survey Professional paper 1,433-A, p. 1-28.

Best, M.G., McKee, E.H., and Damon, P E., 1980, Space-time-composition patterns of late Cenozoic mafic volcanism, southwestern Utah and adjoining areas: American Journal of Science, v. 280, p. 1,035-1,050.

Billingsley, G.H., 1993, Geologic map of the Grandstand Quadrangle, northern Mohave County, Arizona: U.S. Geological Survey Open-File Report 93-588.

Billingsley, G.H., 1994, Geologic map of the Antelope Knoll quadrangle, northern Mohave County, Arizona: U.S. Geological Survey Open-file Report 94-449.

Bird, P., 1998, Kinematic history of the Laramide orogeny in latitudes 35"-49“N, western United States: Tectonics, v. 17, p. 780-801.

Bohnenstiehl, D.R. and Kleinrock, M.C., 2000, Evidence for spreading-rate dependence in the displacement - length ratios of abyssal hill faults at mid-ocean rides: Geology, v. 28, p. 395-398.

Bowers, W.E., 1991, Geologic map of Bryce Canyon National Park and vicinity, southwestern Utah: United States Geological Survey Miscellaneous Investigations Series Map 2180, 1:24,000.

Brune, J.N., 1996, Precariously balanced rocks and ground-motion maps for southern California: Bulletin of the Seismological Society of America, v. 86, p. 43-54.

Bruhn, R.L., Gibier, P R., and Parry, W.T, 1987, Rupture characteristics of normal faults: an example from the Wasatch fault zone, Utah, in Coward, M.P., Dewey, J.F., and Hancock, P.L., eds.. Continental extension tectonics: Geological Society of America Special Publication 28, p. 337-353.

Bruhn, R.L. and Wu, D., 1993, Fault structure and earthquake properties: EOS, Transactions, American Geophysical Union, v. 74, p. 63.

Burchfîel, B.C., Cowan, D.S., and Davis, G A., 1992, Tectonic overview of the Cordilleran orogen in the western United States, in Burchfiel. B.C., Lipman, P.W., and Zoback, M L., eds.. The Cordilleran orogen: conterminous U.S.:

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 123 Geological Society of America, The Geology of North America, v. G-3, p. 407- 480.

Cartwright, J.A., Mansfield, C , Trudgill, B., 1996, The growth of normal faults by segment linkage, in Buchanan, P.G. and Nieuwland, D.A., eds.. Modem Developments in Structural Interpretation, Validation and Modeling: Geological Society of London Special Publication 99, p. 163-177.

Cartwright, J.A., Trudgill, B.D., and Mansfield, C.S., 1995, Fault growth by segment linkage: an example for scatter in maximum displacement and trace length data from the Canyonlands Grabens of SE Utah: Journal of Structural Geology, v. 17, p. 1,319-1,326.

Christenson, G.E. and Nava, S.J., 1992, Earthquake hazards of southwestern Utah, in Harty, K.M., ed.. Engineering and Environmental Geology of Southwestern Utah: Utah Geological Association Publication 21, p. 123-138.

Cladouhos, T.T. and Marrett, R., 1998, Are fault growth and linkage models consistent with power-law distributions of fault lengths?: Journal of Structural Geology, v. 18, p. 281-293.

Cowie, P.A. and Shipton, A.K., 1998, Fault tip displacement gradients and process zone dimensions: Journal of Structural Geology, v. 20, p. 983-997.

Cowie, P.A. and Scholz, C.H., 1992, Growth of faults by accumulation of seismic slip: Journal of Geophysical Research, v. 97, p. 11,085-11,095.

Crider, J.G., 2001, Oblique slip and the geometry of normal-fault linkage: mechanics and a case study from the Basin and Range in Oregon: Journal of Structural Geology, V. 23, p. 1,997-2,009.

Crider, J.G. and Pollard, D.D., 1998, Fault linkage: three-dimensional mechanical interaction between echelon normal faults: Journal of Geophysical Research, v. 103, p. 24,373-24,391.

Crone, A.J. and Haller, K.M., 1991, Segmentation and the coseismic behavior of Basin and Range normal faults: examples from east-central Idaho and southwestern Montana, U.S.A.: Journal of Structural Geology, v. 13, p. 151-164.

Cronin, V S., and Sverdrup, K.A., 1998, Preliminary assessment of the seismicity of the Malibu Coast Fault Zone, southern California, and related issues of philosophy and practice, in Welby, C.W., and Go wan, ME., eds., A Paradox of Power: Voices of Warning and Reason in the Geosciences: Geological Society of America Reviews in Engineering Geology, v. XII, p. 123-155.

Davis, G.H., 1978, Monocline fold pattern of the Colorado Plateau, in Matthews, V., ed..

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 124 Laramide folding associated with basement block faulting in the western United States: Geological Society of America Memoir 151, p. 215-233.

Davis, G.H., 1999, Structural geology of the Colorado Plateau region of southern Utah: Geological Society of America Special Paper 342, 127 p.

Davison, I., 1986, Listric normal fault profiles: calculation using bed-length balance and fault displacement: Journal of Structural Geology, v. 8, p. 209-210.

Dawers, N.H. and Underhill, J.R., 2000, The role of fault interaction and linkage in controlling synrift stratigraphie sequences: late Jurassic, StatQord East Area, northern North Sea: American Association of Petroleum Geologists, v. 84, p. 45- 64.

DeCelles, P.G., Lawton, T.F., and Mitra, G., 1995, Thrust timing, growth of structural culminations, and synorogenic sedimentation in the type Sevier orogenic belt, western United States: Geology, v. 23, p. 699-702.

dePolo, C M. and Slemmons, D.B., 1990, Estimation of earthquake size for seismic hazards, in Krinitzsky, E.L., and Slemmons, D.B., eds.. Neotectonics in earthquake evaluation: Geological Society of America Reviews in Engineering Geology, v. 8, p. 1-28.

dePolo, C M., Clark, D.G., Slemmons, D.B, and Ramelli, A.R., 1991, Historical surface faulting in the Basin and Range province, western North America: implications for fault segmentation: Journal of Structural Geology, v. 13, p. 123-136.

Dickinson. W.R., Klute, M.A., Hayes, M.J., Janecke, S.U., Lundin, E.R., McKittrick, M.A., and Olivares, M.D., 1988, Paleographic and paleotectonic setting of Laramide sedimentary basins in the central Rocky Mountain region: Geological Society of America Bulletin, v. 100, p. 1,023-1,039.

Dickinson, W.R. and Snyder, W.S., 1979, Geometry of subducted slabs related to San Andreas Transform: Journal of Geology, v. 87, p. 609-627.

Doelling, H.H, Davis, F.D., and Brandt, C.J., 1989, The geology of Kane County, Utah, Geology, Mineral Resources, Geologic Hazards: Utah Geological and Mineral Survey Bulletin 124, 192 p.

Dumitru, T.A., Duddy, LA., and Green, P.F., 1994, Mesozoic-Cenozoic burial, uplift, and erosion history of the west-central Colorado Plateau: Geology, v. 22, p. 499-502.

Eardley, A.J. and Beutner, E.L., 1934, Geomorphology of the Marysvale Canyon and vicinity, Utah: Proceedings of the Utah Academy of Sciences, Arts and Letters, v. 11, p. 149-159.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 125 Engdahl, E.R. and Rinehart, W.A., 1991, Seismicity map of North America project, in Slemmons, D.B., Engdahl, E.R., Zoback, M.D., and Blackwell, D.D., eds.. Neotectonics of North America; Geological Society of America, p. 21-27.

Faulds, J.E. and Varga, R.J., 1998, The role of accommodation zones and transfer zones in the regional segmentation of extended terranes; Geological Society of America Special Paper 323, p. 1-45.

Ferrill, D.A., Stamatakos, J.A., and Sims, D., 1999, Normal fault corrugation; implications for growth and seismicity of active normal faults: Journal of Structural Geology, v. 21, p. 1,027-1,038.

