<<

ABSTRACT

STRATIGRAPHY AND TRANSMISSIVITY OF THE FAN, VISALIA,

The Kaweah River fan is located in the Tulare Basin of the of California. This fluvial fan supplies for several farms and cities including Visalia. The geology in this region is well studied and is generally similar to all major river fans that flow into the San Joaquin Valley from the . However, the hydrogeology of the area is not quantified. The objectives of this thesis were to: (a) Correlate stratigraphic units to hydrologic units in the region; (b) identify the spatial extent and stratigraphy of the lithologic units; (c) use correlations to identify the aquifers; and (d) measure and estimate the aquifers’ hydraulic conductivities and transmissivities. Field core samples were taken from soil surface to 132 feet (40.2 m) below surface. In this process, a new method for sampling unlithified-core for laboratory testing was created to make this study possible. The results show that stratigraphy described by Marchand and Allwardt (1981) is found throughout the study area. Together, the upper Turlock Lake and Riverbank Formations are bound in the Layer 3 aquifer which is a major host for groundwater in the Kaweah River fan and is characterized by a transmissivity of 2492 ft2/day (232 m2/day). In conclusion, this study demonstrates that the San Joaquin Valley has two different models for confining beds which are the lacustrine and swamplands deposits for the terminal basin and paleosols for the fluvial fans.

Dustin White May 2016

STRATIGRAPHY AND TRANSMISSIVITY OF THE KAWEAH RIVER FAN, VISALIA, CALIFORNIA

by Dustin White

A thesis submitted in partial fulfillment of the requirements for the degree of Master of Science in Geology in the College of Science and Mathematics California State University, Fresno May 2016

© 2016 Dustin White APPROVED For the Department of Earth and Environmental Sciences:

We, the undersigned, certify that the thesis of the following student meets the required standards of scholarship, format, and style of the university and the student's graduate degree program for the awarding of the master's degree.

Dustin White Thesis Author

Zhi Wang (Chair) Earth and Environmental Sciences

John Wakabayashi Earth and Environmental Sciences

Mara Brady Earth and Environmental Sciences

For the University Graduate Committee:

Dean, Division of Graduate Studies AUTHORIZATION FOR REPRODUCTION OF MASTER’S THESIS

X I grant permission for the reproduction of this thesis in part or in its entirety without further authorization from me, on the condition that the person or agency requesting reproduction absorbs the cost and provides proper acknowledgment of authorship.

Permission to reproduce this thesis in part or in its entirety must be obtained from me.

Signature of thesis author: ACKNOWLEDGMENTS My thesis is a result of countless hours of research yet without the guidance and patience of Dr. Zhi (Luke) Wang this thesis would not have been possible. Therefore, I would like to first thank my advisor Dr. Wang, who along the way always encouraged the development of new ideas backed up by credible research and taught the skills needed to insure my success. Furthermore, I would like to thank my committee members Dr. John Wakabayshi and Dr. Mara Brady for agreeing to work with me and giving excellent advice and guidance necessary to completion of this work. The several hours of philosophical discussions with Dr. Wakabayshi helped formulate many ideas in this thesis in addition to develop and grow. Without Dr. Brady, the stratigraphy would not have been complete. It was her guidance that helped merge field experience with academics. Most importantly, along the way, my family (especially my mom) and close friends gave me the support necessary to finish this thesis. I don’t think I would have been successful without all of their encouragement during the difficult periods. I am truly thankful to everyone who helped with insuring the completion of my thesis. TABLE OF CONTENTS Page

