<<

Inferring bedrock uplift in the Province from river profile analysis and digital topography

By

Timothy Kirt Anderson, B.A., B.S.

A Thesis In Geosciences

Submitted to the Graduate Faculty of Texas Tech University in Partial Fulfillment of the Requirements for Degree of

Masters of Science

Approved

Dr. Aaron Yoshinobu Chairperson of the Committee

Dr. Calvin Barnes

Dr. David Leverington

Accepted

Fred Hartmeister Dean of Graduate School

December, 2008

Texas Tech University, Timothy Anderson, December 2008

Acknowledgements

I would like to thank my advisor, Dr. Aaron Yoshinobu for the opportunity to work on an exciting project, and am grateful for his enthusiastic approach to science. I would like to thank my committee members, Dr. Cal Barnes and Dr. David Leverington, for their thoughtful reviews of this manuscript. I would like to thank Dr. Jeff Lee and Linda Jones of the Department of Economics and Geography for the role they played in my financial support as a teaching assistant. I would like to acknowledge the Department of

Geography and Dr. Yoshinobu for allowing me access to the hardware and software necessary to undertake this research. I would like to thank Dr. Don Elder for providing me with his ‘latest and greatest’ digital data of the Klamath Mountains. Finally, I’d like to thank my friends and family for their understanding, support, encouragement, and willingness to listen to my ideas.

ii Texas Tech University, Timothy Anderson, December 2008

Table of Contents

Acknowledgements ii Abstract v List of Figures vi List of Acronyms xii Chapter 1. Introduction 1 2. Tectonic Setting of the KMP and its Geologic History 10 2.1. Tectonic Setting 10 2.2. beneath the Klamath Mountains Province 12 2.3. Geologic Background: Neoproterozoic to Cretaceous 15 2.4. Cenozoic of the KMP 24 2.5. Development of the Klamath 24 2.6. Cenozoic Paleogeography of the KMP 29 2.7. Late Cenozoic Climate Change 38 3. Analysis of Digital Topography and Klamath Peneplain 39 3.1. Research Methods 39 3.2. Describing KMP Topography 41 3.3. Creation of Klamath Peneplain Surface 54 3.4. Klamath Peneplain Topographic and Landscape Observations 58 3.5. The Klamath Peneplain Interpolation Surface (KPS) 63 3.6. Creation of and Paleotopography Surfaces 64 3.7. Paleotopography 68 3.8. Calculated Erosion from the Klamath Peneplain Erosion Surface 71 4. River Profile Analysis 78 4.1. Introduction 78 4.2. Historical Background of Longitudinal River Profiles 78

iii Texas Tech University, Timothy Anderson, December 2008

4.3. Modern River Incision Models 81 4.4. Detachment-limited stream power model 82 4.5. River Profile Construction Methods 89 4.6. Rivers in the KMP and Adjacent Regions 94 4.7. Longitudinal River Profiles 95 4.8. Rivers in the Southern Klamath Mountains 97 4.9. Rivers in the Central Klamath Mountains 100 4.10. Rivers in the 110 4.11. Rivers in the Northern and Southern Coast Range 117 5. Discussion 178 5.1. Discussion of Origin, Uplift and Erosion of Klamath Peneplain 178 5.2. Volumetric Analysis of Rates and Duration of Uplift and Erosion 182 in KMP

5.3. Evaluation of Miocene and Pliocene Paleotopography 184

5.4. Inferred Faults in the KMP 186 5.5. Uplift and Topographic Evolution based on River Profile 187 Analysis

5.6. Landscape Development and Evaluation of Uplift and Erosion in 192 local KMP

5.7. Uplift Models 211

5.8. Future Tests 223 6. Conclusion 224 Literature Cited 227 Appendix 236

iv Texas Tech University, Timothy Anderson, December 2008

Abstract

The Klamath Mountains Province (KMP), /Southern , is situated at the juncture of the Mendocino Triple Junction, the southern boundary of the

Cascades and the Juan de Fuca- convergent margin, and the western boundary of the . KMP topography extends from sea level to over 2.5 kilometers of elevation. The mean elevation of the central KMP is greater than one kilometer, defining a regional dome of elevated topography. Surface uplift and rock exhumation have been ongoing since the Late Pliocene. A quasi-planar, regionally-extensive erosional surface termed the Klamath Peneplain (KP) is exposed in coastal regions at/near sea level and at elevations in excess of 2 kilometers more than 100 kilometers inland. Pleistocene marine and non-marine deposits have aggraded on the KP, thus preserving the surface. The geometry of the peneplain and the average amount of uplift and erosion may be calculated by interpolating a westward-dipping surface through the basal peneplain exposures. This assumes the peneplain was continuous over the entire western KMP. Time-averaged long-term uplift rates reach a maximum of 0.4mm/yr in the east. The total amount and rate of erosion since formation of the KP may be calculated by subtracting this interpolated surface from modern topography. These time-averaged results indicate that approximately 3850 km 3 of material has been removed since the

Pliocene at a rate of 0.00077 km 3/yr. The maximum long-term erosion rate is 0.29 mm/yr

( River). Our maps suggest that less than sixteen percent of western KMP topography existed before the Pliocene. Paleotopography with elevations greater than a

v Texas Tech University, Timothy Anderson, December 2008 kilometer must have existed in the eastern KMP at the time of maximum peneplanation.

Longitudinal river profiles in the central KMP are 2 to 3 times steeper than those in adjacent areas. River profile knickpoints suggest that the central KMP could be experiencing baselevel fall in some locations. Modern surface uplift and rock exhumation in the KMP may be attributed to one or more of the following; 1) northward migration of the Blanco F.Z., 2) recent duplexing of the or other accreted beneath the KMP, 3) development of a serpentine wedge beneath the

KMP, and 4) climate driven isostatic rebound. This last interpretation seems most favorable given the spatial correlation between uplift and erosion.

vi Texas Tech University, Timothy Anderson, December 2008

List of Figures

1.1 Klamath Mountain Province and Surrounding Regions Topography 8

1.2 Geographic Locations of the KMP and Adjacent Terranes 9

2.1 Tectonics of Northwest USA 12

2.2 Mesozoic Terranes of the KMP 23

2.3 The Klamath Peneplain (from Aalto, 2006) 27

2.4 Paleogeographic Map of Eocene 33

2.5 Paleogeographic Map of Oligocene Geology 34

2.6 Paleogeographic Map of Late Miocene Geology 35

2.7 Paleogeographic Map of Early Pliocene Geology 36

2.8 Paleogeographic Map of Late Pliocene Geology 37

2.9 Paleogeographic Map of Pleistocene Geology 38

2.10 Paleogeographic Map of Late Pleistocene Geology 39

3.1 Elevation of Northern California and Southern Oregon by Latitude 46 and Longitude

3.2 Longitudinal and Latitudinal Relationship between Topography and 47 Precipitation

3.3 Longitudinal Relationship between Topography and Precipitation in 48 Northern Area

3.4 Longitudinal relationship between Topography and Precipitation in 49 Central Area

3.5 Anomalous Topography of the Central Klamath Mountains and 55 Adjacent Areas

3.6 Precipitation and Hillslopes in the Klamath Mountains Province and 56

vii Texas Tech University, Timothy Anderson, December 2008

Adjacent Areas

3.7 Interpolation Surfaces of the Klamath Peneplain Unit 3.8 Distribution 60 of Erosion Remnant Surfaces

3.8 Distribution of Erosional Remnant Surfaces 62

3.9 Geographic Location and Elevation of Individual Erosional Remnant 63 Surfaces

3.10 Slope and Aspect of Erosional Remnant Surfaces 64

3.11 Elevation of Erosional Remnant Surfaces and Topography by Longitude 65

3.12 Interpolation Surface of Erosional Remnant Surface Elevations 68 (Natural Neighbor)

3.13 Modern Topography Vs. Interpolation Surface 70

3.14 Topography Relative to Interpolation Surface 71

3.15 Map of Paleotopography 76

3.16 Map of Calculated Erosion 80

3.17 Map of Calculated Erosion and Rivers 81

4.1 Northern California and Southern Oregon Subbasins 130

4.2 Equilibrium Vs. Non-Equilibrium Streams 131

4.3 Effect of Basin Shape on Hack Gradient 132

4.4 Parameters used in River Profile Analysis (from Wobus et al., 2006) 133

4.5 Channel Concavity Vs. Channel Steepness 134

4.6 River Profile Shapes 135

4.7 Knickpoint Migration 136

4.8 Map of Normalized Channel Steepness of River Segments of the KMP 137

viii Texas Tech University, Timothy Anderson, December 2008

4.9 Map of River Concavity Measurement in the KMP 138

4.10 Major Rivers in the KMP 139

4.11 Upper Eel Subbasin Streams 140

4.12 Middle Eel Subbasin Streams 141

4.13 Lower Eel Subbasin Streams 142

4.14 Longitudinal Profile of Middle Fork 143

4.15 Longitudinal Profile of 144

4.16 Mad-Redwood Subbasin Streams 145

4.17 Longitudinal Profile of Redwood Creek 146

4.18 Longitudinal Profile of 147

4.19 South Fork Trinity Subbasin Streams 148

4.20 Longitudinal Profile of 149

4.21 Trinity Subbasin Streams: Upper Streams 150

4.22 Trinity Subbasin Streams: Central Streams 151

4.23 Trinity Subbasin Streams: Lower Streams 152

4.24 Longitudinal Profile of Coffee Creek (Trinity River Tributary) 153

4.25 Longitudinal Profile of (Trinity River Tributary) 154

4.26 Longitudinal Profile of East Fork (Trinity River Tributary) 155

4.27 Longitudinal Profile of Canyon Creek (Trinity River Tributary) 156

4.28 Longitudinal Profile of Trinity River 157

4.29 Salmon Subbasin Streams 158

4.30 Longitudinal Profile Salmon River 159 ix Texas Tech University, Timothy Anderson, December 2008

4.31 Longitudinal Profile of 160

4.32 Scott Subbasin Streams 161

4.33 Longitudinal Profile of 162

4.34 Upper Klamath Volcanic Streams 163

4.35 Upper Klamath-Siskiyou Streams 164

4.36 Upper Klamath- Downstream Streams 165

4.37 Lower Klamath- Eastern Streams 166

4.38 Lower Klamath- Central Siskiyou Streams 167

4.39 Lower Klamath - Western Streams 168

4.40 Longitudinal Profile of 169

4.41 Applegate Subbasin Streams 170

4.42 Illinois Subbasin Streams 171

4.43 Longitudinal Profile of 172

4.44 Smith Subbasin Streams 173

4.45 Longitudinal Profile Middle Fork 174

4.46 Longitudinal Profile of South Fork Smith River 175

4.47 Lower Rogue Subbasin Rivers 176

4.48 Longitudinal Profile of 177

4.49 Chetco Subbasin Streams 178

4.50 South Umpqua Subbasin Streams 179

4.51 Longitudinal Profile of South 180

x Texas Tech University, Timothy Anderson, December 2008

4.52 Coquille Subbasin Streams 181

4.53 Longitudinal Profile of North Fork Coquille River 182

5.1 Volume of Exhumed and Removed Rock of the Western KMP 186

5.2 Elevation of Klamath Peneplain Surfaces by Longitude 187

5.3 Comparative Analysis of Erosion Rates (from Burbank, 2002) 189

5.4 Lines and Relative Movement of Western Klamath Mountains 194

5.5 Normalized Channel Steepness by Klamath Bedrock 196

5.6 Elevation and Uplift Patterns of the Klamath and Northwest Coast 200

5.7 Knickpoint Analysis of Trinity River Tributaries 208

5.8 Knickpoint Analysis of Salmon and Klamath River Tributaries 209

5.9 Uplift, Knickpoints and Bends in Western Klamath Streams 215

5.10 Orogenic Wedge Diagram 218

5.11 Modern Elevation Vs. Surface Uplift Vs. Erosion 224

5.12 Erosion Compared with River Steepness, Hillslopes, 225 and Pleistocene

xi Texas Tech University, Timothy Anderson, December 2008

List of Acronyms Acronym Name

BFZ Blanco Fracture Zone

BR Basin and Range Province

CMP Central Metamorphic Plate

CRP Coast Range Plate

CT Cascadia Trench

CVA Cascade Volcanic Arc

EKP Eastern Klamath Plate

GP Gorda Plate

JDF Juan De Fuca Plate

KMP Klamath Mountains Province (and adjacent areas)

KP Individual Klamath Peneplain surfaces, or individual erosional remnant

surfaces

KPES Klamath Peneplain Erosional Surface; Surface that measures depth of

erosion based on KPS

KPS Klamath Peneplain Surface; Interpolation Surface based on KP elevations

MTJ Mendocino Triple Junction

NAP North America Plate

SN

WJP Western Jurassic Plate

WPT Western Paleozoic and Triassic Plate

xii Texas Tech University, Timothy Anderson, December 2008

Chapter 1

Introduction

In the last two decades, the emerging field of tectonic has demonstrated that the first order development of landscapes in an active orogenic system is modulated by tectonic and climatic processes. Molnar and England (1990) generated a controversial and productive debate when they suggested that climate change to wetter conditions in the Late Cenozoic may have promoted increased erosion rates which would drive bedrock uplift through isostatic rebound. In their formulation, parameters such as relief and maximum mountain range elevation may increase while river incision may effectively remove , resulting in a positive feedback between erosion and mountain building such that maximum elevations and relief may increase while mean elevations decrease due to isostatic uplift. The principal mechanisms of relief production and erosion in this model are bed rock river incision (e.g., Whipple, 2004 Ann. Rev. in

Ear. & Planetary Sci. paper).

The term ‘uplift’ refers to the vertical displacement of an object in the opposite direction of gravity (Molnar and England, 1990; Abbott et al., 1997). Measurement of uplift is categorized into three different forms 1) surface uplift, 2) rock uplift, and 3) exhumation. The former two terms are referenced to the geoid, whereas the latter is referenced to a predefined plane (Molnar and England, 1990; Ring et al., 1999).

Surface uplift is the upward displacement of an aerially extensive surface (~1000 km 2) relative to the geoid (Molnar and England, 1990; Small and Anderson, 1995; Abbott

et al., 1997). Few examples of surface uplift can be demonstrated. The interpolation 1 Texas Tech University, Timothy Anderson, December 2008 surface produced from Klamath peneplain elevation data (see Chapter 3) describes surface uplift. It is possible to measure surface uplift magnitude and rate of this feature because it is mantled by littoral of the Wimer Formation (Stone, 1992).

Rock uplift refers to the displacement of a rock relative to the geoid (Molnar and

England, 1990; Abbott et al., 1997). Examples of rock uplift include the vertical displacement of highway benchmarks (Kelsey et al., 1996; VanLaningham, 2006), elevation of marine terraces (Kelsey and Carver, 1988; Kelsey and Trexler, 1989;

Merritts and Vincent, 1989), upward exhumation of metamorphic rocks, and tectonically created knickpoints in river profiles.

Exhumation is the upward displacement of rocks by tectonic denudation or geomorphic erosion (Abbott et al., 1997; Ring et al., 1999; Burbank, 2002). Thus, the exhumation rate is defined as the removal rate of material referenced to a predefined surface (Molnar and England, 1990). Barometric and thermochronometric data provide information on the unloading and exhumation of rocks from deep within the crust (Ring et al., 1999; Burbank, 2002). Because exhumation requires the removal of overlying burden by climatic and/or tectonic influences, exhumation is commonly accompanied by isostatic rebound.

Isostatic rebound has been modeled to occur at several scales including those of individual basins (e.g., Small and Anderson, 1998), mountains (e.g., Farley et al., 2001), continents (e.g., Pelletier, 2004), and globally (e.g., Molnar and England, 1990). Isostatic rebound is a function of the heterogeneity within the crust (Pratt Isostacy), the contrast in densities between the crust and (Airy Isostacy), and flexural rigidity of the crust 2 Texas Tech University, Timothy Anderson, December 2008

(Keller and Pinter, 2002). According to Molnar and England (1990), an isostatically responsive landscape can restore a denuded to 5/6ths (83%) its original height.

Numerical models demonstrate a coupling between surficial, climate driven processes and orogenic internal structural processes, and suggest that these boundary conditions are fundamental to the geodynamic evolution of active orogens in nature

(Beaumont et al., 1992; 2001; Koons, 1989; Koons et al., 2002; Willet and Brandon,

2002). The notion that surficial and structural (i.e., tectonic) boundary conditions control the overall geometric evolution of orogens has long been known from studies of orogenic wedges as exemplified in foreland fold and thrust belts and modern accretionary prisms

(e.g., Davis et al., 1983; Dahlen and Suppe, 1988). In such examples, the critical taper – or shape – of the wedge is controlled by structural or erosional denudation of the upper surface when it steepens beyond taper and internal structural thickening when it shallows below critical taper (Davis et al., 1983). Therefore, denudational processes such as erosion and normal faulting may affect the evolution of an active orogen (Ring et al.,

2001).

At the same time, theoretical and field-based studies demonstrated that bedrock channel networks (or mixed bedrock-alluvial channels) in active orogens may control relief production and erosion rate (Howard and Kerby, 1983; Howard, 1994; Howard et al., 1994; Stock and Montgomery, 1999; Whipple and Tucker, 1999; Whipple et al.,

2000; Snyder et al., 2003; Whipple, 2004; VanLaningham et al., 2006). Therefore, bedrock river incision rates may control the overall patterns and rates of regional denudation in the absence of regional extensional tectonism (Whipple, 2004). Hillslope 3 Texas Tech University, Timothy Anderson, December 2008 forming processes in active orogens cut by bedrock or mixed bedrock-alluvial drainages must play a second-order role in spatial erosion patterns because the hill slope angle is governed by the boundary condition set by bedrock river incision. A critical question may be resolved: How may erosion rates in bedrock drainages be quantified at different spatial and temporal scales in an active orogen?

In addition to controlling patterns of erosion in active orogens, bedrock and mixed bedrock-alluvial drainages may reflect spatial patterns and rates of rock uplift in the form of their longitudinal profiles. The longitudinal profile of a bedrock channel reflects the interplay between local changes in rock uplift rate relative to a fixed external , and channel incision rate (Howard, 1994). However, channel incision rate is sensitive to variations in climate and bedrock lithology as well as bed sedimentation (Stock and

Montgomery, 1999; Whipple and Tucker, 1999; Sklar and Dietrich, 2001; Duvall et al.,

2004). Therefore, it is necessary to evaluate the roles that lithologic variations and climate may play in bedrock incision. Where these variables may be constrained, comparing bedrock drainage longitudinal profiles in different catchments may represent a geomorphic signal of rock uplift (Kirby et al., 2003; Kirby and Whipple, 2002; Whipple,

2004). Investigation in localities such as the Tibetan Plateau (Kirby and Whipple, 2003), the Yunnan Province, China (Schoenbohm et al., 2004), the Andes (Montgomery et al.,

2001), the Ranges (Seidl and Dietrich, 1992; Vanlaningham et al., 2006), the Mendocino Triple Junction of northern California (Merrits and Vincent, 1989), and

South Carolina (Marple and Talwani, 1993) have demonstrated that river profiles provide a proxy for identifying patterns of bedrock uplift. Furthermore, the spatial distribution 4 Texas Tech University, Timothy Anderson, December 2008 and zoning of longitudinal profile knickpoints and high channel steepness can indicate regions of differential bedrock uplift (e.g. Kirby and Whipple, 2003).

Although rarely identified in nature, surface uplift, or uplift of all points within an aerially-extensive region, such as a paleo-erosional surface, provides an ideal reference frame to evaluate rates of erosion and patterns and rates of uplift (e.g., England and

Molnar, 1990; Abbott et al., 1997). A vast, contiguous, flat, and dateable geologic marker surface is ideal. Uplift of the marker with respect to an external reference frame and erosion via down-cutting through the marker can be quantitatively analyzed by comparing modern topography with the paleo erosion surface (e.g. Abbot et al., 1997,

Pederson et al., 2002; Gani et al., 2007).

In this thesis, surface uplift, rock uplift, and exhumation are all investigated in an ideal location. The normalized steepness, concavity, and location of knickpoints along longitudinal profile of four major rivers and several hundred tributaries that transect a diverse range of geologic elements are mapped. The longitudinal and latitudinal variation in elevation, hillslope, and precipitation across a large region is also performed. Finally, the extraction of a rare surface demonstrates a significant amount of recent uplift and significant erosion in this actively deforming orogenic province. If river channels with high normalized steepness values are spatially connected to high relief valleys that measure steep hillslopes, KMP river incision is responsible for sculpting the topography of the KMP into its present youthful state. These fluvial studies can be compared to latitudinal and longitudinal measurements of KMP topography and precipitation patterns to realize any influence the aforementioned geomorphologic controls have on the KMP 5 Texas Tech University, Timothy Anderson, December 2008 block. A dated surface that was once near sea level can be extracted to measure uplift and erosion rates in the western KMP, in addition to identifying regions of pre-existing topography (Chapter 3). If geomorphologic erosion is generating uplift, one would expect the greatest amount of calculated erosion to coincide with the greatest amount of uplift, the steepest channel gradients, hillslopes, and relief.

The Klamath Mountain Province (KMP) (Figure 1.1) represents an anomalous region of elevated topography, deep river incision, and recent exhumation in northern California and southern Oregon. The KMP is located above the down-going Gorda slab and is surrounded by the Mendocino Triple Junction to the southwest, the Cascades Arc/Basin

& Range extensional corridor to the East, and the Cascades convergent margin to the west. A regional erosional surface termed the Klamath Peneplain (Diller, 1902) beveled the KMP during the Miocene and early Pliocene (Irwin, 1981; Aalto, 2006). The peneplain was mantled by Upper Miocene to Lower Pleistoence marine to fluvial deposits. Since the middle Pleistocene, the Klamath Peneplain (KP) has been tilted to the west; elevations of the peneplain are near sea-level along the coast and >2000 m in the central-eastern portions of the KMP. The aerial distribution of the Pliocene erosional surface is consistent with westward tilting and ~2 km of exhumation of the central part of the province has occurred since the Pliocene. Tectonic exhumation due to various forms of underplating and erosion processes has been suggested, but climatic influences are also possible. High seasonal precipitation and the presence of inland quaternary glacial deposits demonstrate that climate may be the most influential role in landscape development. Because the extent and geometry of recent uplift of the KMP is poorly 6 Texas Tech University, Timothy Anderson, December 2008 constrained, it is unknown what tectonic and/or climatic driving forces are responsible for recent exhumation.

Questions over the evolution of the KMP still remain; how much uplift or flexure has occurred in the KMP since the Pliocene, and at what rate? How much and at what rate is erosion occurring within the KMP? Is uplift a result of exhumation by tectonic forces or isostatic response to high denudation rates? If exhumation is tectonically driven, what surrounding tectonic regimes are responsible for uplift?

In the following chapters, this thesis summarizes the Mesozoic and Cenozoic geologic history of the KMP (chapter 2), the recent uplift of the KP and its erosion implications (Chapter 3), the topographic relationships of the KMP (chapter 4), and a river profile analysis of all major streams within the KMP (chapter 5). The discussion

(chapter 6) is an attempt to integrate all results into an appropriate tectonic model that accounts for KMP tectonism and climate.

7 Texas Tech University, Timothy Anderson, December 2008

Figure 1.1 This figure comprises several maps that identify the main topographic features of Northern California and Southern Oregon. DEMS are at 10 m resolution.

8 Texas Tech University, Timothy Anderson, December 2008

Figure 1.2 Because the geographic boundaries for each segment of the KMP and adjacent regions are ill-defined, this map delineates the geographic boundaries referred to in the text. The geographic province of the Coast Range includes mapped Coast Range bedrock geology (Irwin, 1997), Franciscan terranes (Elder, 2008) and adjacent Mesozoic terranes of the KMP. Large subbasins that have filled with Quaternary alluvium have been isolated (shaded brown) for discussion. 9 Texas Tech University, Timothy Anderson, December 2008

Chapter 2

Tectonic Setting of the KMP and its Geologic History

2.1 Tectonic Setting

The Klamath Mountains Province (KMP, latitudes 40 °N to 43 °N and longitudes

124 °W to 123 °W) forms the southwestern segment of the Cascadia arc. Figure 2.1 summarizes the magnitude and direction of tectonic plates surrounding the KMP. Here, three tectonic plates are joined at the Mendocino Triple Junction (MTJ): 1) the Pacific plate (PP), 2) the Juan de Fuca plate (JDF), and 3) the North America plate (NAP)

(Atwater, 1970). In its broadest sense, Late Cenozoic deformation within the KMP is a result of tectonic instability between these plate motions. The northward migration of the

MTJ, at a rate of 6.5 cm/yr, is a consequence of the tectonic instability between the three plates (Atwater, 1970; Merritts and Vincent, 1989; McCrory, 1989, Freymueller et al.,

1999). The JDF and PP are separated by a ridge that currently spreads at 17 mm/yr

(McCrory, 1989). The JDF plate is converging with the NAP in a northwest direction at

40 mm/yr at the northern segment of the Cascade Trench (CT) (Engebretson et al., 1985;

Mitchell et al., 1994; Kelsey et al., 1994; Aalto, 2006). The NAP-PP right-lateral motion, at a rate of 64 mm/yr, is accommodated along 4 major transform faults including the San

Andreas Transform (Atwater, 1970; McCrory, 1989) and is 40 mm/yr in northern

California (Freymueller et al., 1999; Furlong and Scwartz, 2004).

There are at least seven active tectonic elements that affect KMP topography including: 1) subduction of the JDF at the Cascadia Trench (CT) and resulting Cascade

10 Texas Tech University, Timothy Anderson, December 2008

Volcanic Arc (CVA), 2) subduction of the Gorda Plate (GP) at the CT, 3) northward migration of the projected Blanco Fracture Zone (BFZ) (McNutt, 1983; Kelsey et al.,

1994; Vanlaningham, 2006), 4) northward migration of the MTJ (McCrory, 1989; Aalto,

2006), 5) east-west extension of the Basin and Range Province (BR), 6) northwest migration of the Sierra Nevada microplate (SN) (Argus and Gordon, 1991), and 7) clockwise rotation of the Oregon microplate (Wells and Heller, 1998). If KMP uplift results from either of these tectonic elements, then the geometry, rate, and magnitude of uplift should by kinematically compatible with any corresponding tectonic force.

Figure 2.1 Tectonic motion in the . Gorda-Pacific spreading rate from McCrory, 1989. Local Pacific-North America right lateral movement from Freymueller et al., 1999. North America-Juan de Fuca plate motion from Engebretson et al., 1985 and Riddihough, 1984. Oregon Forearc rotation and Basin and Range Extension from Wells et al., 1998. Sierra- North America movement from Argus and Gordan, 1991. 11 Texas Tech University, Timothy Anderson, December 2008

2.2 Subduction beneath the Klamath Mountains Province

A large proportion of the KMP overlies young, hot, and buoyantly subducting GP . The GP is a southern subset of the Juan De Fuca (JDF), divided by the Blanco Fracture Zone (BFZ) (Figure 2.1). At the Cascadia Trench (CT), 8

Ma GP crust juxtaposes 15 Ma JDF crust to the south (McNutt, 1983). Seismic activity at the Cascadia Trench (CT) has been virtually absent since human settlement (Wells et al.,

1998). Hypotheses to explain the absence of seismic-activity include 1) low slip rates, 2) aseismic creep of the JDF by serpentinization and/or lubrication, and 3) the locking of the

NAP accretionary complex (Fluck et al., 1997, Bostock, 2002; Brocher et al., 2003). At

Cape Mendocino where seismicity is pervasive, however, Jachaens and Griscom (1983) determined that the Gorda plate plunges as low as 9º toward S60ºE at the CT.

McNutt (1983) and Jachens and Griscom (1983), based on negative gravity

readings, concluded that lithosphere of the KMP and coastal regions are isostatically

rising. McNutt hypothesizes that the KMP lithosphere is elastically supported from the

subducted segment of the GP, but the CVA is supported by local bodies. Casana

suggested that the southward decrease in MOHO depth is a function of the Gorda plate

northeast-oblique subduction. He suggests that the increase in subduction angle beneath

longitudes around the KMP is not a result of eclogitization (e.g. Bostock et al., 2002) but

rather a consequence of the overlying dense KMP material.

The northward migration of the BFZ likely has tectonic and structural

implications in the overlying NA plate. Carver (1987) suggested that the northward

sweep of the BFZ subsequently created Plio-Pleistocene folds and faults south of the 12 Texas Tech University, Timothy Anderson, December 2008

Klamath River. McNutt (1983) inferred that if uplift of the KMP is due to the subduction of the buoyant GP, then NAP lithosphere overlying JDF crust would experience smaller uplift rates or even subsidence. Kelsey and others (1994) and Vanlaningham (2006) compared uplift rate of marine terraces and survey benchmarks to elevated topography to support McNutt’s theories. McCrory (1989) examined uplift of the Eel River basin, a regionally extensive forearc basin, by relating Miocene vector motions of the GP to NAP and GP to NAP margin.

Isostatic uplift of the KMP could be driven by serpentinization of the upper

mantle (Brocher et al., 2003; Yoshinobu et al., 2006). When is hydrolized,

olivine minerals are subject to a volumetric increase. Because a serpentinized mantle is

less dense than surrounding mantle, it is subject to isostatically rise to the base of the

crust. In the KMP, water sources required for serpentinization can be collected from

thermal dehydration of the GP, dehydration of underplated sediments, dehydration of

accreted terranes within the NAP, or water directly from the mantle.

Other tectonic influences locally deform KMP regions. Within 40 km north of the

MTJ, the Pacific plate is converging with the NAP producing north-south structures

(Furlong and Schwartz, 2004). Argus and Gordon (1991) used baseline interferometry to

measure the Sierra Nevada (SN) block movement at 11 mm/yr at N50ºW. Gravity

readings suggest that the toe of the SN is at the northern terminus of the

(McNutt, 1983). Because of its motion, the SN could be impinging the southeastern KMP

promoting uplift in the region.

13 Texas Tech University, Timothy Anderson, December 2008

The tectonic influence of Basin and Range (BR) extension in the Pacific

Northwest is enigmatic. Wells and Heller (1988) used paleomagnetic data to interpret the clockwise rotation of the Oregon foreblock as a product of BR extension induced by dextral . McNutt (1983) noted that if stripped of its sedimentary/volcanic cover, gravity readings of the CVA are similar to those of the BR.

2.3 Geologic Background: Neoproterozoic to Cretaceous

As summarized by Snoke and Barnes (2006), the basement geology of the KMP is

divided into four tectonostratigraphic terranes that are bounded by east dipping thrust

faults that are broadly arcuate in shape (Figure 2.2). These terranes can be furtherer

divided into 8 subterranes (Irwin, 1960, 1994, 1998; Jachens et al., 1982; Helper, 1986;

Goodge, 1990; Barnes et al., 1992; Wright and Wyld, 1994; Ernst 1999; Cashman and

Elder, 2002). Terranes are identified by different accretionary episodes. The four

principal terranes, the Eastern Klamath Plate (EKP), the Central Metamorphic Plate

(CMP), the Western Paleozoic and Triassic Plate (WPT), and the Western Jurassic Plate

(WJP) have sutured together in three subduction-related events during the Devonian,

Middle Jurassic, and Late Jurassic (Irwin, 1981; Irwin, 1994; Irwin and Mankinent, 1998;

Irwin and Wooden, 1999). Irwin (1998) divides these lithotectonic belts into eight

lithotectonic terranes, defined by their respective accretionary episodes which are as

follows: 1) The Central Metamorphic Accretionary Episode (Early or Mid Devonian), 2)

Fort Jones Accretionary Episode (Late Triassic), 3) North Fork Accretionary Episode

(Early Jurassic), 4) Eastern Hayfork Accretionary Episode (approximated to be Early to 14 Texas Tech University, Timothy Anderson, December 2008

Middle Jurassic), 5) Western Hayfork Accretionary Terrane (Middle Jurassic), 6)

Rattlesnake Creek Accretionary Episode (Late Jurassic), 7) Western Klamath

Accretionary Episode (Late Jurassic), and the 8) Pickett Peak Accretionary Episode

(Early Cretaceous). The Coastal Range Plate (CRP), primarily composed of the

Franciscan accretionary complex, was sutured to the already-amalgamated Klamath plate in the late Cretaceous (Irwin, 1981, Aalto et al., 1995). The following paragraphs provide a synopsis of the geologic and tectonic history of the individual tectonostratigraphic terranes.

In a general sense, the lithotectonic belts become progressively younger in age westward, where KMP rocks ages range from the Neoproterozoic to late Cretaceous

(Irwin, 1981). Since the Devonian, collision, subduction, , and emplacement of oceanic crust, sedimentary basin assemblages, volcanic arc terrains, and both mantle and crustally derived plutons resulted in the growth and clockwise rotation of the KMP

(Mankinen et al., 1989; Irwin, 1998). Interpretation of whether terrane-bounded thrust sheets represent ancient plate boundaries or intraplate imbrication is still a principal paleogeographic question, and the discontinuous exposure and structural thinness of such terranes could cause uncertainty about the relative volume of certain KMP terranes within the KMP (Miller and Harwood, 1989). Nevertheless, the orientation of paleomagnetic markers from KMP subterranes indicate that the Klamath block moved as one rigid body relative to North America until its accretion (to the North America plate) in the late

Cretaceous (Mankinen et al., 1989). The following paragraphs are a brief synopsis of the tectonic history and the corresponding terranes involved. The context of which this thesis 15 Texas Tech University, Timothy Anderson, December 2008 is written, however, takes greater concern with the lithologic exposure of geologic units.

The tectonic history prior to the Miocene, therefore, is primarily included for completeness.

The Eastern Klamath Plate (EKP) is composed of metavolcanic and easterly dipping metasedimentary rocks of Devonian through Middle Jurassic age (Irwin, 1981;

Barnes et al., 1992; Cashman and Elder, 2002). The Eastern Klamath terrane, considered the early Paleozoic nucleus of the KMP, comprises (from north to south) the Yreka,

Trinity, and Redding subterranes (Irwin, 1994; Irwin 1999) (Figure 2.2). The Yreka subterrane is composed of Late Silurian/Early Devonian sedimentary and metasedimentary rocks (Cashman and Elder, 2002). The Trinity subterrane consists of three major rocks: 1) tectonized, metamorphosed and felsic intrusive rocks of

Neoproterozoic age, 2) Ordovician ultramafic rocks, and 3) mafic intrusive rocks of Early

Silurian to Early Devonian (Cashman and Elder, 2002). Redding subterrane rocks are

Devonian to Middle Jurassic volcanics and sedimentary rocks (Miller and Harwood,

1990; Cashman and Elder, 2002). The longest period of sedimentation, volcanism, and plutonism in the KMP (spanning from the Silurian to Devonian to Jurassic) is recorded in

Redding subterrane (Irwin, 1999). The orientation of EKP sediments is similar to the orientation of the underlying Trinity (440 to 480my) (Irwin, 1981). Extensional faulting and upwarping that juxtaposes the Yreka and Trinity terrane is interpreted by

Cashman and Elder (2002) to be the response to tectonic denudation during the early

Cretaceous.

16 Texas Tech University, Timothy Anderson, December 2008

The Central Metamorphic Plate (CMP) is a highly foliated region of elevated topography and deep river incision. The Central Metamorphic terrane is made of well- foliated amphibolite and schist of upper-greenschist to amphibolite facies (Davis et al.,

1965; Cashman and Elder, 2002). The Central Metamorphic Accretionary Episode is noted by subduction and of volcanic and oceanic sedimentary rocks which is now represented as the Salmon and Abrams Schists along the Bully Choop fault during the Devonian (Lanphere et al., 1968; Irwin, 1981, 1999; Barnes et al., 1992; Irwin and Mankinen, 1998).

The western Paleozoic and Triassic plate is subdivided into the Stuart Fork, North

Fork- Salmon River, eastern Hayfork, western Hayfork, and Rattlesnake Creek terranes

(Irwin, 1972; Blake et al., 1982; Wagner and Saucedo, 1987; Goodge, 1990; Ernst, 1999).

During the Fort Jones Accretionary Episode, the blueschists facies that characterizes the

Fort Jones terrane (also known as Stuart Fork terrane) resulted in an ocean-continent subduction environment near the early Mesozoic continental margin (Goodge, 1990;

Ernst, 1999). Goodge (1990) found geochemical evidence to support a paleomodel where the Fort Jones terrane first formed in a spreading ridge environment west of the Eastern

Klamath volcanic arc. The low and high pressure-temperature metamorphism the Fort

Jones terrane was subducted in the late Triassic, and accreted during the Late middle

Jurassic. The present terrane is found in an antiformal klippe within the Central

Metamorphic terrane (Goodge, 1990; Irwin, 1999, 2003). Mankinen and others (1989) reported a paleomagnetic declination of 110° on volcanic and sedimentary strata within the Redding subterrane. 17 Texas Tech University, Timothy Anderson, December 2008

The North Fork Accretionary Episode is signified by continued subduction and clockwise rotation of KMP terranes. The ophiolite, mafic volcanics, limestone and chert protoliths of the North Fork terrane were overridden in a ‘suprasubduction zone arc’ along the Siskiyou (and correlative) faults (Irwin and Mankinin, 1998; Ernst, 1999). This event is responsible for the metamorphism of sedimentary, basaltic, and ultramafic rocks that are mapped within the terrane (Ernst, 1999).

The Eastern Hayfork Accretionary Episode is founded upon the accretion of mafic volcanics, limestones, cherts, and blocks of schist during the early Jurassic along the Twin Sisters Fault (Irwin and Mankinen, 1998; Ernst, 1999). Rocks of the Eastern

Hayfork terrane comprise strata-disrupted to a “chaotic deformed mélange” of quartzofeldspathic rocks, metasilstones, metashales, carbon-rich argillites, and blocks of metacherts, metalimestones and metabasalts (Wright, 1982; Wright and Fahan, 1988;

Ernst, 1999). Regional metamorphism sutured the Eastern Hayfork and North Fork terranes together (Ernst, 1999).

The Western Hayfork Accretionary Episode occurred by the subduction of volcanic arcs along the Wilson Point fault (Irwin and Mankinen, 1998). Ernst (1999) posits that the accretion of the Western Hayfork terrane to the Eastern Hayfork terrane during the Middle and Early Jurassic was followed by northward migration during the

Late Jurassic. Sediments in this region have a paleomagnetic declination of 60º

(Mankinen and Irwin, 1990; Irwin and Mankinen, 1998).

The Rattlesnake Creek Accretionary episode represents a complex tectonic accretionary episode where subduction of the Rattlesnake Creek terrane subducted 18 Texas Tech University, Timothy Anderson, December 2008 beneath the Western Hayfork terrane along the Salt Creek fault (Irwin, 1972; Wright and

Wyld, 1994; Irwin and Mankinen, 1998). The Rattlesnake Creek terrane is composed of blocks of peridotite, greenstone, amphibolite, pillow , mafic plutonic rocks, metachert, and limestone within a serpentinite matrix bounded by an overlying cover sequence of volcanic, hemipelagic, and clastic sedimentary rocks (Irwin, 1972; Jachens et al., 1986; Wright and Wyld, 1994). After a thorough petrologic investigation, Wright and

Wyld (1994) interpreted three stages of this terrane as 1) the development of the serpentinite matrix mélange basement from the disruption of oceanic lithosphere, 2) the subduction near an oceanic fracture zone, 3) regional deformation and metamorphism during the development of a Middle Jurassic arc complex, 4) rifting that opened the

Josephine ophiolite intra-arc basin, and 5) construction of a Late Jurassic arc complex arc. During this time, the paleomagnetic declination was approximately 40° (Irwin and

Mankinen, 1998).

The Western Klamath Accretionary Episode marks the subduction of the Western

Klamath terrane beneath the Rattlesnake Creek terrane (Irwin and Mankinen, 1998). The

Orleans fault, with a minimum displacement of 110 kilometers, separates the WJP from the WPT (Jachens et al., 1986). The paleomagnetic declination is 30º (Irwin and

Mankinen, 1998).

The Pickett Peak Accretionary Episode marks the subduction of the Pickett Peak terrane along the South Fork fault (Lanphere and Jones, 1978; Aalto et al., 1995; Irwin and Mankinen, 1998). This Early to Mid Cretaceous terrane is primarily composed of the blueschist facies of the South Fork Mountain Schist, Chinquapin Metabasalt (Irwin and 19 Texas Tech University, Timothy Anderson, December 2008

Mankinen, 1998). The paleomagnetic declination is approximately 15° (Mankinen, 1988;

Mankinen and Irwin, 1990). The Pickett Peak terrane and Yolla Bolly terrane, consisting of non-foliated turbidite sandstones and argillites, are the dominant terranes of the

Franciscan Eastern belt (Aalto et al., 1995).

The Coastal Range Plate (CRP) is a tectonically significant region of the study area, despite its low mean elevation. The CRP is dominated by the northern extent of the

Franciscan Complex rocks, a late Cretaceous to Miocene accretionary assemblage divided into three belts: the Franciscan Eastern belt, the Franciscan Central belt, and the

Franciscan Coastal belt (Carver, 1985; Kelsey and Carver, 1988; Aalto et al., 1995). The former two belts are in fault contact (Aalto, 1995). The Franciscan Central belt is a mélange of sedimentary (often turbidites), metamorphic, and plutonic rocks interpreted to be created in an accretionary prism west of the Mesozoic Klamath-Sierran arc (Aalto,

1981, Aalto et al., 1995). The Franciscan Coastal belt is extensively faulted and includes the King Range, Coastal, False Cape, and Yager terranes (McLaughlin et al., 1994) that juxtaposed each other in shear zones due to plate coupling and wedge thickening (Aalto,

1995).

Nearly all regional studies using geophysical data have interpreted high gravity reading of the Central KMP as a region of relatively dense material (e.g. McNutt, 1983;

Fuis et al., 1987; Casana, 2007). Casana (2007) constructed two distinctly different cross sections of the KMP surface using models produced from gravity, aeromagnetic, seismic receiver functions, surface geology, and structural geology data. Casana (2007) produced a cross section through the north Central KMP illustrating a MOHO depth of 40 km 20 Texas Tech University, Timothy Anderson, December 2008 approximately 65 km inland. Another cross section through the south central KMP illustrates a MOHO depth of less than 30 km. In the latter cross section, he defends that all KMP terranes are underlain by a thick Franciscan terrane at 14 km depth. Taking crustal density and topography into effect, McNutt (1983) calculated defended that the base of NA crust descends eastward from 9 km (coast) to 20 km (150 km inland) depth.

21 Texas Tech University, Timothy Anderson, December 2008

Figure 2.2 Map of Mesozoic terranes compiled from Elder (2008) and Irwin (1997). ‘Gaps’ in map are primarily Cenozoic geologic features.

22 Texas Tech University, Timothy Anderson, December 2008

2.4 Cenozoic Tectonics of the KMP Evidence of 10 separate episodes of uplift has been documented in the KMP, along the Oregon Coast, in the CVA, and near the MTJ during the Cenozoic (e.g. Diller,

1902; Jones and Irwin, 1971; McNutt, 1983; Mortimor and Coleman, 1985; Heller et al.,

1987; Kelsey and Carver, 1988; McCrory, 1989; McNutt, 1983; Kelsey et al., 1994;

Aalto et al., 1995; Aalto, 2006). These uplift events are documented by the blockage of ancient drainage systems, the presence of angular , extensive faulting, high sedimentation rates, uplifted erosional surfaces, and the uplift of marine terraces.

This section reviews the Cenozoic tectonic history of the KMP.

In the mid-Eocene, seamounts subducted at the ancient Pacific Northwest subduction zone, resulting in a westward migration and growth of the subduction zone.

Arkosic sediments from the Idaho batholith (e.g. Yager Formation near Cape Mendocino) were transported along the ancient “Snake River” (Heller et al., 1987; Wells and Heller,

1988; McCrory, 1989; Aalto et al., 1995, Aalto, 2006). By the Mid-Late Eocene, magmatism shifted westward, producing a northwest trending magmatic belt that blocked and/or diverted eastern fluvial drainages (Underwood and Bachman, 1986; Heller et al.,

1987; Aalto et al., 1995; Aalto, 2006).

By the late Oligocene, tectonic rotation of crustal blocks, regional strike-slip

faulting, and the formation of metamorphic core complexes were interpreted as

extensional features related to BR in the Pacific Northwest (Heller et al., 1987). At 29

Ma, the JDF plate was created by the subduction of the Farallon ridge. Because of the

new unstable plate geometries between the North America, Juan de Fuca, and Pacific

23 Texas Tech University, Timothy Anderson, December 2008 plate, underthrusting of the Juan de Fuca plate experienced a counterclockwise shift and decrease in subduction rate (McKenzie and Morgan, 1969; Atwater, 1970; Zandt and

Furlong, 1982; Engebretson et al., 1985; McCrory, 1989; Gullick et al., 2002). In the

KMP, this tectonic instability was synchronous with the unroofing of the Trinity ophiolite, formation of the southward-dipping La Grange Detachment Fault, deposition of the inland Weaverville fluvial systems, and possibly dextral shear (along the Grogan

Fault) (McCrory, 1989; Aalto et al., 1995; Cashman and Elder, 2002; Aalto, 2006).

Interestingly, the right-lateral movement of the Grogan Fault (at least 75 km offset) system, may have translated the basinal Hoh unit of southwest with its fluvial equivalent, the Weaverville unit (Aalto et al., 1995).

Between 14 and 5 Ma, the Condrey Mountain Dome, encompassing an area equal to 70,000 km 2, uplifted approximately 7 km (Mortimor and Coleman, 1985). Mortimor and Coleman (1985) interpreted this uplift as a consequence of serpentinization of the upper mantle, driven by recent underplating of subducted material. However, there is no consensus on the timing of metamorphism in and around the Condrey Mountain dome

(e.g., Helper, 1986). McCrory (1989) studied Eel basin sedimentation, and determined that no tectonic component was observed during this time period.

2.5 Development of the Klamath Peneplain

The development of the erosional remnant surface classically referred to as the

“Klamath Peneplain” (Figure 2.3) is not well understood. There are three questions this section addresses: 1) what is a peneplain, 2) do erosional remnant surfaces meet the 24 Texas Tech University, Timothy Anderson, December 2008 geomorphologic conditions to be classified as a dissected peneplain, and 3) what geologic formations describe the depositional environment, timing, and sequence of events surrounding the development of the Klamath peneplain?

Phillips (2002) lists five surface characteristics a surface must exhibit to be classified as a peneplain. A peneplain: 1) is eroded by fluvial and subaerial mechanisms,

2) is cut to base level, 3) exhibits low relief, 4) truncates all lithologies, and 5) regionally extensive. These conditions require that a region must experience prolonged regional erosion throughout a tectonically quiescent period.

The KP has been dissected by eastern river systems (condition 1), truncates

Miocene and older lithologies (condition 4), and is mapped throughout northern

California and southern Oregon (condition 5). Arguments that the peneplain was at sea level (condition 2) and the regional landscape was quasi-planar (condition 3) are discussed in this thesis. Predicted paleotopographic regions in the eastern side of the

Klamath Mountains that could be responsible for regional beveling (condition 2) are discussed in Chapter 5.

Many authors refer to the Klamath peneplain as a series of erosional remnant surfaces that appear to be laterally conformable (e.g. Irwin, 1997; Aalto, 2006). Davis

(1899) uses the term peneplain as an interpretive term, suggesting that its formation only occurs when a landscape is in its most mature stage. While avoiding the term ‘peneplain,’ most authors agree that the formation of the erosional surface reflect a period of orogenic beveling near base level (e.g., Diller, 1902; Aalto et al., 1995; Irwin, 1997; Aalto, 2006).

In this thesis, the terms ‘erosional remnant surface,’ ‘remnant surface,’ and ‘peneplain 25 Texas Tech University, Timothy Anderson, December 2008 surface’ refer to Diller’s “Klamath Peneplain.”

Figure 2.3 This is a cartoon of Diller’s classic “Klamath peneplain.” Diller believed that Klamath Mountain topography was reduced to a low-relief surface during the Miocene. Image taken from Aalto (2006).

Diller (1902), Irwin (1997) and Aalto (2006) believe that the Klamath Mountains was beveled to a low-lying surface during the Miocene. This feature is most easily recognized by prominent flat-lying surfaces erosional surfaces mantled by thick saprolitic paleosols and/or coarse conglomerates that are characteristic of landscapes that have experienced deep (Aalto, 2006; Barnes personal comm., 2008). The proximity and consistent elevation of erosional remnant surfaces now capping Klamath Mountain ridges (see description in Chapter 3) suggests that the surface was once a contiguous flat surface that has been uniformly uplifted and dissected.

The existence of Klamath Mountain paleo drainage systems has been inferred by provenance studies in the Eel basin sedimentary packages and Wimer estuarine studies

(e.g., Kelsey and Trexler, 1989; McCrory, 1989; Stone, 1992; Aalto et al. 1995; Aalto,

2006). Aalto and others (1995, 2006) uses evidence of anomalously high ratios of potassium feldspar with an age distinctive to Idaho batholith rocks (approximately 55

Ma) to suggest the reestablishment of paleodrainage networks that extended from Idaho,

26 Texas Tech University, Timothy Anderson, December 2008 through the current geographic location of the Klamath Mountains, to the paleo-Pacific

Ocean. Citing that the dissolution of potassium feldspar from Oregon longshore currents was unlikely, Aalto (2006) preferred an interpretation that the Pliocene coast received sedimentation along a fluvial system directly connected to Idaho batholiths (Aalto et al.,

1995; Aalto, 2006) 2006). Even if the sediments are derived from an Idaho batholith source, it is still unclear if the recent uplift of the Klamath Mountains shut off the system.

The age of the remnant erosional surfaces is correlated with the late Mio- early

Pliocene Wimer and St. George Formations that sit on top of the erosional surface (Irwin,

1997; Aalto, 2006). The paleoshoreline identified in the Wimer Formation near Crescent

City has been dated at 5 Ma. In Oregon and Del Norte , a thick saprolitic paleosol is mapped above an erosional that cuts the underlying Yolla Bolly terrane

(Franciscan Complex). This soil is mantled by Wimer fluvial sediments, interpreted to be deposited in a low graded, braided stream (Stone, 1992). Inland and north of the Oregon border, numerous Miocene to Pleistocene-aged laterite soils that mantle ultramafic rocks have been mapped (Irwin, 1997). Saprolitic and nickeliferous lateritic soils are interpreted to have formed by deep weathering of the underlying parent rock (Irwin,

1997; Aalto, 2006). Thus, the age of the erosional surface in Del Norte County must be older than the reported 5 Ma Wimer paleoshoreline, younger than the late Cretaceous

(Yolla Bolly terrane), and received fairly continuous sedimentation and intense weathering as indicated by the presence and characteristics of paleosols (Stone, 1992;

Aalto, 2006).

27 Texas Tech University, Timothy Anderson, December 2008

To the south, the timing of the development of the erosional surface is inferred. In the Klamath Mountains, the surface is dissected by Pleistocene-aged faults. Diller (1902) noted that Oligo-Miocene Weaverville fluvial deposits mantle the surface in the southern

Klamath Mountains. The age of the Weaverville Formation is controversial because it is determined by fossils in lacustrine deposits that are nonconformable with other

Weaverville terrestrial deposits (Elder personal commun., 2008). It is conceivable that given the proximity of Weaverville sediments to Quaternary alluvium, some mapped

Weaverville Formation units may actually be Quaternary alluvium. Thus, if the peneplain beneath what is classified as the Weaverville Formation exists, it may actually underlie Quaternary alluvium deposits. Reconciling the relationship between peneplain units and Weaverville/Quaternary alluvium deposits is paramount in understanding the age of the peneplain and Miocene tectonic setting of this region.

Because the Pullen Formation of the Wildcat Group and the Wimer and St.

George Formations are coeval, it is speculated that an unconformity (demarcated by a saprolitic soil referenced to the Klamath peneplain) underlies all three units (Aalto et al.,

1995; Aalto, 2006). Aalto and others (1995) suggests that the California coast experienced a period of coastal erosion and prolonged subaerial exposure in the Miocene that was then followed by synchronous shoaling of littoral Miocene and Pliocene units.