Ferrill, D A. and Morris, A.P., 2001, Displacement gradient and deformation in normal fault systems: Journal of Structural Geology, v. 23, p. 619-638.

Gibbs, A.D., 1983, Balanced cross-section construction from seismic sections in areas of extensional tectonics: Journal of Structural Geology, v. 5, p. 153-160.

Goldstrand, P.M., 1994, Tectonic development of Upper Cretaceous to Eocene strata of southwestern Utah: Geological Society of America Bulletin, v. 106, p. 145-154.

Gregory, H E., 1951, The geology and geography of the Paunsaugunt region Utah: U.S. Geological Survey Professional Paper 226, p. 1-116.

Gupta, A. and Scholz, C H., 2000, A model of normal fault interaction based on observations and theory: Journal of Structural Geology, v. 22, p. 865-879.

Harris, J.E., 1994, Anderson Junction oil field, Washington County, Utah in Schalla, R.A. and Johnson, E.H., eds.. Oil Fields of the Great Basin: Nevada Petroleum Society Special Publication, p. 377-380.

Harty, K.M., 1992, Landslide distribution and hazards in southwestern Utah, in Harty, K.M., ed.. Engineering and Environmental Geology of Southwestern Utah: Utah Geological Association Publication 21, p. 109-118.

Hecker, S., 1993, Quaternary tectonics of Utah with emphasis on earthquake-hazard characterization: Bulletin 127, p. 1-31.

Hintze, L.F., 1973, Geologic History of Utah: Brigham Young University Geology Studies 20, 181 p.

Hintze, L.F., 1988, Geologic History of Utah: Brigham Young University Geology Studies, Special Publication 7, p. 40-102.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 126 Jackson, G.W., 1990, Tectonic geomorphology of the Toroweap fault, western Grand Canyon, Arizona: implications for transgression of faulting on the Colorado Plateau: Arizona Geological Survey Open-File Report 90-4, p. 1-67.

Janecke, S.U., 1993, Structures in segment boundary zones of the Lost River and Lemhi faults, east central Idaho: Journal of Geophysical Research, v. 98, p. 16,223- 16,238.

Keefer, D.K., 1994, The importance of earthquake-induced landslides to long-term slope erosion and slope-failure hazards in seismically active regions: Geomorphology, V. 10, p. 265-284.

Keefer, D.K., 1984, Landslides caused by earthquakes: Geological Society of America Bulletin, v. 95, p. 406-421.

Keller, E.A. and Pinter, N., 1996, Active tectonics earthquakes, uplift, and landscape: Prentice Hall, Upper Saddle River, New Jersey, p. 266-269.

Kerr, H.G. and White, N., 1996, Application of an inverse method for calculating three- dimensional fault geometries and slip vectors. Nun River Field, Nigeria: American Association of Petroleum Geologists, v. 80, p. 432-444.

Koukouvelas, l.K. and Doutsos, T.T., 1996, Implications of structural segmentation during earthquakes: the 1995 Egion earthquake Gulf of Corinth, Greece: Journal of Structural Geology, v. 18, p. 1,381-1,388.

Larsen, P H., 1988, Relay structures in a lower Permian basement-involved extensional system, east Greenland: Journal of Structural Geology, v. 10, p. 3-8.

Lawton, T.F., Sprinkel, D.A., DeCelles, P.G., Mitra, G., Sussman, A.J., and Weiss, M.P., 1997, Stratigraphy and structure of the Sevier thrust belt and proximal foreland- basin system in central Utah: A transect from the Sevier Desert to the Wasatch Plateau, in Mesozoic to Recent , Link, P.K. and Kowallis, B.J., eds., Brigham Young University Geology Studies, v. 42, Part 2, p. 33-67.

Leeder, M R and Gawthorpe, R.L., 1987, Sedimentary models for extensional tilt- block/half-graben basins, in Continental Extensional Tectonics, Coward, M.P., Dewey, J.F., and Hancock, P.L., eds.. Geological Society of America Special Publication 28, p. 139-152.

Lund, W.R., 1997, Geologic hazards in the region of the Hurricane fault, in Mesozoic to Recent geology of Utah, Link, P.K. and Kowallis, B.J., eds., Brigham Young University Geology Studies, v. 42, p. 255-260.

Lund, W.R., Taylor, W.J., Pearthree, P.A., Stenner, H., Amoroso, L., and Hurlow, H., 2002 in press. Structural development and paleoseismicity of the Hurricane fault.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 127 southwestern Utah and northwestern Arizona: U.S. Geologic Survey Open-File Report.

Lundin, E.R., 1989, Thrusting of the Claron Formation, the Bryce Canyon region, Utah: Geological Society of America Bulletin, v. 101, p. 1,038-1,050.

Mandl, G., 2000, Faulting in brittle rocks: an introduction to the mechanics of tectonic faults: Springer, New York, 434 p.

Manson, D.B., 1996, Earthquake magnitude potential of the Intermountain Seismic Belt, USA, from surface-parameter scaling of late Quaternary faults: Bulletin of the Seismological Society of America, v. 86, p. 1,487-1,506.

Marshak, S. and Mitra, G., 1988. Basic methods of structural geology: Prentice Hall, Englewood Cliffs, New Jersey, 446 p.

Marzolf, J.E., 1988, Reconstruction of Late Triassic and Early and Middle Jurassic sedimentary basis: Southwestern Colorado Plateau to the Eastern , in Weide, D. and Faber, M L. eds.. This extended land: Geological Journeys in the Southern Basin and Range: Field Trip Guidebook: Geological Society of America Cordilleran Section Meeting, UNLV Department of Geoscience Special Publication 2, p. 177-200.

McCalpin, J., 1996, Paleoseismology, Academic Press, New York, 588 p.

McDougall, I. and Harrison, M.T., 1988, Geochronology and thermochronology by the Ar/^^Ar method, Oxford University Press, New York, 212 p.

Menges, C M., 1990, Late Quaternary faults scarps, mountain-front landforms, and Pliocene-Quaternary segmentation on the range-bounding fault zone, Sangre de Cristo Mountains, New Mexico, in Krinitzsky, E.L., and Slemmons, D.B., eds.. Neotectonics in Earthquake Evaluation: Geological Society of America Reviews in Engineering Geology, v. 8, p. 131-156.

Moore, J.M. and Schultz, R.A., 1999, Processes of faulting in jointed rocks of Canyonlands National Park, Utah: Geological Society of America Bulletin, v. Ill, p. 808-822.

Morley, C.K., Nelson, R.A., Patton, T.L., and Munn, S.G., 1990, Transfer zones in the East Africa rift system and their relevance to hydrocarbon exploration in rifts: American Association of Petroleum Geologists Bulletin, v. 74, p. 1,234-1,253.

Mulvey, W.E., 1992, Engineering geologic problems caused by soil and rock in southwestern Utah, in Harty, K.M., ed.. Engineering and Environmental Geology of Southwestern Utah: Utah Geological Association Publication 21, p. 139-144.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 128 Nelson, S T., and Harris, R.A., 2001, The role of rheology in the tectonic history of the Colorado Plateau, in Erskine, M.C., Faulds, J.E., Bartley, J.M., and Rowley, P.O., eds.. The Geologic Transition, High Plateaus to Great Basin - A Symposium and Field Guide: Utah Geological Association Publication 30 and Pacific Section American Association of Petroleum Geologists Guidebook GB 78, p. 189-204.

Nelson, S T. and Tingey, D.G., 1997, Time-transgressive and extension-related basaltic volcanism in southwest Utah and vicinity: Geological Society of America Bulletin, v. 109, p. 1,249-1,265.

Nikonov, A. A., 1988, Reconstruction of the main parameters of old large earthquakes in Soviet Central Asia using the paleoseismogeological method: Tectonophysics, v. 147, p. 297-312.

Papadopoulos, G.A. and Plessa, A., 2000, Magnitude-distance relations for earthquake- induced landslides in Greece: Engineering Geology, v. 58, p. 377-386.