LIST OF TABLES ...... viii

LIST OF FIGURES ...... ix

1. INTRODUCTION ...... 1

1.1 General Statement ...... 1

1.2 Description of Region ...... 1

1.3 Purpose and Scope of This Investigation ...... 5

2. GEOLOGIC BACKGROUND ...... 7

2.1 Cenozoic Formation of the San Joaquin Valley ...... 7

2.2 Glaciation of the Sierra Nevada ...... 13

2.3 Historical Changes in the San Joaquin Valley Water Resources ...... 16

2.4 Soil Basics ...... 18

2.5 San Joaquin Valley Stratigraphic Nomenclature ...... 21

2.6 Late and Stratigraphy of the Eastern San Joaquin .. 22

2.7 Stratigraphy ...... 37

2.8 San Joaquin Valley Hydrogeology ...... 39

3. METHODOLOGY ...... 57

3.1 General Description of Work ...... 57

3.2 Core Drilling ...... 58

3.3 Technique for Sampling Core ...... 64

3.4 Constant Head Permeameter ...... 65

3.5 Falling Head Permeameter ...... 70

3.6 Transmissivity ...... 73

3.7 Microscopy ...... 73 vii vii Page

3.8 Mastersizer 3000 Particle Analyzer ...... 74

3.9 GIS Digital Data ...... 74

4. RESULTS ...... 76

4.1 Geologic Formations of the Kaweah River Fan ...... 76

4.2 Stratigraphy of the Kaweah River Fan ...... 82

4.3 Hydraulic Conductivities of the Kaweah River Fan ...... 89 4.4 Malvern Mastersizer 3000 Particle Size Analyses of Impermeable Sections ...... 91

4.5 Transmissivity of Kaweah River Fan Stratigraphic Units ...... 98

5. DISCUSSION ...... 100

5.1 Recharging Layer 3 Aquifer in the Kaweah River Fan ...... 100 5.2 One Borehole Can Improve Knowledge of an Aquifer on a Sub- Regional Scale ...... 100

5.3 Confining Beds of the River Fans Versus Terminal Basins ...... 102

5.4 Undifferentiated Confining Beds of the Terminal Basin ...... 102

5.5 The Importance of Transmissivity ...... 103

SUMMARY AND CONCLUSIONS ...... 104

REFERENCES ...... 107 APPENDIX: HYDRAULIC CONDUCTIVITY AND TRANSMISSIVITY DATA ...... 115 LIST OF TABLES

Page

Table 1. Summary and Correlations from Geologic Background...... 49 Table 2. Hydraulic Properties of Central Valley Soils (from Bertoldi et al., 1991)...... 56

Table 3. Correlations Between Soil Series and Geologic Formations...... 78

Table 4. Results of Hydraulic Conductivity (K) and Transmissivity (T)...... 99

LIST OF FIGURES

Page

Figure 1. Map of the Central Valley of California...... 2

Figure 2. Map of NE Tulare Basin and surface water...... 4 Figure 3. The Hydrogeologic Provinces of California (from Belitz et al., 2003)...... 8

Figure 4. Late California coastline (modified from Bartow, 1991)...... 10 Figure 5. Kings R. fan Chronostratigraphic Diagram (from Weissmann et al., 2002)...... 15

Figure 6. Soil taxonomy (from Lynn et al., 2002)...... 19 Figure 7. The San Joaquin Valley stratigraphic nomenclature (from Lettis and Unruh, 1991)...... 23 Figure 8. East San Joaquin Valley geologic formations (from Weissmann et al., 2005)...... 24

Figure 9. Duripan soils of the Tulare Basin...... 31

Figure 10. Spatial provinces map (modified from Faunt et al., 2009)...... 41

Figure 11. Cross sections of the San Joaquin Valley...... 43 Figure 12. Lacustrine clay-lenses regional extent map (from Burton and Belitz, (2008)...... 45

Figure 13. Local surface water effects on groundwater...... 55

Figure 14. Core drilling operation...... 58

Figure 15 Core recovered from drilling...... 59

Figure 16. PVC Frame for AMS stainless steel sampler...... 66

Figure 17. PVC . Saturation process of core...... 67

Figure 18. Constant head permeameter...... 68

Figure 19. The tank type constant-head permeameter...... 68

Figure 20. Falling head permeameter design...... 70 x x Page

Figure 21. Falling head permeameters...... 71

Figure 22. Geologic map of the Kaweah River fan region...... 77

Figure 23. Stratigraphic Column of the Kaweah River, CA...... 83

Figure 24. Microscopy of volcanic ash...... 86

Figure 25. Root traces (paleosols) seen in Kaweah River fan...... 88

Figure 26. Hydraulic conductivity of formations and Layer 3 aquifer (blue)...... 90