The unconformity at the interface between the Wildcat Group and the Miocene Bear

River beds and Eocene Yager Formation could be among the final erosional events during this time period. This erosional event is believed to be responsible for the apparent beveling of the western KMP to a low-lying, planar surface (Aalto et al., 1995). 28 Texas Tech University, Timothy Anderson, December 2008

Alternatively, development of the surface may predate the Oligocene as indicated by the age of the Weaverville fluvial deposits. Because the development of a regional erosional surface requires an extensive time frame to bevel topography (Phillips, 2002), it is likely that the development of the surface is time-transgressive and could have developed throughout the Cenozoic. Dating of any mantled sediments in the southern KMP can better constrain timing of development and subsequent uplift of the surface.

Because the surface is preserved atop ridges composed of plutonic rocks, Western terranes, WPT terranes, and Franciscan terranes, ‘peneplanation’ is not terrane specific.

The dissection of this surface is either a consequence of differential erosion and/or unmapped faulting. The Siskiyou Mountains, except where less than 30 km from the coast, are void of remnant surfaces (Irwin, 1997). As discussed later in this thesis, the absence of this surface could also indicate pre-existing remnant topography at the time of peneplanation.

2.6 Cenozoic Paleogeography of the KMP

Seven maps of Cenozoic geologic features and events were schematically drawn

(Figures 2.4 to 2.10). These maps identify uplift and subsidence regions, and rely on mapped geologic evidence documented in other studies. Brief discussions are provided below each map. The purpose of these maps is to illustrate the spatial-temporal relationships of geologic formations relevant to the uplift of Klamath peneplain.

Constructed maps only consider geologic events that occur strictly within northwestern

California and Southern Oregon. The geologic setting of the western and 29 Texas Tech University, Timothy Anderson, December 2008

Pacific Northwest, while important, go beyond the scale of this study. To gain a more complete understanding of the paleotopographic history of the Pacific Northwest and its relationship to the Klamath Mountains, other paleotopographic studies are recommended

(e.g., Atwater, 1970; Nilsen and McKee, 1979; Wells and Heller, 1988; Aalto et al.,

1995; Lock et al., 2006).

The location of the paleoshoreline is inferred by the location of coeval geologic formations extracted from Irwin’s Preliminary Map of Post-Nevadan Features (1997).

Faults are provided by the USGS (www.seamless.usgs.gov/) or supplied by a digitized geologic map of the USFS (Elder, 2006). Paleoshorelines and paleorivers are situated in ideal locations based on the paleogeographic location of geologic features. Blue-shaded regions represent oceanic depocenters and are confined to where marine sediments have been mapped. Green shaded regions represent continental depocenters. Red shaded regions represent areas of interpreted rock uplift. These regions have been interpreted primarily from uplifted marine terraces and faults. Light-blue shaded regions represent features where intense erosion has occurred. This is interpreted from elevated river terraces relative to stream beds and locations where deposits are mapped. Black arrows illustrate the direction of a prograding shoreline. White arrows represent the direction of a retrograding shoreline. Arrows with red arrowheads represent the recorded paleoflow direction of paleorivers recorded by fluvial deposits.

30 Texas Tech University, Timothy Anderson, December 2008

Figure 2.4 Eocene In this map, seven significant geologic features are mapped. The paleoshoreline is drawn east of middle Eocene marine formations and west of fluvial sedimentary units. Sediments within the oceanic Eocene depocenters are interpreted to come from an Idaho batholith source based on high potassium feldspar content (Aalto et al., 1995, 2006). Paleoflow directions are taken from Aalto and others (1995). Two different streams are hypothesized to transport sediments to the coast, however, there is no reported geologic evidence concerning the location of either paleodrainage network. As reviewed in Chapter 2, episodes of volcanism in the Pacific Northwest longitudinally migrated throughout the Eocene (Wells and Heller, 1988). It is suggested that the erosional remnant surface caps the Yager Formation (Aalto, 2006).

31 Texas Tech University, Timothy Anderson, December 2008

Figure 2.5 Oligocene This map illustrates hypothetical events in the Oligocene. The Weaverville Formation (Oligocene to early Miocene) is a continental fluvial geologic formation that is reported to cap the erosional remnant surface, “Klamath Peneplain” (Diller, 1902; Aalto et al., 1995). The formation has deposited in a series of grabens that are bounded by the northeast or east-west striking La Grange Fault (Aalto et al., 1995). Aalto and others (1995) speculated that the missing down-dip estuarine sediments could have been displaced to the north along the right-lateral Grogan Fault. A second hypothesis is that the estuarine deposits have been underthrusted beneath the Coastal Range Fault (Aalto et al., 1995). A final hypothesis is that this missing unit was stripped in the Miocene, as indicated by a prominent unconformity above the Eocene Yager Formation. If the peneplain is preserved by the Weaverville Formation, and if the age of the overlying unit is indeed Oligocene, then this time period marks the earliest appearance of the remnant surface. Faults are illustrated because they might have been active in the Oligocene.

32 Texas Tech University, Timothy Anderson, December 2008

Figure 2.6 Late Miocene The formation of the BFZ at 8.5 Ma is coincident with 10º clockwise rotation of the JDF ridge to N10°W (Wilson et al., 1984; McNeill, 2000). This rotation increased the normal JDF-NA convergence rate (McCrory, 1989). The map representing the late Miocene was drawn at the specific time interval where an 8 m.y. regional angular unconformity between the Yager Formation and Wildcat Formation has been observed onshore in outcrop (Carver, 1988; McCrory, 1989; Stone, 1992; Aalto et al., 1995; Aalto et al., 1996; Gullick et al. 2006) and offshore from seismic lines (McCrory, 1989; McNeil, 2000). Prominent unconformities indicate the marine unit has been exposed above surface for an extended time period. At approximately the same time in northern California and Oregon, saprolitic paleosols formed atop the erosional remnant surface, indicating infrequent erosion and recurrent sedimentation (Irwin, 1997; Aalto, 2006). Plio-Pleistocene soils are also mapped within the Klamath Mountains (Irwin, 1997).

33 Texas Tech University, Timothy Anderson, December 2008

Figure 2.7 Early Pliocene In Figure 5.13, geologic formations (labeled) onlap the surface of the aforementioned unconformity (Stone, 1992; Aalto et al., 1995; Aalto et al., 1996; Aalto, 2006). By 5 Ma, the BFZ rotated to a more oblique convergence vector of N25ºE, and despite a decrease in spreading rate (to less than 17 mm/yr), the NAP-GP convergence rate increased to 58 mm/yr (McCrory, 1989). Study of the late Miocene- Early Pliocene geologic units, reveals rapid coastal deepening of Pacific Northwest coast (Aalto, 2006; Gullick et al., 2006). This glacio- eustatic eastward shift in the shoreline resulted in sedimentation of the coeval Wimer Formation, St. George Formation, and Pullen Formations were deposited (Aalto et al., 1995, Aalto et al., 1996). Aalto and other (1995) suggest this time period reflects the westward growth of the California and Oregon coast. Rapid deepening of the sea level is evidenced by short events that deposited logs atop the peneplain surface (Aalto, 2006). Potassium feldspar-rich sediments were transported from an Idaho batholith source (Aalto, 2006). Sedimentation in the fluvial section of the Wimer Formation, however, is interpreted to have been deposited in a low gradient, braided stream (Stone, 1992).

34 Texas Tech University, Timothy Anderson, December 2008

Figure 2.8 Late Pliocene The breakage of the northern Explorer plate c.a. 4 m.y. destabilized plate motions. This event increased Pacific Plate motion from 57 to 64 mm/yr and rotated the PP-JDF convergence clockwise 20° (McCrory, 1989; McNeill et al., 2000). This adjustment changed the PP-NAP boundary from transtension to transpressive (Riddihough, 1984; Cox and Engebretson, 1985; McCrory, 1989; McNeil et al., 2000). This could be responsible for a sharp decrease in convergence rate at c.a. 3.5 Ma (McCrory, 1989; McNeill et al., 2000), when the Gorda plate motion ceased (Riddihough, 1984). During the Late Pliocene, fluvial flow direction in the Wimer Formation shifted to the northwest from the Siskiyou Mountains, fluvial deposition of clastic sediments in the Prairie Creek Formation initiated, deposition in the estuarine Falor and littoral upper Wildcat Formations initiated. Sediments are reported to be derived locally or from the Western Jurassic terranes (Kelsey and Trexler, 1989; Stone, 1992; Aalto et al., 1995). The late Pliocene-early Pleistocene Formation has exceptionally high potassium feldspar content, inferring sedimentation from the Idaho batholith source could be re-established (Aalto, 2006). Activity along the Grogan fault is documented by the juxtaposition of right-lateral offset between the Prairie Creek Formation and Franciscan assemblages (Kelsey and Trexler, 1989). Stone (1992) interpreted the influx of gravels to the fluvial section of the Wimer Formation as a response to the initial uplift of the Siskiyou Mountains. McCrory (1989) believes that the surplus of sedimentation in the Eel basin, usually in the form of turbidite deposition, is a result of on-land uplift.

35 Texas Tech University, Timothy Anderson, December 2008

Figure 2.9 Pleistocene The KMP during the Mid Pleistocene experienced a large amount of uplift (Aalto, 2006). To the north, southern and central Oregon platforms have uplifted at high rates and rotated clockwise along left-lateral faults (Kelsey et. al, 1996). During this interglacial period, the Pleistocene margin prograded and filled the CT (Carver, 1987). Perched stream terraces are present in the Trinity and Klamath subbasins (Irwin, 1997). McCrory (1989) believes that Eel basin was not exposed above sea level until after 1 Ma, as suggested by an angular unconformity. During this time, braided fluvial deposits of the Carlotta Formation conformably onlap the littoral Scotia Bluff sediments. The presence of the upper non-marine Falor Formation reflects the prograding sea level (Kelsey and Carver, 1988). Gravels of the Surpur Creek unit have been interpreted to come from the nearby Surpur Fault (Kelsey and Trexler, 1989). During this time, streams from the Upper and Middle Fork Eel subbasin transported sediments south of the current Mendocino Triple Junction (Lock et al., 2006).

36 Texas Tech University, Timothy Anderson, December 2008

Figure 2.10 Late Pleistocene During the late Pleistocene, deposition of marine and fluvial terraces are regularly mapped (Irwin, 1997). The Eatons Rough Fault Zone (ERFZ) and northern Grogan Fault have exhibited dextral movement resulting in pull-apart basins (Kelsey and Carver, 1988). The marine Skunk Cabbage unit, near the mouth of the Klamath River, is now more than 500 m in elevation and deposited above an unconformity (Kelsey and Trexler, 1989). Activity of the Del Norte Fault has offset the Wimer and St George Formation by at least 210 m (Stone, 1992). Gullick and others (2002) suggest that the influence of the northward migration of the MTJ is at least 80 kilometers to the north, as indicated by northward dipping Pleistocene offshore strata. On shore, streams near the coast increased in stream gradient (Merritts and Bull, 1989) and inland streams were captured by the development of drainage divides (Lock et al., 2006). The presence of glacial deposits and elevated Pleistocene terraces above modern rivers demonstrate recent erosion processes. Erosion in the Eel basin resulted in a thickening of strata packages since 2 Ma, interpreted to be the onset of Pleistocene glaciation (McCrory, 1989).

37 Texas Tech University, Timothy Anderson, December 2008

2.7 Late Cenozoic Climate Change The climate of the KMP throughout the late Cenozoic has fluctuated. The mid-

Miocene global climates are a time of thermal highs, where warm-wet conditions prevailed in Oregon, as evidenced by isotopic data in paleosols of the Ironside Formation

(Retallack, 2004). Low carbonate isotope data in the Pliocene-Pleistocene ocean suggests a transition from greenhouse to icehouse conditions (Retallack, 2004). During the late

Miocene and Pliocene, a rapid deepening of the continental shelf to bathyal-abyssal depths occurred, indicating the progradation of the paleo-shoreline approximately 20 kilometers from the present shoreline (evidenced by the presence of the Wimer

Formation) (Aalto, 2006). Trends defined by the global eustatic sea level curve have demonstrated an exit from the Pleistocene Ice Age. Nevertheless, the sea level was again inland from the present day shoreline during three Holocene times (at 125 or 230 ka, 105 ka, and 80 ka), as evidenced by Holocene marine terraces near Crescent City (Aalto,

2006). The causation of sequence stratigraphic stacking patterns is still an ongoing debate

(e.g. Aalto et al., 1996).

38 Texas Tech University, Timothy Anderson, December 2008

Chapter 3

Analysis of Digital Topography and Klamath Peneplain

3.1 Research Methods

Topographic analysis of the KMP relies on digital data. As opposed to working with topographic maps or directly measuring phenomena in the field, working with preprocessed digitized maps facilitates the acquisition, organization, and processing of large datasets in a more expedient, more economical, and less tedious way. Because processing tools in software packages already exist, the manipulation of data can now be accomplished in greater magnitudes at large scales. Therefore, the use of digital data in this study is justified and appropriate.

Data

Digital Elevation Models (DEMs) use National Elevation Datasets (NEDs) to

create a continuous rasterized surface of elevation values. NEDs are digitized from a

multitude of physical elevation sources, commonly 7.5” quadrangles. All quantitative

investigations using elevation data in this research rely on 1/3 arc-second resolution

(approximately 10 meters) DEMs acquired from the USGS website

http://www.seamless.usgs.gov/. These data have a vertical accuracy of approximately 7

meters (USGS website). A DEM from coarser (30 m) Shuttle Radar Topography Mission

(SRTM) data was also extracted for quality control.

The National Hydrography Dataset (NHD) is digital data representing hydrologic

features in a format designed for hydrologic modeling. For research concerning the KMP,

the features of study are primarily rivers, lakes, and shorelines. NHD data were compiled 39 Texas Tech University, Timothy Anderson, December 2008 by the U.S. Environmental Protection Agency and U.S. Geological Survey (NHD website). The data were collected from a multitude of websites including the Water

Resources Division, U.S. Department of , Natural Resources Conservation

Service, and the Watershed Boundaries Dataset. NHD data is considered ‘high resolution,’ created at the scales 1:24,000 or 1:12,000 (“http://nhd.usgs.gov/”).

Precipitation data were obtained from the PRISM Group of Oregon State

University, at http://prism.oregonstate.edu . This seamless data generates gridded

precipitation estimates by implementing a model referred to as Parameter-elevation

Regressions on Independent Slopes Model (PRISM). The calculated precipitation values

are based on a model that takes into account elevation, the rain shadow effect,

temperature inversions, coastal humidity, and other climate-influenced variables (PRISM

website). The digital precipitation data, with a spatial resolution of 30 arc-seconds (800

m), is a based on precipitation data collected between 1971 and 2000 (Figure 3.6).

Four digital geologic maps of the KMP and surrounding areas were used in this

study: Irwin’s Preliminary map of pre-Nevadan geology in the KMP (Irwin, 1997),

Elder’s map (2005) dedicated to geology within the Klamaths, the Geology Map of

Oregon (Walker and MacLeod, 1991), and the Geology Map of California (Wagner and

Saucedo, 1971).

All four maps record surficial geology at different scales of different regions.

Irwin’s map, at the scale of 1:250,000, is most useful for Cenozoic geology in the western

KMP, including the Klamath peneplain. Elder’s map (2005) is a compilation of digitized geologic units produced by a multitude of mapping projects by the United States Forest 40 Texas Tech University, Timothy Anderson, December 2008

Service. Map features have been digitized from other maps that range in scale between

1:25,000 and 1:250,000. This map is most useful for understanding lithology within the

KMP. The geology map of Oregon (1:500,000) and California (1: 750,000) cover the broadest geographic range, and are referred to where the former two maps lack data.

3.2 Describing KMP Topography

Investigation of the Klamath Mountains Province (KMP) can be subdivided into topographically-based geographic provinces (Figure 1.2). This section qualitatively and quantitatively describes key topographic and precipitation trends expressed in Figures 3.1 through 3.6.

The central Klamath Mountains are best defined as a broad topographic dome that only exceeds 2000 m in elevation at its most elevated ridges (Figure 3.5). Bounded by the

Klamath and Trinity Rivers to the north and south, this topographic dome exhibits distinctly high elevations even at a continental scale (Figures 3.1 and 3.2). The mean elevation of the central Klamath Mountains is always greater than 1000 m, but can exceed 1500 m at specific longitudes (Figure 3.4). Thus, the close proximity between mean and maximum elevations measured in the central Klamath Mountain may reflect the influence elevated ridges have on measured values of mean topography. Maximum elevations are 700 meters greater than mean elevations within the central Klamath

Mountains (Figure 3.4), and 1500 meters greater than regional topography (Figure 3.2).

The Klamath River is a deeply incised river with headwaters at Klamath Lake in the Oregon Cascades. In its descent to the Pacific Ocean, the river traverses Mesozoic- 41 Texas Tech University, Timothy Anderson, December 2008 aged Klamath terranes via the central Klamath Mountains and California Coast Range.

Upon entering the Klamath Mountains, the river carves deep, narrow valleys that exhibit more than a kilometer of relief. A major tributary to the Klamath River, the northwest flowing Salmon River dissects the heart of the central Klamath Mountains and exhibits extremely high basin relief as well.

The widest and most elevated ridges in the central and eastern Klamath

Mountains form the Trinity , and are north of the upstream reaches of the Trinity

River. The southern boundary of these glaciated ridges north of the Trinity River is characterized by a northeast striking lineament (Figure 3.5). This lineament is mapped in close proximity to the (Oligocene?) La Grange Fault, however its association with the older fault is not well defined (Elder, personal comm., 2008). Trinity River elevations are significantly lower than ridges of the .

Longitudinal and latitudinal swaths across KMP topography describe first-order relationships of topography and precipitation. In this thesis, swaths were made of the

KMP and adjacent areas (Figures 3.1 and 3.2). This includes a northern region of the

KMP including the parts of the Oregon and California Coast Range, elevated topography east of the Coast Range, the Siskiyou Mountains, and the Applegate and Illinois subbasins (Figure 3.3), and 3) a latitudinally thin section from the Pacific Ocean, through central Klamath Mountains and into the (Figure 3.4). In calculations of mean elevations, all non-positive elevation values were assumed to represent the Pacific

Ocean and were omitted.

42 Texas Tech University, Timothy Anderson, December 2008

The mean and maximum elevations plotted in Figures 3.1 and 3.2 were measured at intervals of every hundredth latitudinal decimal degree between 39.36º N and 43.38º N (403 total latitudinal measurements). Average precipitation values were measured between latitudes 39.75º N and 43.38º N (Figure 3.2).

Elevation, precipitation, and hillslope trends were studied in a northern and central region for precipitation interests. Measurements illustrated in Figures 3.3 and 3.4 describe longitudinally and latitudinally-averaged swaths calculated at spatial increments of three tenths of a decimal degree. The northern region of interest (Figure 3.3) is a rectangular area between the coast and 122.22º W, and the central study area is between the coast and 121.25º W (Figure 3.4).

43 Texas Tech University, Timothy Anderson, December 2008

Elevation of Northern California and Southern Oregon by Latitude

Figure 3.1a This figure graphs the mean and maximum elevation of northern California and southern Oregon by latitude. Geographic provinces are colored by their latitudinal position. Topography is noticeably more elevated in the central Klamath Mountains than surrounding areas.

44 Texas Tech University, Timothy Anderson, December 2008

Elevation of Northern California and Southern Oregon by Longitude

Figure 3.1b This figure graphs the mean and maximum elevations northern California and southern Oregon by its longitudinal position. Where longitudinal coordinates are coincident with the central Klamath Mountains, mean topography is greatest. Maximum elevations that exceed 3000 m represent the Cascade volcanic edifices.

45 Texas Tech University, Timothy Anderson, December 2008

Figure 3.2 This figure superimposes mean precipitation (in inches) on the elevation graphs of Figure 3.1.

46 Texas Tech University, Timothy Anderson, December 2008

Figure 3.3 This figure illustrates the relationships between topography and precipitation in the Coast Range and northern Siskiyou Mountains.

47 Texas Tech University, Timothy Anderson, December 2008

Figure 3.4 This figure describes the mean precipitation, mean elevation and maximum elevation of coastal areas and the central KMP.

48 Texas Tech University, Timothy Anderson, December 2008

The Cascade Range and Scott Valley area are characterized by abrupt decreases in regional elevations (both mean and maximum) and hillslopes (Figure 3.6). East of the

Scott Valley area, a 731 km 2 featureless regional low-lands form the Shasta and upper

Klamath subbasins. The Cascade Range is a topographic feature composed of Tertiary,

Miocene, and Holocene volcanic rocks. Its volcanic edifices describe the topographic trend of the range as a north-northwest striking arc (black dashed line in Figure 3.6).

Rivers appear to be deflected around High Cascade volcanic edifices (e.g. Rogue River).

North of the Klamath River, the zigzagged peaks of the Siskiyou Mountain exceed 2000 m in elevation. The range strikes northeast until just west of Condrey

Mountain, when its orientation shifts towards a more easterly direction towards its juncture with the Cascade Range. Similarly, the strike of inland ridges north of the

Siskiyou Mountains are characterized by a series of sub-parallel northeast striking ridges that terminate where younger Cascade volcanic rocks drape over them (Fig. 3.5).

The southern Siskiyou Mountains and are at elevations greater than 1200 m at distances less than 35 kilometers from the Pacific Ocean (Figures

3.1, 3.2, and 3.3). A prominent northeast striking topographic lineament (yellow dashed line in Fig. 3.5) signifies an abrupt decrease in elevation fifty to one hundred kilometers east of the coast. East of this lineament, alluvium-full regional lowlands of the Rogue and

Illinois rarely exceed 500 m in elevation. This basin is best characterized by its low gradient hillslopes (Figure 3.6).

The deep-set rivers of the South Fork Trinity, Mad, and Eel Rivers flow from elevated domes to the southeast (Figure 3.5). These rivers are bounded by narrow 49 Texas Tech University, Timothy Anderson, December 2008 northwest striking ridges at elevations usually between 1200 and 1500 m (faint white lines in Figure 3.5). The aforementioned dome exceeds 2000 meters in elevation and is approximately 120 kilometers southeast of the MTJ (dark transparent shade in Figure

3.5).

Elevated topography near the coast and in the central Klamath Mountains are characterized by high precipitation rates. The greatest amounts of precipitation occur in the Coast Range and Siskiyou Mountains. Longitudinally, mean precipitation is as great as 90 inches in elevated coastal regions. The spatial correlation between elevated topography, high relief, and high amounts of precipitation in the geographic locations near the coast indicates a strong orographic effect. Moisture transport is largely accommodated by the south-southeast directed California Current. In Oregon, an abrupt decrease in precipitation 50 kilometers inland spatially correlates with an abrupt decrease in elevation. is situated in a rainshadow (Figure 3.6).

Despite having similar topographic characteristics and also receiving moisture from the California Current, the central Klamath Mountains receive 50 annual inches less than the Siskiyou Mountains (Figure 3.6). Thus, the central Klamath Mountains are situated in the Siskiyou Mountain rain shadow. Despite the receiving a relatively low amount of moisture, the central Klamath Mountains receive more than 100 inches of annual precipitation.

Because of its distance from the ocean, moisture at the northern terminus of the

Great Valley is anomalous. The easiest explanation for these seemingly anomalous moist

conditions involves convergence of air east flow Klamath Mountain and north-flowing 50 Texas Tech University, Timothy Anderson, December 2008

Great Valley air masses. Any moisture-laden air that manages to traverse the southern

Klamath Mountains from the Pacific Ocean must precipitate when converging with the

Great Valley air masses.

Pleistocene alpine glaciers could be a main driver of erosion in the central

Klamath Mountains. Alpine glacial erosion rates are an order of magnitude greater than fluvial erosion rates (Molnar and England, 1990; Hallet et al., 1996; Small and Anderson,

1998; Burbank, 2002; Pelletier, 2004). In Irwin’s Preliminary Map of Post-Nevadan

Geologic Features (1997), Irwin has mapped 122 locations where Quaternary glacial deposits are situated. Eighty-three of these glacial deposits are associated with the Trinity

Alps, deposited along the northern ridge of the Trinity subbasin at elevations of 1500 m or greater. To the north, twenty-six glacial deposits are also mapped at elevations greater than 1800 m east of Wooley Creek batholith and west of Condrey Mountain. Ten glacial units are mapped on both sides of the drainage divide of the western ridge of the Siskiyou

Mountains. The three most eastern glacial deposits are located east of 124° W in the

Coast Ranges at elevations less than 1000 m. No glacial deposits have been mapped west of the Trinity River.

The bottom map in Figure 3.6 illustrates hillslope measurement in the KMP and adjacent regions. The steepest regions, as defined by hillslope values, are situated in the

Coast Range, central Klamath Mountains, and Siskiyou Mountains. In support of these qualitative observations, averaged measurements taken along longitudinal swaths in the northern and central regions are greater than 25° in the Coast Range and Siskiyou

Mountains (northern area), and in the central Klamath Mountains (central study area). 51 Texas Tech University, Timothy Anderson, December 2008

Figure 3.5 This figure delineates ridge lines that exceed 1000 m in the KMP and adjacent areas. Topographic ridges that define an elevated dome-like appearance are shaded black. On the right, oblique perspectives are provided.

52 Texas Tech University, Timothy Anderson, December 2008

.

Figure 3.6 Annual precipitation and calculated hillslopes of northern California and southern Oregon are illustrated in this figure. On the right, graphs illustrate precipitation and hillslope by latitude and longitude.

53 Texas Tech University, Timothy Anderson, December 2008

3.3 Creation of Klamath Peneplain Surface

Extracted from Irwin’s preliminary map (Irwin, 1997), the erosional remnant surface (os) was converted to a shapefile in ARCGIS v. 9.2. Using an already prepared

DEM of the KMP, a subset DEM of the isolated remnant surfaces was created by masking Irwin’s digitized ‘os’ unit. This subset DEM was then converted into a 1 dimensional (point) shapefile. This converted every pixel of the DEM (10 m resolution) into a point with an elevation value. Because only the contacts of the mapped KP surfaces accurately describe the KP surface, points that intersected KP perimeters were queried.

Because this query entails more than 100,000 points, processing time outlasted six days.

After exporting the isolated contact points as a new shapefile, several interpolation methods could be applied to create a surface connecting individual remnant surfaces.

Three different interpolation methods were considered for analysis and compared to a rasterized Triangulated Irregular Network (TIN) surface (Figure 3.7). A TIN surface is a continuous surface of non-overlapping triangles, connected at vertices defined by input data points. Data points in this interpolation are the peneplain nodes. The rasterized

TIN surface can be created by a natural neighbor or linear interpolation technique.

Results in different methods are insignificant, and the cell size is as great as great as 962 m (the largest polygon created). The inverse distance weighted interpolation (IDW) method linearly weighs the interpolation of a pixel value in accordance to its distance from an input point. The created IDW surface, set at a low power value (the distal nodes are less influential in the resulting interpolation) has a cell size of 414 m. The spline interpolation method creates a surface by using curvature technique, and requires the 54 Texas Tech University, Timothy Anderson, December 2008 resulting interpolation to pass through all data points. The spline interpolation (Figure

3.7) has a maximum cell size of 414 m. The natural neighbor interpolation method, also known as the “Sibson” interpolation method, calculates an elevation value for each pixel by weighing the percentage of overlap of pre-constructed Voronoi polygons (TINs) to that of Voronoi polygons of that particular pixel. Because every individual pixel is created based on a resulting interpolation, the resulting cell size of a Natural Neighbor interpolated surface is the same as the DEM it was constructed from (10 m resolution).

The size of constructed polygons used in this interpolation, however, is dependent on the proximity of extracted pixels.

The interpolation methods in Figure 3.7 illustrate how manipulation of the data can change the predicted elevation of absent peneplain surfaces. Because it is impossible to quantitatively determine the accuracy of an interpolation surface, the question of how

‘realistic’ each surface cannot be quantitatively answered. The accuracy of the resulting interpolation surfaces, however, is presumably more suspect in the Siskiyou Mountains and central Klamath Mountains, where Klamath peneplain units have been mapped. The

Natural Neighbor interpolation surface was selected for in-depth analysis because it appears to be the most accurate. Throughout the rest of this thesis, this interpolation surface will be referred to as the Klamath peneplain surface, or KPS.

The spatial extent of the KPS is restricted to somewhat odd dimensions determined by its ‘convex hull.’ Extrapolation beyond the most distal erosional surfaces cannot be performed using the Natural-Neighbor interpolation method. Alternative interpolation methods (e.g. spline or IDW) can extrapolate to more extensive values 55 Texas Tech University, Timothy Anderson, December 2008

(Figure 3.7), but the usefulness of resulting interpolation values determined at greater distances is suspect.

56 Texas Tech University, Timothy Anderson, December 2008

Figure 3.7 This figure illustrates four interpolation surfaces generated by different built- in ArcGIS interpolation methods . Interpolations are based on erosional remnant surface elevations (western Klamath Mountains).

57 Texas Tech University, Timothy Anderson, December 2008

Even though the KPS and individual remnant surfaces are analyzed using the same data, they fundamentally describe two different uplift processes. The KPS describes the average surface of the all combined peneplain surfaces. Individual erosional remnant surfaces do not describe a surface, as classified by Molnar and England (1990). As will be explained later, the uplift of the KPS refers to uplift of a surface, whereas uplift of individual remnant surfaces refer to localized rock uplift.

3.4 Klamath Peneplain Topographic and Landscape Observations

The Klamath peneplain has been mapped between latitudes 40.32º N and 42.5º N and longitudes 124.2º W and 123.26º W. According to Irwin’s map (1997) (Figures 3.8), this once broad peneplain has been dissected into 248 isolated surfaces ranging between

18279 m 2 and 987.76 km 2 in area, where the mean area is 2.8 km 2 with a significant standard deviation of (8.8 km 2) (Irwin, 1997). Fifty-six surfaces have an area of less than a square kilometer (Figure 3.8). Ninety-six percent of all surfaces have areas less than 10 km 2. Perimeter measurements are just as variable, ranging between 614 m and 114,931

m2, with a mean perimeter and standard deviation of 8,946 m and 12,695 m, respectively.

The mean elevation of the 4,368 extracted KPS points (pixels) that underlie the mapped peneplain unit is 1033 m, with a standard deviation of 370 m. The most elevated erosional surface is greater than 2018 m atop the English Peak pluton in the Salmon subbasin and is 70 kilometers from the least elevated remnant surface (3 meters).

Figures 3.9, 3.10 and 3.11 describe the elevation of individual erosional remnant surfaces. One important observation in these figures is the eastward increase in 58 Texas Tech University, Timothy Anderson, December 2008 elevations. A best-fit line for values east of 123.7º W indicates dips to the west at 2.9º.

Although trends in erosional surface elevations are obscure west of123.7º (and will be discussed in Chapter 5), it is interesting to note that 2.9º is also the dip angle between the coast and the most elevated remnant surfaces in the east.

Figure 3.8 This figure uses data from Irwin (1997) and describes the average perimeter and area of mapped erosional remnant surfaces in this study. It is assumed that these remnant surfaces were once a single, contiguously connected surface.

59 Texas Tech University, Timothy Anderson, December 2008

Figure 3.9 The figure on the left illustrates the contact between the remnant surface and juxtaposing terranes. Elevation of these individual remnant surfaces (right) were created by interpolating DEM extracted pixels that underlie the mapped outline of the remnant surface (left).

60 Texas Tech University, Timothy Anderson, December 2008

Elevation (m)

0 - 200 200. - 400 400 - 600 600 - 800 800 - 1,000 1,000 - 1,200 1,200 - 1,400 1,400 - 1,600 1,600 - 1,800 1,800 - 2,000

Stereonet of Individual Remnant Surfaces

Kilometers 0 15 30 60 Figure 3.10 The map on the left illustrates the elevation of the erosional remnant surface (see text for details). On the right, the map and stereonet illustrate the mean dip angle and dip direction of the individual erosional remnant surfaces.

61 Texas Tech University, Timothy Anderson, December 2008

Figure 3.11 This figure illustrates the longitudinal relationship of mean and maximum elevations of topography with erosional remnant surface elevations. The map in the top left corner highlights the location of eastern erosional remnant surfaces. 62 Texas Tech University, Timothy Anderson, December 2008

3.5 The Klamath Peneplain Interpolation Surface (KPS)

The Klamath Peneplain Interpolation Surface (KPS) (Figures 3.12) interpolates a surface using elevations extracted from the contact pixels of Irwin’s digitized ‘erosional remnant surface unit.’ The resulting surface is a contiguously connected raster that spans the dimensions of all erosional remnant surfaces. As will be discussed in Chapter 5, its corresponding altitudes reflect surface movement of the western Klamath Mountains.

The KPS is at sea level near the coast and at approximately 2000 m in elevation in the central KMP (Figures 3.12). The mean elevation and standard deviation of the entire surface is measured at 1,027 and 400 meters, respectively. West of 124.1º W, surface elevations are generally less than 200 m in altitude. Near the coast, the surface defines two northwest striking ridges just north of the Klamath and Mad Rivers. These ridges form the most significantly measured altitudinal changes in the interpolation surface.

The dip angle of the KPS reflects the dip angle between remnant surfaces and

DEM-recorded dip angle of actual peneplain surfaces. The mean dip of the entire KPS is

2.4º to the west, although KPS angles can measure as great as 20.5º. The KPS dip angle is approximately 5.0 º near the coast. At more inland longitudes, dip angles are less, but can measure exceedingly high where anomalous surficial depressions are situated.

Anomalous depressions in the KPS correlate to anomalously low peneplain surfaces. Two of these surficial depressions are located at 41.28º N, 123 ºW (north of

Ironside Mountain) and 40.5º N, 123.2º W (western Trinity subbasin). There also exists a

100 km long, northeast striking surficial depression that correlates with a topographic lineament (yellow dashed line in Figure 3.5). The latter surficial depression in the eastern 63 Texas Tech University, Timothy Anderson, December 2008

Illinois subbasin is not well constrained, as its values are a consequence of interference in the in the Smith subbasin. Chapter 5 provides a detailed discussion of these depressions.

3.6 Creation of Erosion and Paleotopography Surfaces

Assuming the KPS (Figure 3.12) accurately describes a surface representing a contiguously connected erosional surface (Diller’s “Klamath peneplain”), its relationship to modern topography can be used to identify where erosion has occurred, where pre- existing topography existed, and/or net aggradation has occurred.

Figure 3.13 illustrates the spatial relationship between the KPS and modern topography. This figure is an oblique perspective of how the interpolation surface

(shaded black and transparent) projects through modern topography. As observed in the figure, modern ridges and valleys project above and beneath the KPS. Where modern topography is greater than the KPS, pre-existing topography and/or net aggradation is inferred. Where modern topography is less than the interpolation surface, net erosion of the western Klamaths are inferred.

Figure 3.14 explores the quantitative relationship between the KPS and modern topography. The surface on the left was calculated by subtracting elevation values of the

KPS from modern topography. Where values are large (blue), modern topography is less than the KPS and represents erosion. Where values are negative (white), paleotopography is inferred. Where values are near zero (yellow), peneplain surface elevations are similar to modern topographic elevations. Such calculations measure the vertical depth of erosion, amount of aggradation, and/or height of paleotopographic features. 64 Texas Tech University, Timothy Anderson, December 2008

Figure 3.12 This figure illustrates the natural-neighbor based interpolation surface (KPS) of individual erosional remnant surfaces. 65 Texas Tech University, Timothy Anderson, December 2008

Figure 3.13 This figure illustrates the interpolation surface (KPS of Figure 3.14) as a ‘tinted window’ and projects its through modern topography. Where this surface is greater than modern topography, net erosion since peneplanation is inferred. Where elevations of modern topography is greater than elevations of the interpolation surface occurs, pre-existing paleotopography is inferred.

Three E-W cross-sections (lines A, B and C in Figure 3.16) are constructed to

better illustrate the relationship between modern topography and the KPS. In these

constructed cross sections, red lines illustrate the relative location of the remnant

surface relative to modern topography (shaded red). The blue-shaded regions in the

cross-section reveal the vertical amount of removed material beneath the KPS.

66 Texas Tech University, Timothy Anderson, December 2008

Figure 3.14 This map was created by subtracting the KPS from modern topography. Positive values (blue) represent erosion, and negative values (white) represent paleotopography. E-W cross-sections are drawn on the right to illustrate relationships between modern topography (red) and the interpolation surface (bold red line).

67 Texas Tech University, Timothy Anderson, December 2008

Where modern topography (shaded red) is greater than the KPS (red-line), paleotopography is interpreted. Cross sections are vertically exaggerated in this figure.

3.7 Paleotopography

Positive paleotopography, in this thesis, is defined as topography at higher altitudes than projected elevations of the interpolation surface (KPS). Figure 3.15 is a map illustrating the location, vertical height and name of rocks situated in regions of suggested paleotopography. At least six regions may have been exposed prior to the development of the erosional remnant surface. The accuracy in identifying and measuring antecedent topography depends on the accuracy of the KPS. Because the spatial distribution of remnant surfaces fundamentally affect how reliable any interpolation is, confidence in the accuracy of the interpolation surface can be invested in regions west of the central Klamath Mountains.

Because remnant surfaces cap most ridges in the western Klamath Mountains, it is not surprising that paleotopography is absent where surfaces are densely mapped.

Likewise, the presence of paleotopography should be expected where the peneplain surfaces are absent. The following paragraphs offer arguments for and against the presence of paleotopography.

The largest paleotopographic feature more elevated than the KPS is the Siskiyou

Mountains. This includes the Bear Mountain mafic intrusive complex and WPT peridotite

(Figure 3.15). According to this map, the suggested antecedent topography forms a series of northwest striking ridges that exceed 900 m in two locations. Because there are no 68 Texas Tech University, Timothy Anderson, December 2008 remnant surfaces mapped near the range, paleotopography measured in the Siskiyou

Mountains is poorly constrained.

The existence and measurement of Siskiyou Mountain paleotopography is calculated using remnant surfaces in the central KMP, Coast Ranges, and southern

Siskiyou Mountains. An abrupt disappearance of remnant surfaces in the eastern Illinois subbasin makes it difficult to predict the location of the northern rim of the paleo-

Siskiyou Mountains. This abrupt disappearance of peneplain surfaces is coincident with a northeastern topographic lineament (yellow dashed line in Figure 3.5). It is possible that this linemanet is a high angle, northeast striking normal fault. If true, eastern remnant surface are predicted to exist beneath alluvium in the eastern Illinois subbasin.

It is also possible that the peneplain surfaces did mantle the Mesozoic-aged terranes and plutons of the Siskiyou Mountains, but is not mapped. If true, Siskiyou

Mountain surface uplift would exceed 2000 m in some locations. Given the steep increase in elevation of remnant surfaces mapped in coastal regions just north of the Klamath

River, exceedingly high rates of Siskiyou Mountain uplift is reasonable.

Figure 3.16 indicates the existence of paleotopography along a northwest striking ridge in the central Klamath Mountains at 41.25º N and 123.45º W where metavolcanic and metasedimentary rocks of the Western Hayfork terrane are mapped. Nearby remnant surfaces are mapped atop plutons 15 kilometers to the northeast and west, providing good constraint on the interpolation used.

69 Texas Tech University, Timothy Anderson, December 2008

The most southern region of suggested paleotopography is a northwest striking strip of the Pickett Peak Formation. Surrounding remnant surfaces are mapped in close proximity, indicating that the interpolation surface is well-constrained. The absence of a remnant surface along this ridge is unexpected because an erosional surface is continuously mapped along northern segments of the same ridge composed of the same geologic unit. Irwin (1997) mentions that unmapped remnant surfaces exist south of his map. Like the Siskiyou Mountains, it is possible that this southern ridge is capped by an unmapped surface. If such is the case, ridge elevations indicate a continued southward increase in uplift rates and calculated incision in the South Fork Trinity River.

Other regions of positive paleotopography include small lenses of the Josephine

Ophiolite in Oregon, metasedimentary rocks of the Franciscan complex west of the

Klamath-Trinity River junction, and a small hill of the Ironside Mountain pluton north of

Hayfork Creek. Because these topographic features are calculated to be less than 200 m in paleo-elevation, such features are insignificant and could be recorded in error.

Arguments against the existence of antecedent topography are not justified well enough to discredit these paleotopographic features. Furthermore, there is no sedimentological evidence that can justifiably discredit the existence of these features.

Evidence involving, the existence of paleotopography does not, as of yet, conflict with previous geologic studies.

70 Texas Tech University, Timothy Anderson, December 2008

3.8 Calculated Erosion from the Klamath Peneplain Erosion Surface

Erosional remnant surfaces cap many inland ridges that bound deeply carved

valleys. As a result, the KPS is at elevations much greater than modern topography, a

consequence of erosion. The Klamath peneplain erosion surface (KPES) (Figures 3.16

and 3.17) is created by subtracting the KPS (Figure 3.12) from modern topography and

extracting only positive values. All positive values in Figure 3.14 represent regions that

have been eroded beneath the peneplain.

The greatest depths of calculated erosion are spatially coincident with major

drainage systems today (Figure 3.17). As illustrated in the figure, the greatest amounts of

erosion are typically east of the junction between the Klamath and Trinity Rivers. The

maximum depth of erosion is 1,470 m in the Salmon River.

Southern rivers not in the Klamath Mountains that have significantly eroded into

their basement rock are the Mad River and Redwood Creek. The Mad River is locally

eroded downstream of a prominent knickpoint where river makes an abrupt westward

shift in its flow direction (knickpoints are discussed in Chapter 4). In this location of the

valley, the stream bed is nearly 800 m lower than juxtaposing ridges capped by remnant

surfaces. Near the coast, a remnant surface present in the Mad River valley at an

elevation similar to the river demonstrates a minimal amount of erosion. Only small

amounts of erosion is calculated upstream of the knickpoint, where the river bed

overflows Quaternary alluvium.

71 Texas Tech University, Timothy Anderson, December 2008

Figure 3.15 Map illustrates the existence of all elevations above the KPS. This is interpreted to be paleotopography. The name of the bedrock at the location of this paleotopography is labeled. Data is provided by Elder (2008).

72 Texas Tech University, Timothy Anderson, December 2008

With exception to its headwaters, the bed of the Redwood Creek is 500 to 600 m below the KPS throughout its entire course. The valley width of measured erosion is also uniform throughout the river’s course (usually less than 5 kilometers).

Although the South Fork Trinity River incises through rock in a narrow valley for its first 30 kilometers flow, its greatest erosive segments are downstream of its confluence with . Fifteen kilometers downstream of Hayfork Creek, the river is more than 1,000 m lower than the KPS. The amount of inferred erosion increases steadily to more than 1,100 m at the junction between the Klamath and Trinity River.

Near the headwaters of the South Fork Trinity River, the stream bed can be greater than

800 m below the KPS.

Large amounts of rock have been removed in the Trinity valley and joining tributaries (Figure 3.18). Erosion depth in the Trinity valley (and tributary valleys) is consistently greater than 1,000 m until the river is 15 kilometers south of the Klamath

River. River segments southwest of peneplain surfaces that cap Ironside pluton are usually at altitudes more than 1,100 meters below the KPS.

The deepest depths of calculated erosion are situated in the Salmon River,

Wooley Creek, and upstream Klamath River. These calculated depths of erosion are based on remnant surfaces that are at altitudes between 1400 m and 2000 m. Figures 3.18 and 3.19 demonstrate that nearly the entire subbasin has been significantly denuded beneath the peneplain. The upstream river segments of the Salmon River and

Wooley Creek are consistently at elevations 1,300 meters below the KPS. 73 Texas Tech University, Timothy Anderson, December 2008

The depth of the Klamath River below the KPS varies throughout its course.

Where the Klamath River is north of Wooley Creek, the river is 1,300 meters lower than the KPS. In contrast, less than 700 meters of erosion is calculated west of the confluence between the Klamath and Trinity Rivers. Because no remnant surfaces are mapped in the

Siskiyou Mountains, it is difficult to confidently constrain erosion in the Klamath River and contributing tributaries. Nevertheless, undated fluvial terraces in the Klamath valley at high elevations indicate Quaternary erosion in the Klamath Valley (Yoshinobu personal comm.., 2008).

Near the coast and north of the Klamath River, four rivers have experienced deep depths of erosion. The , the South Fork Smith, Middle Fork Smith and Illinois

Rivers flow in valleys at elevations significantly lower than nearby ridges capped by remnant surfaces. Blue Creek, a Siskiyou Mountain tributary flowing westward into the

Klamath River, is at elevations more than 900 m below the KPS. The greatest amount of erosion in the South Fork and Middle Fork Smith Rivers are in the Siskiyou Mountains.

Erosional depths in the South Fork and Middle Fork Smith Rivers exceed 900 and 700 meters, respectively. Prominent knickpoints appear to be associated with these deeper depths of erosion in all of the aforementioned tributaries.

The eastern Illinois subbasin is calculated to have experienced a large amount of erosion, however a lack of data makes this calculation speculative. East of a northeast striking topographic lineament (yellow dashed line in Figure 3.5), the Illinois River is

900 m lower than remnant surfaces in Coast Range topography. Increased erosion is coincident with a prominent knickpoint in the Illinois River downstream of where it 74 Texas Tech University, Timothy Anderson, December 2008 enters elevated topography of the Oregon Coast Range. Similarly, segments of the Chetco

River are more than 900 m lower than nearby remnant surfaces.

75 Texas Tech University, Timothy Anderson, December 2008

Figure 3.16 This figure illustrates the difference in elevation between the KPS (Figure 3.14) and topography. The location of the erosional remnant surfaces are shown in black and paleotopography is shown in white. Subbasins are labeled and paleotopography.

76 Texas Tech University, Timothy Anderson, December 2008

Figure 3.17 This figure illustrates the difference in elevation between modern topography and the KPS (Figure 3.14). Major rivers are also labeled. On the right side of the figure, oblique perspectives of the amount of erosion are shown with labeled rivers superimposed above the eroded terrane.

77 Texas Tech University, Timothy Anderson, December 2008

Chapter 4

River Profile Analysis

4.1 Introduction

Anomalously elevated topography in the Klamath Mountains Province has been deeply incised by rivers. Figure 4.1 illustrates the location of major rivers and subbasins.

Evidence of relatively deep incision is observed in the presence of deep, narrow valleys that bound major drainages, high erosion rates (Figure 3.16), and elevated fluvial terraces

(Merritts, et al., 1994; Irwin, 1997). Because a river’s incision rate is fundamental in governing the lowering of a landscape by mass removal, the shape of a longitudinal profile may reflect climatically and/or tectonically induced signals (Snyder et al., 2000).

These signals can be explored by identifying and mapping knickpoints and steep river segments.

4.2 Historical Background of Longitudinal River Profiles

In 1875, J.W. Powell introduced the concept of base level; the limit by which a landscape where no potential erosional energy exists (Powell, 1875; Stock and

Montgomery, 1999; Burbank and Anderson, 2001). In fluvial systems where a mature river is at steady-state equilibrium with its landscape, the river bed will not uplift, denude, or aggrade (Bull, 1991; Snyder et al., 2000; Burbank and Anderson, 2001).

Under these conditions, the longitudinal profile of a river will be smooth and concave up

(see Figure 4.2) (Wobus et al., 2006).

78 Texas Tech University, Timothy Anderson, December 2008

Hack (1957) established early fundamental geomorphologic concepts that were, in part, founded upon the shape of a river. Studying profiles of alluviated streams in

Virginia and Maryland, Hack documented relationships between a stream’s distance, drainage area, and particle size on the steepness of a river. He concluded that rivers, like the landscape they drain, will decrease in gradient in accordance with its length from the drainage divide, and the diameter of particles can exert control on the stream’s gradient.

Hack’s longitudinal profile (1) tracks the change in elevation of a stream by a

stream gradient referred to as the Hack gradient. Hack’s gradient (2) is the product of the

stream’s length and local slope at a stream reach. Analyzing rivers near the Mendocino

Triple Junction, Merrits and Vincent (1989) arrived at three conclusions relating uplift

and the Hack gradient: 1) a stream experiencing a low uplift rate results in high Hack

gradients only in its upper segment, 2) a stream experiencing an intermediate uplift rate

results in high Hack gradient values at the mouth, but is variable near its upper reach, and

3) a stream experiencing a high uplift rate is characterized by a convex profile and high

Hack gradient values throughout the entire river profile.

A stream segment, in this thesis, is defined as the portion of a stream bounded

breaks in stream gradient-drainage area patterns in log-log space.

Hack’s Longitudinal profile

z(x) = k h ln(x) (1)

Hack’s Gradient (1957)

kh = Slope * Length (2)

z – elevation of point on profile at distance (x) from drainage divide 79 Texas Tech University, Timothy Anderson, December 2008 x – distance of a point on the stream from the divide. h – an empirical constant approximately equal to 1.7

Hack’s Law

h A(x)=k ax (3)

A(x) - area at x distance from divide ka - slope * area h – an empirical constant approximately equal to 1.7

The utilization of Hack’s gradient is most appropriate in basins where fluvial is consistently proportional to stream length (Figure 4.3). The inconsistent width of most drainage basins makes the application of Hack’s gradient problematic. For example, a basin with a large upstream drainage area may result in misleading Hack gradient values upstream. For a better analysis of a stream’s gradient, Hack’s Law (1957) was developed.

The relationship between drainage area, stream profile concavity, and channel steepness of a particular stream at steady-state equilibrium was not derived, however, until the introduction of Flint’s Law, from which modern profile models are based. Under steady-state conditions (where uplift and erosion are equal), there exists a balance between local channel steepness (ks) and upstream drainage area. Under such conditions, the local channel gradient of a stream will decrease as a power law function because of a larger drainage area. The relationship results in a concave-up longitudinal profile. In

80 Texas Tech University, Timothy Anderson, December 2008 logarithmic space, drainage area and channel steepness are linearly and negatively related. The expression of a stream at equilibrium is written as:

-θ Se = k s *A (4)

ks - channel steepness

A- drainage area

θ- concavity

Channel Steepness

1/n ks = (U/K) (5)

U = uplift

K = erosion coefficient n = slope exponent

An important difference between Flint’s Law and Hack’s Law is the introduction

of channel steepness, ks. Channel steepness is the relationship between net uplift and net

erosion. A more in-depth discussion of channel steepness is developed in the following

section.

4.3 Modern River Incision Models

In the most recent advances of profile analysis, some workers developed

computer programming scripts that extract, calculate, and normalize river profile

information for quantitative and comparative analysis of potential tectonic signals

(Wobus et al., 2006). Modern study of river incision is divided into two fundamentally

different models; a detachment-limited model and a transport-limited incision model. The 81 Texas Tech University, Timothy Anderson, December 2008 former model is more appropriate for bedrock incising channels, whereas the latter is more appropriate for alluviated channels (Howard and Kerby, 1983; Seidl and Dietrich,

1992; Howard, 1994; Seidl and Dietrich, 1998; Snyder et al., 2000; Kirby and Whipple,

2000). The detachment-limited model assumes that channel erosion is accomplished by the detachment of basal bedrock in response to fluvial shear stress. Because of the assumption that discharge is proportional to drainage area, larger drainage areas will incise bedrock at higher rates. The transport-limited incision model assumes river incision is weighed by the influence of the sediment load within a river. In this model, an increase in sediment flux will increase incision rates due to the presence of sedimentation

‘tools’ (Sklar and Dietrich, 2001). If the sediment flux increases beyond a river’s transport capacity, a stream’s bedload prevents erosion, know as the ‘shield effect’ (Sklar and Dietrich, 2001).

It is likely that incision in most rivers involves both mechanisms. The detachment-limited incision model is more popular because it is easier to constrain with empirically determined constants. Uncertainty of results using either model exists. As pointed out by Kirby and others (2003), at least four stream incision uncertainties need to be resolved: 1) the relationship between channel gradient and incision, 2) channel width adjustment to differential rock uplift, 3) the extent and nature of upstream transitions to incision, and 4) the role of sediment flux and armoring within a channel bed.