Parsons, T., 1995, The Basin and Range Province, in Olsen, K.H., ed.. Continental Rifts: Evolution, Structure, Tectonics, p. 277-324.

Pavlides, S B., Zouros, N.C., Zhongjing, F., Shaoping, C., Tranos, M.D., and Chatzipetros, A.A., 1998, Geometry, kinematics and morphotectonics of the Yanqing-Huailai active faults (northern China): Tectonophysics, v. 308, p. 99- 118.

Peacock, D.C.P. and Sanderson, D.J., 1991, Displacements, segment linkage and relay ramps in normal fault zones: Journal of Structural Geology, v. 13, p. 721-733.

Peacock, D.C.P. and Sanderson, D.J., 1994, Geometry and development of relay ramps in normal fault systems: American Association of Petroleum Geologists, v. 78, p. 147-165.

Peacock, D.C.P., Price, S.P., Whitham, A.G., and Pickles, C.S., 2000, The world's biggest relay ramp: Hold with Hope, NE Greenland: Journal of Structural Geology, v. 22, p. 843-850.

Pechmann, J.C., Arabasz, W.L., Nava, S.J., 1992, The St. George, Utah, earthquake of September 2,1992: A normal-faulting earthquake with very weak aftershock activity: EOS (Transactions, American Geophysical Union), v. 73, p. 399.

Perrin, N.D. and Hancox, G.T., 1992, Landslide-dammed lakes in New Zealand; preliminary studies on their distribution, causes and effects, in Bell, D.H., ed.. Proceedings of the Litemational Symposium on Landslides, v. 6, p. 1,457-1,465.

Peterson, P R., 1974, Anderson Junction field: Utah Geological and Mineral Survey Oil and Gas Field Studies 13, 8 p.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 129

Proüiero, D.R. and Schwab, F., 1996, Sedimentary geology an introduction to sedimentary rocks and stratigraphy, W.H. Freeman and Company, New York, p. 47.

Rahn, P.H., 1996, Engineering geology; An environmental approach, Prentice Hall, Upper Saddle River, p. 162-205.

Ramsay, J.G., 1967, Folding and fracturing of rocks, McGraw-Hill, New York, 188 p.

Reber, S., Taylor, W.J., Stewart, M., and Schiefelbein, I.M., 2001, Linkage and reactivation along the northern Hurricane and Sevier faults, southwestern Utah, in Erskine, M.C., Faulds, J.E., Bartley, J.M., and Rowley, P.M., eds.. The geologic Transition, High Plateaus to Great Basin - a symposium and field guide; Utah Geological Association Publication 30 and Pacific Section American Association of Petroleum Geologists Guidebook GB 78, p. 379-400.

Rojstaczer, S.A. and Wolf, S.C., 1992, Permeability changes associated with large earthquakes; an example from Loma Prieta, California; Geology, v. 20, p. 211- 214.

Rowland, S.M. and Deubendorfer, E.M., 1994, Structural analysis and synthesis, a laboratory course in structural geology, 2"** edition: Blackwell Scientific Publications, Boston, 279 p.

Rowley, P.D., Anderson, J.J., and Williams, P.L., 1975, A summary of Tertiary volcanic stratigraphy of the southwestern High Plateaus and adjacent Great Basin, Utah: U.S. Geological Survey Bulletin 1405-B, p. B1-B20.

Rowley, P.D., Steven, T.A., Anderson, J.J., and Cunningham, C.G., 1979, Cenozoic stratigraphie and structural framework of southwestern Utah: U.S. Geological Survey ^ofessional Paper 1149, p. 1-22.

Rowley, P.D., Steven, T.A., and Mehnert, H.H., 1981, Origin and structural implicatoins of upper Miocene rhyolites in Kingston Canyon, Piute County, Utah: Geological Society of America Bulletin, v. 92, p. 590-602.

Sargent, K.A. and Philpott, B.C., 1987, Geologic map of the Kanab quadrangle, Kane County, Utah, and Mohave and Coconino Counties, Arizona: U.S. Geological Survey Map GQ-1603.

Schiefelbein, I.M. and Taylor, W.J., 2000, Fault development in the Utah Transition Zone and High Plateaus Subprovince: Geological Society of America Abstracts with Programs, v. 32, p. 194.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 130 Scholz, C.H., 1990, The Mechanics of Earthquakes and Faulting, Cambridge University Press, New York, 438 p.

Schroder, J.F., 1971, Landslides of Utah: Utah Geological and Mineralogical Survey Bulletin 90, S1 p.

Smith R.B. and Arabasz, W.J., 1991, Seismicity of the Intermountain Seismic Belt, in Slemmons, D.B., Engdahl, E.R., Zoback, M.D., and Blackwell, D.D., eds.. Neotectonics of North America: Geological Society of America, p. 185-228.

Smith, R.B. and Sbar, M L., 1974, Contemporary tectonics and seismicity of the western United States with emphasis on the Intermountain Seismic Belt: Geological Society of America Bulletin, v. 8, p. 1,205-1,218.

Smith, E.I., Sanchez, A., Walker, D., Wang, K., 1999, Geochemistry of mafic magmas in the Hurricane Volcanic Field Utah, implications for small- and large-scale chemical variability of the lithospheric mantle: Journal of Geology, v. 107, p. 433-448.

Song, T. and Cawood, P. A., 2001, Effects of subsidiary faults on the geometric construction of listric normal fault systems: American Association of Petroleum Geologists, V. 85, p. 221-232.

Stenner, H.D., Lund, W.R., Pearthree, P.A., and Benjamin, L.E., 1999, Hurricane fault in northwestern Arizona and southwestern Utah: Arizona Geological Survey Open- File Report 99-8, p. 1-136.

Stenner, H.D., Lund, W.R., Pearthree, P.A., and Everitt, B.L., 1999, Paleoseismologic investigations of the Hurricane fault in northwestern Arizona and southwestern Utah, Arizona Geological Survey Open-File Report 99-8, 137 p.

Stewart, M E. and Taylor, W.J., 1996, Structrual analysis and fault segment boundary identification along the Hurricane fault in southwestern Utah: Journal of Structural Geology, v. 18, p. 1,017-1,029.

Stewart, M E., Taylor, W.J., Pearthree, P.A., Solomon, B.J., and Hurlow, H.A., 1997a, Neotectonics, fault segmentation, and seismic hazards along the Hurricane fault in Utah and Arizona: an overview of environmental factors in an actively extending region: Brigham Young University Geology Studies, v. 42, p. 235-260.

Stewart, M E., Taylor, W.J., Pearthree, P.A., Solomon, B.J., and Hurlow, H.A., 1997b, Field guide to neotectonics, fault segmentation, and seismic hazards along the Hurricane fault in southwestern Utah and northwestern Arizona: Brigham Young University Geology Studies, v. 42, p. 261-273.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 131 Stokes, W.L., 1988, Geology of Utah, Utah Museum of Natural History, University of Utah and Utah Geological and Mineral Survey Department of Natural Resources, 280 p.

Sveringhaus, J. and Atwater, T., 1990, Cenozoic geometry and thermal state of the subducting slabs beneath western North America, in Wernicke, B.P. ed.. Basin and Range Extensional Tectonic near the latitude of Las Vegas, Nevada: Geological Society of America Memoir 176, p. 1-22.

Taylor, W.J., Stewart, M E., Omdorff, R.L., 2001, Fault segmentation and linkage: examples from the Hurricane fault, southwestern U.S.A., in Erskine, M E., Faulds, J.E., Bartley, J.M., and Rowley, P.D., eds.. The geologic Transition, High Plateaus to Great Basin - a symposium and field guide: Utah Geological Association Publication 30 and Pacific Section American Association of Petroleum Geologists Guidebook GB 78, p. 113-126.

Taylor, W.J., 1993, Stratigraphie and lithologie analysis of the Claron Formation in southwestern Utah: Utah Geological Survey Report 93, 152 p.