Figure 27. Histograms from particle size analysis using Mastersizer 3000...... 92

1. INTRODUCTION

1.1 General Statement Tulare Basin groundwater reserves are vital to the productivity of the agriculture, dairy industries, and cities since they all rely on groundwater during seasonal dry periods when dry up and river discharge perishes. , domestic, and industrial industries rely principally on groundwater pumped from wells in the Hanford and Visalia area as either a supplement or as an only source (Croft and Gordon, 1968). It is evident from groundwater equal elevation level maps produced by the California Department of Water Resources (DWR) over the last 100 that the groundwater table fluctuates and is being depleted with increased speed. Therefore, studies on groundwater reserves are vital to the prolonged productivity of the San Joaquin Valley of California. Hydrogeology focuses mostly on aquifers and their boundaries, whereas, stratigraphy explains their origins and spatial extent. By combining these two disciplines, it will result in a more complete understanding of the groundwater systems.

1.2 Description of Region This thesis focuses on the Kaweah River fan which is found within the Tulare Basin also called the Tulare-Buena Vista Lakes Watershed (USGS) seen in Figure 1. The Tulare Basin is the southern-half of the San Joaquin Valley whereas the San Joaquin Basin is its northern-half. In Figure 1, the rivers from the upper watersheds flow into the San Joaquin (colored red and labeled) and Sacramento Valleys (blue river complexes of basin) of the Great or Central Valley of California.

2 2

Figure 1. Map of the Central Valley of California. Note: The ( with blue rivers) and the San Joaquin Valley (drainage basin with red rivers) are within the Central Valley of California. Drill core was recovered from the City of Visalia which is in the Tulare Basin of the San Joaquin Valley. Visalia was built on the Kaweah River fan. 3 3

Generally, the Coast Ranges do not discharge much water into the San Joaquin Valley. Instead, most of the water in the Central Valley is a result of Sierra Nevada river discharge. The perennial streams of the Tulare Basin are: Kings, Kaweah, Tule, and Kern rivers. The Kaweah River fan sediments are composed of Sierra Nevada detritus since they are related to deposits from the Upper Kaweah Watershed (USGS) that erode into the Marble, North, East, South, and Middle Forks of the Kaweah River which drains into Reservoir at Terminus . All rivers of the Tulare Basin, including the Kaweah River, terminate in the historical shallow and swampy bed which is now farmland since the construction of and reservoirs on the rivers of the western Sierra Nevada. Most of the water received by the Kaweah River fan aquifers is transported from the upper forks of the Upper Kaweah Watershed where the heavy precipitation (including snowfall) occurs versus the semi-arid valley floor. When the Kaweah River reaches San Joaquin Valley, it then breaks up into five lesser channels called the Saint Johns River, Mill, Packwood, Cameron, and Deep Creeks (Figure 2) which eventually would either be absorbed into the ground or drain into the Tulare Lake bed (Preston, 1981). The creeks are now used as and still have their original names. Other creeks that deposit onto or near the Kaweah River fan are from north to south: Lime Kiln (next to Dry Creek Drive), Cottonwood, and Yokohl Creeks. Thompson (1892) used Lime Kiln to name what is now commonly called Dry Creek (USGS). These streams bring water and deposit sediments into the valley. Consequently, all creeks dry up during the summer and are therefore intermittent streams.