4.4 Detachment-limited stream power model

In this study, all rivers studied are assumed to follow the detachment-limited stream power model. The evolution of a longitudinal profile is mathematically expressed 82 Texas Tech University, Timothy Anderson, December 2008 in the following equation (Howard, et al., 1994; Whipple and Tucker, 1999; Whipple et al., 2000a; Kirby and Whipple, 2001; Seidl and Dietrich, 2002; Kirby et al., 2003; Wobus et al., 2006; Delong et al., 2007):

(dz/dt) = U(x,t) - KA mSn (6)

(dz/dt) – the rate of change of channel elevation

U – rock uplift rate relative to the geoid

K – a dimensionless coefficient of bedrock erodibility

A – upstream drainage area

S – local channel slope (change of elevation over change of stream distance) m, n – basin hydrology, hydraulic geometry, and erosion coefficients (empirically determined).

In this equation, there are three empirically defined coefficients, K, m, and n. The coefficient K is inversely related to a stream’s (or landscape) erodibility. Under steady state equilibrium conditions, uplift, U, is equal to erosion, K (Snyder et al., 2000).

Snyder and others (2000) determined that uplift could affect K by as much as 6 orders of magnitude. Large values of K correspond to less resistant lithologies and/or higher uplift rates. Stock and Montgomery (1999) determined that lithology variation affects K by as much as 5 orders of magnitude between 10 -2 -10 -5 m 0.2 /yr for volcaniclastic rocks and mudstones of Japan and California and 10 -6 – 10 -7 m 0.2 /yr for granitic rocks and metasedimentary rocks of Australia. Using three-dimensional modeling that accounts for

83 Texas Tech University, Timothy Anderson, December 2008 drainage density, elevation, relief, and hypsometry, Delong and others (2007) determined

K to range between 0.3 and 0.09 m .2-.4 k.y. -1.

The exponents, m and n, are affected by the relationship of channel erosion with upstream drainage area (m) and local channel slope (n) (Howard et al., 1994; Stock and

Montgomery, 1999; Whipple and Tucker, 1999). If incision is proportional to stream power, m and n are approximately equal to 1 (Seidl et al., 1994; Stock and Montgomery;

1999; Burbank and Montgomery, 2001). If erosion rates are proportional to shear stress

(bedrock channels), m is equal to 0.3 and n is equal to 0.7 (Stock and Montgomery,

1999). Stock and Montgomery (1999) calculated that the value of m ranges between .3 and .5, and is dependent on fluvial factors such as a stream’s width, a basin’s hypsometry, and bankfull discharge. Empirical values of the slope exponent, n, vary between 2/3 and 5/3 (Howard and Kerby, 1983; Hancock et. al, 1998; Stock and

Montgomery, 1999; Whipple et al., 2000a; Kirby et al., 2003) depending on erosion mechanisms. A large slope exponent results in a reduced channel steepness value, and by extension a reduced channel equilibrium gradient. The slope exponent value, however, is controversial and difficult to constrain because the relationship between incision rate and channel gradient is poorly understood (Kirby et al., 2003). The ratio of m and n describe the concavity of an equilibrium profile and is typically much easier to calculate (Seidl and Dietrich, 1992; Snyder et al., 2000).

Figure 4.5 illustrates river profiles of different concavity measurements.

Concavity is the mathematical integration of a channel’s slope and measures the distribution of relief along a river profile. Where the concavity of a stream is zero, the 84 Texas Tech University, Timothy Anderson, December 2008 slope along a river is constant. Where concavity equals 1, a stream is steepest upstream.

Where concavity is negative (convex), a stream is steepest near its base level.

Concavity is described as three measurements: 1) the concavity index, 2) the

intrinsic concavity of the system ( θ = m/n), and 3) the reference concavity, θref . The

former two concavities are only equal under steady-state conditions (Snyder et al., 2000;

Kirby et al., 2000).

The concavity index of a user-selected stream segment is calculated from a

regression line. Studies have determined that lithology, regional precipitation, and uplift

influence concavity. In coastal Oregon, Vanlaningham (2006) determined that concavity

of a stream overflowing less resistant sedimentary rocks to be greater than that of

resistant volcanic rocks. Kirby and Whipple (2002) measured convexities in river profiles

that cross the Siwalik Hills anticline in Central Nepal and high concavities upstream and

downstream of the anticline. Roe and others (2002) demonstrated that increased

orographic precipitation can increase the concavity of a longitudinal profile.

The intrinsic concavity ( θ = m/n) represents basin hydrology and channel hydraulic geometry parameters, and is fundamentally defined as the ratio between m and n. While erosion affects the actual concavity of a stream, the intrinsic concavity is not affected by erosion (Seidl and Dietrich, 1992; Kirby and Whipple, 2001; Kirby et al.,

2003). Empirically determined concavity values typically range between .35 and .6

(Kirby et al., 2003; Wobus et al., 2006). If the stream profile is influenced by transport- limited mechanisms, a similar range of values exists (Whipple and Tucker, 2002).

85 Texas Tech University, Timothy Anderson, December 2008

The reference concavity, θref , is a concavity index value representative of all

streams within an area. This value is necessary for normalizing the channel steepness

index for a comparative analysis between all streams.

Channel steepness (eq. 5) is represented as the relationship between uplift, erosion, and channel slope. Figure 4.5 illustrates the affect of a channel’s steepness on a river profile. Increases in channel steepness can be manifested as an increase in rock uplift rate, and/or a decrease in erosive efficiency (Snyder et al., 2000; Kirby and

Whipple, 2001; Kirby et al., 2003; Duvall et al., 2004; Vanlaningham, 2006). There is no apparent relationship between channel steepness and concavity exists (Kirby et al., 2003).

The normalized channel steepness sets a reference concavity in an equilibrium channel gradient equation to calculate the relative steepness of a river segment. The normalized channel steepness (7) is equal to a channel’s local steepness multiplied by the midpoint drainage area taken to a power equal to the difference between the reference concavity and regressed concavity. Channel steepness between all river segments can be compared to one another, where greater values represent steeper rivers. Normalized channel steepness is expressed as the following:

(θ – θ) ksn = k sAcent ref (7) ks = channel steepness, determined from regression

θref = reference concavity, default is .45

θ = concavity of regression line

(logA +logA )/2 Acent = 10 max min (8) 86 Texas Tech University, Timothy Anderson, December 2008

Acent – midpoint drainage area along regression segment

Amax – maximum drainage area of user-selected regression segment

Amin – minimum drainage area of user-selected regression segment

Channel steepness and concavity are calculated from a regression line between user-selected upper and lower bounds of a stream segment. When the actual concavity is significantly larger or smaller than the a user-defined reference concavity (e.g. [ θ-θref ] >

.2), the resulting normalized channel steepness value is less meaningful (Wobus et al.,

2006) because it represents an average steepness of the entire river segment, but does not

capture local gradient changes within a river segment (Kirby et al., 2003). For this

reason, normalized channel steepness of rivers that exhibit complicated shapes is difficult

to confidently measure. Nevertheless, Duvall and others (2004) demonstrated that

calculated ksn values of river segments that exhibit a wide range of θref values varied less

than 2% ratio. In summary, an excellent way to analyze the spatial variation of channel

steepness of rivers of varying drainage area across a region is to compare the k sn values of different streams.

A knickpoint is defined as an oversteepened reach of a stream (Howard et al.,

1994). A knickpoint can be the result of a discrete tectonic event (e.g. a fault), the contact between two underlying lithologies that vary in resistance, an abrupt increase in discharge (e.g. at the confluence between two rivers), or a transient response due to base level fall (Hack, 1957; Gardner, 1983; Seidl and Dietrich, 1992; Seidl et al., 1994;

Howard et al., 1994; Montgomery and Buffington, 1997). Numerous studies have described and modeled knickpoints by their size, shape, spatial distribution, and 87 Texas Tech University, Timothy Anderson, December 2008 migration characteristics in response to base level fall (e.g. Gardner, 1983; Howard et al.,

1994). Knickpoints within this thesis are identified by anomalously abrupt changes of stream gradient in logarithmic space which always correlate with abrupt knickpoints and/or broad convexities in a longitudinal river profile. Figure 4.6 illustrates the shape of a river profile with and without knickpoints.

There are three models for the development of fluvial knickpoints: 1) parallel retreat, 2) slope replacement, and 3) differential erosion without propagation (Gardner,

1983; Seidel and Dietrich, 1992).

In the first model, the angle between the downstream river segment and the oversteepened knickpoint lip remains constant. This model explains the expected migration style at the downstream contact of an extremely resistant substrate. Howard and others (1994) determined that this style of knickpoint migration occurs when stream power is the dominant erosion mechanism (m and n are equal to 1).

According to the second model, slope replacement, the angle between the downstream river segment and the oversteepened knickpoint lip increases during knickpoint migration. A knickpoint is expected migrate according to this model if a stream is overflowing a homogeneous lithology. Howard and others (1994) simulated this style of knickpoint migration when he assumed the erosion rates are proportional to bed shear stress (m =.3 and n =.7).

In the third model, a knickpoint does not migrate. The less resistant lithology will continue to erode indefinitely.

88 Texas Tech University, Timothy Anderson, December 2008

Where a knickpoint is oversteepened, stream power increases and induces erosion of the outer lip of the knickpoint (Gardner, 1983; Seidl and Dietrich, 1992; Howard et al.,

1994; Stock and Montgomery, 1999; Burbank and Anderson, 2001; Vanlaningham,

2006). Erosion of the knickpoint outer lip initiates the upstream migration of the knickpoint. A knickpoint will migrate at higher rates in rivers of larger drainage area and will translate from stream trunks into tributaries at respective confluences (Merritts and

Vincent, 1989).

4.5 River Profile Construction Methods

The construction and quantitative measurement of longitudinal profiles was performed using ArcDesktop and Matlab software packages. The necessary ArcGIS toolbars and Matlab codes were acquired from http://geomorphtools.org/about.htm . The development and preparation of these tools have only recently been integrated into a platform involving Matlab codes originally written by Noah Snyder and Kelin Whipple

(Snyder et al., 2000). During the last decade, Daniel Sheehan and Eric Kirby intergrated these codes intoArcGIS (Wobus et al., 2006; Kirby et al., 2003).

Using functions from the Arc GIS spatial analyst toolbox, a Profiler DEM

Preparation tool box, and a series of Matlab scripts that use the Matlab Statistics toolbox , the creation and quantitative evaluation of river profiles could be accomplished in five steps: 1) DEM preparation and partitioning, 2) the creation of river flow direction and accumulation grids using Arc/Info built-in tools, 3) stream identification, extraction, and

89 Texas Tech University, Timothy Anderson, December 2008 smoothing using ArcGIS tools and Matlab codes 4) stream profile analysis using Matlab codes, and 5) the conversion of analysis results back into an ArcGIS platform.

DEM extraction is described in Chapter 3.1. Because of the processing memory required to run codes in Matlab are roughly limited to about a 5000 pixel by 5000 pixel grid, memory limits are exceed for a DEM with even moderate resolution. To overcome this dilemma, the master DEM of the entire KMP at 10 m resolution was divided into large National Hydrography Dataset (NHD) subbasins. DEMs were then extracted into subbasin watersheds that are. The extraction of watershed subsets were accomplished by clipping a DEM to newly created watersheds. Watersheds were created using an add-in toolbox called the “Hydrology Module,” developed by Peter Isaacson. In this study hundreds of watershed-sized DEMs were prepared. To investigate longer rivers (e.g. the

Klamath and Trinity River), subbasin DEMs were resampled to a coarser resolution. To gain confidence that river profile data was not lost by DEM resampling and other data manipulation processes, longitudinal profiles were also created by extracting elevation data from major National Hydrography Dataset (NHD) rivers of raw DEMs and were found to be similar in shape to modern stream profiles.

A series of grids providing information of flow direction and upstream drainage area were created using built-in ArcGIS functions. Before creating a flow direction grid, a grid that ‘fills’ pits in the DEM was first generated so that rivers would continuously flow through lakes, and unnatural depressions. Grids that provide flow direction and upstream drainage area were created using arc built-in tools. Accumulation and flow direction grids were then exported as ascii files, a text format that can be read in Matlab. 90 Texas Tech University, Timothy Anderson, December 2008

The resulting files must be saved under the nomenclature file_namedem.txt and file_nameacc.txt. These functions can be performed using Spatial Analyst hydrology tools and conversion tools in the ArcGIS toolbox. Alternatively, the development of a

“Profiler DEM Preparation” toolbox creates all necessary grids and conversions.

It is strongly encouraged that all digital data, matlab codes, and resulting profiles are saved in a directory structure that is organized in the following system, as outlined by shortcourse manual. Three separate folders should be created. A folder containing all

written matlab codes should be saved in a separate directory. This folder will be

designated the Matlab working directory. The other two folders, a matlab and arcmap

directory, should be saved in a folder dedicated to a particular subbasin. It is important to

name these folders ‘matlab’ and ‘arcmap’ because matlab codes will call upon these

directories during processing. The latter directory should contain the created text files.

Prior to stream selection, parameter information must be specified. Figure 4.4 is

the parameter window that includes all the necessary parameters for profile calculation.

In addition, the values in Figure 4.4 are the suggested values for analysis on a DEM with

a cellsize of 10 m.

The working folder in the parameters window must be set to the arcmap directory.

The raw DEM should be selected in the “Select your DEM” option. The project name

should follow some nomenclature scheme that refers to the selected streams. The Theta

Ref option refers to the reference concavity mentioned in section 4.2 has a direct affect

on the normalized steepness values of streams. In the following option, steps are

removed from the profile. If this option is selected, smoothing operations cannot be 91 Texas Tech University, Timothy Anderson, December 2008 performed. The spike remover will remove spikes in the river profile which commonly occur in DEMs with lots of pits. The size of the smoothing window is the length in meters of a moving average window. Streams measured from DEMs of 10 m cellsize were smoothed over a 250 m window. Smoothing over larger windows was performed for larger rivers that required coarser DEMs. The contour sampling interval is the vertical distance raw slopes are calculated from. The “Auto k_sn Window” refers to the width of the window along a stream for calculating the normalized stream gradient between user- selected regression limits. The “Search Distance” selection refers to the number of cells downstream from the user-selected stream point to confirm the stream is in an actual channel. To define where the uppermost stream pixel is located, one can specify the minimum number of inflowing pixels in the “Minimum Accumulation” option. Stream selection in this investigation tentatively applied two lines of criteria; 1) the upstream drainage area was greater than 1,000,000 square meters and 2) the river was located in a spatially valuable region of a watershed. The location and parameter information will then be saved within the arcmap directory.

Because files in the matlab format are more efficiently processed in Matlab, ascii files are converted into their matlab format equivalent using the ‘arcdemtxt2matlab’ code.

The matlab code, ‘profile51,’ uses these created matlab files to generate stream profiles for analysis.

Stream Profile Analysis

To quantitatively measure streams and/or stream segments, a user must specify the upper and lower limits in which to run a regression that will best represent the stream. 92 Texas Tech University, Timothy Anderson, December 2008

These regression limits can be selected along a longitudinal river profile (elevation versus distance from mouth), a profile illustrating stream gradient and drainage area relations in logarithmic space, or a profile illustrating stream gradient and stream distance relations in log space. Between these regressed limits, the channel steepness, normalized channel steepness calculated using a reference concavity (See Chapter 4.2) and fitted concavity can all be measured.

If a stream segment is experiencing uniform uplift in a single substrate, then there exists 3 expected relationships among three different river profiles: 1) a river segment will appear smooth and concave in its longitudinal profile, 2) a river segment will have a constant negative slope in gradient versus drainage area in logarithmic space, and 3) a constant negative gradient versus river distance (Wobus et al., 2006). Abrupt deviations in these expected relationships along a river profile, also known as knickpoints, are at the boundary of two stream segments that experience different incision rates relative to their rock uplift rate (Chapter 4.3). Knickpoints and regression limits can be identified at such locations of the river profile.

As an alternative to user-selected upper and lower regression limits, the normalized channel steepness of every stream in a catchment greater than a user-defined drainage area will be created using the ‘profile51_batch’ matlab code. Because regression limits are not selected, the subjectivity of profile analysis is limited.

Using the Profiler Toolbar in ArcGIS, the individual stream segments and

knickpoints can be imported and converted into shapefiles. The resulting shapefiles can

93 Texas Tech University, Timothy Anderson, December 2008 then be classified by their steepness and concavity attributes and studied from a spatial perspective.

Another useful approach to profile analysis among multiple rivers is to overlay their longitudinal profiles and slope-area plots. Thus, the profile shape, gradient patterns, and knickpoint elevations can be compared. This is best accomplished using Adobe

Illustrator.

4.6 Rivers in the KMP and Adjacent Regions

Streams within the Klamath Mountains Province and adjacent areas typically have drainage areas greater than 1,000,000 square meters and stream lengths that range between 15 and 450 km. Most stream segments that flow into major rivers are less than

40 km in length. To the south, rivers are largely bounded by narrow spaced ridges oriented to the northwest (e.g. the Eel, Mad, Redwood, and South Fork Trinity Rivers).

The Central KMP is largely incised by the Klamath and Trinity Rivers, however many

Central KMP tributaries of 50 km or longer (e.g. Wooley Creek, Salmon River, New

River, Scott River) are responsible for the incision of Klamath Mountain topography.

The Siskiyou Mountains are drained to the south by smaller tributaries of the Klamath

River. The northern Siskiyou Mountains are drained by the Applegate and Illinois subbasins. The Coastal Regions include the Smith, Chetco, Rogue and Coquille Rivers.

In northern California/ southern Oregon, the Cascades Volcanic Arc (CVA) is largely drained by the Rogue, Klamath, and Sacramento Rivers.

94 Texas Tech University, Timothy Anderson, December 2008

Normalized stream gradients are significantly steeper in the central Klamath

Mountains than surrounding regions (Figure 4.8). Here, normalized stream gradients

(using a reference concavity of 0.45) are greater than 100. Stream segments that are south or north of the central Klamath Mountains exhibit lower normalized stream gradients, often less than 60. Normalized gradients of streams within the Coast Ranges are steeper in the Chetco and western Illinois subbasins. North of the Rogue River and east of the

Klamath Mountains, normalized gradients of rivers are typically less than 60, but locally greater than 100.

Measurement of concavity indices of stream segments in the KMP are illustrated in Figure 4.9. There do not appear to be any observable correlations in the map. This may be related to the overabundance of knickpoints.

4.7 Longitudinal River Profiles

Streams of significant length and drainage area transect the topography of the

KMP in deeply incised valleys. Traversing different lithologic, climatic, and tectonic boundaries, the longitudinal profiles of the Eel, South Fork Trinity, Trinity, Klamath,

Smith, Illinois, and Rogue rivers exhibit different characteristics (Figure 4.9 ). All streams

flow westward into the Pacific Ocean over different lithologies and climatic regimes.

Most major rivers have headwater elevations between 1500 m and 2000 m. The

majority of stream relief is within the first 10 to 20% of stream length in longitudinal

profiles of the rivers. At elevations around 500 m, many rivers (e.g. Klamath, Rogue,

Trinity, Eel, Illinois, and Mad Rivers) decrease in channel gradient. Many major rivers in 95 Texas Tech University, Timothy Anderson, December 2008 the KMP have been dammed. The Trinity River, for example, has two major dams. The

Klamath River has seven.

All rivers flow across different lithologies. The Illinois and Rogue River overflow alluvial sediments for significant distances. Particularly when traversing alluvium, rivers have extremely low channel gradients. The upstream segments of the Klamath and Rogue

Rivers traverse volcanic rocks. Where traversing volcanic landscapes, river channels can have low gradients, but convex and/or chaotic longitudinal profile shapes near volcanic edifices. The Klamath, Trinity, South Fork Trinity, and Smith Rivers overflow bedrock of differing KMP terranes. Where traversing bedrock, stream profiles can be exceedingly steep. In the Trinity and Klamath Rivers, longitudinal profiles become less concave or very nearly convex when overflowing rocks of the WPT and CM terranes. The Eel,

South Fork Trinity, and Mad Rivers overflow Franciscan rocks throughout the majority of their paths. Longitudinal profiles of rivers in the Franciscan Complex can vary in shape. It is not uncommon for rivers to have large knickpoints or broad convexities that could correlate to mapped faults or contacts between large knockers.

River profile analysis is organized into four geographic regions: 1) the southern

KMP, 2) the Central KMP, 3) the Siskiyou Mountains, and 4) the Coast ranges. For illustrative purposes, tributaries have been color-coded based on their geographic, geomorphologic, and geologic characteristics. Knickpoints in longitudinal profiles are identified by blue cross hairs. The size of the cross-hair symbols correlate to the size of the knickpoint.

96 Texas Tech University, Timothy Anderson, December 2008

4.8 Rivers in the Southern Klamath Mountains Province

The Eel Subbasins: Figures 4.11, 4.12, and 4.13

The northwest striking Eel subbasins are the most southerly subbasins within the study area. Tributaries that flow into the northwest bound Eel River are presented in this section in three large subbasins: 1) the upper Eel subbasin, the middle Eel subbasin, and the lower Eel subbasin.

The upper Eel subbasin and middle Eel subbasin descend from a northward striking ridge that defines the Great Valley western boundary. This ridge can exceed 2300 m of elevation within the Middle Eel subbasin. Although hillslopes within the upper subbasins can exceed 30 degrees, valley floor hillslopes are moderate or exceedingly low at valley floors.

The deeply incised streams that descend from the eastern ridges are relatively steep (k sn > 70) and have deep valleys with more than a kilometer of relief. In contrast, western streams within the Upper Eel and Middle Eel subbasins have very low stream gradients (k sn < 40). All streams flow over different Franciscan Complex rock units.

Longitudinal profiles exhibit numerous knickpoints at varying elevations.

The Lower Eel subbasin (Figure 4.13) is a narrow northwest trending basin that comprises northern tributaries that feed into the Eel River. All tributaries in this region have moderate or very low stream gradients (k sn < 50). Knickpoints are preserved in most profiles at elevations of 750 m. The lowest knickpoints are preserved in coastal streams within the Yager Formation.

Middle Fork Eel River: Figure 4.14 97 Texas Tech University, Timothy Anderson, December 2008

The flows from an elevatred topographic dome that

defines the northeastern terminus of the Great Valley (Figure 4.14). The Middle Fork

drains into the Eel River at an elevation of 262 m. Hence, the majority of the stream relief

is preserved within the southwest directed, upper Middle Fork Eel River.

The main trunk of the Eel River can be characterized as a highly concave ( θ=

0.98), gently steeping (k sn = 50) river bounded by narrow, northwest trending ridges.

Taking the middle and lower stream segments together, the river is moderately steep

(k sn = 85) and moderately concave ( θ= 0.81).

For the first 50 kilometers of the stream (Middle Fork Eel River), the river

overflows fault-bounded greywackes of the Franciscan Complex. An abrupt increase in

profile steepness occurs at 1300 m elevation where the river overflows diamictite

knockers. In this location, the river profile is two to three times steeper (k sn = 130) than upstream and downstream river segments (k sn < 50). Between 150 and 250 km from base level, the longitudinal profile exhibits a subtle convexity characterized by a diverse range of Franciscan knockers. The Eel River overflows Quaternary alluvium for its least elevated 140 kilometers.

Van Duzen River: Figure 4.15

The west-northwest flowing Van Duzen River descends from elevations of 1250 m into the Eel basin. The profile shape of the Van Duzen River is characterized by a very large convexity. The knickpoint at 750 m elevation is located 53 km from the ocean, where the Van Duzen River to the west. Where the river is 30 km from the coast, the river flows over Franciscan rocks. Mapped geology of this basin is provided by Irwin 98 Texas Tech University, Timothy Anderson, December 2008

(1997) until 50 km from the coast. Upstream of the knickpoint, the localized concavity is low ( θ=.48) and gently steeping (k sn = 37), whereas downstream the channel steepness is much higher (k sn = 79) and exceedingly concave ( θ= 5.0). Downstream of the Franciscan

rocks, the river overflows the Wildcat and Yager sandstones and Quaternary alluvium,

has a very low channel steepness (k sn < 45), and is exceptionally concave ( θ> 5.0).

The Mad-Redwood Subbasin: Figure 4.16

The narrow, northwest striking Mad-Redwood subbasin is a coastal subbasin west of the central KMP that descends from northwest striking ridges that exceed 1800 meters in elevation. The subbasin encompasses the northwest flowing Mad and Redwood rivers.

The Mad watershed is a deeply incised, narrow, northwest striking watershed that begins its headwaters more than 125 km from its base. Drainage divides on either side of the watershed are usually less than 15 km apart, and valley floors can be more than a kilometer lower than adjacent ridges. Precipitation throughout the subbasin is approximately 70 annual inches, however a local topographic bulge that results in a southwestward offset of the river receives more than 100 inches of annual precipitation.

To the north, the Redwood Creek headwaters flow from this bulge into a narrow, north- northwest striking watershed. Despite topography within the 80 km long watershed exceeding 1600 m, the majority of watershed elevations are less than a kilometer. The ridges that bound Redwood Creek are less than 10 km apart and relief within the watershed is usually less than 500 m. Coastal streams are also included within the subbasin and are similar in shape to the Redwood Creek.

Redwood Creek: Figure 4.17 99 Texas Tech University, Timothy Anderson, December 2008

The northwest flowing Redwood Creek descends from a local topographic dome

that is a segment of a northwest striking ridge. Flowing over Franciscan rocks

throughout its entire course, the longitudinal profile of the river is concave and

moderately steep ( θ= 0.80, k sn = 59). The lowermost segments of the profile can be characterized by a number of small, abrupt knickpoints in coastal streams.

Mad River: Figure 4.18

The northwest flowing Mad River descends from, parallels and is bounded by

northwest striking ridges. A large knickpoint at an elevation of 600 m divides the river in

two segments. Upstream of the knickpoint, the river has a low concavity and gradient ( θ=

0.48, k sn = 37). Here, the stream at it traverses a regional concealed fault, resulting in a series of abrupt ‘stair-step’ knickpoints preserved within the upper segment of the river profile. Downstream from the knickpoint, the river is steep (k sn = 79) and concave ( θ= 5.0). The longitudinal profile increases in gradient and concavity where the river overflows less resistant Franciscan mods.

4.9 Rivers in the Central Klamath Mountains Province

South Fork Trinity Subbasin: Figure 4.19

The northwest flowing South Fork Trinity subbasin is the most southern subbasin within the Central Klamath Mountains. The uppermost Trinity River headwaters flow from an arcuate-shaped ridge in the southeast. A northwest striking ridge divides the southern Klamath Mountains into the southern Hayfork watershed and northern South

Fork Trinity watershed. 100 Texas Tech University, Timothy Anderson, December 2008

The South Fork Trinity River and Hayfork Creek have headwaters along a

northeast trending ridge that defines the southern corner of the subbasin. The deeply

incised South Fork Trinity River is always less than 17 km in wide. The western ridge

that bounds the South Fork Trinity River is commonly at elevations greater than 1500 m,

although elevations of an interior ridge composed of WPT Rattlesnake diamictite rarely

exceeds 1300 m. Although the western drainage divide is usually at elevations more than

100 m greater than the South Fork River, hillslopes are consistently steep (hillslope> 30º)

east of the river. These steeper hillslopes are spatially coincident with the intrusive and

cataclastic rocks of Galice Formation. West of the South Fork Trinity River, the less

resistant South Fork Mountain Schist can be tracked by its anomalously low hillslopes

north of the Klamath River. The interior ridge is capped by the largest erosional remnant

surface discussed in Chapter 3. Annual precipitation within the South Fork Trinity

watershed is usually 65 inches.

The more northern Hayfork Creek watershed is distinctly different from the South

Fork Trinity watershed. Upper stream segments of the Hayfork watershed flow from elevated ridges into an inner, low relief (hillslope< 1º) valley. This low relief landscape is at an elevation of 700 m and is overlain by the Tertiary-aged, Weaverville sandstone and Quaternary alluvium (Irwin, 1997). Most parts of the Hayfork watershed receive 45 inches of annual precipitation.

Upstream of the confluence between the South Fork Trinity and Hayfork Rivers, both rivers ‘flatten out’ at an elevation of 700 m. Tributaries within the Hayfork Creek watershed (both up and downstream of alluvium deposition) have very low stream 101 Texas Tech University, Timothy Anderson, December 2008

gradients (k sn < 55). Tributaries that feed into the South Fork Trinity River have high or extremely high channel gradients (70

Trinity River-Hayfork Creek junction, tributaries are steep and less concave (k sn > 85, θ<

0.45). Although numerous knickpoints are present in most tributaries within the

subbasin, most knickpoints cluster at an elevation of 1000 m.

South Fork Trinity River: Figure 4.20

The northward flowing South Fork Trinity River (Figure 4.12) descends from

2000 m elevation to 150 m at its confluence with the Trinity River. From map view, the north-northwest flowing South Fork Trinity River parallels a thin veneer of fault-bounded

Galice cataclasites and metasedimentary rocks. The entire length of the South Fork

Trinity River exhibits a less concave profile that is steep ( θ =0.51, k sn > 72). The steep and less concave upper stream segment of the profile (ksn = 83, θ = 0.61) overflows Franciscan rocks and into Galice cataclasite rocks. The steepest subset of this river segment is a less concave (k sn = 109, θ = 0.48) section of the river that overflows a series of thrust-bounded

Franciscan rocks. Downstream in the Galice Formation, a subtle convexity in the profile is located at 700 m elevation, where the river is deflected to the south by the nearby Bear

Wallow pluton. A second subtle knickpoint is located at an elevation of 360 m in diamictite rocks of the WPT Rattlesnake terrane. Downstream of this knickpoint, the river is classified as a very concave and gently flowing stream (θ = 0.97, k sn = 56).

Upstream of the latter two knickpoints, Quaternary alluvium is deposited.

Trinity Subbasin: Figures 4.21, 4.22, 4.23 102 Texas Tech University, Timothy Anderson, December 2008

The Trinity subbasin is an arcuate shaped basin that descends from ridges with elevations of more than 2700 m in the Trinity Alps to the confluence between the Trinity and Klamath River. The valley width in the eastern and western areas of the subbasin is less than 40 km, but exceeds 60 km in the central subbasin. This change in valley width is coincident with a change in the Trinity River flow direction from southwest to northwest.

A majority of the drainage area in the subbasin is north of the Trinity River.

Subbasin hilllslopes regularly exceed 25º, but occasionally exceed 40º. Steeper regions correlate to erosion-resistant terranes of the KMP. Examples of shallow hillslopes in this subbasin are mapped within the following 5 lithologic boundaries: 1) the erosional remnant surface, 2) the Weaverville formation, 3) glacial deposits, 4) WPT diamictites, and 5) Quaternary alluvium.

Forty of the one hundred and twenty-two mapped Quaternary glacial deposits in the Klamath Mountains are located in the Trinity subbasin (Irwin, 1997). Many of these mapped deposits are located in u-shaped valleys that radially extend from a topographic dome west of . The most eastern glacial deposits mapped in the study area are located in the upper reaches of the Little Trinity and East Fork Trinity Rivers.

Trinity subbasin streams are some of the steepest and most deeply incised streams in the KMP. Throughout the watershed, streams north of the Trinity River appear to be either steep or very steep (90> k sn > 160) in deeply carved valleys with more than 1000 m of relief. Comparison of channel steepness between streams within southern watersheds (e.g. Browns Creek watershed or even the Hayfork watershed) and streams in

103 Texas Tech University, Timothy Anderson, December 2008

northern watersheds (e.g. Canyon Creek) reveals a strong spatial contrast in stream

gradients (see Figure 4.27).

Nearly every stream in the Trinity subbasin exhibits a knickpoint. From map view, knickpoints are located along the upper reaches of the upstream tributaries of the

Trinity River. The longitudinal profiles of these streams reveal that many knickpoints are at elevations of 1500 m. Numerous knickpoints at different elevations are mapped on stream profiles along streams that overflow the Canyon Creek pluton. Another pluton, the Ironside Mountain pluton exhibits numerous knickpoints at elevations of 1000 m. The pluton, in this region of the subbasin, is capped by an extensive, low relief erosional remnant surface (Klamath peneplain).

Trinity Tributaries: Figures 4.24, 4.25, 4.26, and 4.27

Coffee Creek is a northern tributary that descends from the east (Figure 4.24). The upstream reaches of the river overflow glacial deposits, evidence that at least part of the valley was carved by glacier deposits. The stream overflows fault bounded CM rocks before overflowing an intrusive rock at a major knickpoint at 1300 m elevation. The lower segment of the stream, despite overflowing alluvium along a significant of its stream length is exceedingly steep and concave (k sn = 144, θ= 0.91).

Southwest of Coffee Creek, the southward flowing Canyon Creek is a steep and

concave stream whose headwaters are in the Canyon Creek pluton at elevations greater

than 2000 m. The creek traverses glacial deposits mapped at several elevations as low as

800 m. Downstream of the final knickpoint, the river overflows Salmon Schist. Where

the stream overflows the CM terrane, the river profile is quite smooth. Throughout the 104 Texas Tech University, Timothy Anderson, December 2008

river profile, the stream is steep (k sn = 104) and both lithologies are very concave. Many

knickpoints are mapped in the dioritic pluton. Where the Stuart Fork stream overflows

the same pluton along the eastern side, similar shaped knickpoints at high elevations are

mapped.

The southwest flowing New River descends more than 1500 m through mostly

WPT metasedimentary units. The channel is moderately steep and is not concave (k sn =

82.5, θ= .42). Because this smooth profile contains a continuous exposure of WPT metasedimentary rocks, the longitudinal profile is ideal for quantitative measurements for this particular lithology. Downstream of the Wilson Point Thrust, the river flows over the regional Ironside pluton.

The East Fork Trinity River, located on the southeastern side of the Trinity subbasin, descends from more than 2000 m elevation to Trinity Lake (720 m). The river overflows metasedimentary rocks and peridotite of the EK terrane throughout its course.

Glacial deposits at 1500 m are spatially coincident with a knickpoint that segments the profile into two segments.

All tributaries of the upper Trinity subbasin exhibit knickpoints at an elevation of 1500 m. The consistency in knickpoint elevations are discussed in the following chapter.

Trinity River: Figure 4.28

The Trinity River flows towards its confluence with the Klamath River from elevations greater than 2000 m along a 350 km stretch, traversing EK, CM, WPT, and

Western terranes. For the entire length of the Trinity River, the stream is not very 105 Texas Tech University, Timothy Anderson, December 2008

concave and moderately steep ( θ= 0.44, k sn = 79). Three segments of the longitudinal profile of the Trinity River are described individually. The upper 100 km of the Trinity

River, before the Trinity Lake, exhibits a very smooth, concave and very steep longitudinal profile ( θ=.74, k sn =132). Downstream from the Trinity and Lewiston dams, the longitudinal profile is convex where traversing CM and WPT rocks. Fifty kilometers downstream of the Lewiston , the flow direction of the Trinity River shifts northward. Where the river traverses CM rocks, the stream gradient is shallow (k sn = 34).

The channel gradient increases k sn = 63 where the river jogs westward into WPT rocks.

Because of the presence of large dams, it is difficult to understand the meaning of this

convexity. It is possible that, in the absence of dams, the middle and upstream segments

of the Trinity River are at equilibrium, and no convexity is present.

A prominent knickpoint within the Trinity River subbasin is preserved in a narrow north-trending valley that parallels the Orleans thrust fault. At the cusp of the knickpoint, the river parallels the Orleans fault. From the Orleans thrust downslope, the stream is highly concave, but gentle ( θ =1.3, k sn =57).

Salmon Subbasin: Figure 4.29

The northwest draining Salmon subbasin is located in the heart of the Central

KMP to the north of the Trinity subbasin. Topography of the subbasin can exceed 2700

m in the Trinity Alps to the southeast. The majority of the drainage area is north of the

Salmon River, where the distance between the southern drainage divide and Salmon

River is 3 times less (10 km) than the northern drainage divide (30 km).

106 Texas Tech University, Timothy Anderson, December 2008

Hillslopes mapped in the Wooley Creek watershed and along the southwestern

ridge of the subbasin are extremely high (hillslopes > 30º). Within the subbasin, a flat

erosional remnant surface (Klamath peneplain) is mapped above the Wooley Creek and

English Peak batholiths. Anomalously flat regions are also located south of the North

Fork Salmon Creek in WPT metavolcanic rocks and along the China Creek pluton that is

traversed by the Little South Fork Salmon River.

The mapped bedrock geology of the Salmon subbasin characterizes many

geologic features of the Klamath Mountains including several large intrusive bodies (e.g.

Wooley Creek batholith, English Peak batholith, Deadman Peak pluton), and northward

striking WPT, Stuart Fork, and CM terranes of metamorphosed volcanic and sedimentary

rocks.

All lengthy tributaries are steep (k sn > 100) and usually moderately concave ( θ<

0.6). Prominent knickpoints at an elevation of 1400 m within the Wooley Creek watershed and North Fork Salmon watershed are exhibited in almost every longitudinal profile.

Receiving more than 100 inches of annual precipitation, western regions of the

Salmon subbasin are twice as wet as central and eastern sections. Eight of the nine mapped Quaternary glacial units are mapped on the northern slopes of the Trinity Alps.

Salmon River: Figure 4.30

At the heart of the Central KMP, on the northern slopes of the Trinity Alps, the northwest flowing Salmon River overruns a multitude of lithologies in its descent to the

Klamath River. The is the upstream tributary of the Salmon River. 107 Texas Tech University, Timothy Anderson, December 2008

Traversing the Canyon Creek pluton, and metavolcanic, metasedimentary, and peridotite

rocks of the Stuart Fork, CM, and WPT, the concave profile of the entire Salmon and

Little Salmon River is steep and moderately concave (k sn = 111, θ= 0.6). Two subtle

convexities in the river profile are located where Quaternary gravels accumulate at the

Siskiyou fault and at a fault-bounded transition between the South Fork metasedimentary

rocks and CM Salmon metavolcanic rocks. Local steepness and concavity measurements

were taken between these convexities. The former segment, which is restricted to flow

over Stuart Fork metasedimentary rocks, is extremely steep and highly concave (k sn =

155, θ = 2.3). The latter profile segment overflows a diverse range of CM rocks, is steep

(k sn = 111), and has a more normal concavity (θ= 0.6).

Wooley Creek (Salmon Tributary): Figure 4.31

North of the Salmon River, Wooley Creek descends from topography at elevations greater than 1500 m to the Salmon River over a distance less than 40 kilometers. Similar to the Salmon River, Wooley Creek is very steep and less concave

(k sn = 116, θ= 0.58). The most elevated 500 meters of the stream, strictly overflowing

rocks the Wooley Creek batholith, are exceptionally steep and concave (k sn = 138, θ =

1.8).

Scott Subbasin: Figure 4.32

The Scott subbasin is an eastern, elevated, and alluvium-rich subbasin that flows northward into the Klamath River. The subbasin is most elevated and steepest along its drainage divides to the south and west. Its western drainage divide is a north striking ridge that consistently exceeds 2000 m in elevation. The inner nucleus of the subbasin is 108 Texas Tech University, Timothy Anderson, December 2008

at an elevation of 850 m and is characterized by Quaternary alluvium deposits and

extremely flat landscape (hillslope< 0.50º). Though subtle, hillslopes of the Eastern

Klamath terrane are lower than other Klamath terranes. In the northern regions of the

Scott subbasin, the river direction abruptly shifts around a west-oriented ridge, before

flowing into the Klamath River at an elevation of 470 m.

The Scott subbasin is relatively arid and usually receives 24 inches of annual

precipitation. An exception to this is the northwestern divide where annual precipitation

rates can be greater than 100 inches. In this corner and along the southern subbasin

drainage divide, glacial deposits have been mapped in 24 different locations.

Streams that flow from less elevated and less steep topography of the eastern

Scott subbasin (EKP) exhibit lower channel gradients (k sn < 60) than streams that flow

from the steeper, more elevated western ridges (k sn > 100). Where streams flow into the flat subbasin, streams have very low channel gradients (k sn = 35). All streams exhibit knickpoints of varying shapes and elevations.

Scott River: Figure 4.33

Northeast of the Salmon River, the northward flowing Scott River descends from a topographic ridge more than 2000 m in elevation to the Klamath River at an elevation of 470 m. The longitudinal profile illustrates the bedrock geology of the river until an elevation of 800 m. Duck Lake Creek, a southern tributary has headwaters that overflow the Pluton. The river possesses a major convexity at 1700 m elevation within the granitic pluton. Downstream of this convexity, the Scott River is concave

109 Texas Tech University, Timothy Anderson, December 2008

(θ=.89) and moderately steep (k sn = 78.1). The stream gradient abruptly flattens where the

river overflows Quaternary alluvium (k sn < 40).

4.10 Rivers in the Siskiyou Mountains

In this thesis, the Siskiyou Mountains are defined as the range north of the

Klamath River. For simplification purposes, this section includes tributaries that flow into the Klamath River. Thus, rivers described in this section actually overflow volcanic rocks of the Cascades and rivers south of the Klamath River. Thus, the reader is advised to refer to the provided maps of the subbasin being described.

The Upper Klamath Subbasin: Figure 4.34, 4.35, and 4.36

The Upper Klamath subbasin is a narrow westward draining subbasin that can locally exceed elevations of 2500 m (Cascadia volcanoes). In this subbasin, the Klamath

River flows into the KMP from an eastern volcanic landscape (Cascadia Volcanic Arc).

The upper subbasin can be partitioned into a western and eastern section by its geomorphologic, topographic, and lithologic characteristics. The width of the basin narrows westward from 50 km to 25 km, transitioning from a flat (hillslope< 10º), volcanic, relatively arid (annual precipitation is less than 20 inches) landscape to an older, metamorphic, wetter (annual precipitation is greater than 50 inches), and steeper

(hillslope> 25º) landscape.

The most eastern 60 km of the subbasin is composed of basalt from the High

Cascades and andesite of the Western Cascades. Longitudinal profiles of many streams 110 Texas Tech University, Timothy Anderson, December 2008

flowing from volcanoes are very concave (.73< θ< 3.0) and have moderate to low

channel gradients (40

River, or at elevations of 1400 m.

While Condrey Mountain hillslopes are not very steep (hillslopes < 15º) its streams are not. Streams that overflow greenschists and blueschists of Condrey Mountain are very steep and usually very concave ((k sn > 120, θ= 1.30). West of Condrey Mountain, streams have deeply incised the topography, are steep and less concave. Where the

Klamath River exits the Upper Klamath subbasin, it is at an elevation (420 m) over 1000 m lower than surrounding ridges. In these regions of the subbasin, Klamath tributaries are moderately concave and extremely steep ( θ= 0.50, k sn >130).

The Lower Klamath Subbasin: Figures 4.37, 4.38, and 4.39

The Lower Klamath subbasin comprises a large coastal drainage network that encapsulates the Klamath River and contributing tributaries. Basin divides of the subbasin are bounded by the Siskiyou Mountains (north), the Wooley Creek batholith

(southeast), the Ironside Mountain (south), and a northwest trending ridge near the coast

(southwest). When entering the subbasin from the east, the Klamath River is at an elevation of 420 m. The river flows south for the majority of the remainder of its course

(220 km) parallel to a thrust fault that exposes metasedimentary rocks of the Galice

Formation. Throughout its course, the Klamath subbasin drainage area substantially increases at its confluence with the Salmon and Trinity Rivers. At its junction with the

Trinity River, the stream shifts its flow direction from southwest to northwest.

111 Texas Tech University, Timothy Anderson, December 2008

Streams in the lower subbasin traverse a diverse range of lithologies including the

Wooley Creek batholith, WPT metavolcanics, Galice metasedimentary rocks, WPT and

Galice peridotite, and Franciscan rocks.

The Lower Klamath subbasin receives a large amount of precipitation. Although

ridge elevations rarely exceed 1400 m west of the Salmon River, ridges in the Siskiyou

Mountains and at the Wooley Creek batholith can exceed 2000 m in elevation. Glacial

deposits have been mapped in eleven locations.

With exception to coastal regions, streams throughout the entire Lower Klamath

subbasin are steep (ksn > 100). Streams in watersheds that flow from the Wooley Creek

batholith is very steep (k sn > 120). Coastal streams and watersheds west of the Klamath

River are typically more gentle and more concave (ksn < 80, θ= 0.70).

River profiles along eastern ridges (e.g. Indian Creek, Clear Creek, Dillon Creek,

Elk Creek, Independence Creeks) exhibit knickpoints at elevations around 1450 m and

1000 m. River profiles within southern watersheds (Red Cap watershed) also exhibit

knickpoints at 1000 m. Knickpoints are common at elevation of 550 m also. From map

view, the spatial organization of knickpoints suggests that there exists a series of

northeast striking lineaments along the eastern slopes of the Siskiyou Mountains and

Wooley Creek batholith. These lineaments appear to align with regional hillslope

gradients.

Klamath River: Figure 4.40

The deeply incising Klamath River descends from 1500 m to sea level over a

distance of 400 km. During its course, the river traverses five different terranes and seven 112 Texas Tech University, Timothy Anderson, December 2008 dams. The Klamath River can be divided into its upstream and downstream components by its lithologic, geomorphologic, and river profile characteristics. For the first 80 km upstream, the river flows southwest over Cascadia volcanic rocks, where river segments are very concave. When entering the KMP, the flow direction of the river shifts west.

Profile segments are less concave, and the local stream gradient becomes varies. West of

Condrey Mountain, the river’s flow direction shifts to the southwest and parallels exposed rock of the Galice Formation. Near its junction with the Trinity River, the

Klamath River shifts its flow direction to the northwest and overflows Franciscan rocks.

As a whole, the entire Klamath River is not concave and moderately steep (0.42, ksn = 63). Because of the river’s rigid and irregularly convex shape, lower segments of the Klamath River are difficult to quantitatively analyze. Nevertheless, normalized steepness and best fit concavity values were calculated and compared using a wide range of smoothing techniques. Whether increasing the smoothing window to 600 m, the contour sampling interval to 30 m, or fixing the reference concavity to 0.25, resulting values were always approximately the same (k sn = 70, θ= 0.25). The Condrey schist and

WPT units are less steep (k sn = 50) than the Galice Formation (k sn = 75). Upstream of the

Irongate Reservoir, the river is steep (70< k sn <85) and highly concave.

Applegate Subbasin: Figure 4.41

The southern headwaters in the Applegate subbasin flow from the sinuous northern slopes of the Siskiyou Mountains into the Coast Ranges. Siskiyou topography can exceed 2250 m in elevation. Within the subbasin, all tributaries flow into the

Applegate River, which in turn flows into the Rogue River east of the Coast Ranges at an 113 Texas Tech University, Timothy Anderson, December 2008

elevation of 260 m. The westward flowing overflows Quaternary

alluvium throughout its length, where basin hillslopes are lowest (hillslopes< 1 º).

Upstream watersheds have moderately steep hillslopes (hillslopes> 20º) and relief

(usually 500 m). Glacial deposits are mapped within the Middle Fork Applegate

watershed. Although precipitation is higher in elevated topography, a majority of the

subbasin receives less than 30 inches of annual precipitation.

Streams that flow from Condrey Mountain and an eastern ridge (WPT peridotite)

are relatively steep in the subbasin and similar to rivers south of Condrey Mountain (k sn >

100). Downstream of Applegate Lake, streams overflow Quaternary alluvium. In central and eastern regions of the subbasin, tributaries have low channel gradients and also overflow alluvium.

Illinois Subbasin: Figure 4.42

The Illinois subbasin is an odd shaped basin that flows from the northern slopes of

the Siskiyou Mountains and elevated topography of the Coast Ranges. The subbasin can

be at elevations greater than 2100 m, but it is very rare for topography in the north Illinois

subbasin to exceed 1400 m. All streams flow into the Illinois River (or its downstream

equivalent) which drains into the Klamath River in the central Coastal Ranges at an

elevation of 30 m. A regional northeast striking topographic ridge divides the subbasin

into a northern and southern component. In this study, the elevated topography west of

this lineament is defined as Coastal Range topography.

The southern regions of the Illinois subbasin are topographically, climatically, and

lithologically different. Quaternary alluvium and gravels are extensively deposited on a 114 Texas Tech University, Timothy Anderson, December 2008

flat (hillslope < 1º) valley floor less than 500 m in elevation. Eastern tributaries, with low

or intermediate channel gradients (45< k sn < 75) flow across this alluvium along. Annual precipitation is usually less than 70 inches in the southern subbasin.

Topographic ridges of the northern Illinois subbasin are deeply incised by steep streams (k sn > 90). At least nine of such ridges are capped by an erosional remnant surface. Tributaries that flow into the Illinois River exhibit more than 1000 m of immediate relief. Hillslopes in the Coast Range are high in the east and very high in the northwestern Illinois subbasin. Rocks within the Coast Ranges include intrusive rocks, peridotite, or Franciscan rocks. Throughout the northern Illinois subbasin, annual precipitation is greater than 100 inches. Knickpoints in eastern and western sections of the subbasin are recorded at elevations of 800 m and 500 m.

Illinois River: Figure 4.43

The Illinois River flows to the northwest across a northeast striking ridge into the

Rogue River in the central Coast Ranges, 35 km west of the coast. The Illinois subbasin

is a deeply incised river that flows at elevations more than 1000 m lower than juxtaposing

ridges. The entire Illinois River can be described as a less concave and moderately

dipping stream ( θ = 0.43, k sn = 70). At elevations slightly greater than 500 m, the stream

flows through quaternary alluvium for more than 40 km, where the normalized steepness

decreases and high concavity values ( θ =.93, k sn = 42.8). Where the stream overflows

prenevadan rocks (peridotite) the longitudinal profile is convex and steeper (k sn = 60).

The steepest channel segment of the stream is less concave and located where the channel

overflows WPT serpentinite ( θ = 0.49, k sn = 86). 115 Texas Tech University, Timothy Anderson, December 2008

116 Texas Tech University, Timothy Anderson, December 2008

4.11 Rivers in the Northern California and Southern Oregon Coast Range

The Coast Range subbasins include those that flow from the southern Siskiyou

Mountains, elevated topography that overflow Coast Range Rocks, and selected subbasins in central Oregon. Figure 1.2 demarcates the geographic regions that include the Coast Range.

Smith Subbasin: Figure 4.44

The Smith subbasin drains Coastal Range topography and the western slopes of the Siskiyou Mountains. Although elevations in the southern subbasin can reach 1900 m

(Siskiyou Mountains), elevations within the northern watersheds rarely exceed 1200 m.

Covered entirely by Quaternary alluvium, the downstream trunk of the Smith River is less than 100 m in elevation. The upstream and deeply incised South, Middle, and North

Fork Smith Rivers debouche their sediment into the Smith River from a drainage area greater than 2000 km 2. To the east, the Crescent City platform is a low elevated, flat

platform that extends 6 to 10 km inland. In this study, the Smith subbasin DEM was

resampled to 15 m resolution. Annual precipitation within this subbasin is particularly

high, almost always greater than 100 inches and can locally exceed 150 inches along

drainage divides. With exception to the 75 mapped erosional remnant surfaces that cap

various lithologies, hillslopes throughout the subbasin are high (hillslopes> 25º).

Although all streams traverse a diverse range of lithologies, all major rivers

overflow different sections of the Josephine Ophiolite. Rivers in the central and southern

regions of the subbasin flow from Siskiyou ridges and traverse a diverse suite of WPT

117 Texas Tech University, Timothy Anderson, December 2008

rocks. In contrast, streams in the northern subbasin almost entirely overflow the Galice

Formation (the North Fork Smith River has headwaters within the Franciscan Complex).

With local exceptions, stream segments throughout the subbasin have moderate or

low channel gradients (k sn < 55). From map view, knickpoints in the southern regions

align in a northeast striking pattern. . Knickpoints within the longitudinal profiles are

preserved at elevations of 500 m.

Middle Fork Smith River: Figure 4.45

The westward bound Middle Fork Smith River descends from 1500 m to the sea.

Twenty kilometers from the coast where the river overflows Quaternary alluvium, the

Smith River adjusts its flow direction to the northwest. The longitudinal profile of the

entire river is concave, exhibits two or three knickpoints, and varies in steepness. The

river can be divided into three or four segments based on lithology and/or profile shape.