Taylor, W.J., 1992, Paleogene fluvial and lacustrine deposits in southwestern Utah and their tectonic implications: examination of the Claron Formation; stratigraphie and lithologie analysis of the Claron Formation in southwestern Utah: Utah Geological and Mineral Survey Miscellaneous Publication, p. 1-33.

Trudgill, B. and Cartwright, J., 1994, Relay-ramp forms in normal-fault linkages, Canyonlands National park, Utah: Geological Society of America Bulletin, v. 106, p. 1,143-1,157.

Van Kooten, G.K., 1988, Structure and hydrocarbon potential beneath the Iron Springs laccolith, southwestern Utah: Geological Society of America Bulletin, v. 100, p. 1,533-1,540.

Vendeville, B., 1991, Mechanism generating normal fault curvature: a review illustrated by physical models, in Roberts, A.M., Yielding, G., and Freeman, B., eds.. The Geometry of Normal Faults: Geological Society Special Publication 56, p. 241- 249.

Walsh, J.J., Watterson, J., Bailey, W.R., and Childs, C., 1999, Fault relays, bends and branch-lines: Journal of Structural Geology: v. 21, p. 1,027-1,038.

Wannamaker, P.E., Bartley, J.M., Sheehan, A.F., Jones, C M., Lowrey, A.R., Dumitru, T.A., Ehlers, T.A., Holbrook, W.S., Farmer, G.L., Unsworth, M.J., Hall, D.3., Chapman, D.S., Okaya, D.A., John, B E., and Wolfe, J.A., 2001, Great Basin- Colorado Plateau transition in central Utah: an interface between active extension and stable interior, in Erskine, M E., Faulds, J.E., Bartley, J.M., and Rowley, P.D., eds.. The geologic Transition, High Plateaus to Great Basin - a symposium

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 132 and fîeld guide: Utah Geological Association Publication 30 and Pacific Section American Association of Petroleum Geologists Guidebook GB 78, p. 1-38.

Wheeler, R.L., 1989, Persistent segment boundaries on basin-range normal faults: U.S. Geological Survey Open File Report OF 89-0315, p. 423-444.

White, N. and Yielding, G., 1991, Calculating normal fault geometries at depth; theory and examples, in Roberts, A.M., Yielding, G , and Freeman, B., eds.. The Geometry of Normal Faults: Geological Society Special Publication 56, p. 251- 260.

White, N., 1992, A method for automatically determining normal fault geometry at depth: Journal of Geophysical Research, v. 97, p. 1,715-1,733.

Williams, G. and Vann, I., 1987, The geometry of listric normal faults and deformation in their hanging walls: Journal of Structural Geology, v. 9, p. 789-795.

Wu, D. and Bruhn, R.L., 1994, Geometry and kinematics of active normal faults. South Oquirrh Mountains, Utah; implication for fault growth: Journal of Structural Geology, v. 16, p. 1,061-1,075.

Young, M.J., Gawthorpe, R.L., and Hardy, S., 2001, Growth and linkage of a segmented normal fault zone; the Late Jurassic Murchison - Statfjord North Fault, northern North Sea: Journal of Structural Geology, v. 23, p. 1,933-1,952.

Zampieri, D., 2000, Segmentation and linkage of the Lessini Mountains normal faults. Southern Alps, Italy: Tectonophysics, v. 319, p. 19-31.

Zhang, P., Slemmons, D.B., and Mao, F., 1991, Geometric pattern, rupture termination, and fault segmentation of the Dixie Valley-Pleasant Valley active normal fault system, Nevada, U.S.A.: Journal of Structural Geology, v. 13, p. 165-176.

Zhang, P., Mao, F., and Slemmons, D.B., 1999, Rupture terminations and size of segment boundaries from historical earthquake ruptures in the Basin and Range Province: Tectonophysics, v. 308, p. 37-52.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. VITA

Graduate College University of Nevada, Las Vegas

Hsa M. Schiefelbein

Address: UNLV Department of Geoscience 4505 Maryland Pkwy Box 454010 Las Vegas, Nevada 89154-4010

Degrees: Bachelor of Science, Geology, 1998 University of Wisconsin - Milwaukee

Special Honors and Awards: AAPG Research Grant, GSA Research Grant, Sigma Xi Research Grant, Arizona/Nevada Academy of Science Research Grant, UNLV Graduate Student Association Research Grant, UWM Travel Grant to the GSA Cordilleran Section Meeting, Two GSA Cordilleran Section Student Travel Grants, UWM Athletic Scholarship, Indiana University Field Camp Scholarship, UWM Field Camp Scholarship, ARCO Undergraduate Research and Travel Grant, Elk’s Club Outstanding Student Scholarship, Jefferson School Parent/Teacher Outstanding Student Scholarship, UNLV Geoscience Dept. Scholarship, UNLV Graduate College Summer Session Scholarship, Two GSA Engineering Geology Division Shlemon Awards

Publications: Schiefelbein, I.M., Buck, B.J., Lato, L., and Merkler, D., 2002 in press. Salt micromorphology in arid soils of southern Death Valley, California, USA: Proceedings in the 17"’ World Congressional Soil Conference, Bangkok, Thailand. Schiefelbein, I.M., 2002, Fault segmentation, linkage, and earthquake hazards along the Sevier fault, southwestern Utah: M.S. Thesis, University of Nevada Las Vegas. Reber. S., Taylor, W.J., Stewart, M E., Schiefelbein, I.M., 2001, Linkage and reactivation along the northern Hurricane and Sevier faults, southwestern Utah: The Geologic Transition, High Plateaus to Great Basin - A Symposium and Field Guide, Utah Geological Association Publication 30 and Pacific Section AAPG Guidebook GB 78, p. 379-400. Schiefelbein, I.M. and Taylor, W.J., 2001, Development of fault-parallel folds in relay ramps within overlapping normal fault linkage zones, southwestern Utah: AAPG and Society of Exploration Geophysists Student Expo Abstracts.

133

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. Schiefelbein, I.M. and Taylor, WJ., 2001, Fault geometry and linkage along normal faults using the Sevier fault, southwestern Utah, as a case study: GSA Abstracts with Programs, v. 33, p. A25. Schiefelbein, I.M. and Taylor, WJ., 2001, Fault segmentation, linkage, and earthquake hazards along the Sevier fault, SW Utah: GSA Abstracts with programs, v, 33, p. A-38. Taylor, WJ., Criscione, JJ., Gilbert, J.J., Justet, L., Kula, J.L., Schiefelbein, I.M., Sheely, J.C., and Stickney, E.K., 2001, Geometry and neotectonics of the northern California Wash fault, southern Nevada: GSA Abstracts with programs, v. 33, p. A-58. Schiefelbein, I.M., 2000, Fault segmentation, linkage, and earthquake hazards along the Sevier fault, southwestern Utah: AAPG Bulletin, v. 84, p. 1873. Bailley, T.L., Gilbert, J.J., Howley, R.A., Rees, M.N., Schiefelbein, i.M., Taylor, W.J., and Van Hoesen, J.G., 2001, Geologic guide to Red Rock Canyon: Integrating research and education: Arizona-Nevada Academy of Science Abstracts with Programs, v. 45, p. 19. Schiefelbein, I.M. and Taylor, W.J., 2000, Fault development in the Utah Transition Zone and High Plateaus Subprovince: GSA Abstracts with Programs, v. 32, p. 194. Bass, C., Schiefelbein, I.M., Cronin, V S., and Dill, R.F., 1999, Near-shore seismic survey along eastern Malibu coastline, southern California: GSA Abstracts with Programs, v. 31, p. 76. Schiefelbein, I.M., Walsh, K., Boldt, J., Riedel, J., Cronin, V S., and Dill, R.F., 1998, Looking for unmapped strands of the Malibu Coast Fault, southern California; GSA Abstracts with Programs, v. 30, p. 46.