4 4

Figure 2. Map of NE Tulare Basin and surface water. Note: The Upper Kaweah Watershed includes all five forks of the Kaweah River. Below Lake Kaweah Reservoir, the Kaweah River breaks up into four creeks and a lesser-river. The Kaweah River fan is a result of deposition of these streams

5 5 1.3 Purpose and Scope of This Investigation This thesis is intended to provide a detailed record of both hydrologic properties and the stratigraphy of the Kaweah River fan based on borehole core data drilled from eastern Visalia, California. Laboratory measurements of vertical hydraulic conductivity were used to calculate transmissivity. Identifying the stratigraphic units help associate the Kaweah subbasin hydrogeology with stratigraphy identified elsewhere in the San Joaquin Valley and will give depth intercepts to relate to regional groundwater models. Furthermore, measuring the hydrologic properties allow for more accurate estimates for recharge, seepage from rivers and canals, drawdown effects from pumping, and overall character of the aquifers of the Kaweah River fan. The purpose of this thesis is to correlate the stratigraphic and hydrogeologic units based on previous reports and measured hydrologic properties, to identify the stratigraphic formations, and then correlate these formations with the aquifers of the Central Valley of California. This thesis offers new insights into correlations of the stratigraphy and hydrogeological units for the NE San Joaquin Valley, and identifies the geology of the Kaweah River fan which until now was mostly undifferentiated. This resulted in a detailed geologic map, a stratigraphic column, and characterization of the uppermost aquafer for the Kaweah River fan. Results from literature review and drill data suggest that the aquifers of the San Joaquin Valley should be divided into two models: (a) Aquifer layers defined by confining hydrologic units composed of lacustrine and deposits which is in use today; (b) Aquifer layers for the fluvial fans where paleosols are the confining beds Laboratory tests were conducted on drill cuttings in order to measure vertical saturated hydraulic conductivity which was used to estimate transmissivity. This allowed for correlations between historical stratigraphic data 6 6 versus the transmissivity values. New methods for measuring hydraulic conductivity of soil and sediment core were created to make this thesis possible. Transmissivity and hydraulic conductivity data were not available for this region until now. This study contributes a summary and interpretive correlation of stratigraphy and hydrogeological units and shows their relationships since one explain the origins and the other quantifies the aquifer properties for water management. In addition, the geologic map of the Kaweah River fan helps to identify the spatial extent of the stratigraphic formations in addition to completing a detailed map of the Kaweah River fan which was undifferentiated in Weissmann et al. (2005). New technique for sampling core cuttings can be used to measure hydraulic properties of preexisting core samples in repositories using permeameters. The technique is inexpensive and effective. Saturated hydraulic conductivity helps characterize the hydrologic units whereas transmissivity estimates relate to the drawdown of an aquifer and are used to estimate its ability to produce water from pumping. Characterization of the uppermost aquafer for the Kaweah River fan is important for water management of groundwater and might be helpful for future groundwater models.

2. GEOLOGIC BACKGROUND

A rather lengthy background study or literature review is provided here to help build a more complete model of the aquifers in the San Joaquin Valley based on existing publications, which is rarely available so far. Thus, this section is probably longer than those appear in other MS theses. However, the contents could be useful resources for future researchers in the interdisciplinary areas of hydrology and geology.

2.1 Cenozoic Formation of the San Joaquin Valley The San Joaquin Valley is the southern portion of the greater Central Valley of California (Figure 3). This is a brief summary describing the Cenozoic evolution of the San Joaquin Valley of California. In a more detailed report on the formation of the San Joaquin Valley, Bartow (1991) concluded that during the Paleogene (23-66 Ma) a tectonic regime shaped the underlying structure of the San Joaquin Valley, yet geography and climate change strongly influenced its sedimentary record. Prior to time, this part of California was a convergent plate margin, and, unlike the Cascade margin to the north, this region was mostly submerged (including the Coast Ranges) and what has become the Central Valley was then a marine forearc basin. These marine sediments are buried deep beneath the Tulare Basin and yield water of poor quality. This section is designed to help understand how the San Joaquin Valley formed and how it transitioned from a marine into a lacustrine environment. The formation of the San Joaquin Valley began with the creation of the Sierra Nevada.