The upper 10 km of the river gently overflows WPT peridotite and/or Quaternary glacial

deposits, and is not very steep or concave (k sn = 49.4, θ= 0.37). At the Middle Fork Fault,

the stream gradient and channel concavity abruptly increases (k sn = 149, θ = 2.0) for a

length of 10 kilometers into the Galice Formation. A subtle knickpoint in the longitudinal

profile is located at 650 m elevation. Downstream of this knickpoint, the longitudinal

profile is less steep. The lithologic transition from metasedimentary rocks of the Galice

Formation to peridotite of the Josephine Ophiolite Sequence along the river profile occurs

across a 10 km fault zone. Between the upstream knickpoint and peridotite of the

Josephine Ophiolite, the stream is fairly concave and moderately steep ( θ = 0.94, k sn =

60.8). Taking these segments together, the middle river segment of the Middle Fork 118 Texas Tech University, Timothy Anderson, December 2008

Smith River is steep and very concave (k sn = 86, θ =1.2). The lower segment of the

Middle Fork Smith River overflows peridotite and quaternary alluvium, is less steep and very concave (k sn = 48.3, θ =1.3).

South Fork Smith River: Figure 4.46

In map view, the westward bound South Fork Smith River traverses through the southern Siskiyou Mountains. Its river profile exhibits a moderately steep and less concave profile (k sn =82, θ= 0.48). The stream traverses five lithologies and exhibits at

least 2 anomalous knickpoints. The headwaters of the river descend from 1750 m

elevation in the Bear Mountain gabbro into WPT metasedimentary rocks. A knickpoint at

534 m elevation, just upstream of the Orleans Thrust Fault, marks the contact between

WPT metavolcanic rocks and siltstone. Downstream of this knickpoint, the stream is

exceptionally steep (k sn = 133). The first appearance of the Galice Formation is not until

an elevation of 400 m at the Orleans Thrust Fault. Where the river overflows the heavily

faulted Galice Formation, the longitudinal profile is convex. The Smith River crosses

Josephine peridotite and Franciscan rocks within 60 km of its base level. Where the river

overflows the Josephine Ophiolite, the stream is considerably steep (k sn = 32) and highly concave.

Lower Rogue Subbasin: Figure 4.47

The Lower Rogue basin is a coastal north of the Smith and

Applegate subbasins. The Rogue River enters the subbasin from the east along into an extensively flat landscape. Similar to the more southern Illinois subbasin, the westward bound Rogue River enters the Coast Ranges by transecting the northern terminus of a 119 Texas Tech University, Timothy Anderson, December 2008

regional northeast striking ridge. This ridge topographically divides the dry (annual

precipitation< 40 inches), low-lying (hillslopes < 5 º), and alluvium-full eastern Rogue

subbasin from the deeply incised, steep (hillslopes > 25 º), and wet (annual precipitation>

100 inches) Coastal Range.

Tributaries in the eastern subbasin flow from an arcuate shaped, eastern topographic ridge that extends into the Cascade Range. Though not well mapped, these ridges are likely the most northern extent of slithers of WPT peridotite. Streams in this region can have low stream gradients (k sn < 50) and are presumably transport-limited near their base.

Tributaries within the Coastal Ranges flow into the Rogue River from drainage divides less than 20 km apart at elevations lower than 1200 m elevation. The Rogue

River is at an elevation less than 230 m when entering the Coast Range. Despite its presence in a valley of more than a kilometer of relief, the Rogue River overflows alluvium only twice. Tributaries in the heart of the Coast Ranges are have deeply incised the topography (relief >500 m) and are steep (k sn > 85). Western tributaries have lower

stream gradients (k sn < 65). Most streams throughout the entire subbasin exhibit knickpoints of various size, shape, and elevation. Nevertheless, knickpoins appear to cluster at an elevation of 500 m.

Rogue River: Figure 4.48

The westward flowing Rogue River descends from 1800m in elevation to the

Pacific Ocean. The Rogue River is the longest river in the northern regions of the subbasin, traversing volcanic rocks and alluvium throughout a majority of its course. 120 Texas Tech University, Timothy Anderson, December 2008

The river initially flows southwest for its first 100 km along a topographic ridge

composed of older Tertiary-aged basalt flows. Upstream of the Coastal Range at an

elevation of 400 m, the river overflows Holocene gravels along a very flat and extensive

landscape (hillslope < 5º) that is characterized by numerous swamps. The channel

gradient increases crossing into the Coast Range (from ksn =42 to k sn =67).

Because of the odd profile shape, it is extremely difficult to calculate meaningful

values. A poorly fit regression line for the entire Rogue River describes the river as a

very gentle and less concave river ( θ = 0.28, k sn = 33). The upstream segments of the river

exhibit large convexities that coincide with dams and different volcanic terranes.

Chetco Subbasin: Figure 4.49

The coastal Chetco subbasin is north of the Smith subbasin. Watersheds within

the subbasin are divided by Coastal Range ridges that can exceed 1500 m in elevation.

Many tributaries flow into the deeply incised , where its valley floor is

usually 700 m lower than nearby ridges. Hillslopes are very steep in the southern and

eastern watersheds (hillslopes> 30°). Like the Smith subbasin, annual precipitation is

always greater than 100 inches, and occasionally exceeds 150 inches. Glacial deposits

have been mapped upstream of the Box Canyon Creek (Chetco River tributary).

Stream gradients are very gentle (k sn < 40) in southern and central watersheds of the subbasin. Northern coastal streams (e.g. the ) are relatively steep (k sn >

75). Although all streams overflow Coastal Range rocks for a majority of their path, upstream tributaries within the Chetco watershed can overflow a diverse range of

121 Texas Tech University, Timothy Anderson, December 2008

lithologies. As is characteristic of with most coastal streams, longitudinal profiles within

this subbasin are not smooth.

South Umpqua Subbasin: Figure 4.50

The inland South Umpqua subbasin is the most northern subbasin in the study area. Having an odd shape, the subbasin is widens to the south to include watersheds that drain the Cow and South Umpqua Rivers. Maximum elevations of eastern watersheds

(e.g. South Umpqua watershed) exceed 2000 m. In contrast, the less elevated topography in western watersheds does not exceed 1200 m in elevation. Hillslopes are steep in the southwestern regions of the subbasin where Coastal Range rocks are mapped. In the northern South Umpqua subbasin, hillslopes angles are shallower. More than two thirds of the 1500 South Umpqua lakes are mapped in the northern section of the subbasin.

Eastern streams that overflow volcanic rocks and Tertiary sedimentary rocks have moderately steep stream gradients (k sn > 80). Western streams that overflow coastal range rocks (Franciscan rocks) and Tertiary sediments have exceedingly low channel gradients

(k sn < 30). The Cow Creek and South Umpqua Rivers overflow quaternary alluvium upstream and downstream of their confluence.

South Umpqua River: Figure 4.51

The South Umpqua River flows westward from volcanic rocks into alluvium from elevations of 1750 m. The overall river profile exhibits a less concave shape ( θ = 0.54).

The river can be divided into an upstream volcanic and downstream alluvium section.

Most of the river steepness is within the upstream volcanic section (k sn = 79), despite

122 Texas Tech University, Timothy Anderson, December 2008

having a fairly low concavity ( θ =.44). In contrast, the alluvium section has a more gentle

gradient (k sn = 25.7).

North Fork Coquille Subbasin: Figure 4.52

The Coquille subbasin is a northern drainage network in coastal Oregon. The 30

most inland kilometers of the subbasin is a vast, flat (hillslope< 10º) and lowly elevated

coastal landscape covered with Holocene to Tertiary-aged sandstones and gravels.

Streams descend from Eocene-aged sandstones and siltstones of the Coastal Range

topography at elevations as high as 1200 m. Hillslopes within the elevated topography of

the Coast Range are usually less than 20º.

There are no steep streams within the Coquille subbasin. The steepest stream

segments are located in the eastern regions of the subbasin (k sn = 55). Northern and

coastal streams have very low channel gradients (k sn < 30). Many knickpoints are mapped

inland at elevations around 250 m.

North Fork Coquille River: Figure 4.53

The North Fork Coquille River flows southwestward from elevations around 750

m to the Coquille. The longitudinal profile can be divided into two segments upstream

and downstream of a knickpoint at an elevation of 300 m. Upstream of the knickpoint,

the river that overflows marine sandstone is nearly flat (k sn = 17) and the river has a moderate concave shape ( θ =.66). Downstream of the knickpoint, the concavity increases

(θ =2.1) and despite a double in stream gradient (k sn = 38), the river still has a fairly flat gradient. The inflection point of the knickpoint is likely defined by the transition from

123 Texas Tech University, Timothy Anderson, December 2008 marine sandstone to siltstone. Landslides have also been mapped south downstream of the knickpoint.

124 Texas Tech University, Timothy Anderson, December 2008

Figure 4.1 These three maps illustrate the location of the subbasins and major rivers within the KMP and adjacent areas. Subbasin data are collected from the National Hydrography Dataset. Rivers are constructed from DEM manipulation (see text for details).

125 Texas Tech University, Timothy Anderson, December 2008

Figure 4.2 This figure illustrates the elevation, drainage area and stream gradient characteristics of a stream out of equilibrium (Hunter Creek) and a stream in equilibrium (4 th of July Creek). In both cases, measurements of concavity and stream gradient are taken by fitting regression lines (colored lines) to the actual data (black plus-signs). If a longitudinal profile exhibits a non-concave shape (Hunter Creek), the stream is either overflowing different substrates or is not at equilibrium. In either case, it is more geologically meaningful to divide the stream into smaller channel reaches for profile investigation. This implies that despite the entire stream not being in steady-state equilibrium with the surrounding landscape, its individual stream segments are (Wobus et al., 2006). How the profile is segmented (in this example, the stream is divided into 3 concave river profile segments and one convex river profile segment) is dependent on the scale of which the investigation is carried out. In this thesis, channels are selected by focusing on the linear relationship between drainage area and stream gradient in log-log space. In logarithmic space, the stream gradient of a profile at steady-state equilibrium stream gradient will decrease at a negative linear slope when plotted to drainage area in logarithmic space (see 4 th of July Creek gradient-drainage area plot). Channel segments are illustrated by color. Within the KMP, stream segments with drainage areas smaller than 100,000 m 2 are assumed to represent a colluvial based sediment-transportation environment. Abrupt increases in drainage area indicate a confluence with another stream. The 4 th of July Creek is considered to be at equilibrium, exhibiting a smooth concave profile.

126 Texas Tech University, Timothy Anderson, December 2008

Figure 4.3 This figure illustrates the effect of different basin shapes on the Hack Gradient. Wider basins will contribute increase a stream’s discharge at shorter distance from the drainage divide. The opposite expected for very narrow drainage basins. Pictures represent examples of differently shaped basins in the KMP.

127 Texas Tech University, Timothy Anderson, December 2008

Figure 4.4 This window illustrates the various parameters that can be entered for profile analysis. Unless otherwise mentioned, the default settings (above) were used.43

128 Texas Tech University, Timothy Anderson, December 2008

Figure 4.5 This schematic illustrates the affect of varying concavity and channel steepness on the shape of a river profile in normal and logarithmic space. Note that the measurement of concavity is the slope of the channel gradient.

129 Texas Tech University, Timothy Anderson, December 2008

Figure 4.6 This figure conceptually illustrates the different forms of a longitudinal profile. During steady-state equilibrium, a river profile is expected to exhibit a smooth concave profile (blue). When a river is not at equilibrium, then it will exhibit a convexity (green) or knickpoint (red).

130 Texas Tech University, Timothy Anderson, December 2008

Figure 4.7 This figure schematically illustrates the upstream migration of a knickpoint. The presence of a knickpoint can indicate that a stream flows over lithologies of variable erosional resistance.A knickpoint could also imply that the river has experienced base-level fall. Eustatic changes in sea level and bedrock uplift are examples of base-level fall. In this illustration, a stream has a downstream knickpoint (T= 0). Because stream power is greater at the lip of the knickpoint, the lip is more susceptible to erosion. Erosion of the downstream lip will result in an upstream migration of the knickpoint. When a knickpoint reaches a confluence between two streams (T= 2), the knickpoint will diverge and continue to migrate upstream via tributaries. If all streams erode at a similar rate, knickpoint migration rate would theoretically be equal and knickpoint elevations in different streams would be at similar.

131 Texas Tech University, Timothy Anderson, December 2008

Map 4.8 The normalized channel steepness of stream segments within the KMP can be utilized to compare the steepness of channels that traverse the different lithologic, climatic, and tectonic settings that comprise the KMP. This diagram illustrates how streams within the central KMP are relatively steeper than streams north and south of the KMP.

132 Texas Tech University, Timothy Anderson, December 2008

Map 4.9 This figure maps concavity measurements of individual stream segments.There are no apparent trends in the concavity of streams in the KMP. This is likely a result of the common occurrence of profile knickpoints and convexities.

133 Texas Tech University, Timothy Anderson, December 2008

Figure 4.10 This diagram illustrates the longitudinal profile and gradient-drainage area relationships of all major rivers within the Klamath Mountain.

134 Texas Tech University, Timothy Anderson, December 2008

Figure 4.11 River profiles of the Upper Eel subbasin. River profiles illustrate path of rivers to and via the Eel River. The subbasin has been divided into two segments, defined by color. Rivers from the eastern and western segments do not ever intersect in profile display. Arrows illustrate flow direction.

135 Texas Tech University, Timothy Anderson, December 2008

Figure 4.12 Tributaries of the Middle Eel subbasin. Tributaries of different regions do not intersect with each other.

136 Texas Tech University, Timothy Anderson, December 2008

Figure 4.13 Tributaries within the Lower Eel Subbasin. Major rivers are labeled by color. Green tributaries are shifted approximately 50 km to the left of the graph for illustration purposes.

137 Texas Tech University, Timothy Anderson, December 2008

Figure 4.14 The Middle Fork Eel River overflows Franciscan rocks throughout the majority of its path. Different lithologies and possibly active thrust faults have modified the shape of the longitudinal profile in two convex zones.

138 Texas Tech University, Timothy Anderson, December 2008

Figure 4.15 The Van Duzen River flows over Franciscan rocks throughout the majority of its path. Abrupt changes in elevation appear to correlate with less resistant Franciscan rocks.

139 Texas Tech University, Timothy Anderson, December 2008

Figure 4.16 The Mad-Redwood subbasin comprises the Mad River, Redwood Creek, and a series of coastal streams to the south. The Mad River exhibits a prominent knickpoint throughout the majority of its profile. The Redwood is a relatively steep river.

140 Texas Tech University, Timothy Anderson, December 2008

Figure 4.17 Redwood Creek is a concave gently flowing stream that overflows Franciscan Rocks.

141 Texas Tech University, Timothy Anderson, December 2008

Figure 4.18 The Mad River overflows Franciscan rocks throughout its entire course. A prominent knickpoint divides its profile into two segments. Changes in stream elevation appear to be coincident with less resistant Franciscan mods.

142 Texas Tech University, Timothy Anderson, December 2008

Figure 4.19 South Fork Trinity Subbasin is largely drained by the Hayfork Creek and South Fork Trinity River.

143 Texas Tech University, Timothy Anderson, December 2008

Figure 4.20 The longitudinal profile of the South Fork Trinity River traverses three terranes.

144 Texas Tech University, Timothy Anderson, December 2008

Figure 4.21 The Upper Trinity subbasin is defined by the subbasin upstream of Trinity Lake. A broad knickzone exists around 1500 m elevation.

145 Texas Tech University, Timothy Anderson, December 2008

Figure 4.22 The Central Trinity subbasin includes streams that feed into the Trinity River just downstream of the Trinity Lake.

146 Texas Tech University, Timothy Anderson, December 2008

Figure 4.23 The lower tributaries of the Trinity subbasin exhibit knickpoints at approximately 1000 m where traversing the Ironside pluton. Channel steepness is greater in more northern tributaries.

147 Texas Tech University, Timothy Anderson, December 2008

Figure 4.24 Coffee Creek is a short tributary that debouches its sediment into the Trinity Lake. It begins its descent within the Trinity Alps. This stream characterizes many streams in the upper Trinity subbasin with a broad knickpoint at approximately 1300 m elevation

148 Texas Tech University, Timothy Anderson, December 2008

Figure 4.25 The New River overflows WPT metasedimentary rocks throughout most of its course. The river is very smooth, less concave, and steep.

149 Texas Tech University, Timothy Anderson, December 2008

Figure 4.26 The East Fork Trinity River almost entirely overflows peridotite of the Eastern Klamath terrane. A prominent knickpoint is spatially coincident with mapped glacier deposit.

150 Texas Tech University, Timothy Anderson, December 2008

Figure 4.27 The Canyon Creek River is an extremely steep and concave river. Upstream, where the river overflows the Canyon Creek Pluton, the profile exhibits multiple knickpoints. Downstream, the river is smooth and concave.

151 Texas Tech University, Timothy Anderson, December 2008

Figure 4.28 The Trinity River overflows 5 different terranes in its descent from the Trinity Alps to the Klamath River.

152 Texas Tech University, Timothy Anderson, December 2008

Figure 4.29 The Salmon Subbasin is a subbasin with very steep streams. Wooley Creek and the North Fork Salmon Creek exhibit knickpoints in their profiles at 1400m elevation.

153 Texas Tech University, Timothy Anderson, December 2008

Figure 4.30 The Salmon River traverses many different lithologies but exhibits a fairly smooth profile nevertheless. Its concavity and normalized channel steepness values are representative of other drainages within the subbasin.

154 Texas Tech University, Timothy Anderson, December 2008

Figure 4.31 Wooley Creek is a steep stream that is moderately concave. In the Salmon subbasin, knickpoints are particularly common at 1400 m. In this particular example, the knickpoint is located at the contact between metasedimentary rocks and a gabbroic pluton.

155 Texas Tech University, Timothy Anderson, December 2008

Figure 4.32 The Scott subbasin is primarily composed of short and steep rivers that debouche their sediment within the flat subbasin.

156 Texas Tech University, Timothy Anderson, December 2008

Figure 4.33 The headwaters of the Scott River, where any significant stream relief is located, are in the Russian Peak Pluton.

157 Texas Tech University, Timothy Anderson, December 2008

Figure 4.34 Upstream tributaries of the upper Klamath subbasin overflow volcanic rocks of the Western and High Cascades. Large knickpoints in the uppermost tributaries (red) are always located just upstream of the incising Klamath River. Knickpoints are also located along volcanic topographic rises.

158 Texas Tech University, Timothy Anderson, December 2008

Figure 4.35 Tributaries in this portion of the subbasin flow southward from an eastern ridge of the Siskiyou Mountains. Eastern streams tend to be higher in gradient.

159 Texas Tech University, Timothy Anderson, December 2008

Figure 4.36 Streams in the lower section of the upper Klamath Subbasin are extremely steep, and typically less concave.

160 Texas Tech University, Timothy Anderson, December 2008

Figure 4.37 The Eastern section of the Lower Klamath subbasin comprises streams grouped into the northern Elk watershed and southern Red Cap watershed. Tributaries within the Elk watershed exhibit knickpoints at 1400 m and 1000 m. Red Cap Creek exhibits knickpoints at 1000 m also.

161 Texas Tech University, Timothy Anderson, December 2008

Figure 4.38 The tributaries feeding from the Siskiyou Mountains are steep streams that deeply incise their watersheds. Knickpoints are common at elevations of around 1000 m, and map along strike across the ridge.

162 Texas Tech University, Timothy Anderson, December 2008

Figure 4.39 The western section of the lower Klamath subbasin includes coastal tributaries (orange), southern and streams from the souther Siskiyou Mountains (black and brown). Knickpoints in Siskiyou streams tend to be at elevations above 500 m.

163 Texas Tech University, Timothy Anderson, December 2008

Figure 4.40 The longitudinal profile of the Klamath River is a very chaotic. Large knickpoints in the profile are observed upstream where the river overflows volcanic rocks. Downstream the Irongate Reservoir, the river overflows WPT, Condrey Mountain, and Galice Formation rocks.

164 Texas Tech University, Timothy Anderson, December 2008

Figure 4.41 The Applegate subbasin is divided into three grouped watersheds. Streams begin at topographic highs in the Siskiyou Mountains to the south. Most streams in this subbasin overflow Quaternary alluvium.

165 Texas Tech University, Timothy Anderson, December 2008

Figure 4.42 The odd-shaped Illinois subbasin comprises gently steeping streams in the south and steep streams in the northern . Knickpoints are generally mapped at 800 m elevation and 500 m elevation.

166 Texas Tech University, Timothy Anderson, December 2008

Figure 4.43 The Illinois River flows from WPT rocks and into Quaternary alluvium before entering the Coastal Range Mountains.

167 Texas Tech University, Timothy Anderson, December 2008

Figure 4.44 The Smith subbasin is a coastal subbasin with 6 major rivers. All streams have low channel gradients. Knickpoints appear in all profiles at an elevation of 500 m.

168 Texas Tech University, Timothy Anderson, December 2008

Figure 4.45 The Middle Fork Smith River is a westward flowing river that traverses several lithologies in the Coastal Regions. The middle segments of the river are relatively steep.

169 Texas Tech University, Timothy Anderson, December 2008

Figure 4.46 The South Fork Smith River is a westward flowing river with headwaters in Bear Mountain. The river is among the steeper rivers in the coastal subbasin.

170 Texas Tech University, Timothy Anderson, December 2008

Figure 4.47 The Lower Rogue Subbasin comprises tributaries within the Coastal Ranges that feed into the Rogue subbasin. Streams within the central regions of the subbasin (light blue and dark green) tend to be relatively steep in this subbasin.

171 Texas Tech University, Timothy Anderson, December 2008

Figure 4.48 The Rogue River overflows Cascadia volcanic rocks throughout the majority of its course. Very pronounced knickpoints are located upstream of the Lost Creek Lake. The Rogue River is characterized by its low channel gradient for the final 250 km of its course. The stream gradient is exceedingly low where it overflows Holocene gravels, but steepens when flowing in the Coast Range.

172 Texas Tech University, Timothy Anderson, December 2008

Figure 4.49 The Chetco Subbasin is a small subbasin in the Coastal Regions north of the Rogue River.

173 Texas Tech University, Timothy Anderson, December 2008

Figure 4.50 The South Umpqua subbasin is an inland drainage network north of the KMP. Streams have very low gradients in the western subbasin.

174 Texas Tech University, Timothy Anderson, December 2008

Figure 4.51 The South Umpqua River traverses volcancics upstream and alluvium downstream.

175 Texas Tech University, Timothy Anderson, December 2008

Figure 4.52 The Coquille Subbasin is the most northern subbasin within this study. A pronounced knickpoint at 500 m is located in many longitudinal profiles.

176 Texas Tech University, Timothy Anderson, December 2008

Figure 4.53 The North Fork Coquille River is the most northern river studied in this investigation. It overflows different sedimentary rocks throughout its course.

177 Texas Tech University, Timothy Anderson, December 2008

Chapter 5

Discussion

This chapter integrates datasets and results from the previous chapters to evaluate the tectonic and topographic evolution of the KMP. The discussion is organized into the following 9 sections:

1) Discussion of the origin, uplift, and erosion of the Klamath peneplain;

2) Volumetric analysis of magnitude and rate of erosion in the KMP;

3) Evaluation of Miocene and Pliocene paleotopography;

4) Prediction of Quaternary faults

5) Evaluation of river profile results

6) Evaluation of local geomorphologic structures that reflect rock uplift

7) The western Klamath Mountains represented as a critically tapered wedge;

8) Models that describe uplift and erosion of the KMP since the late Miocene.

9) Future studies to test models

5.1 Discussion of the Origin, Uplift and Erosion of the Klamath Peneplain

The remnant erosional surfaces that define the Klamath peneplain (Diller, 1902;

Aalto, 2006) are rare geologic feature that meets the criteria for classification as a

surface. There are very few examples of geologic features that can serve as a surface

because they must be conformable across a sizeable area and referenced to the geoid

(Molnar and England, 1990; Abbot et al., 1997; Phillips, 2002). Based on the assumption

that the remnant surfaces represent a conformable, low-relief surface that was near sea 178 Texas Tech University, Timothy Anderson, December 2008 level at the end of the Miocene, the interpolation surface (KPS of Chapter 3, Figure 3.12) is a surface, and is ideal for study of surface uplift of the western Klamath Mountains.

If uplift of the Klamath peneplain and erosion in California subbasins initiated in the early Pliocene and continued to the present, surface movement within the western

Klamath Mountains occurs as a long wavelength tectonic cycle (>1,000,000 yr). A landscape that experiences steady-state equilibrium during this time frame maintains its mean elevation, despite short(er) cycles of uplift and erosion (Kooi and Beaumont, 1996;

Burbank, 2002; Bracken and Wainwright, 2006). Because the mean elevation of the western and central Klamath Mountains has increased, this particular region is not in steady-state equilibrium.

The most elevated remnant surface (2018 m) caps the English Peak Pluton in the central Klamath Mountains, and is 70 kilometers from the least elevated remnant surface

(3 m) at the mouth of the Klamath River. Altitudinal trends of remnant surfaces indicate that the western Klamath Mountains have been tilted or flexured westward ~2º. If erosional remnant surfaces were at sea level at 5 Ma, the maximum surface uplift rate of the western Klamath Mountains is 0.4 mm/yr (Figures 5.1). If uplift initiated at 3 Ma, as suggested by some studies (McCrory; 1989; Stone, 1992; Gullick, 2002), its corresponding maximum uplift is 0.67 mm/yr. If the uplift began at the onset of deformation of the upper Falor unit (0.7 Ma), the maximum uplift rate is 2.9 mm/yr. The bottom table in Figure 5.1 (taken from Burbank, 2002) compares KMP exhumation

(uplift) rates of this thesis with other rates around the world.

179 Texas Tech University, Timothy Anderson, December 2008

Although the erosional remnant surfaces collectively tilt westward at an angle of

2°, elevation differences between some erosional remnant surfaces demonstrate a tilt that is two to five times steeper. Therefore, it may be more appropriate to describe vertical motions of isolated regions within the Klamath Mountains that could be acting as a localized decoupled block, or experiencing complicated short wavelength uplift/subsidence patterns.

Figure 5.2 illustrates the elevation of individual remnant surfaces by their

corresponding longitudinal coordinate. Remnant surfaces share a similar altitude at

similar longitudinal coordinates, and increase in elevation westward. A dichotomy

between two, seemingly decoupled, uplift regimes can be distinguished: 1) uplift of the

Coast Range and southern Siskiyou Mountains (western remnant surfaces; west of 123.5º

W), and 2) uplift of the central Klamath Mountains (eastern remnant surfaces; east of

123.5º W). When grouped by longitudinal coordinates, the collection of western remnant

surfaces define a west tilted regional surface that dips at angles between 4º and 6º

(Figures 3.12 and 5.2). A 4º and 6º tilt is consistently measured east of Mad River

subbasin. This pattern indicates that western remnant surfaces were uplifted in a single,

rigid block and tilted westward at relatively steep angles greater than 4º.

180 Texas Tech University, Timothy Anderson, December 2008

Figure 5.1 The top table presents a comparison of rock volume and aerial exposure of uplifted and removed rock by different interpolation methods. The bottom table compares exhumation rates of the KMP with other regions in the world. Three exhumation ranges are calculated to represent the potential time ranges uplift could have occurred. The bottom table is modified from Burbank (2002).

181 Texas Tech University, Timothy Anderson, December 2008

Remnant surfaces east of the South Fork Trinity and Trinity Rivers (Figure 5.2, east of the red dashed line) lack uplift patterns. A lack of uplift pattern could result from different tectonic geomorphologic forces inland, or simply a lack of mapped erosional remnant surfaces inland, particularly within the Siskiyou Mountains (Figure 3.10).

Figure 5.2 This figure illustrates the elevations of individual remnant surfaces at their respective longitudinal coordinates. Maps illustrate the location of individual remnant surfaces and an interpolation surface, proposed faults, and regions that have locally uplifted (orange) or subsided (blue). The dashed red line segregates two populations of remnant surfaces based on elevation trends seen in the plot. Proposed faults are identified by anomalous offset of remnant surfaces in close proximity. 5.2 Volumetric Analysis of Rates and Duration of Uplift and Erosion in KMP

The amount of western Klamath Mountain crust (geographically defined by the

dimensions of the natural-neighbor interpolation surface, or KPS) is volumetrically 182 Texas Tech University, Timothy Anderson, December 2008 massive. Within this constrained geographic region (area of 13, 460 km 2), 11,373.8 km 3

of crust is elevated above sea level, and the mean elevation is 760 m. If the constructed

paleogeographic map (Figure 3.15) is correct, much of the western Klamath Mountains

were at sea level, and the volume of uplifted western Klamath crust is equal to the

volume between the KPS and modern sea level. The upper table in Figure 5.1 lists six

rows of aerial and volumetric measurements of the amount of western Klamath crust

removed and uplifted according to different interpolation types (see Chapter 3). The

spline and inverse distance weighted (IDW) interpolation methods extend across a larger

geographic range, so volumetric calculations reflect the larger area.

According to the natural neighbor interpolation method in Figure 5.1, 13, 820 km 3 of western Klamath crust has been exhumed above sea level. Note that this value is volumetrically greater than the volume of crustal material calculated in the modern western Klamath Mountains. This discrepancy can be accounted for in the volume of erosion (Figure 5.1). If it is assumed that the of the Klamath Mountains has a density of 2700 kg/m 3, a total crustal mass of 37, 285, 880 kg been exhumed.

Figures 5.1 and 5.3 are tables of quantitative measurements of erosion in the western KMP. Results from volumetric calculations predict that 3,860 km 3 of material

has been removed beneath the interpolated peneplain surface (Figure 5.1, natural

neighbor interpolation method). Of the total volume of rock uplifted (13, 460 km 3), 29% of the material has been removed. This calculation ignores the amount of eroded material in antecedent mountains, but assumes all peneplain surfaces were at sea level. The

183 Texas Tech University, Timothy Anderson, December 2008 greatest volume of erosion occurred in the Klamath, Salmon, Trinity, and South Fork

Trinity subbasins (Figure 3.19).

Figure 5.3 is a table that compares maximum erosion rates of the Klamath

Mountains with different orogens in the world. The Klamath Mountain erosion rates are a magnitude lower than river incision, landslide, and glacial erosion rates in major orogens.

Figure 5.3 This figure compares erosion rates of the western Klamath Mountains with erosion rates around the world. Modified from Burbank (2002).

5.3 Evaluation of Miocene and Pliocene Paleotopography

Assuming all mapped peneplain surfaces were near sea level during the late

Miocene, Figure 3.15 illustrates at least six paleotopographic mountains that stood above

low-lying areas in the late Miocene. Limited evidence from studies of late Miocene and

early Pliocene coastal sedimentary units indicate that these paleotopographic features

were separated by shallow marine waters and/or low gradient braided streams (Stone,

1992; Aalto et al;, 1995, Aalto, 1996). For illustration purposes, Figure 3.15 illustrates

184 Texas Tech University, Timothy Anderson, December 2008 regions with insignificant paleo-altitude in blue. The omnipresence of paleosols (Irwin,

1997; Aalto, 2006) in northern California and southern Oregon suggest topography was above sea level, but in a mature geomorphologic stage. Large antecedent features include the Siskiyou Mountains, a ridge in the heart of the central Klamath Mountains, and South

Fork Mountain.

The paleo-Siskiyou Mountains is the largest paleotopographic feature in Figure

3.15. Reaching paleo-elevations greater than 900 m in two locations, the paleo-Siskiyou

Mountains comprise two or three northeast striking ridges at the present-day Oregon border. This map infers that the Klamath, Illinois, and Rogue River systems never connected since the late Miocene. The lack of mapped remnant surfaces in the Siskiyou

Mountains is mysterious and could have important implications about the details of late

Miocene Siskiyou paleotopography and its uplift history. If erosional remnant surfaces

(from which paleotopographic maps in this thesis are constructed) are absent in the

Siskiyou Mountains and buried beneath alluvium in the Illinois subbasin, then the

Siskiyou Mountains were massive mountains that could have been more than 2,000 m in elevation during the late Miocene, but has slightly subsided since. On the other hand, if erosional remnant surfaces are present in the Siskiyou Mountains, then exhumation of modern Siskiyou Ridges is post-Miocene.

South Fork Mountain is the second largest paleotopographic feature located west of the southern Klamath Mountains (see Figure 1.2). This paleo-ridge of approximately

30 kilometers exceeds half a kilometer in paleoelevation, is capped by South Fork

Mountain schist, and is orientated parallel to adjacent modern ridges that are terrane- 185 Texas Tech University, Timothy Anderson, December 2008 defined. If the series of northwest striking ridges in the MTJ region are geomorphologically related, it is unclear why South Fork Mountain is the only ridge that was subaerially exposed in the late Miocene.

The southwestern drainage boundary for the Salmon subbasin is the largest central KMP Miocene paleotopographic feature predicted by the paleotopographic map

(Figure 3.13). Although the northwest or northeast striking ridges are dwarfed by the paleo-Siskiyou Mountains, the ridges routinely exceed a half kilometer in elevation. This paleotopographic mountain is surrounded by the most elevated erosional remnant surfaces mapped by Irwin (1997).

As discussed in Chapter 3, the predicted antecedent topography in northern

California and Southern Oregon are based on assumptions and the accuracy of the interpolation surface. Assuming the erosional remnant surface was indeed at sea level and quasi-planar, four critical questions affect the location, shape, and size of antecedent features including: 1) why are erosional remnant surfaces absent along certain ridges, 2) how much denudation of antecedent features occurred, 3) have unknown antecedent features been completely removed, 4) what sedimentological evidence supports the existence of the predicted paleotopographic features (and the absence of other)?

5.4 Inferred Faults in the KMP

Locations ‘A’ through ‘E’ of Figure 5.4 are exposures of erosional remnant surfaces with anomalous elevations. If a region comprises relatively elevated remnant surfaces, this region experienced uplift relative to surrounding areas and is shaded red 186 Texas Tech University, Timothy Anderson, December 2008

(Figure 5.4, location E). If erosional remnant surfaces within a region are anomalously less elevated, this region experienced less uplift or subsidence and is shaded blue (Figure

5.4, locations A, B and C). The interpolation surface, KPS, illustrates these locations as depression or bulges. The existence of these features is explained easiest by faults (or tightly folded monoclines), and are illustrated as bold black lines in Figure 5.4.

In Figure 5.4, twelve faults have been proposed. Fault lines (bold lines in the map) are drawn at the midpoint between vertically dismembered erosional remnant surfaces. Faults are drawn where remnant surfaces are vertically offset 300 to more than

1,000 meters. Relative motion is denoted with a U (up) or D (down).With exception to the most eastern fault (parallel to Wooley Creek, between locations B and E) fault lines are less than 3 kilometers from mapped remnant surfaces. Red-shaded regions are areas where remnant surface altitudes exceed 1500 m and are typically vertically offset 500 meters or more from nearby remnant surfaces. Blue-shaded regions represent areas that have subsided 500 to 1,000 meters relative to adjacent erosional remnant surfaces.

Proposed faults have not been observed in the field. Nevertheless, these faults are in geomorphologically youthful regions that exhibit high relief, steep hillslopes, and steep streams. Given the complicated and aged bedrock geology, there is little field evidence of post-Miocene faults. These faults are explored in detail, later in this chapter.

5.5 Rock Uplift and Topographic Evolution based on River Profile Analysis

A collection of more than 600 stream segments in the KMP and adjacent areas exhibit characteristics that reflect differential uplift patterns in the KMP. As discussed in 187 Texas Tech University, Timothy Anderson, December 2008 chapter 4, a mountain’s fluvial network is sensitive to tectonic and/or climatic forcing, and preserves signals of such forcing in a river’s longitudinal river profile.

The most robust findings in KMP river profile analysis are the classification and measurement of river profile knickpoints and normalized channel steepness. This evidence is observed best in drainage area-channel gradient logarithmic plots.

Streams that are relatively steep with respect to their drainage area may have experienced increased incision or uplift rates (Wobus et al., 2006). Streams that exhibit prominent knickpoints could be adjusting to erosion/uplift conditions (Kirby and

Whipple, 2001; Crosby and Whipple, 2005; Wobus et al., 2006), or a result of differential strength of the lithologic substrate a river is incising (Duvall et al., 2004; Vanlaningham,

2006), increased erosion due to precipitation and glaciation at the onset of global cooling

(Roe, 2002). The following paragraphs consider these influences separately.

188 Texas Tech University, Timothy Anderson, December 2008

Fault Lines and Relative Movement of the Western Klamath Mountains

Legend

Fault Line

U Uplifted

D Downdropped

B Downdropped Area

Uplifted Area

Letters indicate regions referred to in text

Figure 5.4 This figure illustrates is modified from Figure 5.1. Proposed fault lines and regions of relative uplift or subsidence are interpreted from the elevation of individual remnant surfaces.

Changes in lithologies are recorded within a longitudinal river profile as a knickpoint, but may also affect channel steepness. Because different lithologies typically exhibit different erosion rates, and a stream’s gradient is dependent on the bedrock incision rate, a change in stream gradient could be produced by a change in rock strength.

Most stream segments in the Klamath Mountains overflow multiple terranes, making it difficult to isolate profile effects of an individual bedrock terrane. There have been no detailed investigations concerning the effects of incision rates of different bedrock materials and their consequential effect on channel steepness. 189 Texas Tech University, Timothy Anderson, December 2008

To remove terrane-specific erosion biases in mapping of river profiles, rivers within specific terranes were isolated and studied separately. Channel steepness of stream segments whose centroids are in the Western and Western Paleozoic and Triassic terranes are illustrated in Figure 5.5. According to this figure, streams in the central Klamath

Mountains are two to three times steeper than northern and southern streams (note the legends in the two maps). It follows that streams in the central Klamath Mountains are climatically or tectonically more disturbed than elsewhere.

Increased steepness of channel segments in the central Klamath Mountains could be attributed to climate change in the late Pleistocene. The current moist conditions and presence of Pleistocene glaciers are two well-documented erosion drivers in the central and eastern Klamath Mountains and Siskiyou Mountains (Irwin, 1997). If these external climate forces are primary contributors in removing massive amounts of rock in the central Klamath Mountains, high measured amounts of erosion in this thesis should spatially correlate with high amounts of precipitation and localized isostatic rebound.

190 Texas Tech University, Timothy Anderson, December 2008

Figure 5.5 This figure illustrates the variation of normalized channel steepness of stream segments that have their centroids in the Western and Western Paleozoic and Triassic terranes. In both examples, streams tend to be steeper near the central KMP. Note the difference in color scheme in the legends.

Channel gradients will consistently change shape when a Klamath terrane juxtaposes intrusive rock or alluvium. When overflowing alluvium, stream gradients are always exceedingly low and followed by downstream knickpoints (e.g., Hayfork Creek and the Scott, Illinois, and Rogue Rivers).

Accumulation of alluvium indicates that the corresponding fluvial system cannot efficiently flush out sediment. The Scott valley is bounded to the west by the Scott

191 Texas Tech University, Timothy Anderson, December 2008

Valley Fault and to the north by the Greenhorn Fault (Elder, 2008). Given the high angle and northerly strike of the Scott Valley Fault, the seemingly youthfulness of tributaries that feed into the Scott River, and moderate thickness of alluvium fill, the valley could be a poorly drained (half?) graben associated with the extensional Basin and Range Province

(Elder personal comm.., 2008). The northern Greenhorn fault may also be responsible for perturbing flow to the Klamath River, and generation of the Scott River knickpoint. A seismic line that documents subsurface bedrock relationships beneath the Scott valley would illuminate ideas about the tectonic development of the basin.

5.6 Landscape Development and Evaluation of Uplift and Erosion in local KMP

Arguments for landscape development in this section rely on topographic data, river profile analysis, the presence of elevated erosional remnant surfaces, and the mapped deposition of Quaternary sediments. Local regions of the KMP and adjacent areas exhibit exceptional or complicated geomorphologic characteristics. Recognizing that northern California and southern Oregon is geomorphologically diverse, this section categorizes landscape development into four regions, which are north of the KMP, the central KMP, the southern KMP, and west of the southern KMP.

192 Texas Tech University, Timothy Anderson, December 2008

Northern Klamath Mountains and Oregon Coast Range

The northern Klamath Mountains comprise the northern slopes of the Siskiyou

Mountains, the Northwestern Klamath Mountains, the Area, and the Coast

Range in Oregon (Figure 1.2). These geographic provinces include the Smith, Illinois,

Chetco, lower Rogue, and South Umpqua subbasins (Figure 4.1). The topography and fluvial network of the Oregon Coast Range is jagged, elevated, and of high relief. The

Northern Klamath Mountains, or regions north of the Siskiyou Mountains and east of the

Oregon Coast Range, exhibit relatively gentle topography and low gradient streams. The following paragraphs discuss the Oregon Coast Range and Northern Klamath Mountains.

According to coastal geodetic measurements taken by Kelsey and others (1996),

the Oregon Coast Range has uplifted (determined from marine terraces) (yellow and

purple lines in Figure 5.6). VanLaningham and others (2006) suggested that the odd-

shaped river profiles in southern Oregon reflect a disequilibrated geomorphologic stage.

The Illinois and lower Rogue subbasins are anomalous because the major rivers

descend to low elevations prior to entering the elevated topography of the Oregon Coast

Range. East of the Coast Range, the Illinois and Rogue Rivers flow through wide, low

relief, alluviated valleys at low channel gradients. In the Oregon Coast Range, the Rogue

and Illinois Rivers, despite being at elevations less than 500 m, flow through narrow,

high relief, deeply incised valleys. Because a river will either be blocked or deflected by

an older topographic feature, Oregon Coast Range topography must be younger than the

Illinois and Rogue Rivers. Furthermore, the Illinois and Rogue River must be incising at

a rate at least equal to the uplift of the Oregon Coast Range (steady-state equilibrium). In 193 Texas Tech University, Timothy Anderson, December 2008 this region, where ridge peaks are a kilometer higher than valley floors, the measured hillslopes (Figure 3.6) are likely near their critical threshold angle of failure. The few mapped remnant surfaces in elevated topography near the Coast Range are at elevations that exceed 1200 m, approximately 900 meters above valley floors.

If these remnant surfaces were at sea level, this valley relief reflects the magnitude of erosion the antecedent Illinois River is responsible for. The longitudinal profile of the Rogue and Illinois River display a prominent knickpoint where the rivers traverse the high relief coastal topography. Deep incision, steep hillslopes, elevated erosional remnant surfaces and prominent river profile knickpoints in the Illinois and

Rogue Rivers indicate that the Southern Oregon Coast Range is uplifting, however the region is experiencing steady-state equilibrium.

The vertical measurement and rate of uplift using preserved erosional remnant surfaces in Del Norte County and Oregon Coast Range are constrained by the age of the marine-nonmarine transition of the Wimer Formation. Stone (1992) demarcates the paleoshoreline (aged at 5 Ma) at elevations greater than 500 m. Ignoring eustatic changes, uplift of the paleoshoreline was at a long-term rate of 0.1 mm/yr. Nearby ridges underlain by Wimer Formation marine sediments exceed 1100 m in elevation. The magnitude and uplift rate of Smith subbasin could be more than twice that calculated for the shoreline.

Erosional remnant surfaces near Crescent City are at 150 m elevation,which contrasts with other surfaces at elevations greater than 1000 m, 20 kilometers inland.

194 Texas Tech University, Timothy Anderson, December 2008

Elevation and Uplift Patterns of the Klamath Mountains and Northwest Coast

Figure 5.6 This figure superposes topographic and rock uplift data of the coastal Pacific Northwest and Klamath Mountains. Data is collected from this study), and VanLaningham and others (2006).Topographic measurements not taken from this thesis are calculated from within the Coast Range.

195 Texas Tech University, Timothy Anderson, December 2008

This indicates the Smith subbasin has tilted 5% to the west. If this tilt is consistent northward, remnant surfaces that are hypothetically unmapped in the Chetco subbasin in coastal Oregon (Irwin, 1997) are predicted to be at lower elevations.

Streams within the Smith River subbasin are similar, with moderate channel gradients and knickpoints at an elevation of 500 m. As deduced from calculated erosion depths (KPES, Chapter 3), knickpoints in the Smith and lower Klamath subbasin (e.g.,

Blue Creek, Crescent City Fork, Turwar Fork) are spatially correlative with large depths of erosion. Because these knickpoints do not appear to correlate with lithology contacts,

(Figures 4.36 and 4.37) the knickzone could relate to a deformation zone or a eustatic fall in sea level. The latter hypothesis is not true because knickpoints are not consistently mapped at 500 m elevation in all northern California and southern Oregon coastal streams. If differential uplift of the Coast Range is accommodated by a tectonic process, the exact site of surface motion is unclear.

Vertical movement of the Smith River subbasin may be accommodated by high angle faults. The USGS reported at least 65 northwest striking Quaternary thrust fault segments offshore in the current Cascadia accretionary wedge ( www.seamless.usgs.gov/ ).

Seismic reflections suggest that nearshore faults are subvertical (Gullick et al., 2002), and one offshore fault near Crescent City has vertically displaced strata at least 800 meters

(Stone, 1992). Stone (1992) determined that that the Wimer Formation and St. George

Formation were offset vertically at least 210 meters along the Del Norte Fault. Both geologic units preserve the peneplain surface (Stone, 1992; Aalto, 2006).

196 Texas Tech University, Timothy Anderson, December 2008

To the north, remnant surface elevations near the northwest flowing Klamath

River are at elevations less than 300 m (location ‘A’ in Figure 5.4), which are 1100 m lower than Siskiyou remnant surfaces, 25 kilometers to the east. Thus, peneplain remnant surface elevations document the rise of the southern Siskiyou Mountains. The development of the southern Siskiyou Range is interpreted by Stone (1992) to have occurred in the Pliocene based on stratigraphic coarsening of Wimer Formation sediments.

Relief of the paleo-Siskiyou Mountains is documented by Plio-Pleistocene near- shore fluvial deposits of the Wimer and Prairie Creek Formations (Kelsey and Trexler,

1989; Stone, 1992; Aalto, 2006). Stone (1992) suggested the Klamath-derived Wimer

Formation conglomerates were transported from a Klamath fault. Kelsey and Trexler

(1989) suggest Klamath-derived coarse fluvial sediments of the Prairie Creek Formation subsided below sea level before uplifting along the northwest striking Grogan and Surpur

Faults in the late Pleistocene (Kelsey and Trexler, 1989; Irwin, 1997). Although both fluvial systems are interpreted to be in low-lying braided plains, no connection between the two paleo-rivers has been suggested. It is possible that orographic precipitation from a

Pliocene rise of the Siskiyou Mountains contributed to both drainage systems. The rise of

Grogan and Surpur Fault ridges prevents the modern Klamath River from draining west.

In the most northern regions of the study area, no remnant surfaces are mapped.

Landscapes in the Coquille and South Umpqua subbasins are characterized by low

channel gradients (Figure 4.8). The inland streams of the South Umpqua subbasin are

similar to the Rogue, Applegate, and Illinois Rivers, because they regularly deposit 197 Texas Tech University, Timothy Anderson, December 2008

Quaternary alluvium in many locations. South Umpqua tributaries exhibit smooth concave profiles, which is distinct from river profiles of the Coquille subbasin. No tributaries in the Coquille subbasin exhibit a smooth river profile, nor descend from drainage divides greater than 1000 m. Although some knickpoints are present at lithology contacts (e.g., Figure 4.44), many other knickpoints are mapped in elevated regions overlain by the Tyee Formation. Selander (2004) studied knickpoints in rivers that overflow sedimentary units of central Oregon and determined that they are associated with fracture planes associated with active faults. Irregular-shaped streams in the Chetco subbasin could be related to active faulting that has induced river incision of the tributaries, increased hillslopes and generated landslides.

Central Klamath Mountains

Erosional remnant surfaces are more elevated in the central Klamath Mountains than other KMP locations. As a result, the central Klamath Mountains are interpreted to have experienced the greatest amount of surface uplift in western California and Oregon.

The steepest streams, greatest relief, steepest hillslopes, and largest amount of erosion also support this interpretation. The geomorphology of the central KMP is complicated, however. The following section describes anomalous river profiles and uplifted surfaces.

Faults in the Central Klamath Mountains

If erosional remnant surfaces were once contiguously connected to create a quasi- planar surface, any dramatic offset in elevation between two adjacent remnant surfaces could be caused by a fault or antiform. Near the Klamath and Trinity River confluence, remnant surfaces at elevations of approximately 1000 m cap three different Klamath 198 Texas Tech University, Timothy Anderson, December 2008 terranes (Figure 5.4, location C). Five kilometers south of this site, a broad west-tilted remnant surface overlies the Ironside Mountain pluton at 1800 m elevation. This offset of remnant surfaces is easiest explained by a high angle northeast striking fault north of the

Ironside pluton (Figure 5.4, south of ‘C’). This predicted fault cuts Red Cap Creek, possibly through a river knickpoint.

Located in the Salmon subbasin to the east, a remnant surface atop the English

Peak batholith exceeds 2000 m in elevation (Figure 5.4, location E). Another remnant surface that caps the western portion of the Wooley Creek batholith (Figure 5.4, location

B) is approximately 500 meters lower in. This offset could reflect a steep northwest tilting flexure. The preferred interpretation within this thesis is a fault present in the approximate location and orientation of Wooley Creek.

The greatest depths of calculated erosion are in the Salmon River, Wooley Creek, and Klamath River. Erosional remnant surfaces of adjacent ridges are at altitudes between 1,400 m and 2,000 m. The Salmon subbasin has been significantly eroded beneath the KPS except at the confluence between the Klamath and Salmon Rivers where only small depths of erosion are calculated. Because no remnant surfaces are mapped in the Siskiyou Mountains, it is difficult to confidently constrain erosion of the Klamath subbasin. Nevertheless, undated fluvial terraces in the Klamath River valley indicate erosion has been active in the Quaternary (Yoshinobu personal comm.., 2008).

South of the Ironside Mountain batholith in the Trinity subbasin, erosional remnant surfaces decrease in elevations to just greater than 1300 m. Four exceptions include remnant surfaces south and southeast of Ironside Mountain batholith. Some 199 Texas Tech University, Timothy Anderson, December 2008 peneplain surfaces are 500 meters or higher than others. This observation is used to justify fault interpretation (Figure 5.4).

In the same subbasin, alluvium is present near Trinity Lake. This location is a site of recurrent deposition, having accumulated Weaverville basin sediments (Oligocene?) and Quaternary alluvium. Two hypotheses could explain deposition of alluvium including 1) recently renewed sedimentation generated by structural activity and 2) increased sedimentation as a result of Pleistocene glaciation. It is also possible that the growth of topography has inhibited sediment evacuation.

Despite several episodes of deposition, the Trinity subbasin has actually been eroded. Figure 3.19 demonstrates the depth and width of erosion in the Trinity subbasin since the late Miocene. High level Pleistocene stream terraces hundreds of feet above the

Trinity River predict recent river incision (Irwin, 1997). Pleistocene glaciers erosion has been partly responsible for the carving of tributary valleys. Furthermore the relief and depth of incision in Trinity tributary valleys suggest that the entire subbasin is eroding at a high rate.

Many large knickpoints and convexities correlate to locations where rivers overflow intrusive rocks. Where streams drain the Canyon Creek pluton, for example, the shape of the longitudinal profile resembles a stair case pattern (Figure 4.27). The longitudinal river profile is smooth and concave where the same stream flows over

Salmon Schist. The repeated appearance of knickpoints in intrusive rocks could reflect differential weathering patterns in plutonic rock and/or fracture patterns. Alternatively,

200 Texas Tech University, Timothy Anderson, December 2008 these knickpoints could be mapping differential uplift patterns that are best preserved in plutonic rocks.

Wobus and others (2006) demonstrate that knickpoints of similar elevations could reflect a transient stage in landscape adjustment or faulting. There are two geographic locations where baselevel may have fallen. Streams that overflow the Eastern Klamath terrane, Central Metamorphic terrane, and plutonic rocks north and east of the Trinity

River exhibit knickpoints at elevations of approximately 1400 m (Figure 5.7). Many of these knickpoints could be a consequence of differential rock strength in plutonic rocks or the presence of glacial till, but the consistent appearance of knickpoints at similar elevations is a signal that the landscape is responding to base-level fall (Crosby and

Whipple, 2005). If base level fall has occurred, the fluctuation in water-fill and/or construction of the dammed Trinity Lake may correspond to base level changes. This hypothesis is unlikely because it invokes unrealistically high knickpoint migration rates

(greater than a kilometer per year). Instead, the lowering of the Trinity River downstream of its convexity is a more reasonable hypothesis, and is evidenced by increased erosion between Pleistocene fluvial terraces (Irwin, 1997).