Thesis Title: Fault Segmentation, Linkage, and Earthquake Hazards Along the Sevier Fault, Southwestern Utah

Thesis Examination Committee: Chairperson, Dr. Wanda J. Taylor, Ph.D. Committee Member, Dr. Eugene I. Smith, Ph.D. Committee Member, Dr. Terry Spell, Ph.D. Graduate Faculty Representative, Dr. Barbara Luke, Ph.D.

134

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. NOTE TO USERS Oversize maps and charts are microfilmed in sections in the following manner

LEFT TO RIGHT, TOP TO BOTTOM, WITH SMALL OVERLAPS

UMI

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. Plate 1 - Geologic

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. c Map of the Northern ! lisa !

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. p Section of Long Valle a Schiefelbein 2002

N

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. ey, Southwestern Utah

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 1

i t

I t

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. t {( 4 V/. x :

rX: A irr Û' I :

x A . X // M r// J % ri 9: .-X' m x ' X-. //: I ! % 1 >/ X r / / / r MU / / ,// uli( V. -T-y ■ - - -T 'M Qas m i

T f | x 8 enc h ; i " 81 -1 \ u i r r w 17 : X.'V )) "'ÀA \/ * * - u. A \\ 1 A U \ X, / V.J O n X. / /y N V \\ % o y }

—C-. 1 X. ■ \ A.m : ! 7 1 X u ; / X //f/ L 'A « (eijii! #/ 1 1 y r % ^ :.x A A \ X: X— (U / .Û 'i / y '/ ) k Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. A

«

Q

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. s

% Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. N A

f

Kilometers 0 1 & Miles 0.5 1

Scale = 1:12,000

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. v\l V \ \¥ to i,g n ''

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. »

W ié

Area of overlap

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. Kilometers 0 0.5 1

Mlles 0.5

Scale = 1:12,000

p with Plate 2 is below dashed black line.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. NOTE TO USERS Overeize maps and charts are mkreMlmed in sections in die following manner

LEFT TO RIGHT, TOP TO BOTTOM, WITH SMALL OVERLAPS

UMI

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. Plate 2 - Geologic M

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. Map of the Southern Si lisa Scl 2

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. Section of Long Valley, >chiefelbein 2002

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. ey, Southwestern Utah

1 } I

Area of overlap with Plate 1 is above ti

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. stem Utah

erlap with Plate 1 is atx>ve dashed black line.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. é

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. i

1

644.4

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. m

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. m

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. N À

Kilometers 0 0.5 1

Miles 0.5

Scale = 1:12,000

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. NOTE TO USERS Oversiz» maps and charts ars microfilmed In sections In the following manner

LEFT TO RIGHT, TOP TO BOTTOM, WITH SMALL OVERLAPS

uivn

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. Descrii

Surficial Deposits

Quaternary Landfill Q lf Kane County Landfill. Actively used.

Quaternary Alluvium Qa Weakly indurated to nonindurated deposits in active washes and floodpiains. Deposit types v from sandstone (yellow, pinkish orange, white), limestone (yellowish gray to gray, pink), and sii (brown) to sand, gravel, and developed soils. Clast (within the deposit) sizes range from silt to boulder and shapes vary from rounded to angular. Much of this unit is occupied by humans o actively cultivated. The contacts are sharp. Thickness of beds or unit is unknown.

Quaternary Alluvium Second Qas Weakly indurated to nonindurated deposits in active washes and floodpiains. Deposit is a developed soil with active washes. The parent material is basalt. Clast (within the deposit) size range from silt to pebble and shape is well-rounded. The pebbles are derived from the basal Tertiary Claron Formations. Much of this unit is occupied by humans or actively cultivated. Th< contacts are sharp. Thickness of beds or unit is unknown.

Quaternary Landallde Qls Includes all slope failures. Size ranges from 228.6 - 243.8 m (750 - 800 ft) In diameter. Associate with steep slopes or weakly indurated units, internal blocks range from silt size to greater than m (3 ft). Clasts are generally rotated, and trees and bushes are uprooted creating hummocky topography. Classified as a rotational slump.

Quaternary Colluvium 1 Qcl Quartz sand. Color is pink to white. Grain size is medium to fine. Unconsolidated. Eroded fronr adjacent Navajo Sandstone cliffs.

Quaternary Colluvium 2 Qc2 Unconsolidated clasts of platy limestone. Color Is gray to pale yellow on both the fresh and w eathered surfare«. Clasts are 5 - 10 rm 10 - A ini ufiHe and —9 rm r~n 7S inl thlrlr AnmiUr Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. escriptîons of Map Units

Cretaceous KaiparowitsyWahvreap Formations Undivided Kkw The unit consists of alternating sandstone, mudstone, siitstone, and conglomerate color is pale yellow to white on both the fresh and w< cemented. Contains quartz, mafic crystals, chert, feldspar, and clay. < • 12.7 cm (2 - 5 in) thick. Also contains ripple marks. The siitstone to color on both fresh and weathered surfaces. Thickness varies from ( alternating with the sandstone and siitstone to mudstone are congl conglomerate layers are 0.2 - 0.9 m (05 - 3 ft) thick. Layers are disco )sit types vary lenses. Generally clast supported. Clasts are very well-rounded and nk),and sllt diameter. Clasts are dominantly chert, but also limestone and doios rom sllt to Dominantly a cliff forming unit. However, some of the mudstones a humans or Total thickness = 457 m (1500 ft).

Cretaceous Straight Ciifte Formation Ksc The upper rock type in the unit Is a lithic sandstone. Colors are pale bsit is a drab yellow on the weathered surface. Coarse to medium grained ieposit) sizes well as shrimp burrows, cephlapods, pelecypods,oysters, and shark the basal in) iron concretions. Upper contact is sharp. Forms steep cliffs. Roc Itlvated. The lower unit are siitstone interbedded with coal. The siitstone is drab weathered surfaces with dark gray to black coal beds. The coal bee -305 cm (1 ft) thick. The lower contact is gradational. Poorly lithifi Total thickness = 73.2 m (240 ft).

>r. Associated Cretaceous Tropic Shale Kt Mudstone, siitstone, and sandstone. The top unit is sandstone. Coli jreater than 0.9 weathered surfaces. Grain size is medium to fine. The lower part o ummocky mudstone and siitstone. Olive gray to drab green on both fresh an coarsens upward. The upper and iower contacts are gradational, poorly exposed 05 m (15 ft) thick bed of septrian nodules in a yel Nodules range from 10.2 - 30.5 cm (4 -12 in) in diameter and the o Some of the nodules are fractured. The interior contains yeliow ca Eroded from size. The formation is poorly lithified and weathers to form slopes m (760 ft).

Cretaceous Dakota Sandstone Kd Limestone, mudstone, and shale. The upper part of the unit is lime both fresh and weathered surfaces. Abundant oysters. The upper resh and bed Is sharp. Generally forms float covered surfaces. The lower pa \nauiarto fin é t Alt kAth fm ch anH uüaathArarl ci iifarat f*nntainc tum di« Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. Plate 3

ndivided Itstone.and conglomerate. Sandstone and Fresh and weathered surfaces. Calcite ir,and clay. Cross beds are low angle and 5.1 siitstone to mudstone is pink to purple in raries from 0.9 - 3.7 m (3 -12 ft) thick. Also ie are conglomerate layers and lenses. The fs are discontinuous and commonly form )unded and 13 - 25 cm (05 -1 in) in e and dolostone. Contacts are sharp, ludstones and siltstones weather to a slope.

ors are pale yellow on the fresh surface and m grained. Contains low angle cross-beds as ,and shark teeth. Contains 2.5 - 5.1 cm (1 - 2 3 cliffs. Rock types of the poorly exposed 3ne is drab gray on both fresh and coal beds are splintery, thinly bedded, and Dorly lithified and weathers to form a slope.

stone. Color is yellowish gray both fresh and /ver part of the unit is alternating layers of h fresh and weathered surfaces. Grain size iational. 183 m (60 ft) from the base is a es in a yellowish gray marine mud matrix, and the outer parts are light gray limestone, yellow calcite crystals up to 1.3 cm (05 in) in m slopes and valleys. Total thickness = 231.7

nit is limestone. Yellowish green to gray on he upper and lower contacts of the oyster lower part is shale to mudstone. Color is 1 C tu r n r llc m n tim irki ic m a l h a H c T k a ^ n a l Ic [ Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. Legend

Lithologie contact. Dashed where approxim; where concealed.