8 8

Figure 3. The Hydrogeologic Provinces of California (from Belitz et al., 2003). Note: These provinces share many of the same boundaries and names as the Physiographic (Wahrhaftig and Birman, 1965) and Geomorphic Provinces (Jenkins, 1938; Hinds, 1952) based on topographic relief, climate, and geology (mostly faults) of their regions.

9 9 2.1.1 The Sierra Nevada At the end of the Mesozoic, the Coast Ranges did not exist. Instead, the San Joaquin Valley was in its infancy and was open to the Pacific Ocean that was part of a greater marine forearc basin created at the west edge of North America (Bartow, 1991). Figure 4 shows the paleogeography of the San Joaquin Valley about 59 Ma from Bartow (1991). However, the plutons in the Sierra Nevada were forming as far back as 400 Ma, yet range from 160 Ma to 85 Ma with the most voluminous emplacement of plutons was from 100 to 85 Ma. The Sierra Nevada plutons crystallization age range from 110 to 85 Ma (Wakabayashi and Sawyer, 2001) and were uplifted two times, first during and after the (60-100Ma) and then uplift and tilting during the late Cenozoic (Unruh, 1991; Wakabayashi and Sawyer, 2001; Wakabayashi, 2013). Wakabayashi (2013) proposed that the uplift events were superimposed on a preexisting range in the south in contrast to a gentle west slope (sloping westward from central Nevada) in the north. Late Cenozoic uplift events are thought started at 20 Ma in the Kern and migrated northward, kicking off between 10 and 6 Ma in the San Joaquin, about 4 Ma in the drainage, and 3 Ma at Donner summit and northward. Moreover, the Sierra Nevada is the largest mountain range in California and defines the eastern boundary and basement rock of the eastern San Joaquin Valley. The Sierra has been the principal sediment source for the Central Valley from the time of accumulation in the forearc basin (marine) to when the Central Valley became subaerial.

10 10

Figure 4. Late Paleocene California coastline (modified from Bartow, 1991). Note: The figure above represents the San Joaquin Valley about 59 Ma and shows how the basin was open to the Pacific Ocean. This resulted in deposition of marine sediments which typically yield poor water quality in today’s aquifers. Right slip for the San Andreas Fault was unspecified during the late Paleocene therefore it was named proto-San Andreas Fault. It was not until the Pleistocene (about 2.6Ma or earlier) that the San Joaquin Valley was cut off from the Pacific Ocean by the Coast Ranges.

11 11 2.1.2 The Coast Ranges During the Neogene (23-2.6 Ma) the formation of the San Joaquin Valley was controlled again by tectonism; northwestward migration of the Mendocino triple junction and the San Andreas fault complex which uplifted and transported the marine sediment eventually creating the Coast Ranges (Bartow 1991). During the late Mesozoic, the western plate boundary of North America began to change from a convergent-margin (oblique-subduction) setting (Atwater, 1970; Bartow, 1991), responsible for creating the Sierra Nevada plutons, into the transform- margin better known as the San Andreas Fault zone. This transition occurred during the mid-Cenozoic (about 30-8 Ma) yet was not onshore until about 18 Ma (Atwater and Stock, 1998). The conversion to a transform plate boundary began with the arrival of a spreading ridge at the trench with the western side of the spreading ridge having a plate motion parallel to the margin of western North America (Atwater and Stock, 1998). The San Andreas Fault now separates the North American and Pacific Plates. Uplift of the mountains continued which resulted in Miocene (23-5.3 Ma) subsidence of the San Joaquin Valley (Bartow 1991). The Coast Ranges did not begin to emerge from beneath the sea on a large scale until about 8-10 Ma, although there were locally emergent areas that can be viewed as large islands. The rise of the Coast Ranges is now thought to be between 10 and 6 Ma (Argus and Gordon, 2001). What is now the Central Valley was becoming emergent at that time with the exception of some low areas such as the San Joaquin Valley that continued to be marine until Miocene time. The Coast Ranges bound the San Joaquin Valley primarily from the west and to a lesser extent, the southwest corner and are subdivided into the southern and northern half by the Bay (Wahrhaftig and Birman, 1965; Belitz et al., 2003). 12 12 2.1.3 The Tehachapi Mountains The uplift of the Tehachapi Mountains was a result of clockwise rotation possibly related to Late Cretaceous and early Paleogene right slip on the proto-San Andreas fault and then began to uplift about 16-17 Ma (Bartow, 1991). The Tehachapi Mountains consists of Sierran basement that is structurally controlled by the left-lateral White Wolf and Garlock Faults (Brown, 1990). Therefore, they are part of the Sierra Nevada province that attaches to Southern Coastal Range. Together both the Tehachapi Mountains and the Southern Coastal Range southern tip bound the San Joaquin Valley from the southeast corner.