A second example of streams characterized by a common-elevation knickpoint pattern is in the Salmon subbasin (Figure 5.8). Tributaries that feed Wooley Creek,

Salmon River, and Klamath River exhibit knickpoints at approximately 1400 m in elevation. Like the previous example, these knickpoints could form by pulses of base- level lowering, which in this region, is the Klamath River. Indirect evidence of base-level fall includes the presence of elevated river terraces (no age given), steep hillslopes, and 201 Texas Tech University, Timothy Anderson, December 2008 abundance of landslide deposits in the Klamath River (Wagner and Saucedo, 1987;

Barnes personal comm., 2008). Study in the has revealed that although there is no direct relationship between hillslope angle and landslides, hillslopes at their critical angle of failure require base-level fall to generate landslides (Burbank, 2002). Thus, a base-level fall in the Klamath River could generate a pulse of migrating knickpoints that generate landslides at the banks of the Klamath River.

Knickpoints mapped in rivers that overflow metamorphic terranes are more subtle than those that overflow plutonic rocks. It is tempting to associate Mesozoic faults with knickpoints because many streams exhibit knickpoints at Siskiyou, Coast Range, and

Orleans fault contacts. Knickpoints correlate with other faults (chapter 4).

There are three issues that must be addressed in correlating fault lines with the location of river profile knickpoints. First, terrane-bounding thrust faults juxtapose genetically different rocks that could erode at different rates. For example, the Orleans fault juxtaposes the less resistant Galice Formation against WPT rocks and always results in a knickpoint.

202 Texas Tech University, Timothy Anderson, December 2008

Figure 5.7 Tributaries that flow into the Trinity River exhibit prominent knickpoints of varying elevations that cluster around 1400 m. Knickpoints are mapped in a diverse range of lithologies including gabbroic, peridotite, and metamorphic rocks. Knickpoints could relate to a broad fault zone along the eastern side of the central KMP or a transient response to base-level fall.

203 Texas Tech University, Timothy Anderson, December 2008

Figure 5.8 A series of knickpoints are plotted at elevations of approximately 1400 m in the Salmon and Klamath subbasin. One interpretation is that the Wooley Creek batholith and English Peak pluton are experiencing the effects of base-level fall.

204 Texas Tech University, Timothy Anderson, December 2008

Furthermore, stream incision will increase where associated with fracture zones. Selander

(2004) studied knickpoints in the Coast Range of Oregon and concluded that fracturing/jointing exert a powerful control on the location and shape of knickpoints. A second issue is the parallelism between streams and older fault lines. The South Fork

Trinity River, for example, parallels the Coast Range Fault (Figures 4.10 and 4.11) and, consequently, any mapped knickpoint will inevitably lie in the Coast Range Fault Zone.

Third, if a fault zone is reactivated, every stream that overflows the corresponding fault should exhibit a knickpoint.

Disregarding the aforementioned issues in knickpoint-fault line correlation, it seems that Mesozoic faults are reactivated. Where knickpoints and fault lines are coincident, only minor gradient increases are observed. Larger knickpoints, however, do not correlate with faults. Hence, if antecedent faults are reactivated in the KMP, profile deformation is not specifically concentrated along such faults. Furthermore, offset of erosional remnant surfaces do not coincide with antecedent faults in the central and southern Klamath Mountains. While reactivation of antecedent faults could explain uplift of the KMP, inconsistent patterns and erosional surface elevations call upon other mechanisms.

The development of knickpoints in the central KMP is not well-understood. Given the long history of tectonic accretion within the KMP, it is probable that pre-existing fractures and joints facilitate erosion and development of knickpoints. Although knickpoint relations are still being explored, Klamath knickpoint zones could, in part, reflect transient conditions of a landscape adjusting to climatic/tectonic perturbations. 205 Texas Tech University, Timothy Anderson, December 2008

Detailed mapping of erosion near the knickpoints, and the timing of base-level fall determined by stream terraces could illuminate ideas about the origin of many knickpoints within the KMP.

Southern Klamath Mountains

In the South Fork Trinity subbasin, more than a dozen erosional remnant surfaces have been mapped, providing good constraint on surface movement. One large surface at an altitude of approximately 1300 m caps rocks of the Rattlesnake Creek terrane (WPT) atop the ridge between the Hayfork and South Fork Trinity Rivers (Figure 5.4, location

‘D’). Near Corral Creek, twenty-three stream terraces are mapped (digitized map of

Elder, 2008), indicating recently enhanced tributary downcutting. Prominent knickpoints in the South Fork Trinity River and Hayfork Creek and irregularly-shaped tributaries with multiple knickpoints suggest that this basin is not in steady-state equilibrium.

Erosional remnant surfaces north of Hayfork Creek, South Fork Mountain, and near the Rattlesnake Creek terrane are at approximately the same elevation. The consistent elevation of many remnant surfaces in this area suggests that the subbasin was uplifted as a single, non-tilted block (Figures 3.12 and 5.2). In this model of subbasin uplift, lower remnant surfaces in tributary valleys (e.g., Corral and Rattlesnake Creek) may have once been attached to the non-tilted block. Weaverville sediments that could overlap the erosional remnant surface are at elevations lower than 800 m. If the erosional surface does underlie the Weaverville Formation as suggested by Diller (1902), a north- striking fault has down-dropped Hayfork valley a minimum of 500 meters. This value is

206 Texas Tech University, Timothy Anderson, December 2008 calculated from subtracting Hayfork valley elevations from nearby remnant surface elevations (1300 m).

Eel and Mad-Redwood Subbasins

Many publications have referenced rock uplift of the Eel and Mad-Redwood subbasins. (e.g., Riddihough, 1984; Engebretson, 1985; Kelsey and Carver, 1988; Carver and Burke, 1988; McCrory, 1989; Merritts and Bull, 1989; Aalto et al., 1995, McNeill,

2000, Gulick et al., 2002; Aalto, 2006; Lock et al., 2006). These studies inferred rock uplift by structurally offset geological formations (Kelsey and Carver, 1988), raised littoral/bathyal sediments (Merrits and Bull, 1989; McCrory, 1989; Aalto et al., 1995), progradation stratigraphy and coarsening of sediments offshore (McCrory, 1989;

McNeill, 2000; Gulick et al., 20002; Aalto, 2006), and geomorphologic evidence (Carver and Burke, 1988; Kelsey and Carver, 1988; Merritts and Bull, 1989; Lock et al., 2006).

Despite substantial documentation of geologic investigation since the Miocene, the timing and cause of rock uplift is still unclear. Irwin (1997) notes that peneplain surfaces may be present further south.

Studying erosional surface elevations, Mad subbasin drainage divides uplifted half a kilometer relative to its corresponding valley floors (Figure 5.4, region ‘A’). The moderate or low stream gradients of rivers that traverse the subbasin indicate little uplift of the valley floor (Figure 4 . 9). The presence of lower remnant surfaces in the coastal

regions of Mad Valley confirms this indication. Strong faults are interpreted between

peneplain surfaces at valley floors and along valley ridges (Figure 5.4, location ‘A’).

These faults are kinematically compatible with late Pleistocene-to-present northeast- 207 Texas Tech University, Timothy Anderson, December 2008 southwest contractional models within the Mad River Fault Zone (MRFZ) (Kelsey and

Carver, 1988; Carver and Burke, 1988; McCrory, 1989; Aalto et al., 1995; Lock et al.,

2006).

Despite few Cenozoic geologic features of the Mad River Fault Zone and Eaton

Roughs Fault Zone (also in Mad River Valley) has been interpreted to be a translational tectonic regime (e.g. Kelsey and Carver, 1988; Merritts and Vincent, 1989; Furlong and

Scwartz, 2004; Lock et al., 2006). Right-lateral movement of the Eaton Roughs Fault

Zone is produced by high angle faults, the offset of marker beds, and differential uplift/fall of valley floors (Kelsey and Carver, 1988). Despite poor timing constraints, these observations are consistent with transcontractional and transtensional tectonic motions produced by the northward migration of the Mendocino Fracture Zone (Burke and Carver, 1988; Kelsey and Carver, 1988; Aalto, 2006; Lock et al., 2006).

Streams south of Redwood Creek are not in steady-state equilibrium given their longitudinal profile (e.g. , Van Duzen River, Owl Creek, Lawrence Creek, and Larabee Creek) (Figures 4.2, 4.3 and 4.4). Many knickpoints are at the approximate location of abrupt changes in the stream flow direction, usually expressed in map view by westward bends in the river flow direction (Figure 5.9). Lock and others (2006) noted stream bends (fish-hooks) in the , Middle Fork Eel River, and Van

Duzen Rivers. They suggest these morphologic bends are the result of a northward migration of drainage divides that ‘capture’ southward flowing streams and redirect them into the main trunk of the Eel River. These observations support a model of a northward advancing wave of uplift that is related to crustal thickening of the North America plate 208 Texas Tech University, Timothy Anderson, December 2008 by viscous coupling between the subducting Gorda slab and North America plate (Fulong and Scwartz, 2004). If true, the presence of the prominent Mad River and Van Duzen knickpoints could be the northern extent of crustal warping associated with the

Mendocino Fracture Zone.

The most significant depths of erosion in the Mad River are greater than 800 m, and located immediately downstream of a prominent knickpoint (Figure 3.14). Local erosion in the Van Duzen River also occurs downstream of a prominent river knickpoint.

Thus, it appears that erosion mapped using peneplain surface elevations are consistent with increased gradients in the Mad and Van Duzen River.

Red arrows in Figure 5.9 illustrate westward bends of southern Klamath Mountain rivers. Prominent knickpoints (transparent black circles) are located near many westward bends in the river. Smaller knickpoints are not shown. The letter ‘U’ indicates locations where mapped peneplain surfaces are high relative to other peneplain surfaces. From map view, it appears that localized uplift is perturbing northwest striking fluvial systems.

The Eel River is the on-land drainage system of a northeast-striking forearc basin where Mio-Pleistocene deposits unconformably rest on Franciscan basement (Kelsey and

Carver, 1988; McCrory, 1989; Aalto et al., 1995; Gullick et al., 2002). Evidence of uplift of on-land segments did not occur until the Pliocene (McCrory, 1989; Gullick, 2002) or even Pleistocene (Aalto, 2006; Lock et al., 2006). Because the Eel River trunk does not exceed 100 m in altitude until more than 70 km inland and coastal tributaries exhibit shallow channel gradients, it does not appear the local base level is uplifting. No remnant surfaces have been mapped in this area to constrain long-term uplift. 209 Texas Tech University, Timothy Anderson, December 2008

Figure 5.9 This figure illustrates the westward bends (arrows) in river flow directions, and their association with prominent knickpoints preserved within the river profile (transparent red circles).

210 Texas Tech University, Timothy Anderson, December 2008

The tributaries of the Middle Fork Eel River and main stem of the Eel River exhibit knickpoints and steep channel gradients. While the lower Eel subbasin is interpreted to be in steady-state equilibrium, the upper and middle Eel subbasins are not.

This is particularly true for eastern sections of the two basins, where tributaries that descend from steep and elevated topography (southern shaded region in Figure 3.5) are noticeably steeper (Figure 4.2).

5.7 Uplift Models

Figure 5.10 is a schematic diagram of the orographic development of the Klamath

Mountain topography in applying the concept of a critically tapered wedge. According to theory, an orogenic wedge will maintain an equilibrated wedge shape, described by angles measured relative to a horizontal line (Dahlen and Suppe, 1988). The values of these angles, α and β, are proportionally dependent on each other and, as such, an

increase in α must be subsequently followed by a decrease in β. The balance between these critical taper angles is sensitive to the distribution of material by tectonic processes such as subduction erosion, sediment underplating, structural duplexing and/or structural thinning, as well as geomorphologic processes including climatic erosion and depositional subsidence.

Figure 5.10 illustrates a two-step stage of orogenic development: 1) the lowering of the central Klamath Mountains and development of the peneplain surface, and 2) the flexural rise of the central Klamath Mountains. Deep weathering in the formation of the remnant surface, deposition of bathyal marine sediments atop present day topography,

211 Texas Tech University, Timothy Anderson, December 2008 and soil formation are evidence that the upper face of the orogenic wedge (measured by

α) was at a substantially lower angle than today. In the top diagram, it is measured at

0.5º. It is unclear what led to the planation of northern California, however Aalto and others (1996) suggest subduction erosion could have been responsible for the lowering of the basal décollment (increase in angle β). Because the current subduction angle of the

Gorda crust ( β) is 12º (Gullick, 2002), and the projection of the modern tilted peneplain surface is 2.9º, the dip angle of β prior to regional uplift is calculated as 14.9º.

Since erosional remnant surfaces were at the approximate elevation of sea level

during the early Pliocene (Stone, 1992; Aalto et al., 1995; Aalto, 2006), the central

Klamath Mountains has been uplifted a maximum of 2,000 meters, or an increase in the α

angle of a critically tapered wedge. In the diagram, an increase in elevation of the upper

face corresponds to a decrease in elevation of the lower face.

This geometric modeling supports the proposed uplift theories of sediment

underplating and serpentinized wedge models (Aalto et al., 1995; Yoshinbou, 2006).

Similarly, isostatic rebound due to erosion could maintain the shape of the wedge.

212 Texas Tech University, Timothy Anderson, December 2008

Figure 5.10 This is a schematic drawing of an orogenic wedge representing the Klamath Mountains. The top picture represents a time period when the western Klamath Mountains were beveled to a low-lying regional surface. Aalto (2006) suggested that the low elevations during the early Pliocene could be a result of subduction erosion.

213 Texas Tech University, Timothy Anderson, December 2008

Subduction-Related Uplift

There are a number of surface development models that could be closely linked to subduction. At the Cascadia trench, the young oceanic crust of the Gorda slab buoyantly subducts beneath the North American crust at a subduction angle between 9º and 12º

(Smith and Knapp, 1980; McNutt, 1983; Furlong and Schwartz, 2000; Gullick et al.,

2002). This shallow subduction angle has led some to theorize that the Gorda lithosphere and overlying North America lithosphere are coupled, resulting in structural duplexing of the upper North American plate (Kelsey and Carver, 1988). It has also been suggested that the northward migration of the Blanco Fracture Zone is acting as a crustal ‘plow’ that is generating elevated surface topography (McNutt, 1983; Kelsey et al., 1994). Other arguments using sedimentological evidence and domal modeling have suggested subduction erosion and subsequent underplating of sediment to the base of the Klamath block as a hypothesis for Klamath uplift (Mortimer and Coleman, 1985; Aalto et al.,

1996; Aalto, 2006). Some arguments have considered the decrease in normal convergence velocities of the Gorda plate has resulted in the buoyant Gorda slab to isostatically rise towards the North American plate (McCrory, 1989, 1995, 1996).

Uplift Generated by Structural Duplexing

If the Klamath block and adjacent Coast Range are experiencing intraplate contraction in response to regional stresses generated at the Cascadia deformation front, the crust could be thickening by structural duplexing of terranes. Northeast-southwest contraction has been interpreted by field observations of contractional features including older-over-younger strata relations, Mio-Pleistocene sedimentary units truncated by 214 Texas Tech University, Timothy Anderson, December 2008

Franciscan basement at thrust faults, and warping/offset of Holocene marine terraces in the Eel and Mad-Redwood valleys (Burke and Carver, 1988; Carver, 1988; McCrory,

1989; Aalto et al., 1995; McNeill et al., 2000; Gullick et al., 2002; Aalto, 2006). Field observations have documented Miocene-Holocene sedimentary formations in this region have been deformed (Kelsey and Carver, 1988; McCrory, 1989; Aalto et al., 1995; Aalto et al., 1996; McCrory, 1996).

Mortimor and Coleman (1985) suggest that the Condrey Mountain Dome is at the focal point of broad arching of the Cascadia forearc in the Klamath Mountains. High angle faults that radiate away from the dome, tilted hanging wall rocks, and decreasing metamorphic grade from the dome’s central edifice support evidence of a domal uplift model. The eastward tilting Miocene Cascade volcanic beds are unconformably overlain by horizontal High Cascade beds, providing excellent timing-constraints on rock movement in the area (Elder personal comm.., 2008). Streams that overflow Condrey

Mountain Dome are exceptionally steep. Rise of the dome may have increased hillslopes local hillslopes and flooded the Applegate subbasin with alluvium. While remnant surfaces nearest Condrey Mountain are elevated (e.g., Wooley Creek and English Peak

Pluton), elevated surfaces in the Trinity subbasin and atop South Fork Mountain more than 100 kilometers away are at similar elevations. The rise of Condrey Mountain could not uplift the Klamath alone.

Although seismicity is recorded less often north of the Klamath River, Plio-

Pleistocene activity along faults has been documented at the mouth of Redwood Creek

(Kelsey and Trexler, 1989) and near Crescent City (Stone, 1992). North of the Klamath 215 Texas Tech University, Timothy Anderson, December 2008

River, north striking thrust faults have been mapped onshore and offshore (Stone, 1992; www.seamless.usgs.gov/)). The Del Norte Fault vertically offset Pleistocene gravels approximately 40 m (Stone, 1992). Uplift along the Surpur Faults and Grogan Faults offset Pleistocene sedimentary units (Kelsey and Trexler, 1989). To the north, transcurrent Quaternary faults are mapped in the coastal regions of Oregon and could also be generating uplift (Kelsey et al., 1996).

Because of the presence of remnant surfaces in northwest striking valleys and along ridges west of the southern Klamath Mountains, topographic relief is created by differential uplift. These well-defined topographic lineaments are orthogonal to Gorda-

North America convergence, and for this reason, could be the result of contractional deformation. Kelsey and Carver (1988) suggested the most eastern contractional tectonic regimes terminate at thee eastern edge of the Mad River Fault Zone. Uplift generated by transcurrent faulting has been suggested in Oregon (e.g., Kelsey et al., 1996). Because detailed geologic mapping at these latitudes is at a coarse resolution, it is difficult to confidently interpret a close relationship between individual structures and river profile deformation. Other studies combining field mapping and river profile analysis (e.g.,

VanLaningham et al., 2006) suggests the odd-shaped river profiles in Oregon could be a result of tectonic instability.

If uplift is accommodated by structural duplexing, the extent of deformation is difficult to determine. Streams in the relatively aseismic Klamath Mountains exhibit knickpoints, but their association with faulting is not obvious. Significantly offset remnant surfaces in the central Klamath Mountains (Figure 5.4) are not kinematically 216 Texas Tech University, Timothy Anderson, December 2008 compatible with tectonic contraction. Thus, uplift of coastal regions may be influenced by

GP-NAP contraction, but not the central Klamath Mountains.

MTJ Tectonics

Crustal uplift occurs in advance of the migrating Mendocino Fracture Zone

(McNutt, 1983; Kelsey and Carver, 1988; Merritts and Vincent, 1989; Aalto et al., 1995;

Furlong and Schwartz, 2004; Lock et al., 2006). Merritts and Vincent (1989) documented the steepening of coastal streams just north of the projected fracture zone. Lock and others (2006) suggested that the northward advancement of the southern terminus of the

Gorda crust resulted in the uplift of the entire coast since the late Miocene. They suggested that the current drainage patterns in the Middle Fork and Upper Eel basins have changed in response to rapid uplift (McCrory, 1989; Lock et al., 2006). Northward tilting of strata is documented 20 kilometers north of the MTJ (Gullick et al., 2002).

Tectonic perturbances in the form of knickpoints are spatially coincident with

westward bends rivers (Figure 5.9). It is unclear if peneplain surfaces preferentially

increase in elevation to the south or to the east, given their southeast strike. These

geomorphologic features and uplifted remnant surfaces could reflect trans-contraction

generated uplift caused by the northward migration of the Mendocino Fracture Zone.

Uplift Induced by Climate-Driven Erosion

Figure 5.11 compares maps of modern topography with peneplain surface-based

calculations of uplift and erosion. According to the figure, the greatest amount of uplift

spatially coincides with the greatest amount of erosion in the central Klamath Mountains.

Figure 5.12 demonstrates the close relationship calculated erosion has with steep river 217 Texas Tech University, Timothy Anderson, December 2008 channels, steep hillslopes, and Pleistocene glaciers. Observations from both figures support an uplift model where the central Klamath Mountains have isostatically rebounded in response to erosion.

Erosion in the central KMP is evidenced by 1000 meters of valley relief in

Klamath and Trinity valleys, elevated erosional remnant surfaces more than 1500 meters above streams, Pleistocene terrace deposits hundreds of meters above streams (Irwin,

1997), steep hillslopes, and steep channel gradients (Chapter 4). According to Figures

5.19, the maximum depth of erosion is 1,480 meters, respectively. If the central Klamath

Mountain topography isostatically rebounds to restore its elevation to 5/6 th of its original elevation (Molnar and England, 1990), the elevation of paleotopography can be predicted by the depth of removed material within individual subbasins. The subbasin that experienced the deepest depths of erosion, the Salmon subbasin, isostatically rebounded

1,230 meters, or 770 meters less than the amount of uplift predicted by adjacent erosional remnant surfaces. It is possible that the 770 meters could have been accommodated by tectonic uplift.

218 Texas Tech University, Timothy Anderson, December 2008

Modern Elevation Vs Surface Uplift Vs Erosion

Figure 5.11 This figure compares modern topography with the calculated magnitude and rate of uplift and erosion since 5 Ma.

219 Texas Tech University, Timothy Anderson, December 2008

Erosion Compared with River Steepness, Hillslopes, and Pleistocene Glaciers

Figure 5.12 This figure compares calculated erosion with regions where channels are steep (shaded white), hillslope measurements, precipitation, and the location of Pleistocene glaciers.

220 Texas Tech University, Timothy Anderson, December 2008

Because littoral sediments are not deposited in the central Klamath Mountains, it is also possible that erosional remnant surfaces were never at sea level and the Salmon subbasin was at an elevation of approximately 770 m. If this latter case were true, the original peneplain dipped 1.1° westward prior to uplift.

The uplift history of the eastern regions of the Klamath Mountains is unclear.

The presence of Pleistocene glacial deposits (Figure 5.12) in eastern sections of the central Klamath Mountains indicate that the eastern Klamath paleotopography was cool and elevated enough for alpine glacier growth during the Pleistocene. Because of glacier production, Pleistocene cooling generated the greatest amount of erosion, and subsequent isostatic rebound, in the eastern Klamath Mountains first. Steep streams, deeply incised valleys, high relief, and elevated topography of the eastern Klamath Mountains support this argument. The tilt of the western Klamath peneplain could actually be generated by the isostatic rise of the Klamath block, which was partially driven by glaciers.

The Rogue and Illinois Rivers overflow alluvium within wide valleys until reaching the Oregon Coast Range topography at elevations of 230 and 380 m, respectively. The presence of alluvium reflects a channel’s incapacity to transport sediment. In regions north of the Siskiyou Mountains (excluding the Coast Range), fluvial sedimentation rates produced by nearby hillslopes exceed the local base-level lowering rate of major rivers. It can be hypothesized that the Rogue, Applegate, and

Illinois Rivers are incapable of lowering their local base level; a possible consequence of low discharge rates in a relatively arid climate. VanLaningham and others (2006) reported subsiding survey benchmarks in the Rogue Valley area (Figure 5.6), providing 221 Texas Tech University, Timothy Anderson, December 2008 evidence that major drainages are not in a setting that can generate channel gradients steep enough to transport sediments, and/or lower base level by erosion.

In contrast to the Rogue Valley area, the Coast Range topography is elevated, has steep hillslopes, steep channel gradients, and receives very high precipitation. The steepest Coast Range rivers are 50 kilometers inland (Figure 5.12, isolated shaded white region in the north) where Pleistocene glaciers are mapped (Irwin, 1997) and annual precipitation exceeds 150 inches. The few peneplain surfaces record the uplift of the

Coast Ranges by more than a kilometer. Kelsey and other (1996) and VanLaningham and others (2006) record high uplift rates of coastal marine terraces and coastal survey benchmarks (Figure 5.6). Thus, the Oregon Coast Range is currently uplifting today, and is partly responsible for elevated erosional remnant surfaces that were at sea level at 5 Ma

(Aalto, 2006).

Climate may have an influential role in the Rogue Valley area and Oregon Coast

Range. The wet Oregon Coast Range is uplifting (VanLaningham, 2006) and eroding by steep river channels in its deeply incised valleys. Pleistocene glacier deposits have been also been mapped in at least 2 sites (Irwin, 1997). The relatively dry Rogue Valley is subsiding (VanLaningham, 2006) while accumulating sediment on the valley floor.

222 Texas Tech University, Timothy Anderson, December 2008

5.8 Future Tests

Ideas and certainty about the amount of uplift are not well-constrained in this thesis. Given the strong relationship between topography, river profiles, and the elevation of the erosional remnant surface, uplift can be inferred, particularly within the central

Klamath Mountains.

Field observations of local stream characteristics, measurement of stream discharge rates, dating and measurement of fluvial terraces, and bedload/bedrock descriptions at the site of river profile knickpoints could constrain knickpoint interpretation. The relationship between fluvial terraces and erosional remnant surfaces could constrain timing, and may reveal fault activity (e.g., if fluvial terraces are more elevated than remnant surfaces. Further detailed mapping of peneplain surfaces would constrain uplift and erosion calculations, as well as the geographic extent of the Klamath peneplain paleosurface. Confirmation of the absence of the surfaces in critical regions

(e.g., Siskiyou Mountains) could help constrain calculations of paleotopography and erosion. Sedimentological provenance studies of post-Miocene units should correlate with paleotopographic features. Exhumation rates from low temperature thermochronometric data would be a powerful dataset to compare uplift predicted in this thesis.

223 Texas Tech University, Timothy Anderson, December 2008

Chapter 6

Conclusion

The Klamath Mountains Province (KMP) represents an anomalously elevated region that has been deeply incised by the Klamath, Salmon, Trinity, and Smith drainage basins. The topographic features included within the KMP are seemingly youthful, characterized by steep hillslopes, high basin relief, mean elevations of one kilometer, and maximum elevations that exceed two and a half kilometers. The elevated and incised topography of the KMP is largely a post-Miocene mountain range that is evidenced by the uplift of a rare erosional remnant surface that was at sea level at 5 Ma. A hypothesis of climate-driven uplift in the KMP is supported by interpolated surfaces that measure uplift and erosion spatially coincide with geomorphologic features, particularly steep channel gradients. The following bullets are summarized measurements and observations from this thesis:

Evidence of Uplift of the Central Klamath Mountains

• Klamath peneplain and Plio-Pleistocene littoral sediments at elevations

between 0 and 2,018 m (Figure 3.11)

• Exhumation rate is between 0.4 mm/yr (beginning at 5 Ma) and 2.9

mm/yr (beginning at 0.7 Ma) (Figure 5.11)

• Interpolation surface connecting mapped peneplain surfaces increases

eastward (Figure 3.16)

• 13,810 km 3 of rock is exhumed above sea level in western KMP (Figure

5.1) 224 Texas Tech University, Timothy Anderson, December 2008

• The mean elevation of the western Klamath Mountains has increased

from near sea level at 5 Ma to as much as 760 m

• Relatively high channel steepness in the central KMP (Figure 4.8)

• Mean elevations exceed 1,000 m (Figures 3.1 to 3.4)

• Maximum elevations exceed 2,500 m (Figures 3.1 to 3.4)

Evidence of Erosion in Central Klamath Mountains

• Modern topography is as much as 1,482 m below interpolation surface

(Figure 3.9).

• Very high hillslopes, generally greater than 30°

• Valleys have high relief, often exceeds a kilometer (Ch. 4)

• Central Klamath Mountain rivers are 2 to 3 times steeper than rivers in

adjacent regions

• 3,860 km 3 of rock material has been removed from western KMP

o This is 29% of total exhumed rock (Figure 5.1)

• Lack of alluvium in central Klamath Mountain valleys

o Evidence of in Trinity, Salmon and Klamath

valleys

Evidence of Antecedent Topography

• Less than 16% of modern topography existed at 5 Ma

• At least 6 antecedent ridges (Figure 3.15)

225 Texas Tech University, Timothy Anderson, December 2008

o Antecedent Siskiyou Mountain ridges exceeds 900 m in

paleoelevation

o Antecedent South Fork Mountain exceeds 500 m in paleoelevation

• Ridges in central Klamath Mountains exceed 500 m in paleoelevation

Several tectonic models suggest that uplift of adjacent regions of the Klamath

Mountains is assisted by structural duplexing generated by subduction-related processes.

Offset of remnant surfaces near the coast support this interpretation. The offset of remnant surfaces in the Klamath Mountains, however, are inconsistent with duplexing models and call for alternative hypothesis. According to Molnar and England (1990), the depth of erosion in a column of crust can result in an isostatic response of the crust. Using calculated values of erosion (Figure 3.19), the amount of predicted isostatic rebound in the Klamath, Salmon, Trinity, and South Fork Trinity valleys can account for at least half of the interpolated surface-based uplift measurements. Data from this thesis therefore suggests that the uplift of features in the Klamath Mountains to their present-day elevations is largely a function of erosion, engineered by erosion in the Klamath and

Trinity River subbasins. If this interpretation is correct, then the young apatite fission- track ages are predicted to be in the upper Trinity, Salmon, and Lower Klamath subbasins.

226 Texas Tech University, Timothy Anderson, December 2008

Literature Cited

Aalto, K., 1981, Multistage mélange formation in the Franciscan Complex, northernmost California: Geology, v. 9, p. 602-607.

Aalto, K., Moley, K., and Stone, L., 1995, Neogene paleogeography and tectonics of northwestern California, Cenozoic paleogeography of the - II Pacific Section: Society of Sedimentary Geology, v. 75, p. 162-180.

Aalto, K., Moley, K., and Miller, W., 1996, Discussion: evolution of a trench-slope basin within the Cascadia subduction margin: the Neogene Humboldt Basin, California. Sedimentology, v. 43, p.761- 769.

Aalto, K., 2006, The Klamath Peneplain: A review of J.S. Diller’s classic erosion surface, in Snoke, A.W., and Barnes, C.G., eds., Geological studies in the Klamath Mountains province, California and Oregon: A volume in honor of William P. Irwin, Geological Society of America Special Paper 410, v. 22, p. 451-463.

Abbott, L., Silver, E., Anderson R., Smith, R., Ingle, J., Kling, S., Haig, D., Small, E., Galewsky, J., and Silter W., 1997, Measurement of tectonic surface uplift rate in a young collisional mountain belt: Nature, v. 385, p. 501-507.

Allen, P., 2008, From Landscapes into Geological history: Nature, v. 451, no. 17. p. 274- 276.

Anders, A., Roe, G., Hallet, B., Montgomery, D., Finnegan, N., and Putkonen, J., 2006, Spatial patterns of precipitation and topography in the Himalaya, in Willett, S., Hovius, N., Brandon, M., and , D., eds. Tectonics, Climate, and Landscape Evolution: Geological Society of America Special Paper 398, Penrose Conference Series, p. 39-53.

Argus, D. and Gordon, R., 1991, Current Sierra Nevada-North America motion from very long baseline interferometry; implications for the kinematics of the Western United States: Geology, v.19, no. 11, p. 1085-1088.

Atwater, T., 1970, Implications of for the Cenozoic evolution of North America: Geological Society of America Bulletin, v. 81, p. 3513-3536.

Bacon C., Lanphere M., and Champion D., 1999, Late Quaternary slip rate and seismic hazards of the West Klamath Lake fault zone near Crater Lake, Oregon Cascades: Geology, v. 27, no. 10, p. 43–46.

227 Texas Tech University, Timothy Anderson, December 2008

Blake, M., and Jones, D., 1981, The Franciscan Assemblage and Related Rocks in Northern California: A Reinterpretation, in W.G. Ernst ed., The Geotectonic Development of California Englewood Cliffs, NJ, Prentice-Hall, 706 p.

Bostock, M., Hyndmen, R., Rondenay, S., and Peacock, S., 2002, An inverted continental Moho and the serpentiniztion of the forearc mantle: Nation, v. 417, p. 536-538.

Brocher, T., Parsons, T., Tréhu, A., Snelson, C., and Fisher, M., 2003, Seismic evidence for widespread serpentinized forearc upper mantle along the Cascadia margin: Geology, 2003, v. 31, no. 3, p. 267-270.

Burbank, D.W., and Anderson, R.S., 2001, Tectonic Geomorphology: Blackwell Scientific, Oxford , 270 p .

Burbank, D., 2002, Rates of erosion and their implications for exhumation: Mineralogical Magazine, v. 66, p. 26-52.

Bursik, M., and Gillespie, A., 1993, Late Pleistocene glaciation of Mono Basin, California: Quaternary Research, v. 39, p. 24-35.

Carver, G., 1985, Quaternary Tectonics North of the Mendocino Triple Junction, in Kelsey, H., Lisle, T., and Savina, M., eds, Redwood country guidebook: American Geomorphological Field Group 1985 Field Trip, p.155-167.

Cashman, S. and Elder D., 2002, Post-Nevadan detachment faulting in the Klamath Mountains, California: GSA Bulletin, v. 114, 12, p. 1520-1534.

Clark, M., House, M., Royden, L., Whipple, K., Burchfiel, B., Zhang, X., and Tang, W., 2005, Late Cenozoic uplift of southeastern Tibet: Geological Society of America, v. 33, p. 525-528.

Cox, A. and Engebretson, D., 1985, Change in motion of Pacific plate at 5 m.y. B. P.: Nature, v. 313, p. 472-474.

Cox, R.T. 1994, Analysis of drainage-basin symmetry as a rapid technique to identify areas of possible Quaternary tilt-block tectonics: an example from the Mississippi embayment: Geological Society of America Bulletin, v. 106, p. 571-581.

Crosby, B. and Whipple, K., 2005, Knickpoint initiation and distribution within fluvial networks: 236 waterfalls in the Waipaoa River, North Island, New Zealand: Geomorphology, v. 82, p. 16-38.

228 Texas Tech University, Timothy Anderson, December 2008

Dahlen, F. and Suppe, J., 1988, Mechanics, growth, and erosion of mountain belts: GSA special paper, v. 218, p. 161-178.

Davis, G., 1969, Tectonic Correlations, Klamath Mountains and Western Sierra Nevada, California: Geological Society of America Bulletin, v. 80, no. 6, p. 1095–1108.

DeLong, S., Pelletier, J., and Arnold, L, 2007, Bedrock landscape development modeling: Calibration using field study, geochronology, and digital elevation model analysis: GSA Bulletin, v. 119, no. 2, p.157-173.

Diller, J.S., 1902, Topographic development of the Klamath Mountains: Washington, D.C., U.S: Geological Survey Bulletin, v. 196, p. 69.

Duvall, A., Kirby E., and Burbank D., 2004, Tectonic and lithologic controls on bedrock channel profiles and processes in coastal California, Journal of Geophysical Research, v. 109, F03002, doi:10.1029/2003JF000086.

Elder, D., 2008, Digital Map of the Klamath Mountains Province, unpublished.

England, P., and Molnar, P., 1990, Surface Uplift, uplift of rocks, and exhumation: Geology, v. 18, 1173-1177.

Farley, K., Rusmore, M., and Bogue, S., 2001, Post-10 Ma uplift and exhumation of the northern Coast Mountains British Columbia: Geology, v. 29, no. 2, p. 99 – 102.

Finnegan, N., Roe G., Montgomery D., and Hallet, B., 2005, Controls on the channel width of rivers: Implications for modeling fluvial incision of bedrock: Geological Society of America, v. 33, no 3, p. 229-232.

Furlong, K. and Scwartz, S., 2004, Influence of the Mendocino Triple Junction on the Tectonics of Coastal California, Annual Review of Earth Planet: Science, v. 32, p. 403- 433.

Freymueller, J., Murray, M., Segall, P., and Castillo, D., 1999, Kinematics of the Pacific North America plate boundary zone, northern California: Journal of Geophysics Research, v. 104, p. 7419-7441.

Fuis, G., Zucca, J., Mooney, W., and Milkereit, B., 1987, A geologic interpretation of seismic-refraction results in northeastern California: GSA Bulletin, v. 98, p. 53- 65.

Gani, N. and Gani M., 2007, Blue Nile Incision on The Ethiopian Plateau: Pulsed plateau growth, Pliocene uplift, and hominin evolution: Gelogical Society of America Today, v. 17, No 9, p. 4-11. 229 Texas Tech University, Timothy Anderson, December 2008

Gulick, S., Meltzer, A., and Clarke Jr. S., 2002, Effect of the northward-migrating Mendocino triple junction on the Eel River forearc basin, California: Stratigraphic development: Geological Society of America Bulletin, v. 114, No. 2, p. 178–191.

Hack, J., 1957, Studies of Longitudinal Stream Profiles in Virginia and Maryland: Geological Survey Professional Paper 294 – B, p. 45- 95.

Heller P., Tabor R., and Suczek C., 1987, Paleogeographic evolution of the United States Pacific Northwest during Paleogene time: Canadian Journal Earth Sciences, v. 24, p. 1652-1667.

Howard, A., 1967, Drainage analysis in geologic interpretation: a summary: American Association of Petroleum Geologists, v. 51, p. 2246-2259.

Howard, A. and Kerby, G., 1983, Channel changes in badlands: Geological Society of America Bulletin, v. 87, p. 739-752.

Howard A., Siedl M., and Dietrich W., 1994, Modeling fluvial erosion on regional to continental scales, Journal of Geophysical research, v. 99, p. 13,971-13,986.

Holbrook, J. and Schumm, S., 1999, Geomorphic and sedimentary response of rivers to tectonic deformation: a brief review and critique of a tool for recognizing subtle epirogenic deformation in modern and ancient settings: Tectonophysics, v. 305, p. 287- 306.

Hurtrez J, Lucazeau F., Lavé J, and Avouac J., 1999, Investigation of the relationships between basin morphology, tectonic uplift, and denudation from the study of an active fold belt in the Siwalik Hills, central Nepal: Journal of Geophysical Research, v. 104, p. 12,779-12,796.

Irwin, W., 1972, Terranes of the western Paleozoic and Triassic belt in the southern Klamath Mountains, California: U.S. Geological Survey Professional Paper 800-C, p. 103-111.

Irwin, W., 1981, Tectonic accretion of the Klamath Mountains, in W.G. Ernst ed., Prentice-Hall, Englewood Cliffs, NJ, USA, p. 29-49.

Irwin, W., 1997, Preliminary map of selected post-Nevadan geologic features of the Klamath Mountains and adjacent areas, California and Oregon: U.S. Geological Survey Open File Report 97-465.

Irwin, W. and Wooden, J., 1999, Plutons and Accretionary Episodes of the Klamath 230 Texas Tech University, Timothy Anderson, December 2008

Mountains, California and Oregon: U.S Geological Survey Open File Report 99-374.

Irwin, W., 2003, Correlation of the Klamath Mountains and Sierra Nevada: U.S. Geological Survey Open File Report 02-490, 2 sheets.

Keller, E. and Pinter, N., 2002, Active Tectonics: Earthquakes, Uplift, and Landscape, Princeton Hall, Upper Saddle River, New Jersey

Kelsey, H. and Carver, G. 1988, Late Neogene and Quaternary Tectonics Associated With Northward Growth of the San Andreas Transform Fault: Northern California, Journal of Geophysical Research, v. 93, p. 4797-4819.

Kelsey, H. and Trexler, J., 1989, Pleistocene Deformation of a Portion of the Southern Cascadia Forearc: Prairie Creek Formation, Northern California: Journal of Geophysical Research, v. 94, no. 10, p. 14,027- 14,039.

Kelsey, H., Engebretson, D., Mitchell., and Ticknor, R., 1994, Topographic form of the Coast Ranges of the Cascadia Margin in relation to coastal uplift rates and plate subduction: Journal of Geophysical Research, v. 99, p. 12245- 12255.

Kelsey, H., Ticknor, R., Bockheim, J., Mitchell, C., 1996, Quaternary upper plate deformation in coastal Oregon: GSA Bulletin, v. 108, no. 7, p. 843- 860.

Kirby, E. and Whipple K., 2001, Quantifying differential rock-uplift rates via stream profile analysis: Geology, v. 29, p. 415-418.

Kirby, E. and Whipple K., 2003, Distribution of active rock uplift along the eastern margin of the Tibetan Plateau: Inferences from bedrock channel longitudinal profiles: Journal of Geophysical Research, v. 108, v B4, 2217, doi: 10.1029/2001JB000861.

Kooi, H. and Beaumont, C., 1996, Large-scale geomorphology: Classic concepts reconciled and integrated with contemporary ideas via a surface processes: Journal of Geophysical Research, v. 101, p. 3361-3386.

Koons, P., 1994, Three-dimensional critical wedges: Tectonics and topography in oblique collisional orogens, Journal of Geophysical Research, v. 99, no. B6, p. 12301-12315.

Lock, J., Kelsey, H., Furlong, K., and Woolace, A., 2006, Late Neogene and Quaternary landscape evolution of the northern California Coast: Evidence for Mendocino triple junction tectonics: Geological Society of America Bulletin, v. 118, no. 9/10, p. 1232- 1246.

231 Texas Tech University, Timothy Anderson, December 2008

Mankinen, E., Irwin, W., and Grommé, C., 1988, Paleomagnetic results form the Shasta Bally plutonic belt in the Klamath Mountains province, northern California: Journal of Geophysical Research Letters, v. 15, p. 56-59.

Mankinen, E., Irwin, W., and Grommé S., 1989, Paleomagnetic Study of the Eastern Klamath Terrane, California, and implications for the Tectonic History of the Klamath Mountains Province: Journal of Geophysical Research, v. 94, p. 10444-10472.

Marple, R. and Talwani, P., 1993, Evidence of possible tectonic upwarping along the South Carolina coastal plain from an examination of river morphology and elevation data: Geology, v. 21, 651-654.

McCrory, P., 1989, Late Neogene geohistory analysis of the Humbold basin and its relationship to convergence of the Juan de Fuca plate: Journal of Geophysical Research, v. 94, p. 3126-3138.

McCrory, P., 1995, Reply: Evolution of a trench-slope basin within the Cascadia subduction margin: the Neogene Humboldt Basin, California, v. 43, p. 764- 769.

McLaughlin, R., Sliter, W., Frederickson, N., Harbert, and W., McCulloch, D., 1994, Plate motions recorded in tectonostratigraphic terranes of the Franciscan Complex and evolution of the Mendocino triple junction, northwestern California: U.S. Geological Survey Bulletin 1997, 60 p.

McNeill, L., Goldfinger C., Kulm, L. and Yeats, R., 2000, Tectonics of the Neogene Cascadia forearc basin: Investigations of a deformed late Miocene unconformity: Geological Society of America Bulletin, v. 112, no 8, p. 1209-1224.

McNutt, M., 1983, Influence of Plate Subduction on Isostatic Compensation in Northern California, Tectonics, v. 2, No. 4 p. 399-415.

Melnick, D. and Echtler, H., 2006, Inversion of forearc basins in south-central Chile caused by rapid glacial age trench fill: Geological Society of America, v. 34, No. 9, p. 709-712.

Merritts, D. and Vincent, K., 1989, Geomorphic response of coastal streams to low, intermediate, and high rates of uplift, Mendocino triple junction region, California, Geological Society of America Bulletin, v. 101, p. 1373-1388.

Miller, M. and Harwood, D., 1989, Paleozoic and early Mesozoic paelogeographic relations between the Klamath Mountains, northern Sierra Nevada, and western North America, Geology, p. 369-372.

232 Texas Tech University, Timothy Anderson, December 2008

Molnar, P. and England, P., 1990, Late Cenozoic uplift of mountain ranges and global climate change: Chick or egg?: Nature, v. 346, p 29-34.

Montgomery D., Balco G., and Willett, S., 2001, Climate, tectonics, and the morphology of the Andes, Geology, v. 29, no. 7 p. 579-582.

Nilsen, T. and McKee, E., 1979, Paleogene Paleogeography of the Western United States in Cenozoic Paleogeography of the Western United States, Pacific Section: Society of Economic Paleontologists and Mineralogists, p. 257- 276.

Pelletier, J., 2004, Estimate of three-dimensional flexural-isostatic response to unloading: Rock uplift due to late Cenozoic glacial erosion in the western United States: Geology, v. 32, no. 2., p. 161- 164.

Phillips, J., 2002, Erosion isostatic response, and the missing peneplains: Geomorphology, v. 45, p. 225-241.

Reneau, S., Dietrich W., Rubin M., Donahue D., and Jull A., 1989, Analysis of hillslope erosion rates using dated colluvial deposits: Journal of Geology, v. 97, p. 45-63.

Retallack, G., 2004, Late Miocene climate and life on land in Oregon within a context of Neogene global change: Paleogeography, Paleoclimatology, Paleoecology, v. 214 (1-2), p. 97-123.

Ring, U., Brandon, M., Willett, S., and Lister, G., 1999, Exhumation processes, in Ring, U., Brandon, M., Lister, G. and Willett, S. (eds) Exhumation Processes: Normal Faulting, Ductile Flow and Erosion: Geological Society, London, Special Publications, v. 154, p.1- 27.

Roe G., Montgomery D., and Hallet B., 2002, Effects of Orographic Precipitation Variations on the Concavity of Steady-State River Profiles, Geology, v. 30, p. 143-146

Schoenbohm, L., Whipple, K., Buchfiel, B., and Chen, L., 2004, Geomorphic constraints on surface uplift, exhumation, and plate growth in the Red River region, Yunnan Province, China: Geological Society of America Bulletin, v.116, p. 895-909.

Small, E. and Anderson, R., 1995, Geomorphologically Driven Late Cenozoic Rock Uplift in the Sierra Nevada, California: Science, v. 270, p. 277-280.

Seidl, M. and Dietrich, W., 1992, The Problem of Channel Erosion into Bedrock: Catena , v. 23, p.101-124.

233 Texas Tech University, Timothy Anderson, December 2008

Seidl, M.., W. Dietrich, and J. Kirchner, 1994, Longitudinal profile development into bedrock: An analysis of Hawaiian channels: Journal of Geology, v. 102, p. 457-474.

Snyder, N., Whipple, K., Tucker, G., and Merritts, D., 2000, Landscape response to tectonic forcing: DEM analysis of stream profiles in the Mendocino triple junction region, northern California: Geological Society of America Bulletin, v. 112, No. 8 p. 1250-1263, doi: 10.1130/0016-7606.

Strand, R., 1962, Geologic Map of California: Redding Sheet, Department of Conservation, Division of Mines and Geology, 1:250,000, 1 sheet.

Stock, J., and Montgomery D. 1999, Geologic constraints on bedrock river incision using the stream power law, Journal of Geophysical Research: 4983-4993.

Stock, G., Ehlers, T., Farley, K., 2006, Where does sediment come from? Quantifying catchment erosion with detrital apatite (U-Th)/He thermochronometry: Geology, v. 34, no. 9, p. 725- 728.

Stone, J., 1992, Depositional/tectonic history of the Plio-Pleistocene Crescent City/Smith River, Del Norte County, California, in Carver, G., and Aalto, K., eds, Field guide to the Late Cenozoic subduction tectonics and sedimentation of northern coastal California: AAPG, Pacific section, Guidebook, GB-71, p. 51-53.

Tippett, J. and Kamp P., 1995, Quantitative relationships between uplift and relief parameters for the Southern Alps, New Zealand, as determined by fission track analysis: Earth Surface and Processes and Landforms, v. 20, 153-175.

Vanlaningham, S., Meigs A., and Goldfinger C., 2006, The effects of rock uplift and rock resistance on river morphology in a subduction zone forearc, Oregon, USA: Earth Surface Processes and Landforms, v. 31, p. 1257-1279.

Wagner, D. and Saucedo G., 1987, Geologic map of the Weed Quadrangle, California, Department of Conservation, Division of Mines and Geology, 1:250,000, 4 sheets.

Wakabayashi, J. and Sawyer, T., 2001, Stream Incision, Tectonics, Uplift, and Evolution of Topography of the Sierra Nevada, California: Geology, v. 109, p. 539-562.

Wells, R. and Heller, P., 1988, The relative contribution of accretion, shear, and extension to Cenozoic tectonic rotation in the Pacific Northwest, Geological Society of America Bulletin, v. 100, p. 325- 338.

Wells, R., Weaver, C., and Blakely R.J., 1998, Fore-arc migration in Cascadia and its neotectonics significance: Geology, v. 26, p. 759-762. 234 Texas Tech University, Timothy Anderson, December 2008

Whipple, K., Wobus, C., Crosby, B., Kirby, E. and Sheehan, D., 2007, New Tools for Quantitative Geomorphology: Extraction and Interpretation of Stream Profiles from Digital Topographic Data: Geological Society of America Annual Meeting, Shortcourse no. 506, Denver, CO.

Wilson, D., Hey, R., and Nishimura, C., 1984, Propagation as a mechanism of reorientation of the Juan de Fuca Ridge, Journal of Geophysical research, v. 89, p. 9215- 9225.

Wilson, D., 1993, Confidence intervals for motion and deformation of the Juan de Fuca plate, Journal of Geophysical Research, v. 98, p. 16053-16071.

Wobus, C., Whipple K., Kirby E., Snyder N., Johnson J., Spyropolou, K., Crosby B., and Sheehan, D., 2006, Tectonics from topography: Procedures, promise, and pitfalls. Special Paper 398: Tectonics, Climate, and Landscape Evolution: v. 398, no.0, p. 55-74

Zandt, G. and Furlong, K., 1982, Evolution and thickness of the lithosphere beneath coastal California: Geology, v. 10, p. 376- 381.

Zhang, Y., 2005, Global tectonic and climatic control of mean elevation of continents, and Phanerozoic sea level change, Earth and Planetary Science Letters, v. 237, p. 524- 531.

Zhou, W. and Stune K., 1994, Modeling of dynamic uplift, denudation rates, and themomechanical consequences of erosion in isostatically compensated mountain belts, Journal of Geophysical Research, v. 99, p 13923-13939.