1 _ Fault contact. Dashed where approxiamtely I concealed. ? where the fault existance is une the hangingwall.

7,0 Strike and dip of bedding.

A. 20 G" 35 A = Anticline showing trace of axial surface a Syncline showing trace of axial surface and p

Age Correlation of Map Units

Surficial Deposits Qlf Qa Qls Qcl Qc2 Qas H olocene 1 Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. Legend

Lithologie contact. Dashed where approximately located. Dotted where concealed.

Fault contact. Dashed where approxiamtely located. Dotted where concealed. ? where the fault existance is uncertain. Ball and bar are on the hangingwall.

70 Strike and dip of bedding.

B. 35 A = Anticline showing trace of axial surface and plunge of axis. B = Syncline showing trace of axiai surface and plunge of axis.

Age Correlation of Map Units

)eposlts Qcl Qc2 Qas H olocene 1 Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. Quaternary Colluvium 2 Qc2 Unconsolidated clasts of platy limestone. Color Is gray to pale yellow on both the fresh and weathered surfaces. Clasts are 5 -10 cm (2 - 4 in) wide and ~2 cm (-0.75 in) thick. Angular t< subangular. Transport direction is downsiope movement. Source is the nearby Co-op Creek Member of the Carmel Formation.

Long Valley Quaternary Spring Deposit Qs Spring deposit, tan to brown on both fresh and weathered surfaces. Fine grained. Cemented with silica and locally small amounts of cartxmate. Millimeter to centimeter nearly planar beds of silica possibly opaline. Weathers to a resistant high. Associatec springs. Based on field observations is classified as thin-bedded opacline sinter facie; (Jarvis et al., 1998; Campbell et al., 2(X)1 ; Braunstein and Lowe, 2001 ). Total thicknes 12.2 m (0 -4 0 ft).

Young Quaternary Conglomerate Qcg Clast supported conglomerate. Framework Is composed of poorly sorted consolidated conglomerate and sandstone. Clast lithologles Include limestone (yellow, gray, pink), sandstc (yellow to gray), basalt (dark gray to brown), and petrified wood (light to dark brown and bla Clasts are rounded to angular. Size of clasts ranges from sand to greater than 15 m (5 ft). Co 305 cm (1 ft) thick layers of medium to coarse-grained llthic sandstone. Color Is tan on weat surfaces and light tan to yellow on fresh surfaces. Sandstone locally contains low-angle cros Upper and lower contacts are sharp. Weathers to rubbly slope or cliff former. Total thickness 9 m (3-30 ft).

Old Quaternary Conglomerate Qcgl Clast supported conglomerate. Framework Is composed of poorly sorted consolidated sand Clast lithologles Include limestone (yellow,gray, pink), sandstone (yellow to gray),and basalt gray to brown). Clasts are rounded to angular. Size of clasts ranges from sand to -0.5 m (-1. Contains 305 cm (1 ft) thick layers of medium to coarse-grained llthic sandstone. Color Is tar weathered surfaces and light tan to yellow on fresh surfaces. Locally contains low-angle cro< Upper and lower contacts are sharp. Weathers to rubbly slope or cliff former. Outcrops topographically higher than Qcg. Total thickness Is -12 m (-40 ft).

Bedrock

Quaternary Basalt Qb Basalt, brown to black on weathered surfaces and dark gray on fresh surfaces. Fine crystallln Phenocrysts of olivine -5 mm (0.125 In), plagloclase -2.5 mm (0.0625 In), and pyroxene. VesI bottom 0.9 -1 3 m (3 - 6 ft), dense middle 4.6-13.7 m (15-45 ft), and vesicular top 0.9 -1.8 nr 6 ft). Weathers to a resistant cap unconformably overlying Quaternary,Tertiary, and Cretacec rocks. Unit thickness Is 6.1 -183 m (20 - 60 ft).

Tertiary Claron Formation Tc Limestone, conalomerate, and sandstone to siitstone. Subdivided Into two members - uooe Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. Cretaceous Dakota Sandetome Kd Limestone, mudstone, and shale. The upper part of the unit Is llm both fresh and weathered surfaces. Abundant oysters. The uppe fresh and bed Is sharp. Generally forms float covered surfaces. The lower p Angular to olive gray on both fresh and weathered surfaces. Contains two d -op Creek dark gray to black on both fresh and weathered surfaces, splinter contains selenite gypsum. Gypsum crystals occur In 25 - 5.1 cm I size from 25 - 5.1 cm (1 - 2 In) In length. The unit Is generally poc forms valleys. The lower contact Is sharp and the upper contact I: lined. m (320 ft). sntimeter thick ssociated with Iter facies Jurassic Carmsi Formation - Winsor Member Jew I thickness 0 - Conglomerate, silty fine-grained sandstone, and siitstone. The up Pale yellow on both fresh and weathered surfaces. Generally mal sand. Clasts are well-rounded, gravel size, and composed of chert fine-grained, yellow to white sandstone with well-rounded grain; of alternating beds of silty fine-grained sandstone and siitstone. brown. Beds are -0.9 m (-3 ft) thick. Upper and lower unit cont* ated weathers to a slope or forms valleys. Total thickness = 853 m (28 k), sandstone m and black). 1 (5 ft). Contains Jurassic Carmel Formation • Paria River Member n on weathered Limestone, gypsum, shale, and siitstone. The upper unit Is fine-gr angle cross beds, surface and very light gray fresh surface. Abundant pelecypods. thickness Is 0.9 - poorly exposed. Lower unit has three alabaster gypsum layers se siitstone. The gypsum Is dirty white on the weathered surface an fresh surface. Has a sugary texture. Weathers to blocky ledges. E shale and siitstone are greenish gray and reddish brown on both vary In thickness from 2.5 - 7.6 cm (1 - 3 In). Contacts between thi ated sandstone, sharp. Upper and lower contacts are gradational. Weathers to sk md basalt (dark thickness = 73.2 m (240 ft). 0.5 m (-15 ft), lolor Is tan on angle cross beds, Jurassic Carmsi Formation - Crystai Creek Member reps Jcc Fine-grained siitstone, sandy shale to sandstone, and limestone. 1 siitstone with sandy shale to sandstone Interbeds. Colors range fi gray both fresh and weathered surfaces. Weakly Indurated. Abur stringers. Forms valleys and slopes. Upper and lower contacts an thick-bedded, fine-grained limestone. Colors are dark yellow on v on fresh surfaces. Cliff to ledge former. Total thickness = 51.8 m (

! crystalline, Jurassic Carmsi Formation - Co-op Creek Member xene. Vesicular Jco Sandy siitstone to siitstone and limestone. The upper unit Is lime: 0.9-1.8 m (3- and form small cliffs. The basal 13 m (6 ft) are a sandy siitstone tc i Cretaceous and red on both fresh and weathered surfaces. The bedding layei thick. Forms slopes. Under the upper limestone Is a fine grained I yellow on both fresh and weathered surfaces. The limestone Is di of reddish siitstone. Lower limestone layers are platy with calcite between. The bedding thickness Is -2 5 cm (-1 In). Upper conta^ Is sharp. Pentacrlnus, bryozoans, and oysters are common. WeatH ers - upper white thickness = 121.9 m (400 ft). I Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. unit Is limestone. Yellowish green to gray on The upper and lower contacts of the oyster ie lower part Is shale to mudstone. Color Is alns two discontinuous coal beds. The coal is is, splintery, thinly bedded, and typically > - 5.1 cm (1 - 2 in) beds. The crystals range in lerally poorly lithified, weathers to a slope, and r contact is gradational. Total thickness = 975

ie. The upper unit Is a conglomerate layer, lerally matrix supported. Matrix Is clay to fine ;d of chert and quartzlte. The middle unit Is a ded grains. The lower unit rock types consist siitstone. Colors range from brick red to unit contacts are gradational. Member 153 m (280 ft).