2.1.4 The San Joaquin Valley The Great Valley (Jenkins, 1938; Hinds, 1952; Wahrhaftig and Birman, 1965) also called the Great Central Valley (Lettis, 1982) and Central Valley (Belitz et al., 2003) is divided into two separate, lesser valleys called the Sacramento and San Joaquin Valleys, which have significant differences in geology. They are named after the main rivers flowing down their axial troughs, namely the Sacramento and San Joaquin Rivers. It was not until the Pleistocene (about 2.6Ma or earlier) or earlier, that the San Joaquin Valley was enclosed by mountains from the Pacific. Around the beginning of the Pleistocene, the San Joaquin Valley depositional environment transitioned from marine to marsh and lacustrine environments as a result of being cut off from the ocean to the west by the Coast Ranges’ uplift (Bartow, 1991) thus creating a closed basin. Once cut off from the Pacific Ocean by the subaerial exposure of the modern Coast Ranges, alluvium filled the basin from the westward-tilted Sierra Nevada to the east and marine sediments of the Coast Ranges from the west. Note that it doesn’t appear that the greater Central Valley was a closed depression for much time during its subaerial history. Rather it appears to have a fairly consistent outlet, be it at the 13 13 current location (Carquinez Strait) or somewhere to the south. The San Joaquin Basin was an exception, as is the Tulare Basin today. Nonetheless, even with an outlet, however, the Central Valley was still an environment of deposition instead of erosion. The Sierra Nevada (continental) and Coast Ranges (marine) deposition can be distinguished based on composition and are designated different geological formations. Active tectonic subsidence of the axial trough and Tulare Lake basin allowed for water and sediments to collect. The Pleistocene sediments of the San Joaquin Valley typically contain the potable groundwater whereas the older marine sediments usually have dissolved particulates not fit for agriculture or human consumption (Page and LeBlanc, 1969).

2.2 Quaternary Glaciation of the Sierra Nevada Sierra Nevada glaciation has been studied by various authors for over a century. The history and summaries of investigation into the glacial stages of the Sierra Nevada are given by several authors (Blackwelder, 1931; Wahrhaftig and Birman, 1965; and Fullerton, 1986). Both Marchand and Allwardt (1981) and Lettis (1988) correlated San Joaquin Valley Quaternary clastic sedimentary deposits to Sierra Nevada glacial stages. Lettis (1988) correlated San Joaquin Valley geologic formations of Marchand and Allwardt (1981) to the more recent Sierra Nevada glacial stages of Fullerton (1986) which can be seen in Figure 5 from Weissmann et al., (2002). The Kaweah River fan deposits are a result of these glacial cycles (Weissmann et al., 2005). Central Valley geologic formations and glacial age correlations are in order of oldest to youngest and are as follows: (a) the Laguna Formation was formed during the McGee stage and Glacier Point; (b) Turlock Lake Formation deposits occurred during the Donner Lake, Red Meadow for the upper unit and Sherwin, and El Portal for the lower unit; (c) 14 14