235 Texas Tech University, Timothy Anderson, December 2008

Appendix Table 1 Name of River Longitude Latitude Normalized Ksn Ksn Minimum Maximum Steepness (Low σ) (High 2 σ) Drainage Area Drainage Area 122.74552 39.2486 83 93 72 792793 9077854 Rice Fork 122.86633 39.3281 34 36 31 9077854 435166507 Bear Creek 122.87087 39.3518 71 78 64 5280671 358609548 Baechtel Creek 123.42295 39.3635 16 17 15 388000 3837500 Baechtel Creek 123.38883 39.3689 16 18 14 3897400 13799600 Baechtel Creek 123.34778 39.4242 12 17 6 15312800 176837300 Eel River 122.87044 39.4294 88 100 76 294089184 372771936 Eel River 122.96274 39.4555 53 63 42 9208656 496062288 Eel River 123.17134 39.4719 31 34 28 161970192 1227837888 Tomki Creek 123.26562 39.4778 19 19 18 2976624 160576128 Davis Creeek 123.36073 39.4837 21 23 19 824074 508021805 Willits Creek 123.38358 39.5027 19 20 18 101954 777600091 Eel River 122.99317 39.5033 66 77 56 3428640 9120960 Eel River 123.01489 39.5175 58 67 48 1095552 2746080 Sherwood Creek 123.46388 39.5293 7 10 4 84017 23886431 Eel River 123.22858 39.5313 101 109 93 3319009 12134827678 Eel River 122.78966 39.5405 53 55 52 321696 294090048 123.39275 39.5487 45 55 35 26826958 943604310 Eel River 123.26962 39.5582 33 39 28 1228163040 1358218224 Outlet Creek 123.39999 39.5785 20 24 16 33541000 387455200 Elk Creek 123.06207 39.5897 113 121 105 2030255 1116560182 Long Valley Creek 123.45211 39.5964 32 44 20 17528800 33060300 Eden Creek 123.19235 39.6202 27 35 18 4282288 52712703 Baldy Creek 122.98515 39.6346 68 72 64 293611 1772633 Long Valley Creek 123.44832 39.6354 26 37 16 529400 12521400 Baldy Creek 122.97126 39.6502 77 87 67 3264466 8439859 Eden Creek 123.18171 39.6579 137 172 103 54772500 57977400 Eden Creek 123.15415 39.6661 123 145 102 57720500 67853000

236 Texas Tech University, Timothy Anderson, December 2008

Table 1 (Continued) Name of River Longitude Latitude Concavity Concavity Fitted Smoothing Contour within 2 σ Steepness Window Interval Rice Fork 122.74552 39.2486 1.4101 0.2017 97155547.01 250 12.192 Rice Fork 122.86633 39.3281 0.78127 0.13365 12563.3905 250 12.192 Bear Creek 122.87087 39.3518 1.1471 0.11191 11137274.32 250 12.192 Baechtel Creek 123.42295 39.3635 0.5777 0.19386 110.8968 250 12.192 Baechtel Creek 123.38883 39.3689 0.70915 0.47663 1161.4639 250 12.192 Baechtel Creek 123.34778 39.4242 1.0697 0.61576 889608.525 250 12.192 Eel River 122.87044 39.4294 7.1454 4.3046 8.29E+58 250 12.192 Eel River 122.96274 39.4555 0.98996 0.19576 606638.7273 250 12.192 Eel River 123.17134 39.4719 0.52811 0.16363 177.7151 250 12.192 Tomki Creek 123.26562 39.4778 0.53446 0.12081 87.7683 250 12.192 Davis Creeek 123.36073 39.4837 0.6472 0.09905 853.3315 250 12.192 Willits Creek 123.38358 39.5027 0.54191 0.059507 114.1633 250 12.192 Eel River 122.99317 39.5033 1.69 0.47488 15970064778 250 12.192 Eel River 123.01489 39.5175 1.6022 0.37288 920124512.5 250 12.192 Sherwood Creek 123.46388 39.5293 0.82893 0.096076 2567.6658 250 12.192 Eel River 123.22858 39.5313 0.6874 0.058703 6606.544 250 12.192 Eel River 122.78966 39.5405 0.54117 0.03831 265.7251 250 12.192 Outlet Creek 123.39275 39.5487 1.1448 0.15233 23422156.81 250 12.192 Eel River 123.26962 39.5582 9.0268 7.3127 4.99E+79 250 12.192 Outlet Creek 123.39999 39.5785 0.54247 0.40242 142.3069 250 12.192 Elk Creek 123.06207 39.5897 0.82772 0.054528 60262.8157 250 12.192 Long Valley Creek 123.45211 39.5964 4.0264 1.1817 8.93E+27 250 12.192 Eden Creek 123.19235 39.6202 1.3931 0.26179 144282026.6 250 12.192 Baldy Creek 122.98515 39.6346 0.87307 0.090209 21218.3622 250 12.192 Long Valley Creek 123.44832 39.6354 1.2194 0.23681 1776657.512 250 12.192 Baldy Creek 122.97126 39.6502 1.9289 0.61723 7.89684E+11 250 12.192 Eden Creek 123.18171 39.6579 -39.9488 19.5918 0 250 12.192 Eden Creek 123.15415 39.6661 13.8078 2.383 2.45E+106 250 12.192

237 Texas Tech University, Timothy Anderson, December 2008

Table 1 (Continued) Name of River Longitude Latitude Normalized Ksn Ksn Minimum Maximum Steepness (Low σ) (High 2 σ) Drainage Area Drainage Area Cold Creek 122.82016 39.6823 17 19 15 19460 21092363 Cold Creek 122.89509 39.6947 265 278 252 42784700 51426800 Estell Creek 122.9372 39.7024 74 76 73 350400 412689100 Cold Creek 122.86659 39.7041 183 199 167 33888700 39168100 Baldy Creek 122.97962 39.7452 118 124 113 16633915 7987191678 Cold Creek 122.98903 39.7608 96 98 93 51761500 412250900 Town Creek 123.23546 39.7813 47 54 40 1962536 4641588834 Short Creek 123.19552 39.8035 24 29 18 7982600 255418500 Mill Creek 123.23752 39.8052 32 39 25 13499700 250943500 Murphy Creek 123.14998 39.8583 104 112 96 1306237 431876045 Williams Creek 123.13171 39.8626 132 144 121 6219257 131736263 Short Creek 123.26839 39.8712 36 38 34 748100 2767700 Mill Creek 123.30253 39.8862 64 75 52 2863600 10637100 Mill Creek 123.29788 39.9002 29 30 27 202166 3867916 Middle Fork Eel River 123.01582 39.9067 181 221 141 413908300 469549500 Williams Creek 123.11974 39.9207 145 156 134 1897076 6219257 Hulls Creek 123.20057 39.9233 56 62 51 8158348 66841477 Williams Creek 123.10525 39.927 72 78 66 732600 2067500 Middle Fork Eel River 123.02242 39.9499 158 182 134 322882800 413908300 Hulls Creek 123.22139 39.9545 103 110 96 1944567 14059251 Hulls Creek 123.25636 39.9758 161 189 133 90439335 200070081 Middle Fork Eel River 123.0789 39.9776 126 153 99 257012100 282033800 Hulls Creek 123.25124 39.9823 88 105 72 14054391 200273877 Middle Fork Eel River 123.10567 40.0003 127 148 106 230795600 255419300 Middle Fork Eel River 123.09897 40.0295 131 154 108 83889500 227250100 North Fork Eel 123.28749 40.0392 51 63 39 282890151 369319257 North Fork Eel 123.28792 40.0397 54 65 43 282536343 369361620

238 Texas Tech University, Timothy Anderson, December 2008

Table 1 (Continued) Name of River Longitude Latitude Concavity Concavity Fitted Smoothing Contour within 2 σ Steepness Window Interval Cold Creek 122.82016 39.6823 0.30555 0.053085 2.3614 250 12.192 Cold Creek 122.89509 39.6947 6.0171 2.2269 1.68E+45 250 12.192 Estell Creek 122.9372 39.7024 0.47523 0.038702 136.9788 250 12.192 Cold Creek 122.86659 39.7041 -5.6333 3.6022 0 250 12.192 Baldy Creek 122.97962 39.7452 0.76775 0.081009 46705.0077 250 12.192 Cold Creek 122.98903 39.7608 0.94781 0.20495 1597515.488 250 12.192 Town Creek 123.23546 39.7813 1.0951 0.12741 1878753.838 250 12.192 Short Creek 123.19552 39.8035 1.252 0.25299 33562116.24 250 12.192 Mill Creek 123.23752 39.8052 1.4152 0.34306 962939302.8 250 12.192 Murphy Creek 123.14998 39.8583 0.8366 0.086316 46600.5263 250 12.192 Williams Creek 123.13171 39.8626 1.2312 0.12474 69988941.25 250 12.192 Short Creek 123.26839 39.8712 0.63001 0.20035 474.0376 250 12.192 Mill Creek 123.30253 39.8862 -0.73863 0.30526 0 250 12.192 Mill Creek 123.29788 39.9002 0.54468 0.19142 123.651 250 12.192 Middle Fork Eel River 123.01582 39.9067 -12.1041 4.2494 0 250 12.192 Williams Creek 123.11974 39.9207 -1.0701 0.23411 0 250 12.192 Hulls Creek 123.20057 39.9233 1.246 0.28892 52105813.84 250 12.192 Williams Creek 123.10525 39.927 1.2811 0.34042 8309747.872 250 12.192 Middle Fork Eel River 123.02242 39.9499 -9.743 6.074 0 250 12.192 Hulls Creek 123.22139 39.9545 0.83067 0.13894 35999.378 250 12.192 Hulls Creek 123.25636 39.9758 2.7054 0.70406 2.41E+20 250 12.192 Middle Fork Eel River 123.0789 39.9776 14.1531 10.2955 7.25E+117 250 12.192 Hulls Creek 123.25124 39.9823 0.88547 0.26093 196113.1151 250 12.192 Middle Fork Eel River 123.10567 40.0003 9.5607 3.8145 3.28E+78 250 12.192 Middle Fork Eel River 123.09897 40.0295 1.4822 0.44425 33187476343 250 12.192 North Fork Eel 123.28749 40.0392 6.15 3.3667 2.39E+50 250 12.192 North Fork Eel 123.28792 40.0397 5.1873 2.2727 1.52E+42 250 12.192

239 Texas Tech University, Timothy Anderson, December 2008

Table 1 (Continued) Name of River Longitude Latitude Normalized Ksn Ksn Minimum Maximum Steepness (Low σ) (High 2 σ) Drainage Area Drainage Area Hulls Creek_Unnnamed 123.19673 40.0483 80 86 74 398445 8158348 Red Mountain Creek 123.28133 40.0493 105 115 94 26792643 3765958123 Red Mountain Creek 123.20814 40.085 57 63 51 154467 10381527 Salt Creek 123.35043 40.0875 27 28 26 278964 277775892 Middle Fork Eel River 123.01108 40.1029 41 42 40 619294 46023901 North Fork Eel 123.34874 40.1362 34 36 31 20713563 278383635 Bar Creek 123.35737 40.1475 47 49 45 303732 1239895872 North Fork Eel 123.35569 40.155 37 42 32 14557806 265462353 Bluff Creek 123.42114 40.1582 33 36 30 3051432 20709918 Mad 123.16245 40.1854 35 36 33 162841 10314433 Mad 123.2877 40.2277 30 34 26 13102400 309658000 Soldier Creek 123.37082 40.236 44 47 41 814779 11514069 Little Van Duzen River 123.55398 40.3098 87 92 83 319154 2013724 Larabee Creek 123.6899 40.3323 21 22 19 372308 123094716 Little Van Duzen River 123.59071 40.3521 57 62 52 2511886 29648314 Van Duzen River 123.48335 40.361 25 27 24 81658 103752842 South Fork Trinity 123.28806 40.3671 60 62 59 2567136 12889504441 Van Duzen River 123.50766 40.3739 44 47 40 343558 107646521 E. Fork South Fork Trinity River123.28415 40.3757 73 75 70 1722743 4842162612 Hayfork Creek 123.07299 40.3757 40 42 39 2396600 111091600 Mad 123.47485 40.4153 22 30 14 309658000 366716400 Browns Creek 122.96528 40.433 62 64 59 2633438 15033579 Little Van Duzen River 123.64134 40.4334 40 62 18 46096614 52724601 Larabee Creek 124.01133 40.4508 18 29 8 188118369 467851302 Rattlesnake Creek 123.36567 40.4531 72 74 71 389539 25671356563 Salt Creek 123.15317 40.4664 52 57 47 3077436 206519293 Browns Creek 122.93367 40.4736 60 65 55 18528102 54281097

240 Texas Tech University, Timothy Anderson, December 2008

Table 1 (Continued) Name of River Longitude Latitude Concavity Concavity Fitted Smoothing Contour within 2 σ Steepness Window Interval Hulls Creek_Unnnamed 123.19673 40.0483 0.81253 0.087095 13838.5067 250 12.192 Red Mountain Creek 123.28133 40.0493 0.93926 0.19109 769708.8071 250 12.192 Red Mountain Creek 123.20814 40.085 0.86484 0.097257 18519.8548 250 12.192 Salt Creek 123.35043 40.0875 0.44329 0.068467 30.946 250 12.192 Middle Fork Eel River 123.01108 40.1029 0.53677 0.060068 168.9174 250 12.192 North Fork Eel 123.34874 40.1362 0.76415 0.26626 13729.0249 250 12.192 Bar Creek 123.35737 40.1475 0.56322 0.046966 381.2124 250 12.192 North Fork Eel 123.35569 40.155 0.91618 0.23414 247652.7665 250 12.192 Bluff Creek 123.42114 40.1582 1.1375 0.36252 2551657.543 250 12.192 Mad 123.16245 40.1854 0.50574 0.1268 82.0127 400 13.5 Mad 123.2877 40.2277 0.9208 0.17981 142616.5608 400 13.5 Soldier Creek 123.37082 40.236 0.91569 0.14599 51736.5731 250 12.192 Little Van Duzen River 123.55398 40.3098 0.82689 0.16768 14697.1856 250 12.192 Larabee Creek 123.6899 40.3323 0.47966 0.092182 47.1512 250 12.192 Little Van Duzen River 123.59071 40.3521 1.0299 0.13066 629041.6363 250 12.192 Van Duzen River 123.48335 40.361 0.59066 0.083448 325.6169 250 12.192 South Fork Trinity 123.28806 40.3671 0.53161 0.077935 359.3715 250 12.192 Van Duzen River 123.50766 40.3739 0.72067 0.049866 3743.9643 250 12.192 E. Fork South Fork Trinity River123.28415 40.3757 0.58867 0.051682 1093.3126 250 12.192 Hayfork Creek 123.07299 40.3757 0.66434 0.10887 1790.298 250 12.192 Mad 123.47485 40.4153 14.5366 23.4011 6.37E+121 400 13.5 Browns Creek 122.96528 40.433 0.70845 0.15543 3522.3762 250 12.192 Little Van Duzen River 123.64134 40.4334 12.9577 5.2986 8.63E+97 250 12.192 Larabee Creek 124.01133 40.4508 5.6037 1.3391 6.52E+44 250 12.192 Rattlesnake Creek 123.36567 40.4531 0.45268 0.042738 88.1839 250 12.192 Salt Creek 123.15317 40.4664 0.97648 0.098703 437585.1556 250 12.192 Browns Creek 122.93367 40.4736 1.0618 0.39608 2766717.646 250 12.192

241 Texas Tech University, Timothy Anderson, December 2008

Table 1 (Continued) Name of River Longitude Latitude Normalized Ksn Ksn Minimum Maximum Steepness (Low σ) (High 2 σ) Drainage Area Drainage Area Little Van Duzen River 123.66688 40.4772 129 138 121 78757353 210396933 Van Duzen River 123.6669 40.4851 158 201 116 121857777 210295845 Van Duzen River 123.9548 40.4905 28 31 24 302205897 408188808 Van Duzen River 123.90836 40.4931 46 52 39 238074957 408154140 Van Duzen River 123.78172 40.5069 89 113 64 238453713 291509118 Van Duzen River 123.7043 40.5079 193 224 163 212556231 221775570 Vam Duzen river 123.70476 40.5081 182 210 154 212631966 221769414 Indian Valley Creek 123.32763 40.5154 18 20 17 252101 24312341 Plummer Creek 123.42657 40.5252 141 155 127 15304700 882294200 Reading Creek 122.80738 40.5267 62 63 61 449793 12470841 Van Duzen river 123.73446 40.5269 92 185 -1 222867207 237595032 Van Duzen river 123.73472 40.5271 92 185 -1 222874335 237603699 Hayfork Creek 123.13466 40.5309 42 45 38 113695300 731896600 Indian Valley Creek 123.39042 40.5434 169 181 157 35819200 41889300 Owl Creek 123.88219 40.5475 22 24 21 76302 10921554 North Fork East fork Hayfork122.99113 Creek 40.5507 49 51 48 1490144 708587849 Indian Valley Creek 123.41075 40.5513 172 190 154 42254300 51933300 Yager Creek 124.05788 40.5536 41 59 24 181917090 195771249 Lawrence Creek 124.05753 40.5537 41 59 24 181797291 603983466 Yager Creek 124.05499 40.5557 45 61 28 322830927 348538140 Yager Creek 124.05133 40.5567 49 64 35 321199344 348637932 Mad 123.64258 40.5626 44 53 34 407290400 587023600 Owl Creek 124.00383 40.5638 72 84 60 11553111 603979011 Little Creek 122.98202 40.567 76 78 75 78570 2459970 East Fork Browns Creek 122.93011 40.5675 63 65 62 126068 1192501002 North Fork East Fork Hayfork123.20084 Creek 40.5687 51 53 50 1545170 1955863683 Tule Creek 123.25607 40.5717 57 63 51 17445800 799224400

242 Texas Tech University, Timothy Anderson, December 2008

Table 1 (Continued) Name of River Longitude Latitude Concavity Concavity Fitted Smoothing Contour within 2 σ Steepness Window Interval Little Van Duzen River 123.66688 40.4772 0.64389 0.31652 4763.6295 250 12.192 Van Duzen River 123.6669 40.4851 2.4582 1.7361 5.81E+18 250 12.192 Van Duzen River 123.9548 40.4905 6.5249 3.0898 2.99E+53 250 12.192 Van Duzen River 123.90836 40.4931 4.7726 1.2063 2.87E+38 250 12.192 Van Duzen River 123.78172 40.5069 12.4476 4.1756 7.41E+102 250 12.192 Van Duzen River 123.7043 40.5079 31.8138 7.0774 6.12E+263 250 12.192 Vam Duzen river 123.70476 40.5081 33.6926 7.8784 2.85E+279 250 12.192 Indian Valley Creek 123.32763 40.5154 0.56789 0.11834 130.2104 250 12.192 Plummer Creek 123.42657 40.5252 0.79043 0.12348 79694.0614 250 12.192 Reading Creek 122.80738 40.5267 0.39102 0.077955 26.7032 250 12.192 Van Duzen river 123.73446 40.5269 24.0096 33.268 1.38E+199 250 12.192 Van Duzen river 123.73472 40.5271 24.0096 33.268 1.38E+199 250 12.192 Hayfork Creek 123.13466 40.5309 1.1382 0.31648 29934008.94 250 12.192 Indian Valley Creek 123.39042 40.5434 7.2552 2.4543 7.50E+53 250 12.192 Owl Creek 123.88219 40.5475 0.59581 0.086466 250.7417 250 12.192 North Fork East fork Hayfork122.99113 Creek 40.5507 0.56659 0.044586 389.2101 250 12.192 Indian Valley Creek 123.41075 40.5513 4.4821 1.6992 1.62E+33 250 12.192 Yager Creek 124.05788 40.5536 17.6469 4.6244 1.04E+144 250 12.192 Lawrence Creek 124.05753 40.5537 17.6469 4.6244 1.04E+144 250 12.192 Yager Creek 124.05499 40.5557 16.0507 5.2479 4.94E+134 250 12.192 Yager Creek 124.05133 40.5567 14.545 5.6285 7.01E+121 250 12.192 Mad 123.64258 40.5626 4.2577 3.0652 7.76E+34 400 13.5 Owl Creek 124.00383 40.5638 1.1346 0.1388 19333509.35 250 12.192 Little Creek 122.98202 40.567 0.4263 0.069244 55.1284 250 12.192 East Fork Browns Creek 122.93011 40.5675 0.51846 0.037323 200.3193 250 12.192 North Fork East Fork Hayfork123.20084 Creek 40.5687 0.37792 0.060849 18.8682 250 12.192 Tule Creek 123.25607 40.5717 0.62895 0.14638 2338.7355 250 12.192

243 Texas Tech University, Timothy Anderson, December 2008

Table 1 (Continued) Name of River Longitude Latitude Normalized Ksn Ksn Minimum Maximum Steepness (Low σ) (High 2 σ) Drainage Area Drainage Area Indian Creek 123.76042 40.5733 88 99 76 3030271 10585012 Indian Creek 123.95997 40.574 25 30 20 117977229 307516743 Yager Creek 123.96144 40.5741 26 31 21 66140955 177695613 Yager Creek 123.76393 40.5745 65 76 54 2297484 6217479 Browns Creek 122.93931 40.5758 56 61 52 56193831 181133820 Indian Valley Creek 123.43826 40.5819 153 181 125 54838800 880843000 Reading Creek 122.89648 40.5875 65 70 61 17981514 457332073 Big Creek 123.18311 40.5998 54 56 53 834180 761885597 Indian Creek 123.80862 40.6031 52 58 46 12317852 67807186 Yager Creek 123.81255 40.604 52 58 46 6990786 44023581 Lawrence Creek 123.99242 40.6059 48 57 38 63344997 318336642 Indian Creek 122.84166 40.6065 97 99 94 1008930 309276905 Indian Creek 123.8987 40.6118 47 60 35 79242381 114201171 Yager Creek 123.89811 40.6119 53 67 39 44610588 63947151 Lawrence Creek 123.98785 40.6164 62 75 49 35577144 170414523 Little Creek 122.97113 40.6182 75 79 72 2943153 377976491 Hayfork Creek 123.39567 40.6295 118 142 95 93123700 978127600 Grass Valley Creek 122.78254 40.6373 58 60 56 359737 965063709 South Fork Elk Creek 124.03642 40.6411 45 46 43 292000 9347300 Little Grass Valley Creek 122.75619 40.6481 40 42 37 865728 26182926 Corral Creek 123.37759 40.6509 127 143 110 75194600 85175800 Mad 123.82083 40.6571 86 113 60 681771200 733652000 South Fork Elk Creek 124.06067 40.6614 44 59 30 9414200 18163400 Eltapom Creek 123.48137 40.6631 180 203 157 48068300 49950600 Lawrence Creek 123.96202 40.6711 45 56 34 25196427 63371241 Eltapom Creek 123.46931 40.6729 220 245 194 44881200 47396700 North Fork Elk Creek 123.99117 40.6756 40 45 35 2586100 9141000

244 Texas Tech University, Timothy Anderson, December 2008

Table 1 (Continued) Name of River Longitude Latitude Concavity Concavity Fitted Smoothing Contour within 2 σ Steepness Window Interval Indian Creek 123.76042 40.5733 1.2555 0.299 21248191.19 250 12.192 Indian Creek 123.95997 40.574 1.6576 1.1744 2.86653E+11 250 12.192 Yager Creek 123.96144 40.5741 1.6425 1.1652 1.13182E+11 250 12.192 Yager Creek 123.76393 40.5745 1.1638 0.50207 3822762.193 250 12.192 Browns Creek 122.93931 40.5758 1.6317 0.26742 1.97E+11 250 12.192 Indian Valley Creek 123.43826 40.5819 0.92808 0.30254 1657259.714 250 12.192 Reading Creek 122.89648 40.5875 1.2877 0.28986 194059659.5 250 12.192 Big Creek 123.18311 40.5998 0.54551 0.040439 303.1911 250 12.192 Indian Creek 123.80862 40.6031 1.5495 0.16436 8897070105 250 12.192 Yager Creek 123.81255 40.604 1.3365 0.21676 152440032.1 250 12.192 Lawrence Creek 123.99242 40.6059 1.343 0.52108 849160342.3 250 12.192 Indian Creek 122.84166 40.6065 0.62522 0.049606 1771.3739 250 12.192 Indian Creek 123.8987 40.6118 6.1488 2.0801 2.01E+47 250 12.192 Yager Creek 123.89811 40.6119 6.3293 1.7851 1.95E+47 250 12.192 Lawrence Creek 123.98785 40.6164 3.5498 0.76601 3.19E+25 250 12.192 Little Creek 122.97113 40.6182 0.64307 0.083143 1961.1774 250 12.192 Hayfork Creek 123.39567 40.6295 0.71766 0.46214 40293.9514 250 12.192 Grass Valley Creek 122.78254 40.6373 0.54657 0.064599 356.258 250 12.192 South Fork Elk Creek 124.03642 40.6411 0.49222 0.092529 94.3829 250 12.192 Little Grass Valley Creek 122.75619 40.6481 0.73639 0.12175 3848.4023 250 12.192 Corral Creek 123.37759 40.6509 -13.9249 5.3029 0 250 12.192 Mad 123.82083 40.6571 27.9896 12.2959 5.33E+245 400 13.5 South Fork Elk Creek 124.06067 40.6614 2.3599 0.83813 1.90E+15 250 12.192 Eltapom Creek 123.48137 40.6631 27.4413 4.8725 6.44E+209 250 12.192 Lawrence Creek 123.96202 40.6711 1.796 0.863 7.85058E+11 250 12.192 Eltapom Creek 123.46931 40.6729 20.9269 5.8878 1.97E+159 250 12.192 North Fork Elk Creek 123.99117 40.6756 1.1711 0.71716 2571267.753 250 12.192

245 Texas Tech University, Timothy Anderson, December 2008

Table 1 (Continued) Name of River Longitude Latitude Normalized Ksn Ksn Minimum Maximum Steepness (Low σ) (High 2 σ) Drainage Area Drainage Area Lawrence Creek 123.93207 40.6819 46 48 44 49896 34992567 Eltapom Creek 123.46129 40.6823 157 167 147 38232800 41322500 Lawrence Creek 123.91235 40.688 47 50 44 205474 24368639 Eltapom Creek 123.46201 40.6904 99 120 79 26547000 27712600 South Fork Elk Creek 124.13706 40.6949 21 29 12 18233100 121875800 Corral Creek 123.28484 40.6953 27 28 26 271063 12657916 Eltapom Creek 123.45364 40.6993 108 117 100 17605900 26315700 Eltapom Creek 123.43228 40.7005 70 72 69 6644900 17354100 North Fork Elk Creek 124.12379 40.7072 30 37 23 8993258 199314627 Deadwood Creek 122.7568 40.7174 68 69 67 445296 43744821 Grouse Creek 123.62069 40.7222 106 107 104 2486716 233234741 West Weaver Creek 122.95807 40.7296 98 100 95 478120 752398423 East Weaver Creek 122.93777 40.7359 111 116 106 1545928 249852326 Freshwater Creek 123.96527 40.736 58 62 53 908919 4189091 Freshwater Creek 123.98291 40.7418 116 130 102 4029900 7988200 Mosquito Creek 123.617 40.7442 87 88 86 1195459 4873856783 Little Browns Creek 122.89943 40.7457 40 41 40 939665 46969363 Mad 123.88402 40.7511 39 57 20 747097600 923976400 Ryan Creek 124.12476 40.7661 8 10 6 460887 1000000000 Freshwater Creek 124.07659 40.7691 33 42 23 8261344 427894206 Little Freshwater Creek 124.09212 40.7736 13 16 9 357269 175484851 Freshwater Creek_Unnamed 124.08527 40.7824 26 33 20 735122 2769467757 Rush Creek 122.87526 40.787 90 98 83 5762591 102702972 Big East Fork Canyon Creek 123.04343 40.8089 132 137 128 1659882 2265751912 East Fork Big French Creek_unnamed123.28011 40.8223 114 115 114 347169 309276905 Canyon Creek 123.04638 40.8223 104 110 98 36619007 298471917 Rush Creek_unnamed 122.95223 40.8323 84 87 82 86022 5514318

246 Texas Tech University, Timothy Anderson, December 2008

Table 1 (Continued) Name of River Longitude Latitude Concavity Concavity Fitted Smoothing Contour within 2 σ Steepness Window Interval Lawrence Creek 123.93207 40.6819 0.5352 0.055634 183.3138 250 12.192 Eltapom Creek 123.46129 40.6823 2.7581 2.6685 5.68E+19 250 12.192 Lawrence Creek 123.91235 40.688 0.67952 0.047214 1333.4542 250 12.192 Eltapom Creek 123.46201 40.6904 11.9384 9.584 2.89E+87 250 12.192 South Fork Elk Creek 124.13706 40.6949 1.9381 0.54547 4.62031E+12 250 12.192 Corral Creek 123.28484 40.6953 0.41052 0.064874 15.8486 250 12.192 Eltapom Creek 123.45364 40.6993 0.63284 0.91118 2487.2985 250 12.192 Eltapom Creek 123.43228 40.7005 0.54031 0.28772 328.4287 250 12.192 North Fork Elk Creek 124.12379 40.7072 1.3812 0.30746 253707267.9 250 12.192 Deadwood Creek 122.7568 40.7174 0.45786 0.062204 84.3527 250 12.192 Grouse Creek 123.62069 40.7222 0.44028 0.068133 95.5298 250 12.192 West Weaver Creek 122.95807 40.7296 0.66707 0.033203 2904.6021 250 12.192 East Weaver Creek 122.93777 40.7359 0.84597 0.04752 70577.4168 250 12.192 Freshwater Creek 123.96527 40.736 0.84617 0.35323 17296.2155 250 12.192 Freshwater Creek 123.98291 40.7418 -3.0172 0.83553 0 250 12.192 Mosquito Creek 123.617 40.7442 0.42867 0.058195 68.7553 250 12.192 Little Browns Creek 122.89943 40.7457 0.54149 0.059017 188.9721 250 12.192 Mad 123.88402 40.7511 7.8181 5.9015 2.29E+67 400 13.5 Ryan Creek 124.12476 40.7661 0.64297 0.20886 241.722 250 12.192 Freshwater Creek 124.07659 40.7691 2.0656 0.35912 3.85385E+13 250 12.192 Little Freshwater Creek 124.09212 40.7736 0.63575 0.23829 351.6202 250 12.192 Freshwater Creek_Unnamed 124.08527 40.7824 0.93736 0.14451 69895.9263 250 12.192 Rush Creek 122.87526 40.787 1.3441 0.065055 280715143.3 250 12.192 Big East Fork Canyon Creek 123.04343 40.8089 0.67114 0.034841 5049.7851 250 12.192 East Fork Big French Creek_unnamed123.28011 40.8223 0.46257 0.02319 142.9753 250 12.192 Canyon Creek 123.04638 40.8223 1.3755 0.084337 2132670058 250 12.192 Rush Creek_unnamed 122.95223 40.8323 0.31908 0.047444 14.8397 250 12.192

247 Texas Tech University, Timothy Anderson, December 2008

Table 1 (Continued) Name of River Longitude Latitude Normalized Ksn Ksn Minimum Maximum Steepness (Low σ) (High 2 σ) Drainage Area Drainage Area East Fork Big French Creek 123.27162 40.8359 97 99 96 815074 2612091687 Big French Creek 123.2735 40.8487 134 136 132 372759 700744738 Mad 123.98385 40.8536 25 31 19 926690000 1259604400 East Branch East Fork North123.11634 Fork Trinity 40.8635 River 116 121 111 665368 375450497 Big Creek 123.37696 40.8653 106 108 104 461416 163062919 New 123.45394 40.8679 139 145 133 14983947 613272708 Backbone Creek 123.14768 40.8713 123 126 120 1071078 3912418718 Whites Creek 123.15974 40.8726 120 123 118 1544781 7029403755 East Fork North Fork Trinity123.12187 River 40.8746 120 125 114 2579472 144888460 East Fork Willow Creek 123.69774 40.8765 86 88 84 759117 69454253 China Creek 123.41949 40.8818 85 86 83 335040 1282649831 Trinity River 122.76903 40.8825 90 110 70 37973171 133893830 Willow Creek 123.7639 40.8951 33 36 31 103620 7133158 Stuart Fork 122.92836 40.8961 117 121 113 20740860 168074676 Stuart Fork 122.9296 40.8979 119 124 115 34570395 167962410 Trinity River 122.7854 40.9009 157 163 150 7163591 138949549 Rattlesnake Creek 123.14911 40.9023 100 102 98 8394273 397754226 Davis Creek 122.77709 40.9083 102 105 99 555222 133893830 Willow Creek 123.72335 40.9091 104 112 96 9823766 71968567 Bell Creek 123.45166 40.9163 76 81 70 2965896 6931251 East New River_unnamed 123.37195 40.9164 116 119 114 906836 807859632 North Fork Trinity 123.15379 40.921 77 78 75 413335 626926604 East Fork Stuart Fork 122.86344 40.9217 100 102 98 92664 4165911 Bell Creek 123.46351 40.9319 93 97 89 243277 1982884 Hobel Creek 122.75394 40.9333 28 29 27 978804 25432218 South Fork East New River 123.3661 40.9369 122 126 118 1439797 1152858648 Creek 122.87595 40.9521 54 61 48 369321 1957720

248 Texas Tech University, Timothy Anderson, December 2008

Table 1 (Continued) Name of River Longitude Latitude Concavity Concavity Fitted Smoothing Contour within 2 σ Steepness Window Interval East Fork Big French Creek 123.27162 40.8359 0.37772 0.034104 30.8881 250 12.192 Big French Creek 123.2735 40.8487 0.52489 0.023644 464.0041 250 12.192 Mad 123.98385 40.8536 4.6165 5.1892 1.28E+39 400 13.5 East Branch East Fork North123.11634 Fork Trinity 40.8635 River 0.73103 0.026047 10907.9295 250 12.192 Big Creek 123.37696 40.8653 0.31495 0.043886 14.1153 250 12.192 New 123.45394 40.8679 0.38531 0.13231 58.1892 250 12.192 Backbone Creek 123.14768 40.8713 0.53933 0.030826 615.1857 250 12.192 Whites Creek 123.15974 40.8726 0.51813 0.037381 426.9008 250 12.192 East Fork North Fork Trinity123.12187 River 40.8746 0.88774 0.066305 185460.5525 250 12.192 East Fork Willow Creek 123.69774 40.8765 0.52222 0.060719 259.1277 250 12.192 China Creek 123.41949 40.8818 0.3672 0.04849 25.7525 250 12.192 Trinity River 122.76903 40.8825 2.4264 0.81187 1.79E+17 250 12.192 Willow Creek 123.7639 40.8951 0.66767 0.11675 726.0117 250 12.192 Stuart Fork 122.92836 40.8961 0.88343 0.10652 336946.9385 250 12.192 Stuart Fork 122.9296 40.8979 1.0965 0.1234 17324585.32 250 12.192 Trinity River 122.7854 40.9009 1.0992 0.13213 7805307.82 250 12.192 Rattlesnake Creek 123.14911 40.9023 0.58193 0.060765 1291.9547 250 12.192 Davis Creek 122.77709 40.9083 0.64191 0.047928 1886.5384 250 12.192 Willow Creek 123.72335 40.9091 0.83739 0.38526 66036.1911 250 12.192 Bell Creek 123.45166 40.9163 1.8146 0.38161 97209404710 250 12.192 East New River_unnamed 123.37195 40.9164 0.52895 0.037855 507.3874 250 12.192 North Fork Trinity 123.15379 40.921 0.42497 0.037901 58.9794 250 12.192 East Fork Stuart Fork 122.86344 40.9217 0.42959 0.075611 77.8513 250 12.192 Bell Creek 123.46351 40.9319 0.78857 0.10172 8659.0831 250 12.192 Hobel Creek 122.75394 40.9333 0.48973 0.12381 58.1486 250 12.192 South Fork East New River 123.3661 40.9369 0.54593 0.041488 725.7139 250 12.192 Granite Creek 122.87595 40.9521 0.99832 0.26861 95520.6726 250 12.192

249 Texas Tech University, Timothy Anderson, December 2008

Table 1 (Continued) Name of River Longitude Latitude Normalized Ksn Ksn Minimum Maximum Steepness (Low σ) (High 2 σ) Drainage Area Drainage Area East New River 123.34855 40.9535 101 102 99 2129047 807859632 Granite Creek 122.84889 40.9546 64 96 32 5175171 6505596 Granite Creek 122.85759 40.9547 110 135 85 2740473 5157270 Three Creek 123.72768 40.9552 77 79 76 400236 5971202 Slide Creek 123.37588 40.9625 99 102 96 3380318 2265751912 Granite Creek 122.83708 40.9627 111 122 100 6567237 10832292 Cedar Creek 123.53395 40.9634 92 94 90 323335 187988496 Deer Creek 122.91578 40.966 94 98 91 404919 20794158 Granite Creek 122.77073 40.9678 162 167 158 11194848 142135074 Groves Prairie Creek 123.5478 40.9695 120 122 118 3262222 508819453 Swift Creek 122.76959 40.9714 156 160 151 43338807 142493661 Stuart Fork 122.95772 40.9766 60 82 39 17394750 32270238 Rattlesnake 123.07824 40.9803 149 160 138 2269539 6967539 Eagle Creek 123.3725 40.9823 81 82 80 844580 898810120 Canyon Creek 123.03813 40.9866 103 123 83 939665 3038264 East Fork Horse Linto Creek 123.45368 40.9886 106 114 97 3262222 23898926 Eightmile Creek 123.39412 40.9991 83 83 82 189667 1830409506 Sturart Fork 122.97601 41.0003 81 89 74 7242534 16436844 North Fork Swift Creek 122.73243 41.0032 125 134 116 7664382 154582263 Horse Linto Creek 123.45314 41.0051 69 84 55 18308511 22465188 Pine Creek 123.7678 41.0057 46 50 41 1822000 6121100 Swift Creek 122.87335 41.0062 59 62 56 515553 29295493 Pine Creek 123.77855 41.0067 17 20 15 370927 2308319 Horse Linto Creek 123.53814 41.0096 126 131 120 23063986 441354587 Supply Creek 123.7359 41.0111 89 93 84 5265810 22569921 Horse Linto Creek 123.55052 41.0125 106 112 101 26589516 288044415 Horse Linto Creek_Upstream 123.55482 41.0176 169 182 156 8223523 566103387

250 Texas Tech University, Timothy Anderson, December 2008

Table 1 (Continued) Name of River Longitude Latitude Concavity Concavity Fitted Smoothing Contour within 2 σ Steepness Window Interval East New River 123.34855 40.9535 0.51017 0.045644 333.1541 250 12.192 Granite Creek 122.84889 40.9546 6.5402 4.2829 1.45E+43 250 12.192 Granite Creek 122.85759 40.9547 2.6136 1.2438 1.38E+16 250 12.192 Three Creek 123.72768 40.9552 0.60177 0.072617 674.3756 250 12.192 Slide Creek 123.37588 40.9625 0.57986 0.054413 1236.0105 250 12.192 Granite Creek 122.83708 40.9627 2.2833 0.92528 6.07E+14 250 12.192 Cedar Creek 123.53395 40.9634 0.35839 0.03582 2.15E+01 250 12.192 Deer Creek 122.91578 40.966 0.34915 0.059244 24.505 250 12.192 Granite Creek 122.77073 40.9678 0.55469 0.095054 1065.0549 250 12.192 Groves Prairie Creek 123.5478 40.9695 0.64643 0.075234 3.48E+03 250 12.192 Swift Creek 122.76959 40.9714 1.0145 0.2713 4504412.132 250 12.192 Stuart Fork 122.95772 40.9766 3.6494 1.0255 2.36E+25 250 12.192 Rattlesnake 123.07824 40.9803 1.2605 0.30859 33453375.3 250 12.192 Eagle Creek 123.3725 40.9823 0.48175 0.03945 169.3064 250 12.192 Canyon Creek 123.03813 40.9866 2.0323 0.45338 6.2561E+11 250 12.192 East Fork Horse Linto Creek 123.45368 40.9886 -0.84702 0.13652 0 250 12.192 Eightmile Creek 123.39412 40.9991 0.4156 0.034303 50.1864 250 12.192 Sturart Fork 122.97601 41.0003 1.8953 0.50734 1.25682E+12 250 12.192 North Fork Swift Creek 122.73243 41.0032 0.84267 0.12745 104207.3112 250 12.192 Horse Linto Creek 123.45314 41.0051 7.6424 4.4126 2.58E+54 250 12.192 Pine Creek 123.7678 41.0057 -0.96717 1.2836 0 250 12.192 Swift Creek 122.87335 41.0062 0.58919 0.089327 648.6232 250 12.192 Pine Creek 123.77855 41.0067 1.0532 0.28767 77614.041 250 12.192 Horse Linto Creek 123.53814 41.0096 1.0339 0.1275 4187418.408 250 12.192 Supply Creek 123.7359 41.0111 1.0507 0.19125 1876354.959 250 12.192 Horse Linto Creek 123.55052 41.0125 1.4472 0.27425 7.72E+09 250 12.192 Horse Linto Creek_Upstream 123.55482 41.0176 1.0738 0.0659 7876947.285 250 12.192

251 Texas Tech University, Timothy Anderson, December 2008

Table 1 (Continued) Name of River Longitude Latitude Normalized Ksn Ksn Minimum Maximum Steepness (Low σ) (High 2 σ) Drainage Area Drainage Area East Fork Horse Linto Creek 123.42238 41.0179 56 60 53 189667 4038105 East Fork Trinity_Unnamed 122.57708 41.0193 129 136 121 6902942 59242833 Cedar Creek 122.63225 41.0263 93 96 90 2943153 547558574 Redwood 123.88127 41.027 59 63 55 2765006 2569768451 Pine Creek 123.75805 41.0276 34 37 31 6031600 14948300 Horse Linto Creek _Upstream123.47397 41.0304 136 143 128 2390067 7812450 South Fork Salmon 122.93226 41.0316 52 54 49 276282 31120155 North Fork Swift Creek 122.78279 41.0345 148 167 130 4242618 6355260 Supply Creek 123.70432 41.0347 122 132 112 22946814 40754745 North Fork Swift Creek 122.77072 41.0349 185 204 167 6346836 7754616 Boulder Creek 122.82987 41.0475 89 95 82 1748871 8698752 Battle Creek 123.33028 41.0481 107 109 104 1008930 309276905 Horse Linto Creek_upstream 123.46675 41.0494 42 51 33 400236 1846755 Horse Linto Creek 123.43765 41.0514 62 68 57 1601892 5762591 Pine Creek 123.77024 41.0533 89 109 69 14948300 20688600 Tish Tang a Tang Creek 123.57015 41.0543 196 210 182 30272778 63960597 Boulder Creek 122.72709 41.0607 152 160 143 15033579 5656074101 Boulder Creek 122.81608 41.0658 177 188 167 8537481 14825916 Pine Creek 123.76437 41.0665 123 172 74 20688600 21897600 South Fork Salmon 122.94815 41.0671 211 256 166 32224100 33703500 East Fork Coffee Creek 122.73398 41.0679 129 136 123 26212658 1284256911 Smith Creek 122.57733 41.0742 114 117 112 5951794 527635495 Tish Tang a Tang Creek 123.507 41.0755 146 151 141 7554546 26057376 East Fork Trinity 122.57542 41.0758 109 111 108 13451559 314037775 Coffee Creek 122.75651 41.0764 144 149 138 82699909 1284256911 Coffee Creek 122.75651 41.0764 144 149 138 68710346 919989130 Crow Creek 122.5766 41.083 118 120 115 2836066 1284256911

252 Texas Tech University, Timothy Anderson, December 2008

Table 1 (Continued) Name of River Longitude Latitude Concavity Concavity Fitted Smoothing Contour within 2 σ Steepness Window Interval East Fork Horse Linto Creek 123.42238 41.0179 0.77719 0.095852 4950.7785 250 12.192 East Fork Trinity_Unnamed 122.57708 41.0193 1.1131 0.15409 6996593.794 250 12.192 Cedar Creek 122.63225 41.0263 0.93443 0.14111 230306.606 250 12.192 Redwood 123.88127 41.027 0.79796 0.050106 30877.8178 300 13 Pine Creek 123.75805 41.0276 0.51523 1.0325 117.9399 250 12.192 Horse Linto Creek _Upstream123.47397 41.0304 1.252 0.20537 26449286.8 250 12.192 South Fork Salmon 122.93226 41.0316 0.68895 0.10206 2301.7975 250 12.192 North Fork Swift Creek 122.78279 41.0345 3.4229 0.6245 1.20E+22 250 12.192 Supply Creek 123.70432 41.0347 2.9085 0.80815 2.78E+20 250 12.192 North Fork Swift Creek 122.77072 41.0349 -5.6186 2.0555 0 250 12.192 Boulder Creek 122.82987 41.0475 1.0847 0.2521 1442784.41 250 12.192 Battle Creek 123.33028 41.0481 0.56753 0.031801 742.8293 250 12.192 Horse Linto Creek_upstream 123.46675 41.0494 0.93914 0.37066 38262.8368 250 12.192 Horse Linto Creek 123.43765 41.0514 1.1 0.26436 1097746.504 250 12.192 Pine Creek 123.77024 41.0533 -4.9245 2.7561 0 250 12.192 Tish Tang a Tang Creek 123.57015 41.0543 2.0399 0.27383 2.82E+14 250 12.192 Boulder Creek 122.72709 41.0607 0.70882 0.079249 16877.7 250 12.192 Boulder Creek 122.81608 41.0658 -1.619 0.45871 0 250 12.192 Pine Creek 123.76437 41.0665 -63.1229 20.4004 0 250 12.192 South Fork Salmon 122.94815 41.0671 32.1772 9.8465 6.17E+240 250 12.192 East Fork Coffee Creek 122.73398 41.0679 0.79742 0.068815 97709.6037 250 12.192 Smith Creek 122.57733 41.0742 0.56787 0.045606 1015.9817 250 12.192 Tish Tang a Tang Creek 123.507 41.0755 0.83347 0.14299 78927.7445 250 12.192 East Fork Trinity 122.57542 41.0758 0.57744 0.066879 1209.4117 250 12.192 Coffee Creek 122.75651 41.0764 0.90902 0.16141 884739.8651 250 12.192 Coffee Creek 122.75651 41.0764 0.94705 0.14536 1846978.337 250 12.192 Crow Creek 122.5766 41.083 0.51784 0.034421 398.4994 250 12.192

253 Texas Tech University, Timothy Anderson, December 2008

Table 1 (Continued) Name of River Longitude Latitude Normalized Ksn Ksn Minimum Maximum Steepness (Low σ) (High 2 σ) Drainage Area Drainage Area Plummer Creek 123.21813 41.0859 77 78 76 212000 2793300 North Fork Coffee Creek 122.75985 41.0881 155 161 149 50760918 762054318 Pine Creek 123.77554 41.0906 63 75 50 21772500 42121700 Tish Tang a Tang Creek 123.48116 41.0908 72 74 71 418325 5072000 Coffee Creek 122.77596 41.0978 140 145 134 97261398 736725294 Granite Creek 122.77184 41.1003 112 116 107 236726 8193715592 Smith Creek 122.506 41.1025 70 74 67 1004644 5526558 Mill Creek_Upstream 123.63627 41.1063 161 164 157 1456621 832374420 Red Cap Creek 123.44745 41.1063 77 82 73 206351 9829001 Mill Creek 123.65563 41.1073 108 121 94 104074956 125626383 South Fork Salmon 123.08773 41.1144 97 103 92 38910000 477885200 Mill Creek_upstream 123.51231 41.1152 83 88 77 2011716 3907278 Little Pine Creek 123.82384 41.1169 27 100 -46 5858552 16490296 Mumbo Creek 122.57873 41.118 90 92 88 1091232 263683998 Pine Creek 123.77135 41.1189 65 81 50 42210900 46715700 Little Pine Creek 123.81248 41.1199 230 307 154 13334700 20678800 Little Pine Creek 123.80748 41.1246 237 311 164 20864800 23640500 Coffee Creek 122.90882 41.1298 99 110 87 8244666 61200765 North Fork Mill Creeek 123.61101 41.1307 112 116 108 6321160 214129888 Little Pine Creek 123.8361 41.1318 38 42 33 1198538 6277010 Mill Creek 123.57391 41.1349 161 168 154 8302500 91206648 Little Trinity 122.6733 41.1399 111 116 105 15033579 4528304406 Picayune Creek 122.66462 41.1486 104 108 100 26212658 1792755761 Little South Fork Salmon 123.15417 41.1487 111 114 108 496392 2521102234 East Fork Coffee Creek 122.78692 41.1494 123 125 121 3426705 24114753 Red Cap Creek 123.46719 41.1497 82 84 80 10297900 40033100 Tangle Blue Creek 122.67822 41.1498 89 95 84 30637359 737656956

254 Texas Tech University, Timothy Anderson, December 2008

Table 1 (Continued) Name of River Longitude Latitude Concavity Concavity Fitted Smoothing Contour within 2 σ Steepness Window Interval Plummer Creek 123.21813 41.0859 0.51012 0.097855 187.449 250 12.192 North Fork Coffee Creek 122.75985 41.0881 0.74108 0.11224 35536.697 250 12.192 Pine Creek 123.77554 41.0906 4.1804 1.2283 4.02E+29 250 12.192 Tish Tang a Tang Creek 123.48116 41.0908 0.52106 0.086281 203.5005 250 12.192 Coffee Creek 122.77596 41.0978 1.0343 0.14731 10011055.79 250 12.192 Granite Creek 122.77184 41.1003 0.31474 0.037417 11.6353 250 12.192 Smith Creek 122.506 41.1025 0.62673 0.1492 950.6108 250 12.192 Mill Creek_Upstream 123.63627 41.1063 0.55489 0.038153 8.70E+02 250 12.192 Red Cap Creek 123.44745 41.1063 0.078025 0.062448 0.34872 250 12.192 Mill Creek 123.65563 41.1073 7.1507 1.8086 1.13E+56 250 12.192 South Fork Salmon 123.08773 41.1144 0.93992 0.12797 996646.7135 250 12.192 Mill Creek_upstream 123.51231 41.1152 1.2526 0.33789 13147727.84 250 12.192 Little Pine Creek 123.82384 41.1169 1.1403 9.8604 2729705.366 250 12.192 Mumbo Creek 122.57873 41.118 0.31605 0.034158 9.0306 250 12.192 Pine Creek 123.77135 41.1189 17.3093 8.1862 7.92E+130 250 12.192 Little Pine Creek 123.81248 41.1199 -64.3753 21.1777 0 250 12.192 Little Pine Creek 123.80748 41.1246 14.4051 2.8387 6.75E+104 250 12.192 Coffee Creek 122.90882 41.1298 1.0937 0.20197 5466750.976 250 12.192 North Fork Mill Creeek 123.61101 41.1307 0.64248 0.077773 3495.8302 250 12.192 Little Pine Creek 123.8361 41.1318 1.0202 0.34442 189265.0231 250 12.192 Mill Creek 123.57391 41.1349 0.88867 0.1027 295393.0017 250 12.192 Little Trinity 122.6733 41.1399 0.733 0.046606 20167.6674 250 12.192 Picayune Creek 122.66462 41.1486 0.74457 0.0614 26113.4458 250 12.192 Little South Fork Salmon 123.15417 41.1487 0.59839 0.030951 1723.3568 250 12.192 East Fork Coffee Creek 122.78692 41.1494 0.53468 0.088289 522.3717 250 12.192 Red Cap Creek 123.46719 41.1497 0.49396 0.17811 175.6084 250 12.192 Tangle Blue Creek 122.67822 41.1498 0.8481 0.070041 173338.7113 250 12.192

255 Texas Tech University, Timothy Anderson, December 2008

Table 1 (Continued) Name of River Longitude Latitude Normalized Ksn Ksn Minimum Maximum Steepness (Low σ) (High 2 σ) Drainage Area Drainage Area North Fork Mill Creek 123.51291 41.1581 116 118 114 300668 6093412 East Fork Trinity 122.46573 41.1606 48 53 42 1617894 3032883 Pine Creek 123.77187 41.1617 84 91 76 49384800 125559800 Little Pine Creek 123.77958 41.162 89 92 86 22543200 125830200 Trinity 122.65615 41.1632 125 130 120 13959479 1383072896 Scott Mountain Creek 122.6758 41.1634 113 119 106 10341999 722884338 North Fork Coffee Creek 122.87666 41.1703 95 96 94 461300 47430351 West Fork Knownothing Creek123.36761 41.1705 117 118 115 444745 8027357 Roach Creek 123.90495 41.171 47 50 44 580825 1692250 Red Cap_ Unnamed 123.43115 41.1751 105 108 103 346199 2473195 Red Cap Creek 123.50062 41.1793 140 149 132 40000700 76232500 East South Fork Salmon 123.08328 41.1794 93 103 83 80071600 173846100 East Fork South Fork Salmon 123.0825 41.1802 93 103 83 80050000 173707400 South Fork Taylor Creek 123.17744 41.1817 117 119 114 370330 4698596933 Taylor Creek 123.16589 41.1833 121 123 118 1195459 1880848767 Saitne Claire Creek 123.22598 41.1834 99 102 95 1602402 4873856783 Plummer Creek 123.24463 41.1885 140 143 138 4634512 2615140520 West Fork Shadow Creek 123.18528 41.1892 92 93 90 641442 2430445487 Crawford Creek 123.20035 41.192 85 86 84 742635 1748013296 Roach Creek 123.89075 41.1935 41 45 38 1942627 24946162 Middle Fork Red Cap Creek 123.47361 41.1955 140 142 137 346199 20994106 East Fork Knownothing Creek123.30545 41.2075 91 91 90 357013 2343048682 Methodist Creek 123.278 41.2093 97 99 95 1240050 9422259685 Roach Creek 123.92749 41.2116 53 61 45 3069000 6634300 West Fork Knownothing Creek123.31861 41.2117 134 139 129 9640408 134655672 Red Cap_Unnnamed 123.52858 41.2135 134 138 131 2742851 9094974759 East South Fork Salmon 123.0291 41.2142 154 160 148 50981300 79581700