r t Is fine-grained limestone. Gray weathered ecypods. Forms a resistant ledge. Unit Is 1 layers separated by layers of shale and surface and white to locally motteled on the ledges. Each layer Is -0.9 m (-3 ft) thick. The n on both fresh and weathered surfaces. Beds îtween the gypsum, shale and siitstone are thers to slopes and rounded hills. Total

bar nestone. The upper unit Is a alternating rs range from red, brown, yellow, and greenish ted. Abundant gypsum talus and light gray intacts are gradational. The bottom of unit Is ellow on weathered surfaces and pale yellow = 51.8 m (170 ft).

er nit Is limestone. Beds are 0.9 m (3 ft) thick iltstone to siitstone. Colors are grayish green Iding layers vary from 25 - 91.4 cm (1 - 36 In) ? grained limestone. Color Is pale yellow to stone Is divided by a 10.2 cm (4 In) thick layer ith calcite layers 1.3 cm (05 In) thick 3er contact Is gradational and lower contact on. Weathers to a platy talus slope. Total

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. Surficial Deposits

Qlf Qa Qls Qcl Qc2 Qas Holocene

Qcg1 Qb Pleistocene

Unconformity

Bedrock

Tc

Unconformity

Kkw

Ksc

Kt

Kd

K-1 Unconformity

Jew

Jcc

Jco

J-1 Unconformity __ j Jn I 1 I Triassic Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. Deposits

Qcl Qc2 Qas Holocene Q uaternary

Pleistocene

nform lty drock 3 Tertiary nform lty

3 C retaceous 3 3 conformity 3

Jurassic 3 Ç]

onform lty : ] Triassic-Jurassic Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. Bedrock

Quaternary BawK Basalt brown to black on weathered surfaces and dark gray on fresh surfaces. Fine crystalll Phenocrysts of olivine ~5 mm (0.125 in), plagioclase -2.5 mm (0.0625 in), and pyroxene. Ve bottom 0.9-13 m (3-6 ft), dense middle 4.6 -13.7 m (15-45 ft), and vesicular top 0.9 -1.8 6 ft). Weathers to a resistant cap unconformably overlying Quaternary, Tertiary, and Cretacc rocks. Unit thickness is 6.1 -183 m (20 - 60 ft).

Tertiary Claron Formation Limestone, conglomerate, and sandstone to siitstone. Subdivided into two members - upp and lower pink. The white member is a thick to medium bedded fresh water limestone. Co gray on the weathered surface and chalky white on the fresh surface. Contacts are sharp. ( cavities [25 - 7.6 cm (1 - 3 In) In diameter] filled with calcite. Thickness is 91.4 m (300 ft). Th the lower pink unit Is a sandy limestone. Pink to red on the weathered surface and light pir the fresh surface. The lower part of the unit contains scattered pebbles of Precambrian to Paleozoic quartzlte, chert, limestone, dolostone. Pebbles are very well-rounded. The lower | member Is fluvlal-lacustrlne In origin (HIntze, 1988; Doelling et al., 1989). Contacts are shaq; Thickness Is 152.4 m (500 ft). Unit weathers to hoodoos, steep cliffs, or steep slopes. Total tl = 243.8 m (800 ft).

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. auinger». rorms vaiieysana smpes. upper ana lower contacts « thick-bedded, fine-grained limestone. Colors are dark yellow or on fresh surfaces. Cliff to ledge former. Total thickness = 51.8 rr

e crystalline, Jurassic Carmsi Formation - Co-op Crssk Mambsr ixene. Vesicular Jco Sandy siitstone to siitstone and limestone. The upper unit is lim ) 0.9 -1.8 m (3 - and form small cliffs. The basal 13 m (6 ft) are a sandy siitstone id Cretaceous and red on both fresh and weathered surfaces. The bedding lay thick. Forms slopes. Under the upper limestone is a fine graine* yellow on both fresh and weathered surfaces. The limestone is < of reddish siitstone. Lower limestone layers are platy with calcit between. The bedding thickness is -2 5 cm (-1 in). Upper conti is sharp. Pentacrinus, bryozoans, and oysters are common. Weat >ers-upper white thickness = 121.9 m (400 ft). stone. Color is e sharp. Contains 100 ft). The top of Jurassic Navajo Sandstone d light pink on Jn Well-rounded, well-sorted, fine-grained, friable, quartz sandston* brian to and/or calcite. Red, orange, tan, to white on both fresh and weal he lower pink Bedding thickness ranges from 0.9 - 9.1 m (3 - 30 ft). Upper and are sharp. Contains 1.3 -25 cm (0.5 -1 in) red to orange oxidation spots or :s. Total thickness Forms cliffs. Total thickness = 609.6 m (2000 ft).

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. Kontacts are gradational, rne Dotcom or unit is I yellow on weathered surteces and pale yellow Is = 51.8 m (170 ft). nber r unit is limestone. Beds are 0.9 m (3 ft) thick y siitstone to siitstone. Colors are grayish green >edding layers vary from 25 - 91.4 cm (1 • 36 in) ine grain ^ limestone. Color is pale yellow to nestone is divided by a 10.2 cm (4 in) thick layer with calcite layers 1.3 cm (05 in) thick Jpper contact is gradational and lower contact I mon. Weathers to a platy talus slope. Total

z sandstone cemented with hematite, silica, ih and weathered surfaces. Cross-I^ded. Upper and lower unit contacts are sharp, on spots on both fresh and weathered surfaces.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. Jcc

Jco

J-1 Unconformity Jn ]

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. Jurassic iJcc

Jco iconformlty

Jn Tnassic - Jurassic

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. NOTE TO USERS Oversize maps and charts are microfilmed in sections in the following manner

LEFT TO RIGHT, TOP TO BOTTOM, WITH SMALL OVERLAPS

UMI

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. Plate 4a

8000

6000

5000

4000

3000

2000

1000

Reproduced with Pemiission of the copyrnght o'vner. Further reproduction prohibited without permission. 3 - Cross Sections

A’

8000 B 8000 Qb Qb

7000 7000

Qa 60 0 0 Jew 6 0 0 0

Jco 5000 500 0

4000 4000

3000 3000

2000

Jm 2000

1000 1000 permission of the copyright owner. Further reproduction prohibited w ithout ® Mapped Area

Qa Qa

f

Ks

Hi

Ha

Jew Jep Joe Jco

Jn

f^eprod "<=WwWpem,/ssio„ 4» “ pyrigp, owner. further ''^production permission. 8000 Q a

7000

Jew

Joe 6000 Jco

5 0 0 0

4 0 0 0

3000

Jm

2000

,,Zrrepro,u,„o„ 1000

-1000

7000

Qcg 6000 Ks

5000 Kd

JOM JCP Joe 4000 Jco

3000

2000

m Jip. 1000

Jm Jm 0,.ecopv...ow ner. P.«.e.ep™ ..«,o„p.o«eaw».o„.pe™ :ss,on^

Reproducedwith permission l e 1000 1000

Im

-1000 -1000

Kkw Sa Ks Ks Ks Ks Kl

Kl Kl Kl Kd Kd Kd Jew Kd Jew Jcp Jew Jcc Jew Jûp Jeo Jcp Jcc Jcc

rJeo. Jeo Jco

Jn Jn Jn

■JUL J#P Jk Jlp yip j|

Jk Jm Jm

l e l e Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. Jm

Ic

fm

Qa C' ^000

%

Jcc Jco 5000 Jh

Jn 4000

J%*

Jk

Jlp Jm Jk

Jm

Im 1000 '^«produced permission Of,he

permission. 1000

Im

Pern -1000

10

)0

X)

X)