Riverbank Formation was deposited during Tahoe, Mono Basin, and Casa Diablo glacial stages; and finally, (d) the Modesto Formation consist Tahoe, Tenaya, Tioga, and Hilgard glacial stages of the Sierra Nevada (Lettis, 1988; Weissmann et al., 2002). The Tenaya, Tioga and Hilgard all occurred during the Wisconsin glaciation of North America (Birkeland and Janda, 1971; Weissmann et al., 2002). Holocene glacial events were minor in comparison to Pleistocene events. Recess Peak and Matthes tills were deposited during according to Birkeland and Janda (1971). Matthes glaciation (the little ice age) occurred during the Holocene, during historical time. Conversely, there is some controversy on Recess Peak that is now believed to be age by Clark and Gillespie (1997). Raub et al. (2006) inventoried all existing ice bodies within the Sierra Nevada, most of which did not constitute as glaciers, however, a few small glaciers still exist such as the which is the largest within the providence. Matthes till can be found within the Palisade glacial cirque (Raub et al., 2006) on the east side of the Sierra Nevada as well as other localities. The most voluminous deposition took place during these glacial periods. Conversely, soil formation and stream incision of these deposits occurred during interglacial cycles. Generally, the alluvial deposits described by Marchand and Allwardt (1981) show clear evidence (rock flour, reverse graded beds, till deposits at headwater, etc.) of several glacial periods found in sediments as old as the Laguna Formation (Lettis, 1988; and Weissmann et a., 2005). The geologic formations are discussed in more detail in stratigraphy section of this thesis.

15 15

Figure 5. Kings R. fan Chronostratigraphic Diagram (from Weissmann et al., 2002). Note: The fan (North) connects to the Kaweah River fan (south) at a north-south boundary. Both fans share the same stratigraphic nomenclature (Marchand and Allwardt, 1981) that describes Sierra Nevada eastern basin deposits. This diagram correlates Sierra Nevada glacial stages (Lettis, 1988), the magnetostratigraphic scale, and marine oxygen isotope record (Weissmann et al., 2002) to the stratigraphic nomenclature used to describe Kaweah R fan sediments. 16 16 2.3 Historical Changes in the San Joaquin Valley Water Resources The surface water reserves of the San Joaquin Valley have been strongly altered by increased irrigation and agriculture which directly affects groundwater levels of the aquifers. The two main streams of the Central Valley, the Sacramento and San Joaquin Rivers, flow down their axial troughs until they merge and then terminate into the . The flows from the Sierra Nevada until it connects to the axial trough of the San Joaquin Basin and then connects the Fresno, Chowchilla, Merced, Tuolumne, Stanislaus, Calaveras, and Mokelumne Rivers. The San Joaquin River eventually connects to the Sacramento-San Joaquin River Delta which dumps into San Francisco Bay. Rivers of the southern San Joaquin Valley (Tulare Basin) typically drain into the Tulare Lake bed, with exception to the Kings River which has been diverted at the where it then connects to the San Joaquin River. According to Atwater et al. (1986), an -dam found at the toes of the Kings River and Los Gatos Creek (alluvial fan formed at the base of the Southern Coastal Range) fans formed during major glaciation cycles (primarily the Sierra Nevada Tioga and Tahoe glaciation), which created the Tulare Lake no later than 100,000 years before present. Furthermore, the alluvial fan-dam marks the maximum elevation of the Tulare Lake -plain because its spillway into the San Joaquin Basin has been breached several times making the lake level inconsistent (Atwater et al., 1986). This is probably a consequence of the closed depression to which the Kings and southern rivers flow: the sediment has no place to go whereas from the San Joaquin northward the streams still flow to the ocean. Before modern and historical times, the San Joaquin Valley was a much wetter environment. Recall, that the San Joaquin Valley consists of two basins called the San Joaquin (northern region) and Tulare (southern region) Basins. The 17 17