256 Texas Tech University, Timothy Anderson, December 2008

Table 1 (Continued) Name of River Longitude Latitude Concavity Concavity Fitted Smoothing Contour within 2 σ Steepness Window Interval North Fork Mill Creek 123.51291 41.1581 0.29414 0.038239 12.7361 250 12.192 East Fork Trinity 122.46573 41.1606 2.2021 0.2989 7.41959E+12 250 12.192 Pine Creek 123.77187 41.1617 -0.64148 0.604 0 250 12.192 Little Pine Creek 123.77958 41.162 0.33162 0.18546 11.2166 250 12.192 Trinity 122.65615 41.1632 0.81377 0.044494 101704.3367 250 12.192 Scott Mountain Creek 122.6758 41.1634 0.74229 0.0502 22632.4464 250 12.192 North Fork Coffee Creek 122.87666 41.1703 0.40887 0.052205 49.6229 250 12.192 West Fork Knownothing Creek123.36761 41.1705 0.42391 0.052013 81.0917 250 12.192 Roach Creek 123.90495 41.171 0.67782 0.37586 1164.9368 250 12.192 Red Cap_ Unnamed 123.43115 41.1751 0.55449 0.12166 478.1502 250 12.192 Red Cap Creek 123.50062 41.1793 1.0752 0.82654 9867091.625 250 12.192 East South Fork Salmon 123.08328 41.1794 2.1283 0.60737 3.77E+15 250 12.192 East Fork South Fork Salmon 123.0825 41.1802 2.1283 0.60737 3.77E+15 250 12.192 South Fork Taylor Creek 123.17744 41.1817 0.50458 0.024912 308.7099 250 12.192 Taylor Creek 123.16589 41.1833 0.52487 0.028406 463.8433 250 12.192 Saitne Claire Creek 123.22598 41.1834 0.52388 0.049586 443.7677 250 12.192 Plummer Creek 123.24463 41.1885 0.52483 0.045306 534.515 250 12.192 West Fork Shadow Creek 123.18528 41.1892 0.49942 0.035558 253.2167 250 12.192 Crawford Creek 123.20035 41.192 0.43666 0.030723 75.9189 250 12.192 Roach Creek 123.89075 41.1935 0.8355 0.25176 19733.9331 250 12.192 Middle Fork Red Cap Creek 123.47361 41.1955 0.44377 0.063115 121.5036 250 12.192 East Fork Knownothing Creek123.30545 41.2075 0.44807 0.022725 91.2658 250 12.192 Methodist Creek 123.278 41.2093 0.53934 0.035925 463.5808 250 12.192 Roach Creek 123.92749 41.2116 -0.97625 0.82175 0 250 12.192 West Fork Knownothing Creek123.31861 41.2117 0.92349 0.083457 387701.1927 250 12.192 Red Cap_Unnnamed 123.52858 41.2135 0.62939 0.056332 2885.8702 250 12.192 East South Fork Salmon 123.0291 41.2142 1.9756 0.34946 1.54E+14 250 12.192

257 Texas Tech University, Timothy Anderson, December 2008

Table 1 (Continued) Name of River Longitude Latitude Normalized Ksn Ksn Minimum Maximum Steepness (Low σ) (High 2 σ) Drainage Area Drainage Area Roach Creek 123.9406 41.2148 48 59 36 561130 3741396 East Fork South Fork Salmon123.02838 41.216 144 145 142 15362500 79693800 Roach Creek 123.89571 41.2164 35 39 31 6291000 48093100 East Fork South Fork Salmon122.94781 41.2205 75 76 73 839300 46510500 Tangle Blue Creek 122.73526 41.2207 103 113 93 4108489 12962121 North Fork Red Cap Creek 123.49954 41.2242 121 126 115 4763242 17069065 North Fork Red Cap Creek 123.46432 41.2254 127 136 119 1470200 4773200 Tangle Blue Creek 122.70215 41.2284 145 149 141 16502778 30349728 Red Cap Creek 123.55384 41.2294 96 103 89 53283379 661035402 Red Cap Creek 123.55385 41.2298 99 106 92 76248600 156821200 West Fork Shadow Creek 123.08248 41.2316 112 123 100 534114 3097802 North Fork Red Cap Creek 123.55671 41.234 103 110 95 17069065 2135934915 Roach Creek 123.86832 41.2427 70 80 60 20361500 75323800 Nordeheimer Crek_unnamed 123.38452 41.2437 117 119 114 715930 180493361 Metah Creek 123.92804 41.2441 37 45 29 323120 2649848 Sixmile Creek 123.00216 41.245 150 156 144 2523700 12502100 Granite Creek 123.42721 41.2461 147 151 144 1544781 8959552 Tectah Creek 123.97119 41.2473 36 39 33 312163 10173974 Roach Creek 123.8607 41.2512 88 99 78 47160800 75479100 Scott Mountain Creek 122.7022 41.2621 82 82 81 578988 9702747 Bluff Creek 123.67114 41.268 119 131 107 159858100 191910400 Granite Creek 123.38662 41.2725 139 147 130 9293747 250958775 South Russian Creek 122.96846 41.2788 72 74 69 421100 12097600 Metah Creek 123.9107 41.2847 59 62 57 2742851 1925946393 Metah Creek 123.89461 41.2926 77 79 75 1525881 49731231 Metah Creek 123.89452 41.2934 77 80 75 1692250 166331224 Little Trinity 122.63286 41.2957 109 111 106 6078078 16286265

258 Texas Tech University, Timothy Anderson, December 2008

Table 1 (Continued) Name of River Longitude Latitude Concavity Concavity Fitted Smoothing Contour within 2 σ Steepness Window Interval Roach Creek 123.9406 41.2148 1.3475 0.23571 16624711.92 250 12.192 East Fork South Fork Salmon123.02838 41.216 0.4335 0.095744 111.3769 250 12.192 Roach Creek 123.89571 41.2164 0.80272 0.31599 15598.8556 250 12.192 East Fork South Fork Salmon122.94781 41.2205 0.40678 0.061049 41.7256 250 12.192 Tangle Blue Creek 122.73526 41.2207 1.0777 0.2429 2063277.603 250 12.192 North Fork Red Cap Creek 123.49954 41.2242 1.3312 0.15567 178906290.5 250 12.192 North Fork Red Cap Creek 123.46432 41.2254 -0.6054 0.18122 0.000019 250 12.192 Tangle Blue Creek 122.70215 41.2284 1.066 0.27304 5008800.607 250 12.192 Red Cap Creek 123.55384 41.2294 1.811 0.5438 9.27028E+12 250 12.192 Red Cap Creek 123.55385 41.2298 1.7191 0.58556 1.71957E+12 250 12.192 West Fork Shadow Creek 123.08248 41.2316 0.91225 0.14101 80123.8263 250 12.192 North Fork Red Cap Creek 123.55671 41.234 0.79603 0.17572 56991.0554 250 12.192 Roach Creek 123.86832 41.2427 -1.1384 0.74047 0 250 12.192 Nordeheimer Crek_unnamed 123.38452 41.2437 0.53587 0.042111 489.3956 250 12.192 Metah Creek 123.92804 41.2441 1.134 0.20646 493805.2392 250 12.192 Sixmile Creek 123.00216 41.245 0.89717 0.15093 149712.0037 250 12.192 Granite Creek 123.42721 41.2461 0.63878 0.072813 2550.7416 250 12.192 Tectah Creek 123.97119 41.2473 0.79339 0.10188 5313.5321 250 12.192 Roach Creek 123.8607 41.2512 -1.4614 1.661 0 250 12.192 Scott Mountain Creek 122.7022 41.2621 0.44186 0.060309 72.3438 250 12.192 Bluff Creek 123.67114 41.268 -7.5904 4.1046 0 250 12.192 Granite Creek 123.38662 41.2725 0.81561 0.12888 77831.6389 250 12.192 South Russian Creek 122.96846 41.2788 0.51135 0.1389 202.1042 250 12.192 Metah Creek 123.9107 41.2847 -0.12844 0.19215 0.005196 250 12.192 Metah Creek 123.89461 41.2926 0.31697 0.089436 10.4462 250 12.192 Metah Creek 123.89452 41.2934 0.27443 0.087443 5.2103 250 12.192 Little Trinity 122.63286 41.2957 0.46243 0.24227 137.1016 250 12.192

259 Texas Tech University, Timothy Anderson, December 2008

Table 1 (Continued) Name of River Longitude Latitude Normalized Ksn Ksn Minimum Maximum Steepness (Low σ) (High 2 σ) Drainage Area Drainage Area Slate Creek 123.64446 41.2971 119 123 116 878659 204579175 South Russian Creek 123.16603 41.2986 120 126 113 22382088 840336602 Bluff Creek 123.69849 41.3038 132 169 95 150081600 160052500 Little Trinity 122.63481 41.3061 87 96 78 4189644 5950422 Taylor Creek 123.1402 41.3133 116 118 113 1168900 527379700 Tectah Creek 123.92859 41.3135 43 44 42 7642400 50764700 Crapo Creek 123.30519 41.3148 171 172 170 1384054 290549864 Little North Fork Salmon 123.22386 41.3155 134 136 131 9377200 527374100 Little Trinity 122.63585 41.3157 68 70 66 1879200 4682043 Duck Lake Creek 122.9371 41.3183 105 114 96 1159800 5592600 Trinity 122.53283 41.3227 115 128 102 5648616 9120762 Scott_upstream 122.89723 41.329 36 40 33 86900 5840800 Morehouse Creek 123.35693 41.334 242 281 203 20004100 20461600 East Fork Pecwan Creek 123.84091 41.341 250 295 204 29465000 71344000 North Fork Salmon Creek 123.14805 41.345 90 91 88 1111030 3912418718 Morehouse Creek 123.33356 41.3465 178 181 175 3570800 19851900 Pecwan Creek 123.84153 41.3466 350 384 317 31859200 71344000 East Fork Pecwan Creek 123.81073 41.3475 263 280 246 20041200 29989700 Camp Creek 123.59177 41.3484 91 95 86 20282249 204579175 Camp Creek 123.59177 41.3484 91 95 86 20282249 204579175 Bluff Creek 123.7221 41.3488 98 103 93 89379100 153098100 Crapo Creek 123.25344 41.3511 126 128 124 1433900 2123600 Crapo Creek 123.25107 41.3542 171 172 170 1384054 290549864 Pecwan Creek 123.83173 41.3622 376 417 335 13626100 31859200 East Fork Pecwan Creek 123.78108 41.3632 91 95 86 2389336 16490296 Morehouse Creek 123.30971 41.3654 165 181 149 2625000 3590300 Little North Fork Salmon 123.24744 41.3747 67 79 54 1544300 4065000

260 Texas Tech University, Timothy Anderson, December 2008

Table 1 (Continued) Name of River Longitude Latitude Concavity Concavity Fitted Smoothing Contour within 2 σ Steepness Window Interval Slate Creek 123.64446 41.2971 0.60698 0.061515 1359.2828 250 12.192 South Russian Creek 123.16603 41.2986 0.82527 0.066446 159471.7414 250 12.192 Bluff Creek 123.69849 41.3038 -29.9948 12.4515 0 250 12.192 Little Trinity 122.63481 41.3061 -1.5369 0.82913 0 250 12.192 Taylor Creek 123.1402 41.3133 0.49558 0.032139 258.2118 250 12.192 Tectah Creek 123.92859 41.3135 0.28543 0.45905 2.802 250 12.192 Crapo Creek 123.30519 41.3148 0.37911 0.032395 53.6606 250 12.192 Little North Fork Salmon 123.22386 41.3155 0.5758 0.043638 1353.6 250 12.192 Little Trinity 122.63585 41.3157 0.85464 0.17938 29566.9238 250 12.192 Duck Lake Creek 122.9371 41.3183 -0.54932 0.15289 0.000043 250 12.192 Trinity 122.53283 41.3227 -1.6071 0.48325 0 250 12.192 Scott_upstream 122.89723 41.329 0.73131 0.09752 1862.3935 250 12.192 Morehouse Creek 123.35693 41.334 40.2289 16.9876 1.17E+293 250 12.192 East Fork Pecwan Creek 123.84091 41.341 1.8967 1.1485 2.33448E+13 250 12.192 North Fork Salmon Creek 123.14805 41.345 0.50773 0.029723 276.1073 250 12.192 Morehouse Creek 123.33356 41.3465 0.68185 0.070296 7867.6156 250 12.192 Pecwan Creek 123.84153 41.3466 2.2952 0.92719 3.47E+16 250 12.192 East Fork Pecwan Creek 123.81073 41.3475 -4.1884 0.7505 0 250 12.192 Camp Creek 123.59177 41.3484 1.1177 0.1768 13935042 250 12.192 Camp Creek 123.59177 41.3484 1.1177 0.1768 13935042 250 12.192 Bluff Creek 123.7221 41.3488 2.0465 0.55605 6.94E+14 250 12.192 Crapo Creek 123.25344 41.3511 0.72687 0.31724 7062.1034 250 12.192 Crapo Creek 123.25107 41.3542 0.37911 0.032395 53.6606 250 12.192 Pecwan Creek 123.83173 41.3622 -9.4025 1.4382 0 250 12.192 East Fork Pecwan Creek 123.78108 41.3632 0.8586 0.13331 55740.7915 250 12.192 Morehouse Creek 123.30971 41.3654 -4.5637 0.6113 0 250 12.192 Little North Fork Salmon 123.24744 41.3747 3.2774 0.74194 4.85E+19 250 12.192

261 Texas Tech University, Timothy Anderson, December 2008

Table 1 (Continued) Name of River Longitude Latitude Normalized Ksn Ksn Minimum Maximum Steepness (Low σ) (High 2 σ) Drainage Area Drainage Area Little North Fork Salmon 123.22536 41.38 121 126 116 7535300 9548400 Steinacher Creeek 123.28521 41.3841 119 134 104 2512600 5285400 Pecwan Creek 123.79989 41.386 108 115 101 1751643 5858552 Steinacher Creek 123.30504 41.3908 168 176 160 4613800 10244100 Steinacher Creek 123.34047 41.3912 182 187 177 10218600 24437100 Hancock Creek 123.25512 41.3966 56 60 52 641442 2879019 Bluff Creek 123.71575 41.4001 76 90 61 44200500 89955700 Camp Creek 123.59245 41.4019 126 151 100 12952677 24100301 Camp Creek 123.59245 41.4019 126 151 100 12952677 24100301 Slide Creek 123.78348 41.4102 128 134 122 931300 1913600 Haypress Creek 123.39859 41.4125 199 209 188 17320826 626926604 Hancock Creek 123.26055 41.4174 262 279 246 2790000 16653400 Camp Creek 123.58171 41.4189 59 60 57 206351 13407284 Camp Creek 123.58171 41.4189 59 60 57 206351 13407284 Slide Creek 123.79696 41.4223 176 179 174 1911300 10295900 North Fork Salmon 123.19739 41.4227 95 101 90 1780200 8643400 Hancock Creek 123.35542 41.4283 142 149 135 16627700 346205700 Etna Creek 122.99 41.43 94 97 91 3124200 25308600 North Fork Salmon 123.17664 41.4345 96 108 83 7711600 10311600 Bluff Creek 123.6874 41.4353 42 45 40 1284145 51476675 Rock Creek 123.60789 41.438 81 84 77 820083 2389336 Meaks Creek 122.87202 41.4391 122 132 113 1736800 906279900 Slide Creek 123.81152 41.4401 150 156 145 10316000 16242500 Bridge Creek 123.37335 41.4427 171 174 169 9622700 383417500 Slide Creek 123.87577 41.443 74 82 66 16083800 320466700 Rock Creek 123.58927 41.4464 87 90 83 2473195 11679271 Nickowitz Creek 123.84943 41.4469 93 96 89 20282249 1993542360

262 Texas Tech University, Timothy Anderson, December 2008

Table 1 (Continued) Name of River Longitude Latitude Concavity Concavity Fitted Smoothing Contour within 2 σ Steepness Window Interval Little North Fork Salmon 123.22536 41.38 -1.2185 0.46546 0 250 12.192 Steinacher Creeek 123.28521 41.3841 2.71 0.55355 1.22E+17 250 12.192 Pecwan Creek 123.79989 41.386 1.4187 0.15675 199190488.7 250 12.192 Steinacher Creek 123.30504 41.3908 -0.86965 0.13927 0 250 12.192 Steinacher Creek 123.34047 41.3912 1.0781 0.18623 5800528.163 250 12.192 Hancock Creek 123.25512 41.3966 0.92614 0.12173 51225.4117 250 12.192 Bluff Creek 123.71575 41.4001 -1.3172 1.0951 0 250 12.192 Camp Creek 123.59245 41.4019 -3.2307 1.2694 0 250 12.192 Camp Creek 123.59245 41.4019 -3.2307 1.2694 0 250 12.192 Slide Creek 123.78348 41.4102 -1.0227 0.25112 0 250 12.192 Haypress Creek 123.39859 41.4125 0.66946 0.062915 10624.1799 250 12.192 Hancock Creek 123.26055 41.4174 -0.73423 0.099316 0.000001 250 12.192 Camp Creek 123.58171 41.4189 0.42612 0.054092 45.1703 250 12.192 Camp Creek 123.58171 41.4189 0.42612 0.054092 45.1703 250 12.192 Slide Creek 123.79696 41.4223 0.3796 0.084404 57.97 250 12.192 North Fork Salmon 123.19739 41.4227 0.96471 0.22278 212440.3756 250 12.192 Hancock Creek 123.35542 41.4283 0.79153 0.067219 85912.5651 250 12.192 Etna Creek 122.99 41.43 0.7118 0.095238 6881.724 250 12.192 North Fork Salmon 123.17664 41.4345 -6.134 1.2008 0 250 12.192 Bluff Creek 123.6874 41.4353 0.68389 0.068039 1996.1492 250 12.192 Rock Creek 123.60789 41.438 0.3329 0.376 15.679 250 12.192 Meaks Meadow Creek 122.87202 41.4391 0.87319 0.061223 108939.5442 250 12.192 Slide Creek 123.81152 41.4401 -0.71991 0.51706 0.000001 250 12.192 Bridge Creek 123.37335 41.4427 0.57004 0.044447 1407.2127 250 12.192 Slide Creek 123.87577 41.443 0.69511 0.2476 9196.2136 250 12.192 Rock Creek 123.58927 41.4464 0.83586 0.13159 34720.189 250 12.192 Nickowitz Creek 123.84943 41.4469 0.60571 0.15064 1637.1659 250 12.192

263 Texas Tech University, Timothy Anderson, December 2008

Table 1 (Continued) Name of River Longitude Latitude Normalized Ksn Ksn Minimum Maximum Steepness (Low σ) (High 2 σ) Drainage Area Drainage Area Nickowitz Creek 123.84885 41.447 93 96 90 19971400 320474700 Nickowitz Creek 123.71635 41.45 50 53 47 151278 3041908 Nickowitz Creek 123.71635 41.45 50 53 47 156587 3041908 Haypress Creek 123.39484 41.4531 240 256 224 16347100 18168200 Wooley Creek_unnamed 123.30468 41.4532 116 118 115 398472 1748013296 Irving Creek 123.44463 41.4542 153 165 141 9248000 12360600 West Fork Blue Creek 123.90592 41.4568 101 105 96 10531056 1188246017 Irving Creek 123.45831 41.4569 114 133 94 12312400 13659100 Big Elk Fork 123.30867 41.4579 122 125 118 7670400 344928900 Wooley Creek 123.3037 41.4582 116 118 114 5918000 383959500 Blue Creek 123.84652 41.4583 78 82 74 103721100 320468000 North Fork Woolely Creek 123.32915 41.4589 121 122 120 828875 3912418718 Blue Creek 123.84511 41.4591 86 92 80 103221100 320455300 Nickowitz Creek 123.74168 41.4601 65 71 59 3259182 20282249 Nickowitz Creek 123.74002 41.4602 67 73 61 3491975 19594529 Cuddihy Fork 123.33083 41.4615 128 130 126 2639300 346215300 Haypress Creek 123.40262 41.4621 136 163 109 12404900 16330200 Irving Creek 123.42443 41.4658 95 99 91 1625900 9125800 Irving Creek 123.46621 41.4659 148 162 133 13489500 18025600 Irving Creek 123.48589 41.4678 164 177 152 18025600 21917200 Haypress Creek 123.40248 41.4719 160 175 144 11394400 12225800 Crescent City Fork 123.84798 41.4794 86 90 83 1375868 2288498130 Irving Creek 123.41656 41.4797 120 129 111 762600 1611000 Haypress Creek 123.40027 41.4857 123 158 87 9132500 10229500 Duck Lake Creek 122.87196 41.487 78 86 70 5592600 1521257700 Rock Creek 123.55993 41.4897 86 90 82 12952677 219191632 East Fork blue Creek 123.73613 41.4906 86 102 71 32995000 103285500

264 Texas Tech University, Timothy Anderson, December 2008

Table 1 (Continued) Name of River Longitude Latitude Concavity Concavity Fitted Smoothing Contour within 2 σ Steepness Window Interval Nickowitz Creek 123.84885 41.447 0.58345 0.15178 1074.6037 250 12.192 Nickowitz Creek 123.71635 41.45 0.7351 0.057176 2476.8162 250 12.192 Nickowitz Creek 123.71635 41.45 0.7351 0.057176 2476.8162 250 12.192 Haypress Creek 123.39484 41.4531 -12.9249 3.3451 0 250 12.192 Wooley Creek_unnamed 123.30468 41.4532 0.48054 0.018707 208.5907 250 12.192 Irving Creek 123.44463 41.4542 -2.4821 0.85188 0 250 12.192 West Fork Blue Creek 123.90592 41.4568 0.7032 0.095132 7371.1037 250 12.192 Irving Creek 123.45831 41.4569 12.6593 5.7207 7.83E+88 250 12.192 Big Elk Fork 123.30867 41.4579 0.64033 0.039395 4141.7547 250 12.192 Wooley Creek 123.3037 41.4582 0.58447 0.041511 1462.2556 250 12.192 Blue Creek 123.84652 41.4583 1.2624 0.57566 501677947.8 250 12.192 North Fork Woolely Creek 123.32915 41.4589 0.45843 0.022953 147.0645 250 12.192 Blue Creek 123.84511 41.4591 1.7867 0.62973 1.25142E+13 250 12.192 Nickowitz Creek 123.74168 41.4601 1.0398 0.20996 875531.5551 250 12.192 Nickowitz Creek 123.74002 41.4602 1.0713 0.22583 1427863.524 250 12.192 Cuddihy Fork 123.33083 41.4615 0.5122 0.027571 386.8649 250 12.192 Haypress Creek 123.40262 41.4621 4.8226 2.8397 2.27E+33 250 12.192 Irving Creek 123.42443 41.4658 0.79152 0.15847 18070.311 250 12.192 Irving Creek 123.46621 41.4659 -2.4864 1.0657 0 250 12.192 Irving Creek 123.48589 41.4678 4.713 1.2383 2.27E+33 250 12.192 Haypress Creek 123.40248 41.4719 5.007 5.4968 2.77E+34 250 12.192 Crescent City Fork 123.84798 41.4794 0.63539 0.06157 2379.0855 250 12.192 Irving Creek 123.41656 41.4797 -1.33 0.46009 0 250 12.192 Haypress Creek 123.40027 41.4857 -18.5769 9.3688 0 250 12.192 Duck Lake Creek 122.87196 41.487 0.89146 0.082341 170076.6287 250 12.192 Rock Creek 123.55993 41.4897 -0.28344 0.31444 0.000337 250 12.192 East Fork blue Creek 123.73613 41.4906 -2.0934 0.72519 0 250 12.192

265 Texas Tech University, Timothy Anderson, December 2008

Table 1 (Continued) Name of River Longitude Latitude Normalized Ksn Ksn Minimum Maximum Steepness (Low σ) (High 2 σ) Drainage Area Drainage Area Rock Creek 123.57818 41.4939 93 95 91 941419 298989152 Scott 122.86549 41.4972 34 38 29 6284300 1521222900 Haypress Creek 123.39689 41.4993 53 56 50 2181200 9900100 West Fork Blue Creek 123.91826 41.4993 84 119 50 4811700 8915700 Blue Creek 123.76707 41.504 558 631 484 100948500 103625200 Kidder Creek 123.09293 41.5087 73 77 70 191000 13425300 West Fork Blue Creek 123.92657 41.5096 77 85 69 580825 5858552 Wooley Creek 123.13599 41.5108 138 157 120 768000 3428300 Haypress Creek 123.38851 41.5143 39 42 36 554037 1662172 East Fork Blue Creek 123.69943 41.5147 54 59 50 974460 39062524 Bridge Creek 123.33292 41.5152 83 88 78 4634512 7738700 Blue Creek 123.75094 41.5182 189 210 168 30682545 59092938 Bridge Creek 123.33282 41.5285 68 73 62 859792 4003001 Etna Creek 122.88291 41.5287 48 58 39 25555200 1521445800 Big Elk Fork 123.19968 41.538 152 161 144 4114100 7670400 Mill Creek 123.04119 41.5385 139 141 137 811800 15748100 Ukonom Creek 123.39237 41.5414 162 175 150 2793700 11985200 Blue Crek 123.72971 41.5435 78 83 73 12089184 29642178 Big Elk Fork 123.20551 41.5463 95 97 93 1602402 4003001 Cuddihy Fork 123.31523 41.5503 110 138 81 972600 2412000 Crescent City Fork 123.78941 41.551 148 157 139 622311 1474142 Ukonom Creek 123.34553 41.5513 58 77 40 820083 2010809 Turwar Creek 123.9601 41.553 84 88 80 1240603 288851187 Mynot Creek 124.03599 41.561 52 59 45 456222 23283121 Kidder Creek 122.9291 41.5621 69 76 61 12989300 1525524000 Blue Creek 123.70967 41.5638 79 81 76 5040800 10862400 Granite Creek 123.32326 41.5652 94 108 80 792276 2473195

266 Texas Tech University, Timothy Anderson, December 2008

Table 1 (Continued) Name of River Longitude Latitude Concavity Concavity Fitted Smoothing Contour within 2 σ Steepness Window Interval Rock Creek 123.57818 41.4939 0.3244 0.099694 15.895 250 12.192 Scott 122.86549 41.4972 0.69717 0.13281 3660.6218 250 12.192 Haypress Creek 123.39689 41.4993 0.62368 0.30808 912.7354 250 12.192 West Fork Blue Creek 123.91826 41.4993 -2.3558 1.7449 0 250 12.192 Blue Creek 123.76707 41.504 -34.006 19.0339 0 250 12.192 Kidder Creek 123.09293 41.5087 0.57743 0.084846 498.4632 250 12.192 West Fork Blue Creek 123.92657 41.5096 1.1104 0.1228 1079196.066 250 12.192 Wooley Creek 123.13599 41.5108 1.8059 0.31037 23149122561 250 12.192 Haypress Creek 123.38851 41.5143 1.1714 0.54313 885357.1344 250 12.192 East Fork Blue Creek 123.69943 41.5147 0.74838 0.07367 5987.774 250 12.192 Bridge Creek 123.33292 41.5152 1.6922 0.76564 20504397431 250 12.192 Blue Creek 123.75094 41.5182 -4.4428 1.9129 0 250 12.192 Bridge Creek 123.33282 41.5285 1.0338 0.30437 387562.5738 250 12.192 Etna Creek 122.88291 41.5287 0.91193 0.18258 319106.4582 250 12.192 Big Elk Fork 123.19968 41.538 -0.98666 0.25253 0 250 12.192 Mill Creek 123.04119 41.5385 0.47742 0.057153 205.3019 250 12.192 Ukonom Creek 123.39237 41.5414 -1.0095 0.19673 0 250 12.192 Blue Crek 123.72971 41.5435 -0.60615 0.37722 0.000002 250 12.192 Big Elk Fork 123.20551 41.5463 0.52044 0.29731 280.6045 250 12.192 Cuddihy Fork 123.31523 41.5503 2.1041 0.59215 1.21508E+12 250 12.192 Crescent City Fork 123.78941 41.551 -0.87293 0.39723 0.000002 250 12.192 Ukonom Creek 123.34553 41.5513 2.8319 1.2026 3.10E+16 250 12.192 Turwar Creek 123.9601 41.553 0.65482 0.10981 2241.3933 250 12.192 Mynot Creek 124.03599 41.561 0.88233 0.22049 28668.8509 250 12.192 Kidder Creek 122.9291 41.5621 0.88904 0.14158 188484.2767 250 12.192 Blue Creek 123.70967 41.5638 1.0883 0.3741 2176922.957 250 12.192 Granite Creek 123.32326 41.5652 -1.8242 0.5782 0 250 12.192

267 Texas Tech University, Timothy Anderson, December 2008

Table 1 (Continued) Name of River Longitude Latitude Normalized Ksn Ksn Minimum Maximum Steepness (Low σ) (High 2 σ) Drainage Area Drainage Area Ukonom Creek 123.3663 41.5653 124 136 112 1942627 13407284 Goose Creek 123.89058 41.5692 49 51 48 245925 55316025 Copper Creek 123.64159 41.571 114 121 108 4445699 24100301 Dillon Creek 123.54365 41.5722 111 119 103 58377400 188970000 Granite Creek 123.31829 41.5726 101 107 95 3041908 7720412 Dillon Creek 123.54272 41.5743 102 104 99 26727990 204579175 Blue Creek 123.69826 41.5787 62 70 55 1634870 5467990 Elk Creek 123.24061 41.5826 79 80 78 2878500 11393700 Ukonom Creek 123.4406 41.5843 189 195 183 12952677 106222674 Dillon Creek 123.58115 41.5843 114 122 106 58397800 188080300 Ukonom Creek 123.44583 41.5845 181 187 175 12021800 84447500 Dillon Creek 123.58464 41.585 108 113 103 22493650 343226273 Granite Creek 123.31495 41.5881 134 151 116 7657300 14351300 Elk Creek 123.22057 41.589 72 77 67 281474 5103465 Turwar Creek 123.97466 41.5893 58 60 56 5262900 74052500 Duzell Creek 122.79244 41.591 39 40 39 37104 16569747689 Moffett Creek 122.77745 41.5916 49 51 47 1165194 6559963281 Dillon Creek 123.62359 41.5986 111 119 103 58377400 188970000 Hunter Creek 124.03656 41.599 42 48 37 792276 99141317 South Fork Kelsey Creek 123.1952 41.5992 73 79 67 383267 2270580 Elk Creek 123.23735 41.6004 131 135 127 4930420 436967803 Dillon Creek 123.69656 41.6016 73 74 72 523722 5858552 Independence Creek 123.3479 41.6019 125 147 104 2010809 3872710 Hunter Creek 124.03897 41.6028 33 37 29 3373571 166331224 Elk Creek 123.26383 41.6048 121 125 117 14869099 661035402 Dillon Creek 123.58173 41.606 103 108 99 38798900 188392200 Independence Creek 123.36529 41.6081 178 182 174 4551000 12380900

268 Texas Tech University, Timothy Anderson, December 2008

Table 1 (Continued) Name of River Longitude Latitude Concavity Concavity Fitted Smoothing Contour within 2 σ Steepness Window Interval Ukonom Creek 123.3663 41.5653 -0.89011 0.18128 0 250 12.192 Goose Creek 123.89058 41.5692 0.53178 0.063145 196.9442 300 13 Copper Creek 123.64159 41.571 1.031 0.15645 1403615.821 250 12.192 Dillon Creek 123.54365 41.5722 -0.038603 0.75006 0.017271 250 12.192 Granite Creek 123.31829 41.5726 0.97821 0.4 326960.6809 250 12.192 Dillon Creek 123.54272 41.5743 0.34384 0.29608 21.2532 250 12.192 Blue Creek 123.69826 41.5787 -1.1813 0.48134 0 250 12.192 Elk Creek 123.24061 41.5826 0.45778 0.14838 90.7873 250 12.192 Ukonom Creek 123.4406 41.5843 0.81233 0.10848 112209.5497 250 12.192 Dillon Creek 123.58115 41.5843 -0.19508 0.791 0.000897 250 12.192 Ukonom Creek 123.44583 41.5845 0.69981 0.098646 15037.9285 250 12.192 Dillon Creek 123.58464 41.585 0.007434 0.44188 0.042049 250 12.192 Granite Creek 123.31495 41.5881 -2.1576 0.52718 0 250 12.192 Elk Creek 123.22057 41.589 -0.10705 0.1615 0.024638 250 12.192 Turwar Creek 123.97466 41.5893 0.73078 0.23859 7518.7355 250 12.192 Duzell Creek 122.79244 41.591 0.4177 0.041298 26.5605 250 12.192 Moffett Creek 122.77745 41.5916 0.55174 0.072296 343.3363 250 12.192 Dillon Creek 123.62359 41.5986 -0.038603 0.75006 0.017271 250 12.192 Hunter Creek 124.03656 41.599 0.8332 0.09979 18691.6242 250 12.192 South Fork Kelsey Creek 123.1952 41.5992 0.82755 0.10896 13192.1289 250 12.192 Elk Creek 123.23735 41.6004 0.62814 0.048895 3168.5987 250 12.192 Dillon Creek 123.69656 41.6016 0.35379 0.092524 18.8809 250 12.192 Independence Creek 123.3479 41.6019 -5.9239 2.5191 0 250 12.192 Hunter Creek 124.03897 41.6028 1.2801 0.20916 32694593.35 250 12.192 Elk Creek 123.26383 41.6048 0.73985 0.092583 24770.6443 250 12.192 Dillon Creek 123.58173 41.606 -0.003899 0.46788 0.033007 250 12.192 Independence Creek 123.36529 41.6081 0.81966 0.18695 65442.4729 250 12.192

269 Texas Tech University, Timothy Anderson, December 2008

Table 1 (Continued) Name of River Longitude Latitude Normalized Ksn Ksn Minimum Maximum Steepness (Low σ) (High 2 σ) Drainage Area Drainage Area Dillon Creek 123.64784 41.6121 150 152 147 8002700 58667200 Dillon Creek 123.68246 41.6137 81 97 64 4879100 8026900 South Fork Kelsey Creek 123.18324 41.6144 130 141 119 4424613 9636147 Independence Creek 123.42001 41.6285 120 124 115 12378100 46300800 Turwar Creek 123.97144 41.6285 47 55 40 1018200 3272700 Turwar Creek 123.97608 41.6353 84 106 62 3221700 5255800 South Fork Kelsey Creek 123.14375 41.6354 197 202 193 12036020 61479793 Independence Creek_unnamed123.41175 41.6387 117 119 115 714385 211759399 Bear Creek 123.24023 41.6455 132 148 116 3239500 8088700 Kelsey Creek 123.15107 41.646 169 172 166 8944100 45741700 123.21205 41.6477 62 69 55 335040 1205261 123.20834 41.6538 63 68 57 1439797 2932118 West Fork Hunter Creek 124.06034 41.655 57 71 43 334460 3741396 Dillon Creek 123.66374 41.6556 87 95 79 1942627 8271857 McAdam Creek 122.85091 41.661 57 59 54 104742 1930697729 Rattlesnake Creek 122.89572 41.6623 76 80 72 461300 161154880 Dillon Creek 123.62073 41.6637 102 104 99 26727990 204579175 Dillon Creek 123.64293 41.6681 174 183 166 9829001 21730947 Dillon Creek 123.63211 41.6689 166 174 158 14986800 31653500 Doolittle Creek 123.27683 41.6763 134 139 128 1813122 3259182 McAdam Creek 122.8257 41.678 64 67 61 121481 919989130 Doolittle Creek 123.28588 41.6786 109 118 100 3128400 3472800 Elk Creek 123.33122 41.685 125 128 121 8674090 318579263 Doolittle Creek 123.2922 41.6857 108 111 106 5103465 537448777 Southfork Clear Creek 123.54608 41.6885 128 133 124 639238 2193084 Granite Creek 123.3391 41.6906 124 129 119 13877847 298989152 Elk Creek 123.34061 41.6951 106 109 102 36788300 244007200

270 Texas Tech University, Timothy Anderson, December 2008

Table 1 (Continued) Name of River Longitude Latitude Concavity Concavity Fitted Smoothing Contour within 2 σ Steepness Window Interval Dillon Creek 123.64784 41.6121 0.3554 0.16261 30.9922 250 12.192 Dillon Creek 123.68246 41.6137 -5.9005 0.77846 0 250 12.192 South Fork Kelsey Creek 123.18324 41.6144 -1.3931 0.24036 0 250 12.192 Independence Creek 123.42001 41.6285 0.91842 0.22933 394322.4042 250 12.192 Turwar Creek 123.97144 41.6285 1.1728 0.5393 1936781.989 250 12.192 Turwar Creek 123.97608 41.6353 -2.6109 1.0688 0 250 12.192 South Fork Kelsey Creek 123.14375 41.6354 0.70593 0.1375 15330.5647 250 12.192 Independence Creek_unnamed123.41175 41.6387 0.54296 0.048109 552.121 250 12.192 Bear Creek 123.24023 41.6455 -2.2739 0.37561 0 250 12.192 Kelsey Creek 123.15107 41.646 0.67153 0.11315 7315.8853 250 12.192 123.21205 41.6477 0.86207 0.26026 14944.4847 250 12.192 123.20834 41.6538 0.86452 0.39733 26346.1281 250 12.192 West Fork Hunter Creek 124.06034 41.655 -1.1397 0.21208 0 250 12.192 Dillon Creek 123.66374 41.6556 1.1299 0.21537 2627826.141 250 12.192 McAdam Creek 122.85091 41.661 0.50795 0.052359 148.9894 250 12.192 Rattlesnake Creek 122.89572 41.6623 0.75085 0.071933 7619.3787 250 12.192 Dillon Creek 123.62073 41.6637 0.34384 0.29608 21.2532 250 12.192 Dillon Creek 123.64293 41.6681 -3.4515 1.3932 0 250 12.192 Dillon Creek 123.63211 41.6689 1.0637 0.87973 6803256.504 250 12.192 Doolittle Creek 123.27683 41.6763 2.3658 0.4571 2.52E+14 250 12.192 McAdam Creek 122.8257 41.678 0.5197 0.072436 189.8964 250 12.192 Doolittle Creek 123.28588 41.6786 -2.678 2.1669 0 250 12.192 Elk Creek 123.33122 41.685 0.69298 0.068586 10444.941 250 12.192 Doolittle Creek 123.2922 41.6857 0.56824 0.061375 879.4246 250 12.192 Southfork Clear Creek 123.54608 41.6885 1.1243 0.20584 1592280.76 250 12.192 Granite Creek 123.3391 41.6906 0.77057 0.05812 43931.0909 250 12.192 Elk Creek 123.34061 41.6951 0.86577 0.16227 260210.5643 250 12.192

271 Texas Tech University, Timothy Anderson, December 2008

Table 1 (Continued) Name of River Longitude Latitude Normalized Ksn Ksn Minimum Maximum Steepness (Low σ) (High 2 σ) Drainage Area Drainage Area Southfork Clear Creek 123.49325 41.697 73 77 69 5028237 44529585 East Fork Elk Creek 123.28473 41.6972 177 186 168 1443900 1915200 East Fork Elk Creek 123.28743 41.6988 201 214 188 1595100 1867400 East Fork Elk Creek 123.28855 41.6998 177 186 168 1443900 1915200 Bear Creek 123.33773 41.6998 145 151 138 7720412 422151333 East Fork Elk Creek 123.29108 41.7053 151 160 142 1876758 4294957 Tompkins Creek 123.10736 41.7104 155 158 153 834698 144196169 Doolittle Creek 123.34998 41.7199 106 109 104 6199000 243809200 Oak Flat Creek 123.43628 41.7303 85 87 83 622311 48044971 Oak Flat Creek 123.43641 41.7306 104 108 100 1398200 22849200 Westfork Clear Creek 123.64858 41.7319 101 103 99 829696 33503990 Clear Creek 123.55809 41.7358 96 98 94 37720425 367276858 Oak Flat Creek 123.44568 41.7442 90 93 87 985400 22850200 Tenmile Creek 123.55114 41.7457 132 149 114 33503990 52567911 East Fork Elk Creek 123.34133 41.7464 95 98 92 4294957 501619640 Mill Creek 122.9386 41.7475 111 113 110 397738 79690848 Fivemile Creek 123.50683 41.748 133 173 92 17724200 20569600 South Fork Smith 123.73153 41.7482 76 77 75 561825 101079450 123.21908 41.7486 122 123 121 3961900 110416400 Oak Flat Creek 123.46412 41.7519 85 87 83 622311 48044971 Oak Flat Creek 123.44894 41.7536 80 82 79 6646000 22839100 Tenmile Creek 123.5541 41.7631 147 167 126 25423800 30449200 South Fork Smith 123.99743 41.7659 64 68 60 102385800 1873713600 Tenmile Creek 123.55102 41.7701 122 149 95 23432400 25276500 123.21315 41.7719 137 139 136 551204 566103387 Fivemile Creek 123.50181 41.7738 132 134 130 5111200 14200800 122.76724 41.7754 63 63 62 203648 39318288

272 Texas Tech University, Timothy Anderson, December 2008

Table 1 (Continued) Name of River Longitude Latitude Concavity Concavity Fitted Smoothing Contour within 2 σ Steepness Window Interval Southfork Clear Creek 123.49325 41.697 1.0781 0.43989 3182272.427 250 12.192 East Fork Elk Creek 123.28473 41.6972 -1.5543 0.72632 0 250 12.192 East Fork Elk Creek 123.28743 41.6988 -2.1313 1.9472 0 250 12.192 East Fork Elk Creek 123.28855 41.6998 -1.5543 0.72632 0 250 12.192 Bear Creek 123.33773 41.6998 0.75121 0.039927 30065.3169 250 12.192 East Fork Elk Creek 123.29108 41.7053 1.5396 0.24012 1579789108 250 12.192 Tompkins Creek 123.10736 41.7104 0.53107 0.03773 574.7618 250 12.192 Doolittle Creek 123.34998 41.7199 0.55874 0.063584 738.3095 250 12.192 Oak Flat Creek 123.43628 41.7303 0.44699 0.10925 86.589 250 12.192 Oak Flat Creek 123.43641 41.7306 0.8651 0.077659 73602.6607 250 12.192 Westfork Clear Creek 123.64858 41.7319 0.32217 0.066821 13.9701 250 12.192 Clear Creek 123.55809 41.7358 0.4151 0.58324 63.5459 250 12.192 Oak Flat Creek 123.44568 41.7442 0.45776 0.17129 103.5099 250 12.192 Tenmile Creek 123.55114 41.7457 2.4554 6.3531 2.17E+17 250 12.192 East Fork Elk Creek 123.34133 41.7464 0.51793 0.084812 326.379 250 12.192 Mill Creek 122.9386 41.7475 0.51154 0.035486 310.0754 250 12.192 Fivemile Creek 123.50683 41.748 -4.4501 4.0735 0 250 12.192 South Fork Smith 123.73153 41.7482 0.49026 0.04635 161.494 300 13 123.21908 41.7486 0.49061 0.04663 252.7795 250 12.192 Oak Flat Creek 123.46412 41.7519 0.44699 0.10925 86.589 250 12.192 Oak Flat Creek 123.44894 41.7536 0.60105 0.2924 1006.2638 250 12.192 Tenmile Creek 123.5541 41.7631 2.7652 1.8932 2.87E+19 250 12.192 South Fork Smith 123.99743 41.7659 1.0577 0.17235 10919326.89 300 13 Tenmile Creek 123.55102 41.7701 2.1881 17.9834 8.62E+14 250 12.192 123.21315 41.7719 0.43819 0.03621 110.5134 250 12.192 Fivemile Creek 123.50181 41.7738 0.74507 0.15787 15061.3207 250 12.192 122.76724 41.7754 0.45669 0.040979 74.4197 250 12.192

273 Texas Tech University, Timothy Anderson, December 2008

Table 1 (Continued) Name of River Longitude Latitude Normalized Ksn Ksn Minimum Maximum Steepness (Low σ) (High 2 σ) Drainage Area Drainage Area Elk Crek 123.38694 41.7777 108 111 106 5103465 537448777 Oak Flat Creek 123.47214 41.7777 104 108 100 1398200 22849200 Elk Creek 123.38812 41.7785 121 125 117 14869099 661035402 Tenmile Creek 123.54987 41.7817 104 110 98 16812400 23427500 Oak Flat Creek 123.47511 41.7839 90 98 81 3587617 4459038 Goose Creek 124.04272 41.7859 53 56 49 55316025 1873714275 123.16366 41.7879 137 138 136 347169 356552554 Oak Flat Crek 123.48019 41.7893 104 108 100 1398200 22849200 Hurdygurdy Creek 124.02727 41.7916 63 66 59 46540350 1856123775 Preston Creek 123.62345 41.7927 144 152 135 15648400 17284700 Fivemile Creek 123.51226 41.7947 99 101 96 1002968 2193084 Oak Flat Creek 123.4813 41.7969 143 161 126 626300 1398200 Hurdygurdy Creek 123.80105 41.7975 84 87 82 69525 46443825 Tenmile Creek 123.54869 41.7978 110 119 101 7704490 14270532 Preston Creek 123.61174 41.8003 155 163 148 9073100 15648400 Clear Creek 123.63918 41.8063 103 105 100 30682545 65535920 Tenmile Creek 123.54797 41.811 108 111 105 1771704 6373468 122.53145 41.8119 39 41 37 70074 1163154 Preston Crek 123.59897 41.8172 120 127 114 1076899 7889326 Doolittle Creek 123.40334 41.821 114 122 106 4008633 4258076730 Little South Fork Indian 123.49337 41.8235 100 109 91 522700 8302800 Clear Creek 123.65099 41.8291 78 87 68 21011900 34338500 Doolittle Creek 123.45953 41.831 57 66 47 1462300 1731200 122.20537 41.8311 24 26 23 119700 21870600 Doolittle Creek 123.45395 41.8318 95 110 79 1728100 1856300 Doolittle Creek 123.44776 41.8321 144 179 110 1853600 4895700 122.63349 41.8347 83 100 66 12385400 33399000

274 Texas Tech University, Timothy Anderson, December 2008

Table 1 (Continued) Name of River Longitude Latitude Concavity Concavity Fitted Smoothing Contour within 2 σ Steepness Window Interval Elk Crek 123.38694 41.7777 0.56824 0.061375 879.4246 250 12.192 Oak Flat Creek 123.47214 41.7777 0.8651 0.077659 73602.6607 250 12.192 Elk Creek 123.38812 41.7785 0.73985 0.092583 24770.6443 250 12.192 Tenmile Creek 123.54987 41.7817 1.5511 2.0651 11598860235 250 12.192 Oak Flat Creek 123.47511 41.7839 2.1315 2.651 1.16988E+13 250 12.192 Goose Creek 124.04272 41.7859 0.73087 0.14103 13380.4925 300 13 123.16366 41.7879 0.4815 0.022736 224.9615 250 12.192 Oak Flat Crek 123.48019 41.7893 0.8651 0.077659 73602.6607 250 12.192 Hurdygurdy Creek 124.02727 41.7916 0.67842 0.10207 5212.5455 300 13 Preston Creek 123.62345 41.7927 -3.4972 4.6431 0 250 12.192 Fivemile Creek 123.51226 41.7947 0.73817 0.27703 6108.4335 250 12.192 Oak Flat Creek 123.4813 41.7969 -1.5751 0.54784 0 250 12.192 Hurdygurdy Creek 123.80105 41.7975 0.48002 0.058726 130.7602 300 13 Tenmile Creek 123.54869 41.7978 1.308 1.6246 128444495.3 250 12.192 Preston Creek 123.61174 41.8003 2.2853 0.58937 1.20E+15 250 12.192 Clear Creek 123.63918 41.8063 -0.25737 0.87584 0.000389 250 12.192 Tenmile Creek 123.54797 41.811 0.75729 0.16976 11835.8193 250 12.192 122.53145 41.8119 0.74725 0.071637 1789.4544 250 12.192 Preston Crek 123.59897 41.8172 -0.17785 0.07439 0.009791 250 12.192 Doolittle Creek 123.40334 41.821 0.80674 0.12037 42972.7525 250 12.192 Little South Fork Indian 123.49337 41.8235 -0.12783 0.12742 0.016815 250 12.192 Clear Creek 123.65099 41.8291 3.0218 4.874 6.84E+20 250 12.192 Doolittle Creek 123.45953 41.831 -4.124 6.1904 0 250 12.192 122.20537 41.8311 0.3703 0.093187 8.9778 250 12.192 Doolittle Creek 123.45395 41.8318 9.7524 2.6188 1.38E+60 250 12.192 Doolittle Creek 123.44776 41.8321 -1.1663 0.24555 0 250 12.192 122.63349 41.8347 2.6502 0.59352 5.71E+17 250 12.192

275 Texas Tech University, Timothy Anderson, December 2008

Table 1 (Continued) Name of River Longitude Latitude Normalized Ksn Ksn Minimum Maximum Steepness (Low σ) (High 2 σ) Drainage Area Drainage Area 122.71879 41.8377 51 58 43 45328428 508819453 Clear Creek 123.6501 41.8435 58 62 53 17099900 21620700 122.48259 41.8456 27 29 25 1294104 288044415 123.00205 41.8483 137 159 116 26589516 181420876 Middle Fork Smith 124.04909 41.8508 48 55 42 228875625 1873714275 Little South Fork Indian 123.43082 41.8537 94 100 89 9173748 1364053677 East Fork Indian 123.39745 41.8566 102 109 96 30682545 436967803 Clear Creek 123.64076 41.8573 75 82 68 7033200 17099900 122.76482 41.8638 80 105 55 17148100 269535100 123.05821 41.8643 75 81 69 16747046 268269580 123.02797 41.8656 141 150 133 4655364 277981212 122.62852 41.8684 112 114 110 312039 10929745 122.7568 41.8718 90 98 82 9038100 270061000 Clear Creek 123.63465 41.8747 70 79 60 644153 7720412 122.7304 41.8767 105 106 104 2446100 262614200 123.17228 41.8773 131 134 127 461416 163062919 Thompson Creek 123.32344 41.8785 117 122 112 49301100 92805900 Middle Fork Smith 123.66626 41.8811 57 58 57 272342 13482517 122.17004 41.8822 124 132 116 31763657 48669702 123.14592 41.8843 108 111 105 551204 216724164 122.74514 41.8891 146 184 107 14896800 16976100 Indian Creek 123.43427 41.8896 88 90 85 3373571 616967337 Baldface Creek 124.04181 41.8955 52 54 49 68816025 1873855350 122.97788 41.8962 213 219 207 7133158 20005910 122.3568 41.8966 109 113 105 9823766 411054696 Middle Fork Smith 123.77985 41.9011 54 56 51 26383050 228680775 Thompson Creek 123.33726 41.9012 112 117 107 31203872 142283046

276 Texas Tech University, Timothy Anderson, December 2008

Table 1 (Continued) Name of River Longitude Latitude Concavity Concavity Fitted Smoothing Contour within 2 σ Steepness Window Interval 122.71879 41.8377 1.582 0.50571 68604388337 250 12.192 Clear Creek 123.6501 41.8435 1.181 3.6847 13425363.43 250 12.192 122.48259 41.8456 0.39762 0.11119 16.0494 250 12.192 123.00205 41.8483 1.5664 0.68751 43154053546 250 12.192 Middle Fork Smith 124.04909 41.8508 1.2685 0.38173 827107729.6 300 13 Little South Fork Indian 123.43082 41.8537 0.76323 0.076031 29226.0229 250 12.192 East Fork Indian 123.39745 41.8566 0.78132 0.081653 42861.0535 250 12.192 Clear Creek 123.64076 41.8573 -0.58706 0.36146 0.000004 250 12.192 122.76482 41.8638 0.89895 0.4077 266819.3128 250 12.192 123.05821 41.8643 1.4194 0.21751 2459071916 250 12.192 123.02797 41.8656 1.1694 0.12539 23440103.11 250 12.192 122.62852 41.8684 0.56474 0.055149 608.4125 250 12.192 122.7568 41.8718 0.79437 0.20605 36239.4608 250 12.192 Clear Creek 123.63465 41.8747 -0.35394 0.12531 0.000388 250 12.192 122.7304 41.8767 0.51555 0.074949 310.8684 250 12.192 123.17228 41.8773 0.67767 0.036397 4421.249 250 12.192 Thompson Creek 123.32344 41.8785 0.15428 0.72791 0.54807 250 12.192 Middle Fork Smith 123.66626 41.8811 0.43308 0.064872 47.2263 300 13 122.17004 41.8822 1.8482 0.78228 5.28877E+12 250 12.192 123.14592 41.8843 0.65309 0.045871 2932.9144 250 12.192 122.74514 41.8891 13.6149 4.1665 1.06E+97 250 12.192 Indian Creek 123.43427 41.8896 0.63069 0.04561 2182.3039 250 12.192 Baldface Creek 124.04181 41.8955 0.79948 0.12633 50118.749 300 13 122.97788 41.8962 0.79872 0.19235 68003.8603 250 12.192 122.3568 41.8966 0.74932 0.086821 21843.0857 250 12.192 Middle Fork Smith 123.77985 41.9011 0.81347 0.17638 41268.6916 300 13 Thompson Creek 123.33726 41.9012 -0.1737 0.58962 0.001416 250 12.192