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 2000

1 0 0 0 H

Jm Jm

%

-1 0 0 0 H

Im

-2000 -4

-3000

7000 r 70C

Jcc

6000 60C

Kd

Jew 5000 Jew 500

Jcc Jeo Jco 4 0 0 0 4 0 0

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. Jk

Jm

Tk

"Km

Pk

Ph

D' r 7000 7000

cc

CO 6 0 0 0 6000

L 500 0 5000

4 0 0 0 4000

Reproduced wtth permission of ,Pe copyrlgh, owner. Further reproduction prohibited without permission. iiiii* 2000

Jm 1000 Im

« a

1000

Pern •2000

Pph

•3000

E' 7000 7000

6000 ICC 6 0 0 0

CO Qa

5000 5000

U C O i Jew 4000 4 0 0 0

Jco Reproduced with permission of the copyright owner. owner. Further reproduction prohibited - 7000

- 60 0 0

5000

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 7 0 0 0 -1

5 0 0 0 -

4 0 0 0 -

3000 -

2000 -

1000 -

■1000 H

-2000

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 2 0 0 o

'Proc]^

owne, E" 7000 7 0 0 0

6 000 6 0 0 0

iKs Qa

5000 5 0 0 0 Kl

jS L Jew 4000 4 0 0 0

Jco 3000 Jk 3 0 0 0

Jm Jn 2000 2 0 0 0

Tiai 1 0 0 0

Pk /jm / PI

p ^ :P% -1000

2000 -2000

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 7 0 0 0

6 0 0 0

5 0 0 0 fJcOi

4 0 0 0

3 0 0 0 Jm

2 0 0 0

1000

-1000

pph -2000

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. NOTE TO USERS Oversize maps and charts are microfilmed in sections in the following manner

LEFT TO RIGHT, TOP TO BOTTOM, WITH SMALL OVERLAPS

UMI

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 7000

60 0 0

Qcg Qa

5 0 0 0 Jcc Jco

4000

3 0 0 0

2000

1000 Jm

R®P™^«cedw,-,ftpe™3sio„o„,e ^PyMghtov^ner. Further " prohibited Without permission. 4b - Cross

O' 7000

7000

6 0 0 0

6 0 0 0 Qcg

5 0 0 0

500 0 f«teci Jco/ 4 0 0 0 Jco

4 0 0 0

30 0 0

3 0 0 0

2 0 0 0 Jm 2 0 0 0

1000 Jm 1000 Jm

copyright

prohibited Without ’■ Sections for Mapr

R' 7000 7000

6000

5000

4000 4 0 0 0

3000 3 0 0 0

2 0 0 0

2 0 0 0

1000 1000

copyright owner. Further '■«production prohil,,,, ' '"■'Poulpemtission, S’ 7000

7000 Jco

6 0 0 0

50 0 0

5 0 0 0

4000

3 0 0 0

3 0 0 0

2000

2000

1000

1000 7000

6 0 0 0 Jco

5 0 0 0

500 0

4 0 0 0

400 0

3 0 0 0

3 0 0 0

2000

2000

1000

1000 Reprod 1000 ipiiiiiB iiœ i

"Sc

k O

Tim

-1000 Pk

W 7000

6000 Qa

Kd

5000 Jew

Kd Jew Jco 4000

Jco

3000

2 0 0 0 ytc Jm

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 1000

1000

I c j

Tnj 1000

'1000

W 7000

Jco Jco 7000

6000 HU JSL. Jco 6 0 0 0 Jn J

5 0 0 0

4000 Jn Jk 4 0 0 0 Jm 3000

Tfc Jk t c 3000

Jm 2 0 0 0 Jnl

copyrig^jf owner. 2 0 0 0 ■ reproduction Prohibited Without Permission. k 1000 1 0 0 0

Pk

M . Ph -1000 -1000

- 7 0 0 0

— 6 0 0 0

- 5 0 0 0

- 4 0 0 0

k 3000

U 2000

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. r 1000 1 0 0 0 -

0 -

-1000 -1000

Y 7 0 0 0 -1

6 0 0 0 - Qcg — Jco / r Kd

5 0 0 0 -

4 0 0 0 - If

3 0 0 0 -

2 0 0 0 -

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. -1000 -1000

- 7000

- 6000

5000

k 4000

L. 3000

U 2000

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 30 0 0

2 0 0 0 - ^

1000 -

0 -

-1000

L' M 7000 r 70 0 0 7 0 0 0 -I

Jco

600 0 - 6 0 0 0 6 0 0 0 - Jco S2- Kl 5000 - 500 0 5 0 0 0 - Kd Kd JCMT J e w 3. ;■ 4000 40 0 0 Joe 4 0 0 0 - Jco Jco Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 3 0 0 0 Jm 3000 -

Je - 2 0 0 0 2 0 0 0 -

1000 Jl ■fie 1000 - 3» Pk

-I

Ph Ph

PcmPk Pcm -1000 -1000 - L .

N - 7000 7 0 0 0

- 60 0 0 6 0 0 0 Qa

- 5 00 0 5 0 0 0 Kd Jew

- 4 0 0 0 4 0 0 0

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 0 0 3000

l e l e

100 - 2 0 0 0

Jm

300 1000 Pk Pk

m le Ph

Ph Pcm Pcm Pk Pcm 300 -1000

N' r- 7000 7 0 0 0 -I

- 6 0 0 0 6 0 0 0 -

Qa Qa

- 5000 5 0 0 0 -

Kd Kd Jew Jew M - 4 0 0 0 4 0 0 0

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 3000 H ...... 1

2 0 0 0 -

Jk

Jm 1000 H

l e

OH

Im

-1000

- 7 0 0 0 7 0 0 0 -1

6 0 0 0 -

h 5 0 0 0 5 0 0 0 H

h 4 0 0 0 4 0 0 0 -

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. . 3000

- 2 0 0 0

- 1000

-1000

7000 r- 7000

Jco

6000 6 0 0 0 Qa Jco

5000 50 0 0

4000 4 0 0 0

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 7000 - - 7000 M 7 0 0 0 -1

6000 - - 6000 6 0 0 0 - % Kl 5000 H - 5000 5 0 0 0 H Kd

Jew

4000 - i - 4000 4 0 0 0 - Jco

3000 H - 3000 30 0 0 -

Jn 2000 4 - 2000 2 0 0 0

1000 - h 1000 1000

Jm s a m

üm -1000 - -1000 -1 0 0 0

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 7 000 N 7000 7000

600 0 - 6000 6 0 0 0 Qa Qa

500 0 - 5000 5 00 0 Nd Jew 4 0 0 0 4000 4000 Jco

3 000 3000 3000

2000 2000 2000

1000 1000 1000

Jm

üc

1000 üm 1000 -1000

Reproducedwith permission of the copyright owner. Further reproduction prohibited without permission. N' 7000 7000

6 0 0 0 la 6000 Qa Qa Kl Jn 50 0 0 Kl 5000

Ktf JSL. Jew 40 0 0 Jco 400 0 -LÉM ÈÊP^ Jk Jss. Jco Jm 300 0

Jn Üc

2000 2000 ÜC ' m üm jip

Jk 1000 Pk 1 1000 - üm Jm WSSSÊMilSiîiilii Pk ÜC L 0

■■ ü m 1 Pcm ^ Ph ■ -1000 -1000

Reproduced with permission of the copyright owher Further reproductioh prohibited wi Without permission. 7000 7000

6 0 0 0 6 0 0 0 Qa Jco

5000 5000

4 0 0 0 4000

Jm

3000 Jm 3000

i c Jm 2000 2000 Im

1000 1000-1

■Pcm Pcm 1000 -1000

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. 70 0 0

7000 Jco

6 0 0 0

Jco 6 0 0 0 Qa

5 0 0 0

5 0 0 0

4 0 0 0

4 0 0 0

Jm 3 0 0 0

300 0 Jm 2000 2000

1000 -I 1000

i l l i l i i l y

1 Rem

1000

Reproduced with permission of the copyrigm owner. Further reproduction prohlbhed without permission.