277 Texas Tech University, Timothy Anderson, December 2008

Table 1 (Continued) Name of River Longitude Latitude Normalized Ksn Ksn Minimum Maximum Steepness (Low σ) (High 2 σ) Drainage Area Drainage Area Thompson Creek 123.34225 41.9045 119 129 108 62084800 76473700 Dunn Creek 123.57534 41.909 72 80 65 598200 1748700 122.20664 41.9119 85 92 79 435200 11434100 123.01013 41.9132 159 170 148 939665 5366977 Middle Fork Smith 123.6934 41.9138 149 160 138 13710150 25977150 122.58949 41.9153 89 92 86 4230800 256253800 122.14786 41.9158 115 153 76 64798400 71403700 122.73888 41.9167 90 95 84 3824900 8800900 Thompson Creek 123.34977 41.9168 117 122 112 49301100 92805900 122.69048 41.9174 80 103 56 2147500 2449700 122.6519 41.9182 64 67 61 425000 4228200 122.81865 41.9184 137 145 130 126632400 280537100 122.56779 41.9187 51 58 44 28710300 256317300 122.81854 41.92 140 148 132 126608500 280253800 122.97292 41.92 129 136 122 786598 4335764 122.81796 41.9202 139 146 131 126410800 280364800 122.75589 41.9209 55 57 54 23400 14396600 122.68252 41.922 97 100 93 284700 1537200 North Fork Smith River 124.03668 41.9224 49 50 47 375656 25216259914 123.05445 41.9251 187 195 180 5897100 17397000 122.36101 41.9253 94 95 93 1008930 566103387 Thompson Creek 123.34773 41.9256 97 107 88 47753100 61109500 122.95138 41.9279 60 62 59 32100 2180300 122.15955 41.9281 108 149 66 71557000 76660100 122.58096 41.9298 60 67 52 145616800 256323500 122.85839 41.9299 136 139 133 2508500 280405900 122.58111 41.93 60 67 52 145614700 256317300

278 Texas Tech University, Timothy Anderson, December 2008

Table 1 (Continued) Name of River Longitude Latitude Concavity Concavity Fitted Smoothing Contour within 2 σ Steepness Window Interval Thompson Creek 123.34225 41.9045 -5.0235 1.1863 0 250 12.192 Dunn Creek 123.57534 41.909 1.2415 0.26857 3918282.307 250 12.192 122.20664 41.9119 -0.14594 0.071428 0.00889 250 12.192 123.01013 41.9132 -0.36422 0.074591 0.001324 250 12.192 Middle Fork Smith 123.6934 41.9138 1.9719 0.45714 1.86198E+13 300 13 122.58949 41.9153 0.69454 0.078981 5068.5934 250 12.192 122.14786 41.9158 12.4682 4.0243 1.88E+96 250 12.192 122.73888 41.9167 1.1762 0.45351 7845705.513 250 12.192 Thompson Creek 123.34977 41.9168 0.15428 0.72791 0.54807 250 12.192 122.69048 41.9174 6.5314 3.886 4.16E+40 250 12.192 122.6519 41.9182 0.46569 0.17234 86.3851 250 12.192 122.81865 41.9184 1.1462 0.35415 84406101.46 250 12.192 122.56779 41.9187 0.95688 0.61929 1009284.346 250 12.192 122.81854 41.92 1.22 0.35512 338907160.4 250 12.192 122.97292 41.92 0.62576 0.16015 1649.7287 250 12.192 122.81796 41.9202 1.0614 0.36438 16456999.34 250 12.192 122.75589 41.9209 0.34704 0.027601 14.0169 250 12.192 122.68252 41.922 0.60084 0.13216 747.6008 250 12.192 North Fork Smith River 124.03668 41.9224 0.54356 0.031789 272.4953 300 13 123.05445 41.9251 -0.3733 0.26159 0.000234 250 12.192 122.36101 41.9253 0.47295 0.044231 137.5045 250 12.192 Thompson Creek 123.34773 41.9256 -2.9335 3.5068 0 250 12.192 122.95138 41.9279 0.38189 0.072293 25.8 250 12.192 122.15955 41.9281 7.9373 6.7596 1.01E+61 250 12.192 122.58096 41.9298 2.6084 0.76008 5.16E+19 250 12.192 122.85839 41.9299 0.50776 0.045114 360.687 250 12.192 122.58111 41.93 2.6084 0.76008 5.16E+19 250 12.192

279 Texas Tech University, Timothy Anderson, December 2008

Table 1 (Continued) Name of River Longitude Latitude Normalized Ksn Ksn Minimum Maximum Steepness (Low σ) (High 2 σ) Drainage Area Drainage Area 122.58134 41.9302 60 67 52 145533800 256307700 122.5817 41.9305 60 67 52 145533800 256251700 122.58183 41.9306 60 67 52 145231000 256254000 122.58188 41.9307 60 67 52 145533800 256230600 122.84266 41.9307 117 119 116 5069900 280406500 123.05106 41.9342 149 170 128 3187000 6077500 122.727 41.9353 63 65 62 73100 3813600 123.2397 41.9409 94 100 89 13708200 19869600 123.0646 41.9446 88 96 80 43600 5897100 Tho9mpson Creek 123.34038 41.9492 113 116 111 1634870 130648633 122.58429 41.9512 64 78 50 24724600 28677300 122.68988 41.9515 63 64 62 89881 10929745 123.2641 41.9535 73 81 65 3568200 13090500 123.21389 41.9554 78 81 75 19060325 38001834 East Fork Indian 123.41418 41.9566 95 97 92 4639400 33379600 122.22735 41.9638 168 186 149 78070500 919476400 122.21617 41.9643 170 190 150 12510600 881002000 122.64474 41.9663 86 90 82 18857200 25324000 Thompson Creek 123.37261 41.9675 98 100 96 765412 49731231 122.5636 41.9779 67 70 64 9575700 24537600 122.77572 41.9823 86 97 74 54197900 122152800 South Fork Winchuck 124.15044 41.9837 38 45 32 1652369 35599207 123.30101 41.9861 44 47 42 1102415 8737116 122.22335 41.9865 57 73 41 486304200 953018300 Indian Creek 123.52073 41.9889 81 84 77 236882 2230050 123.17955 41.9889 91 94 87 31554500 468543600 122.88227 41.9893 159 175 144 1847600 3939600

280 Texas Tech University, Timothy Anderson, December 2008

Table 1 (Continued) Name of River Longitude Latitude Concavity Concavity Fitted Smoothing Contour within 2 σ Steepness Window Interval 122.58134 41.9302 2.6084 0.76008 5.16E+19 250 12.192 122.5817 41.9305 2.6084 0.76008 5.16E+19 250 12.192 122.58183 41.9306 2.6084 0.76008 5.16E+19 250 12.192 122.58188 41.9307 2.6084 0.76008 5.16E+19 250 12.192 122.84266 41.9307 0.52594 0.054965 484.4713 250 12.192 123.05106 41.9342 1.7228 0.43505 39902211569 250 12.192 122.727 41.9353 0.33916 0.044123 14.4993 250 12.192 123.2397 41.9409 2.8151 1.4185 1.03E+19 250 12.192 123.0646 41.9446 -0.11616 0.064147 0.037403 250 12.192 Tho9mpson Creek 123.34038 41.9492 0.55034 0.041934 651.3344 250 12.192 122.58429 41.9512 9.7528 6.5409 8.17E+70 250 12.192 122.68988 41.9515 0.44182 0.043784 62.0515 250 12.192 123.2641 41.9535 1.3086 0.16807 59022286.28 250 12.192 123.21389 41.9554 1.2433 0.70942 68668875.29 250 12.192 East Fork Indian 123.41418 41.9566 -0.074851 0.12266 0.015357 250 12.192 122.22735 41.9638 1.0285 0.36426 9784479.25 250 12.192 122.21617 41.9643 0.71981 0.13948 24412.6494 250 12.192 122.64474 41.9663 2.2328 1.0549 1.10E+15 250 12.192 Thompson Creek 123.37261 41.9675 0.53636 0.047767 380.8791 250 12.192 122.5636 41.9779 1.1191 0.36942 4749601.424 250 12.192 122.77572 41.9823 2.3049 0.82673 4.99E+16 250 12.192 South Fork Winchuck 124.15044 41.9837 1.1928 0.19256 4820958.553 250 12.192 123.30101 41.9861 0.76103 0.12034 4980.3621 250 12.192 122.22335 41.9865 3.4253 1.2419 9.92E+27 250 12.192 Indian Creek 123.52073 41.9889 0.65966 0.14645 1417.2635 250 12.192 123.17955 41.9889 0.66765 0.18683 4962.0997 250 12.192 122.88227 41.9893 0.90565 0.60908 151750.6598 250 12.192

281 Texas Tech University, Timothy Anderson, December 2008

Table 1 (Continued) Name of River Longitude Latitude Normalized Ksn Ksn Minimum Maximum Steepness (Low σ) (High 2 σ) Drainage Area Drainage Area 123.23884 41.99 132 163 101 24040100 31032300 123.16704 41.9927 99 110 88 44213500 467786600 122.48594 41.9935 56 57 55 301137 136500781 122.47201 41.995 71 73 69 8223523 80070936 123.02992 41.9971 68 106 30 67685300 81427200 122.60696 41.9973 60 61 59 400236 46969363 123.26854 41.9982 66 75 57 12672000 18691800 122.6413 42.0005 45 50 40 30653953 127129724 122.79938 42.005 123 125 122 3822400 122235000 East Fork Indian 123.43957 42.0064 70 74 66 733700 5464200 122.55941 42.007 120 133 107 3345500 8479900 122.95081 42.0074 135 137 133 3352600 66937700 East Fork Indian 123.44743 42.0102 84 87 80 523722 1198538 122.75559 42.0103 91 95 87 4199300 126707900 122.46214 42.0122 86 87 85 2285985 779636013 Bear Creek 124.14261 42.0125 24 27 21 406493 9719732685 122.78448 42.0127 123 128 119 7611600 54077500 122.15772 42.0132 100 115 85 486472900 514892900 122.25192 42.0142 63 79 47 76675600 122895000 122.66046 42.0149 72 74 70 2285985 118402010 Sucker Creek 123.37371 42.0181 49 51 48 257003 11212164 122.86544 42.0189 68 70 67 1205261 3629490 122.67177 42.0195 94 98 90 2285985 118402010 122.56443 42.0195 89 91 87 551204 3148252 122.64011 42.0209 53 54 52 1205261 118402010 Althouse Creek 123.49961 42.0217 76 79 74 102902 18408159 122.1814 42.0221 233 252 214 60245246 118402010

282 Texas Tech University, Timothy Anderson, December 2008

Table 1 (Continued) Name of River Longitude Latitude Concavity Concavity Fitted Smoothing Contour within 2 σ Steepness Window Interval 123.23884 41.99 6.0458 4.3688 5.47E+43 250 12.192 123.16704 41.9927 1.1058 0.28446 19309183.15 250 12.192 122.48594 41.9935 0.48195 0.046324 99.5789 250 12.192 122.47201 41.995 0.70238 0.27575 5422.4958 250 12.192 123.02992 41.9971 16.833 3.3664 7.34E+130 250 12.192 122.60696 41.9973 0.4857 0.092656 103.607 250 12.192 123.26854 41.9982 2.0967 1.6564 5.60292E+13 250 12.192 122.6413 42.0005 0.99039 0.49416 835854.2042 250 12.192 122.79938 42.005 0.48075 0.073591 208.7645 250 12.192 East Fork Indian 123.43957 42.0064 0.71997 0.10136 3362.6511 250 12.192 122.55941 42.007 1.5253 0.33099 1688319487 250 12.192 122.95081 42.0074 0.36672 0.077916 34.5585 250 12.192 East Fork Indian 123.44743 42.0102 1.4375 0.47721 59641117.3 250 12.192 122.75559 42.0103 0.63182 0.10978 2146.5843 250 12.192 122.46214 42.0122 0.59445 0.0686 958.1345 250 12.192 Bear Creek 124.14261 42.0125 0.69522 0.10738 1501.9603 250 12.192 122.78448 42.0127 0.6178 0.16238 2363.9152 250 12.192 122.15772 42.0132 16.7928 13.464 1.55E+144 250 12.192 122.25192 42.0142 -10.4365 2.1455 0 250 12.192 122.66046 42.0149 0.54195 0.089535 322.6594 250 12.192 Sucker Creek 123.37371 42.0181 0.52067 0.082229 142.5704 250 12.192 122.86544 42.0189 0.65625 0.16142 1459.9835 250 12.192 122.67177 42.0195 0.79283 0.057467 22844.179 250 12.192 122.56443 42.0195 0.52827 0.12338 288.849 250 12.192 122.64011 42.0209 0.53085 0.062575 204.8784 250 12.192 Althouse Creek 123.49961 42.0217 0.29356 0.030135 8.1739 250 12.192 122.1814 42.0221 -4.6733 7.2623 0 250 12.192

283 Texas Tech University, Timothy Anderson, December 2008

Table 1 (Continued) Name of River Longitude Latitude Normalized Ksn Ksn Minimum Maximum Steepness (Low σ) (High 2 σ) Drainage Area Drainage Area 4th of July 124.12628 42.0255 31 34 28 163679 2579423390 122.52194 42.0274 63 64 62 571159 7658961 122.62073 42.0328 54 60 48 3502689 22258216 123.07145 42.0344 95 99 92 48219800 77744900 East Fork Winchuck 124.12355 42.0374 31 34 29 473039 486679214 123.291 42.0409 73 75 71 10003600 58895100 122.97077 42.0422 99 101 98 415500 40483300 Wheeler Creek 124.13641 42.0472 29 30 28 1267304 2216557553 122.78148 42.0495 62 63 61 19900 6448000 122.685 42.0581 128 156 101 844580 1782236 122.49347 42.0583 66 67 66 203648 1163154 122.08328 42.0635 64 83 44 378219000 445832900 123.20858 42.0656 56 63 49 60229900 159418500 122.35496 42.0667 39 45 33 269606400 424402300 122.35518 42.0686 39 45 33 269644500 423254500 122.3555 42.0695 39 45 33 269449200 423293700 122.07579 42.0748 85 100 71 368182100 442404500 122.3797 42.0799 83 92 74 54346000 424396600 Baldface Creek 123.87281 42.0828 47 48 46 135225 67617225 122.97507 42.0847 98 104 91 1891790 11107149 122.81675 42.0859 105 107 102 2966967 41580768 122.14717 42.0874 37 39 35 2829681 52257270 122.36296 42.0899 70 76 64 13112900 423407200 Althouse Creek 123.55145 42.0914 67 74 60 24040992 195850852 122.86724 42.0927 127 130 123 4167700 33951800 123.09269 42.1008 41 45 38 578351000 707056400 Dunn Creek 123.63156 42.1014 75 81 69 4510493 1104643150

284 Texas Tech University, Timothy Anderson, December 2008

Table 1 (Continued) Name of River Longitude Latitude Concavity Concavity Fitted Smoothing Contour within 2 σ Steepness Window Interval 4th of July 124.12628 42.0255 0.69003 0.044406 1471.2922 250 12.192 122.52194 42.0274 0.44916 0.12755 66.1453 250 12.192 122.62073 42.0328 1.6458 0.43317 8368680561 250 12.192 123.07145 42.0344 2.6049 0.96735 7.33E+18 250 12.192 East Fork Winchuck 124.12355 42.0374 0.72212 0.087752 2966.251 250 12.192 123.291 42.0409 0.56251 0.15763 512.0188 250 12.192 122.97077 42.0422 0.52003 0.058929 284.6444 250 12.192 Wheeler Creek 124.13641 42.0472 0.61327 0.089351 423.4248 250 12.192 122.78148 42.0495 0.42517 0.046526 44.2738 250 12.192 122.685 42.0581 4.3735 1.6021 1.69E+26 250 12.192 122.49347 42.0583 0.29955 0.1232 9.0316 250 12.192 122.08328 42.0635 11.2378 3.8317 4.30E+94 250 12.192 123.20858 42.0656 1.6384 0.89603 2.13162E+11 250 12.192 122.35496 42.0667 2.7632 1.5522 3.07E+21 250 12.192 122.35518 42.0686 2.7632 1.5522 3.07E+21 250 12.192 122.3555 42.0695 2.7632 1.5522 3.07E+21 250 12.192 122.07579 42.0748 9.2931 2.9886 7.76E+77 250 12.192 122.3797 42.0799 1.0972 0.14861 13720377.76 250 12.192 Baldface Creek 123.87281 42.0828 0.44269 0.04289 45.0178 300 13 122.97507 42.0847 0.95157 0.21683 208928.2421 250 12.192 122.81675 42.0859 0.54274 0.086653 468.5817 250 12.192 122.14717 42.0874 -0.074263 0.30603 0.006551 250 12.192 122.36296 42.0899 0.72331 0.12349 10202.8124 250 12.192 Althouse Creek 123.55145 42.0914 1.7617 0.23617 7.65674E+11 250 12.192 122.86724 42.0927 0.76192 0.069957 19969.103 250 12.192 123.09269 42.1008 2.1034 5.4153 1.66E+16 250 12.192 Dunn Creek 123.63156 42.1014 0.8335 0.062125 67847.9426 350 13

285 Texas Tech University, Timothy Anderson, December 2008

Table 1 (Continued) Name of River Longitude Latitude Normalized Ksn Ksn Minimum Maximum Steepness (Low σ) (High 2 σ) Drainage Area Drainage Area Dunn Creek 123.63184 42.1024 72 77 66 5643555 6542368311 North Fork Chetco 124.23552 42.1027 31 36 27 40367000 857885700 Elk Creek 123.68868 42.1034 40 44 36 438351 14573406421 North Fork Chetco 124.23455 42.1068 52 65 39 19411802 333145342 122.4661 42.1168 56 103 8 51666200 54092500 122.8225 42.1185 98 121 74 22973200 51400300 122.21031 42.1236 21 25 17 1491919 32913534 Emily Creek 124.13403 42.1313 50 53 47 391375 195966154 122.47949 42.1339 58 92 24 44286400 49532300 South Fork Chetco 124.14025 42.135 33 35 32 12793692 1000000000 122.34705 42.1381 48 58 37 192424000 266859500 122.34427 42.1404 65 75 54 148307000 269491100 122.34303 42.1413 68 80 57 148003200 269449200 South Fork Chetco_Upstream123.98504 42.1442 128 130 125 406493 13287913 Sucker Creek 123.55511 42.1447 67 71 64 12571262 4641588834 North Fork Rough and Ready123.7149 Creek 42.1447 48 50 47 44467 3692221821 122.97524 42.1465 122 125 118 1580145 1934839527 123.04677 42.1472 78 80 77 12523331 3126900101 122.48528 42.1589 46 67 24 29038800 43902500 North Fork Chetco 124.22148 42.1615 42 44 39 5481700 17738300 North Fork Chetco 124.27283 42.1633 41 58 24 12082100 39061800 122.40093 42.1636 108 139 78 7800300 10690000 122.95317 42.1662 96 99 93 53330700 1098011500 122.96786 42.1703 90 92 87 35196500 1184033700 122.41004 42.1756 21 23 19 301137 6643452 Little Chetco 123.85723 42.1771 52 55 49 804174 28357954 122.32589 42.1774 51 61 41 28099600 191974100

286 Texas Tech University, Timothy Anderson, December 2008

Table 1 (Continued) Name of River Longitude Latitude Concavity Concavity Fitted Smoothing Contour within 2 σ Steepness Window Interval Dunn Creek 123.63184 42.1024 0.83639 0.065436 70309.9606 250 12.192 North Fork Chetco 124.23552 42.1027 1.7389 0.66976 5.1887E+11 250 12.192 Elk Creek 123.68868 42.1034 0.72103 0.042578 4007.0222 250 12.192 North Fork Chetco 124.23455 42.1068 1.7497 0.4458 5.48279E+11 250 12.192 122.4661 42.1168 26.9944 13.0856 6.34E+206 250 12.192 122.8225 42.1185 7.165 2.8184 2.51E+53 250 12.192 122.21031 42.1236 0.73753 0.37936 2117.6907 250 12.192 Emily Creek 124.13403 42.1313 0.60436 0.092304 593.5504 250 12.192 122.47949 42.1339 9.0337 2.2443 4.76E+67 250 12.192 South Fork Chetco 124.14025 42.135 0.62306 0.14364 796.5809 250 12.192 122.34705 42.1381 5.0903 5.4349 2.36E+40 250 12.192 122.34427 42.1404 3.0452 1.3773 2.66E+23 250 12.192 122.34303 42.1413 2.98 1.1895 7.59E+22 250 12.192 South Fork Chetco_Upstream 123.98504 42.1442 0.57629 0.055219 791.1256 250 12.192 Sucker Creek 123.55511 42.1447 0.76306 0.063904 20841.4919 250 12.192 North Fork Rough and Ready 123.7149Creek 42.1447 0.51909 0.029092 147.2414 250 12.192 122.97524 42.1465 0.64768 0.025004 3426.543 250 12.192 123.04677 42.1472 0.55924 0.067142 576.5036 250 12.192 122.48528 42.1589 9.1826 4.8276 2.91E+67 250 12.192 North Fork Chetco 124.22148 42.1615 0.97454 0.39706 235649.7838 250 12.192 North Fork Chetco 124.27283 42.1633 2.5652 0.93655 1.42E+17 250 12.192 122.40093 42.1636 2.6067 2.3996 1.26E+17 250 12.192 122.95317 42.1662 0.62053 0.1333 2404.8872 250 12.192 122.96786 42.1703 0.84942 0.13803 189612.7392 250 12.192 122.41004 42.1756 0.42942 0.27769 16.8875 250 12.192 Little Chetco 123.85723 42.1771 0.69329 0.071958 2242.194 250 12.192 122.32589 42.1774 1.0283 0.7596 1756682.154 250 12.192

287 Texas Tech University, Timothy Anderson, December 2008

Table 1 (Continued) Name of River Longitude Latitude Normalized Ksn Ksn Minimum Maximum Steepness (Low σ) (High 2 σ) Drainage Area Drainage Area Chetco 123.90549 42.1794 57 58 56 2917565 67807186 123.2148 42.185 47 49 45 3447143 371563096 122.46145 42.1921 18 29 8 2265500 27904500 123.38303 42.1937 134 136 132 704573 3848163 122.26742 42.1938 66 78 55 9028600 144275400 North Fork Chetco_Upstream124.25421 42.1938 69 76 63 547000 11866900 122.38302 42.1948 32 36 27 1953600 8720800 122.27636 42.1952 59 63 54 1042900 144954500 Bravo Creek 124.20261 42.1956 26 33 18 616770 3951007 122.35393 42.1968 49 64 34 9162500 28023100 Tincup Creek 124.12805 42.199 50 56 44 36451200 893254900 122.99077 42.2021 44 47 41 83526 3222131959 Box Canyon 123.9613 42.2043 75 77 74 571744 12317852 Box Canyon 124.09549 42.2048 53 56 50 12317852 18513550367 122.07402 42.2048 49 56 41 27843700 212454900 122.08524 42.2069 86 105 68 94994900 212114500 Bear Creek 124.09535 42.2074 42 45 39 150196800 893390400 Chetco 124.0856 42.2164 42 45 39 150196800 881529800 123.29617 42.2164 73 80 67 3688249 1043357590 Chetco 123.88543 42.2339 76 94 59 88410136 119726228 Chetco 123.88318 42.2374 55 68 41 94037200 136539300 122.24563 42.2564 20 21 18 88200 7575500 123.01127 42.2598 38 39 36 1102415 319805548 123.0506 42.2681 56 57 55 2711595 618772415 Pistol River 124.32031 42.2811 67 74 60 24368639 505479682 123.13713 42.282 65 67 64 193484 69245392 East Fork Pistol 124.29536 42.2868 76 83 68 24368639 451150027

288 Texas Tech University, Timothy Anderson, December 2008

Table 1 (Continued) Name of River Longitude Latitude Concavity Concavity Fitted Smoothing Contour within 2 σ Steepness Window Interval Chetco 123.90549 42.1794 0.50098 0.10037 138.4933 250 12.192 123.2148 42.185 0.59653 0.085622 612.2059 250 12.192 122.46145 42.1921 1.28 0.35215 13061373.68 250 12.192 123.38303 42.1937 0.44993 0.089582 134.4308 250 12.192 122.26742 42.1938 1.255 0.22608 53255954.58 250 12.192 North Fork Chetco_Upstream 124.25421 42.1938 0.91389 0.078004 70168.2476 250 12.192 122.38302 42.1948 0.94997 0.31157 73620.5967 250 12.192 122.27636 42.1952 0.75891 0.086145 6132.9397 250 12.192 Bravo Creek 124.20261 42.1956 1.762 0.37565 5011278582 250 12.192 122.35393 42.1968 2.5505 0.81731 4.05E+16 250 12.192 Tincup Creek 124.12805 42.199 0.81343 0.15293 52957.4687 250 12.192 122.99077 42.2021 0.33214 0.043506 8.4589 250 12.192 Box Canyon 123.9613 42.2043 0.43162 0.089524 57.2887 250 12.192 Box Canyon 124.09549 42.2048 0.70886 0.071804 6603.9559 250 12.192 122.07402 42.2048 1.3077 0.32074 265275392.2 250 12.192 122.08524 42.2069 3.3996 0.87021 5.71E+25 250 12.192 Bear Creek 124.09535 42.2074 1.2549 0.32223 319083497 250 12.192 Chetco 124.0856 42.2164 1.2549 0.32223 319083497 250 12.192 123.29617 42.2164 0.88855 0.049635 130931.3854 250 12.192 Chetco 123.88543 42.2339 5.5519 2.0376 6.47E+42 250 12.192 Chetco 123.88318 42.2374 4.6257 3.0486 2.23E+35 250 12.192 122.24563 42.2564 0.46278 0.13274 29.2812 250 12.192 123.01127 42.2598 0.33276 0.085365 6.2736 250 12.192 123.0506 42.2681 0.54241 0.070535 282.67 250 12.192 Pistol River 124.32031 42.2811 0.95129 0.23277 696598.9012 250 12.192 123.13713 42.282 0.41555 0.046825 42.0805 250 12.192 East Fork Pistol 124.29536 42.2868 0.92054 0.23583 441181.5441 250 12.192

289 Texas Tech University, Timothy Anderson, December 2008

Table 1 (Continued) Name of River Longitude Latitude Normalized Ksn Ksn Minimum Maximum Steepness (Low σ) (High 2 σ) Drainage Area Drainage Area 122.12242 42.2929 70 80 59 6294200 20452700 East Fork Pistol 124.16536 42.2982 100 119 81 9095943 17994676 123.37057 42.2989 86 89 83 269133 111907691 North Fork Pistol 124.28712 42.2999 78 82 75 2074304 346014780 Tincup Creek 123.93431 42.3345 94 97 90 1009521 36974408 East Fork Pistol 124.14831 42.3388 32 39 25 593831 4597814 South Fork Collier Creek_Upstream124.12705 42.3469 83 94 71 1708900 2881400 123.47016 42.3475 50 55 46 218154 78074308 North Fork Pistol 124.26092 42.3536 41 43 40 593831 23462288 Hunter Creek 124.35392 42.3627 77 88 65 25275700 102022300 South Fork Collier Creek 124.0722 42.375 88 91 85 2908100 90557400 123.55339 42.3768 54 57 52 1677860 6234308276 Hunter Creek 124.29589 42.3785 115 146 85 16414100 22583400 South Fork Collier Creek_Upstream124.04578 42.3824 94 97 91 488700 90560500 Hunter Creek 124.27247 42.386 61 77 46 8757635 16681005 South Fork Collier Creek_Upstream124.06577 42.3886 109 114 104 1730196 188522122 Onion Creek 123.72151 42.3944 51 55 48 3263100 169758900 Elko Creek 124.24281 42.396 33 35 32 135421 8757635 Lawson Creek 124.15356 42.3976 59 61 56 187100 33265900 124.20977 42.4142 46 50 43 30600 3060500 123.53085 42.4343 75 85 65 5179475 27552083 123.57289 42.4354 73 73 72 176646 5179475 Lawson Creek 124.13303 42.4476 96 104 88 34074100 59843800 Silver Creek_Upstream 123.74484 42.4521 66 73 59 1181582 4003912 124.23184 42.4549 92 96 87 815074 216724164 Silver Creek 123.87166 42.4589 75 78 73 13059965 4641588834 124.22201 42.4595 65 72 57 6821900 62303600

290 Texas Tech University, Timothy Anderson, December 2008

Table 1 (Continued) Name of River Longitude Latitude Concavity Concavity Fitted Smoothing Contour within 2 σ Steepness Window Interval 122.12242 42.2929 1.2554 0.61335 32125050.93 250 12.192 East Fork Pistol 124.16536 42.2982 3.0588 0.99392 2.83E+20 250 12.192 123.37057 42.2989 0.64097 0.022581 1506.5764 250 12.192 North Fork Pistol 124.28712 42.2999 0.60847 0.079362 1294.0048 250 12.192 Tincup Creek 123.93431 42.3345 0.59246 0.081812 903.725 250 12.192 East Fork Pistol 124.14831 42.3388 1.2315 0.24527 2616241.398 250 12.192 South Fork Collier Creek_Upstream124.12705 42.3469 1.8004 0.1963 31546345561 250 12.192 123.47016 42.3475 0.73471 0.044598 3464.6581 250 12.192 North Fork Pistol 124.26092 42.3536 0.47744 0.11228 74.1535 250 12.192 Hunter Creek 124.35392 42.3627 2.2204 0.35851 3.10E+15 250 12.192 South Fork Collier Creek 124.0722 42.375 0.74615 0.083633 13519.0819 250 12.192 123.55339 42.3768 0.67392 0.074119 2108.234 250 12.192 Hunter Creek 124.29589 42.3785 5.3037 2.1118 2.47E+37 250 12.192 South Fork Collier Creek_Upstream124.04578 42.3824 0.60327 0.052836 1033.9147 250 12.192 Hunter Creek 124.27247 42.386 -7.8219 3.0661 0 250 12.192 South Fork Collier Creek_Upstream124.06577 42.3886 0.71264 0.056749 8188.9059 250 12.192 Onion Creek 123.72151 42.3944 0.79081 0.15787 24269.5269 250 12.192 Elko Creek 124.24281 42.396 0.65824 0.090107 751.6479 250 12.192 Lawson Creek 124.15356 42.3976 0.33739 0.043592 11.5947 250 12.192 124.20977 42.4142 0.27118 0.055875 4.7753 250 12.192 123.53085 42.4343 1.768 0.26656 1.03E+11 250 12.192 123.57289 42.4354 0.52154 0.069872 204.3248 250 12.192 Lawson Creek 124.13303 42.4476 1.7545 0.64284 9.22408E+11 250 12.192 Silver Creek_Upstream 123.74484 42.4521 1.0862 0.29707 720385.1556 250 12.192 124.23184 42.4549 0.63617 0.084196 1665.8403 250 12.192 Silver Creek 123.87166 42.4589 0.58862 0.090198 1025.3798 250 12.192 124.22201 42.4595 1.1601 0.18121 10750791.76 250 12.192

291 Texas Tech University, Timothy Anderson, December 2008

Table 1 (Continued) Name of River Longitude Latitude Normalized Ksn Ksn Minimum Maximum Steepness (Low σ) (High 2 σ) Drainage Area Drainage Area Silver Creek 123.90333 42.4624 81 84 78 53243900 193329900 Silver Creek 123.81526 42.4652 86 100 72 25042600 88374400 123.58407 42.4705 51 52 50 280464 2829681 123.60971 42.4718 202 227 177 1982884 2932118 Lawson Creek 124.11535 42.4747 85 130 40 63129100 68801000 123.53285 42.4869 72 77 67 4007200 27769400 Lawson Creek 124.06742 42.4963 92 106 79 75981600 98555600 123.25578 42.4972 55 57 52 1545928 9480560 Indigo Creek 123.93017 42.4973 88 101 76 125729800 146650100 Indigo Creek 123.8999 42.4989 121 167 75 119647100 125206200 Silver Creek 123.74594 42.4997 58 59 57 115375 24040992 123.62292 42.5028 48 50 45 6187366 48669702 North Fork Silver Creek 123.83125 42.5056 80 85 76 32002000 52975400 East Indigo Creek 123.85465 42.5064 107 139 75 47671000 49386600 North Indigo Creek 123.98368 42.5251 86 97 74 7370464 47762717 West Indigo Creek 123.89898 42.5309 87 88 87 1543142 8544310429 North Fork Silver Creek 123.75354 42.5364 37 40 34 36600 32010900 123.54631 42.5424 17 20 14 400236 2432766272 123.69184 42.5498 87 89 86 203648 6883952 North Indigo Creek 123.94664 42.5506 183 195 170 1601200 6384500 East Indigo Creek 123.80901 42.5533 47 49 45 91700 47818100 124.27105 42.5544 60 66 54 2543346 396693970 123.49385 42.5552 36 39 33 11735405 9068364279 124.20952 42.5609 107 109 105 335040 2206121 123.29945 42.5645 90 98 82 16349500 85299000 123.53367 42.5688 19 22 17 97119300 1067072300 123.52408 42.5753 32 35 30 335040 1766461787

292 Texas Tech University, Timothy Anderson, December 2008

Table 1 (Continued) Name of River Longitude Latitude Concavity Concavity Fitted Smoothing Contour within 2 σ Steepness Window Interval Silver Creek 123.90333 42.4624 0.665 0.21719 4767.2419 250 12.192 Silver Creek 123.81526 42.4652 2.4745 0.51169 3.26E+17 250 12.192 123.58407 42.4705 0.53173 0.082014 160.6655 250 12.192 123.60971 42.4718 5.5349 1.3383 4.40E+34 250 12.192 Lawson Creek 124.11535 42.4747 9.9918 8.5278 3.99E+76 250 12.192 123.53285 42.4869 1.0236 0.17453 805290.9112 250 12.192 Lawson Creek 124.06742 42.4963 6.5343 2.2425 2.24E+50 250 12.192 123.25578 42.4972 0.93241 0.19865 84874.0972 250 12.192 Indigo Creek 123.93017 42.4973 -3.6395 3.9014 0 250 12.192 Indigo Creek 123.8999 42.4989 -33.2806 13.2429 0 250 12.192 Silver Creek 123.74594 42.4997 0.44617 0.057173 54.6133 250 12.192 123.62292 42.5028 0.81385 0.21723 22404.7188 250 12.192 North Fork Silver Creek 123.83125 42.5056 2.1676 0.76697 9.83E+14 250 12.192 East Indigo Creek 123.85465 42.5064 -51.4842 31.0843 0 250 12.192 North Indigo Creek 123.98368 42.5251 1.7659 0.34809 3.2654E+11 250 12.192 West Indigo Creek 123.89898 42.5309 0.43403 0.046403 70.1835 250 12.192 North Fork Silver Creek 123.75354 42.5364 0.21846 0.035102 1.2033 250 12.192 123.54631 42.5424 0.6456 0.064923 618.5905 250 12.192 123.69184 42.5498 0.38959 0.058518 37.2545 250 12.192 North Indigo Creek 123.94664 42.5506 -0.82311 0.20163 0.000001 250 12.192 East Indigo Creek 123.80901 42.5533 0.43306 0.053589 43.5086 250 12.192 124.27105 42.5544 0.71555 0.1141 6028.0369 250 12.192 123.49385 42.5552 0.80593 0.068853 23087.6933 250 12.192 124.20952 42.5609 0.5198 0.12984 272.7957 250 12.192 123.29945 42.5645 1.5067 0.2208 7305374104 250 12.192 123.53367 42.5688 0.7341 0.21444 5854.1011 250 12.192 123.52408 42.5753 0.65105 0.037387 910.587 250 12.192

293 Texas Tech University, Timothy Anderson, December 2008

Table 1 (Continued) Name of River Longitude Latitude Normalized Ksn Ksn Minimum Maximum Steepness (Low σ) (High 2 σ) Drainage Area Drainage Area 123.52408 42.5753 32 35 30 335040 2706652070 123.56092 42.5762 15 17 13 27418500 1067139500 124.24367 42.5805 61 65 57 707002 268269580 124.24787 42.5818 42 45 39 15052415 356552554 123.61874 42.5829 73 75 70 83710 44927230996 124.24588 42.5853 56 61 51 8103800 168434400 123.58976 42.5947 32 41 22 52257270 9068364279 123.97698 42.5949 75 79 72 23923200 83430300 123.97641 42.5951 77 80 73 23920900 83389300 123.60936 42.5978 63 70 56 7936224 3011372491 124.25111 42.5983 42 44 41 323335 2804635636 123.23383 42.5989 27 28 26 2000 15093800 123.70349 42.6031 72 73 72 147871 320473044 123.8704 42.6181 82 97 67 3630100 6562000 123.87102 42.6183 88 103 72 3628500 6765400 123.90639 42.6198 169 173 165 9805300 23736700 123.90581 42.62 172 177 168 9699900 23736700 123.88536 42.6223 116 126 107 6429100 9802200 123.88919 42.6233 153 169 138 7350000 9927500 123.85136 42.624 79 88 70 226575 2285985 123.85163 42.6241 79 87 70 183041 2635417 124.15628 42.6319 98 103 93 1122517 11325413 123.75499 42.6334 109 113 105 4038105 56109287 123.81309 42.6343 102 104 100 155600 3495900 123.38787 42.6367 47 53 41 92310471 241123413 124.07245 42.6439 103 125 81 6187366 52257270 124.16752 42.6447 48 59 37 5422900 7131000

294 Texas Tech University, Timothy Anderson, December 2008

Table 1 (Continued) Name of River Longitude Latitude Concavity Concavity Fitted Smoothing Contour within 2 σ Steepness Window Interval 123.52408 42.5753 0.65105 0.037387 910.587 250 12.192 123.56092 42.5762 0.86218 0.25036 70672.2664 250 12.192 124.24367 42.5805 0.71279 0.031864 4914.5822 250 12.192 124.24787 42.5818 1.0624 0.13088 3011653.478 250 12.192 123.61874 42.5829 0.56846 0.036293 456.4383 250 12.192 124.24588 42.5853 0.91329 0.065755 194155.8261 250 12.192 123.58976 42.5947 1.0453 0.24697 3945794.607 250 12.192 123.97698 42.5949 1.2875 0.36843 237301839.7 250 12.192 123.97641 42.5951 1.3376 0.35268 580737052.8 250 12.192 123.60936 42.5978 0.86192 0.06853 91610.909 250 12.192 124.25111 42.5983 0.57159 0.03587 331.0937 250 12.192 123.23383 42.5989 0.37045 0.043305 9.73 250 12.192 123.70349 42.6031 0.4386 0.037216 65.7177 250 12.192 123.8704 42.6181 1.7785 0.89266 68125232411 250 12.192 123.87102 42.6183 1.8095 0.73675 1.10294E+11 250 12.192 123.90639 42.6198 0.64874 0.31114 4631.9014 250 12.192 123.90581 42.62 0.68339 0.28021 8287.6674 250 12.192 123.88536 42.6223 -2.1358 0.99615 0 250 12.192 123.88919 42.6233 -2.6922 1.0745 0 250 12.192 123.85136 42.624 1.0719 0.23453 347487.3047 250 12.192 123.85163 42.6241 0.87663 0.17922 25075.1936 250 12.192 124.15628 42.6319 0.96661 0.15009 244879.4867 250 12.192 123.75499 42.6334 0.79428 0.17013 31825.3449 250 12.192 123.81309 42.6343 0.41666 0.10965 63.9279 250 12.192 123.38787 42.6367 2.6961 0.61789 9.25E+19 250 12.192 124.07245 42.6439 1.888 0.25702 2.35613E+12 250 12.192 124.16752 42.6447 5.7561 3.1905 6.11E+37 250 12.192

295 Texas Tech University, Timothy Anderson, December 2008

Table 1 (Continued) Name of River Longitude Latitude Normalized Ksn Ksn Minimum Maximum Steepness (Low σ) (High 2 σ) Drainage Area Drainage Area 124.10133 42.6523 207 217 198 4294100 4618800 124.11026 42.6534 140 145 134 1757100 4316500 124.15732 42.6578 87 100 74 1905600 3481900 123.55234 42.6629 49 50 48 58660 2347774443 123.87322 42.6651 143 145 140 1008930 74573903 123.46477 42.6661 47 50 45 189667 2520834892 123.58038 42.674 91 95 88 2253400 1059706100 123.46116 42.675 40 41 38 1719971 1377197225 123.69692 42.6808 83 85 81 1248893 46969363 123.63713 42.6891 86 91 81 9092300 31008800 123.17632 42.6905 36 37 35 1982884 64686077 123.26299 42.6996 58 61 56 103620 1083301 123.65792 42.7159 65 70 60 1439797 8223523 123.57479 42.7244 40 43 37 96506 1659882 123.8514 42.7438 104 110 99 25660577 298471917 123.13657 42.7513 42 46 38 107371 1719971 123.76819 42.7622 58 61 55 4655364 11325413 South Fork Coquille 123.91506 42.8019 21 25 16 135421 73147170 South Fork Coquille 124.08795 42.8881 54 62 46 102883488 796589482 Rock Crek 123.8511 42.901 66 68 64 1652369 14888108 Twelevemile Creek 123.78402 42.9041 45 46 43 438506 9812271 Twelvemile Creek 123.72542 42.9351 28 33 23 11418608 73147170 Rock Creek 124.00042 42.991 45 54 36 22589648 5938305706 Middle Fork Coquille 123.93698 43.0025 38 44 32 139326242 1460861533 Middle Fork Coquille 123.93698 43.0025 38 44 32 144708428 2679066779 Catching Creek 124.19681 43.0111 11 13 9 1419918 88410136 Upper Rock Creek 123.97554 43.0173 42 53 31 54205300 782120700

296 Texas Tech University, Timothy Anderson, December 2008

Table 1 (Continued) Name of River Longitude Latitude Concavity Concavity Fitted Smoothing Contour within 2 σ Steepness Window Interval 124.10133 42.6523 -7.9786 4.397 0 250 12.192 124.11026 42.6534 0.82587 0.31323 37629.2546 250 12.192 124.15732 42.6578 2.328 0.66506 1.02E+14 250 12.192 123.55234 42.6629 0.47849 0.022919 83.0616 250 12.192 123.87322 42.6651 0.50164 0.067217 310.403 250 12.192 123.46477 42.6661 0.58363 0.028603 486.7541 250 12.192 123.58038 42.674 0.71383 0.08971 7991.584 250 12.192 123.46116 42.675 0.62152 0.062151 943.5641 250 12.192 123.69692 42.6808 0.34658 0.092662 18.4781 250 12.192 123.63713 42.6891 0.77433 0.45033 18873.1514 250 12.192 123.17632 42.6905 0.57701 0.092153 320.6506 250 12.192 123.26299 42.6996 0.61335 0.13748 471.5443 250 12.192 123.65792 42.7159 0.97413 0.13454 164510.4552 250 12.192 123.57479 42.7244 0.7629 0.078257 2372.214 250 12.192 123.8514 42.7438 1.372 0.34982 1188925979 250 12.192 123.13657 42.7513 0.79849 0.13765 4037.1714 250 12.192 123.76819 42.7622 1.6099 0.78421 5146129828 250 12.192 South Fork Coquille 123.91506 42.8019 0.88353 0.10341 13711.9507 250 12.192 South Fork Coquille 124.08795 42.8881 2.561 0.3347 2.95E+19 250 12.192 Rock Crek 123.8511 42.901 0.4278 0.10014 49.0642 250 12.192 Twelevemile Creek 123.78402 42.9041 0.72952 0.084222 2768.4562 250 12.192 Twelvemile Creek 123.72542 42.9351 1.5557 0.63136 6735656332 250 12.192 Rock Creek 124.00042 42.991 1.4105 0.18786 2160662154 250 12.192 Middle Fork Coquille 123.93698 43.0025 2.0262 0.61036 1.31E+15 250 12.192 Middle Fork Coquille 123.93698 43.0025 2.1016 0.61969 5.89E+15 250 12.192 Catching Creek 124.19681 43.0111 0.86164 0.16009 12778.2723 250 12.192 Upper Rock Creek 123.97554 43.0173 1.3924 0.24968 3916575669 250 12.192

297 Texas Tech University, Timothy Anderson, December 2008

Table 1 (Continued) Name of River Longitude Latitude Normalized Ksn Ksn Minimum Maximum Steepness (Low σ) (High 2 σ) Drainage Area Drainage Area Catching Creek 124.20068 43.0184 20 24 15 616770 156104577 Sandy Creek 123.96289 43.0276 23 27 20 9411400 782237000 Upper Rock Creek 123.815 43.0387 59 65 54 16377000 49171200 Big Creek 124.0138 43.0454 13 15 10 24456100 782473800 Middle Fork Coquille 123.68587 43.0612 18 20 15 146086 70426584 Bill Creek 124.33748 43.0668 17 19 14 804174 48208939 Hall Creek 124.22447 43.0719 11 13 9 971973 70426584 Upper Rock Creek 123.77531 43.0731 37 44 30 1775600 15550000 Little Bear Creek 124.31702 43.0764 26 30 22 176569 402659798 Sandy Creek 123.84979 43.081 96 109 82 2030600 9310500 Big Creek_upstream 123.89769 43.0845 31 44 18 1652369 13801227 Big Creek 123.91675 43.0933 53 91 15 20540200 23076100 South Fork Elk Creek 124.03981 43.0988 28 34 22 1471059 1490769177 Camas Creek 123.76246 43.1193 46 47 45 62483 41359028 East Fork Coquille 124.01827 43.1236 10 18 1 212561600 725482500 South Fork Elk Creek 123.90552 43.1238 32 36 28 110253 960857 East Fork Coquille 124.01536 43.1248 10 18 1 212194100 724918200 East Fork Coquille 124.00515 43.1309 26 45 7 212454000 717286600 East Fork Coquille 123.70263 43.1406 33 47 19 4235000 11209500 East Fork Coquille 123.8473 43.144 26 44 8 42435200 203602300 Cherry Creek 124.05545 43.1443 17 22 12 4117594 1658262116 East Fork Coquille 123.80437 43.1458 75 81 70 12055900 203600600 East Fork Coquille 123.94558 43.1561 54 83 26 209721800 260126500 West Fork Brummit Creek 123.86937 43.1621 54 67 41 12584400 203593400 Middle Creek 124.03372 43.1674 24 29 20 698121 8486117795 East fork Brummit 123.77492 43.184 27 29 26 77312 9314756 South Fork Cherry 123.92082 43.2051 36 40 33 146455 4117594

298 Texas Tech University, Timothy Anderson, December 2008

Table 1 (Continued) Name of River Longitude Latitude Concavity Concavity Fitted Smoothing Contour within 2 σ Steepness Window Interval Catching Creek 124.20068 43.0184 0.90178 0.095831 31972.1205 250 12.192 Sandy Creek 123.96289 43.0276 0.77571 0.22351 12998.7003 250 12.192 Upper Rock Creek 123.815 43.0387 1.0475 0.56205 1813740.1 250 12.192 Big Creek 124.0138 43.0454 0.4836 0.37111 36.5582 250 12.192 Middle Fork Coquille 123.68587 43.0612 0.65722 0.06012 439.8751 250 12.192 Bill Creek 124.33748 43.0668 1.084 0.12956 343693.8244 250 12.192 Hall Creek 124.22447 43.0719 0.87065 0.41403 12011.7616 250 12.192 Upper Rock Creek 123.77531 43.0731 1.3973 0.30454 116661323.9 250 12.192 Little Bear Creek 124.31702 43.0764 0.81499 0.082886 6561.8402 250 12.192 Sandy Creek 123.84979 43.081 1.5779 0.29618 2402963980 250 12.192 Big Creek_upstream 123.89769 43.0845 1.8206 0.43434 48556102287 250 12.192 Big Creek 123.91675 43.0933 15.7621 4.2195 1.76E+114 250 12.192 South Fork Elk Creek 124.03981 43.0988 1.0218 0.11035 368015.5191 250 12.192 Camas Creek 123.76246 43.1193 0.38301 0.073003 18.9746 250 12.192 East Fork Coquille 124.01827 43.1236 2.6019 1.8 4.46E+19 250 12.192 South Fork Elk Creek 123.90552 43.1238 0.78296 0.26266 2341.6533 250 12.192 East Fork Coquille 124.01536 43.1248 2.6019 1.8 4.46E+19 250 12.192 East Fork Coquille 124.00515 43.1309 12.0185 5.1116 2.20E+98 250 12.192 East Fork Coquille 123.70263 43.1406 2.3997 0.6068 6.71E+14 250 12.192 East Fork Coquille 123.8473 43.144 1.9971 0.82822 8.86326E+13 250 12.192 Cherry Creek 124.05545 43.1443 1.0977 0.16575 1446625.981 250 12.192 East Fork Coquille 123.80437 43.1458 0.73751 0.26761 11519.3682 250 12.192 East Fork Coquille 123.94558 43.1561 20.4517 7.6888 1.03E+169 250 12.192 West Fork Brummit Creek 123.86937 43.1621 1.1803 0.34816 14862555.33 250 12.192 Middle Creek 124.03372 43.1674 0.89488 0.060305 38993.3096 250 12.192 East fork Brummit 123.77492 43.184 0.36673 0.095957 9.4262 250 12.192 South Fork Cherry 123.92082 43.2051 0.86455 0.095152 11613.6205 250 12.192

299 Texas Tech University, Timothy Anderson, December 2008

Table 1 (Continued) Name of River Longitude Latitude Normalized Ksn Ksn Minimum Maximum Steepness (Low σ) (High 2 σ) Drainage Area Drainage Area West Fork Brummit Creek_upstream123.86279 43.2068 37 39 36 500 12657800 North Fork Coquille 124.03915 43.2107 18 20 16 3974013 2926047084 Cunningham Creek 124.15688 43.2118 14 17 11 616770 1120424808 North Fork Coquille 123.86529 43.2885 26 27 24 85998 3835439

300 Texas Tech University, Timothy Anderson, December 2008

Table 1 (Continued) Name of River Longitude Latitude Concavity Concavity Fitted Smoothing Contour within 2 σ Steepness Window Interval West Fork Brummit Creek_upstream123.86279 43.2068 0.4 0.067663 19.2836 250 12.192 North Fork Coquille 124.03915 43.2107 0.91644 0.10911 64343.0366 250 12.192 Cunningham Creek 124.15688 43.2118 0.99402 0.14541 62383.7196 250 12.192 North Fork Coquille 123.86529 43.2885 0.52558 0.12935 78.5095 250 12.192

301 Texas Tech University, Timothy Anderson, December 2008 PERMISSION TO COPY

In presenting this thesis in partial fulfillment of the requirements for a master’s

degree at Texas Tech University or Texas Tech University Health Sciences Center, I

agree that the Library and my major department shall make it freely available for research

purposes. Permission to copy this thesis for scholarly purposes may be granted by the

Director of the Library or my major professor. It is understood that any copying or

publication of this thesis for financial gain shall not be allowed without my further

written permission and that any user may be liable for copyright infringement.

Agree (Permission is granted.)

___Timothy Anderson ______9/25/08 ______Student Signature Date

Disagree (Permission is not granted.)

______Student Signature Date