San Joaquin River (south end) up to the (north end) fed the northern region of the San Joaquin Valley and is called the San Joaquin Basin. The south end of the San Joaquin Valley or the Tulare Basin had three lakes named the Kern, Buena Vista, and Tulare Lakes. The Kern and Buena Vista Lakes are fed by the , whereas, the Tulare Lake is fed by Tule, Kaweah, and the Kings Rivers and other lesser rivers and creeks (Thompson, 1892; Preston, 1981). During periods of flooding, the three lakes would merge along with the Summit Lake of the Kings River and spill into the San Joaquin River Basin and marsh lands to the north (Thompson, 1892). The Tulare Lake was regarded as the largest lake, west of the Mississippi River before the construction of reservoirs and canals built for water management still in use today. The Tulare Basin was characterized by shallow lakes surrounded swamps and swampy sloughs (Preston, 1981). Today, surface water of the San Joaquin Valley is controlled by reservoirs, preexisting river channels, canals, and aqueducts. Moreover, most of the San Joaquin Valley is used for farming including the Tulare Lake bed. The altitude of the water table fluctuates; however, it is clearly being drained from excessive groundwater pumping for use of farming and lack of techniques to recharge the groundwater reserves. The last Sierra Nevada glaciation, though a minor event, occurred during the Matthes till deposition less than 2000 years ago (Birkeland and Janda, 1971; Clark and Gillespie, 1997) and is currently in an interglacial period. California’s Central Valley has had several droughts in the last one hundred years (Paulson et al., 1991) and is currently in a drought during the production of this thesis. The Kaweah River fan terminates into the Tulare Lake bed and is bound from the north and south by the Kings and fans which are all part of the Tulare Basin of the San Joaquin Valley. As a result, many 18 18 private well owners and farmers are drilling wells deeper into the San Joaquin Valley to recover groundwater from greater depths. The hydrogeology section of this thesis will explain groundwater in more detail.

2.4 Soil Basics Soils are derived from altered rock or unconsolidated parent material (including organic material) due to physical, chemical, and biogeochemical weathering at the surface of Earth’s crust and must be able to support plant life. The five factors that influence soil formation and work in conjunction are climate (water and gas exchange), topography, parent material, biota, and time. Climate is most influential and includes effective precipitation (meteoric water that migrates to the groundwater table) and temperature especially in the basin soils described in this thesis. Soil taxonomy can be simplified into two philosophies, the genetic (formation of soil horizons) and morphological approaches (Schaetzl and Anderson, 2009). The layer horizons (genesis) and morphology can be related by the morphological epipedons that are equivalent to the upper part of the soil darkened by organic matter, the upper eluvial horizons, or both, and may include the upper part of the B horizon if it is darkened by organic matter (Schaetzl and Anderson, 2009). Similarly, the endopedon includes the rest of the horizons absent of blackened organic material down to the regolith. The morphological philosophy uses international Soil Survey Staff (2010) convention to classify soils. This convention includes six categories and numerous classes seen in Figure 6 from Lynn et al. (2002). They are organized into six different categories that begin with the 12 soil-orders and end with the more detailed soil series. Soil taxonomy classifies soils based on their characteristics and mineralogy. Soils are 19 19 useful for identifying stratigraphic formations at the surface using taxonomic names and the subsurface as preserved soils better known as paleosols (also spelled palaeosols). Soil series of the Kaweah River region has been correlated to stratigraphic formations that were used to generate the geologic map in the results section of this report. Correlations were made on a basis of previous reports (Marchand and Allwardt, 1981 and Weissmann et al., 2005), soil character, and spatial distributions.

Figure 6. Soil taxonomy (from Lynn et al., 2002). Note: The six categories of soil classification for soil taxonomy are order (soil characteristics), suborder (identifies the order), great group (identifies the suborder and order), subgroup (additional information in two syllables in which the first denotes important properties of the soil), a family (identifies subsets of the subgroup and is a qualifier), and series (soils named after geographical features). This International classification system for soils uses soil properties that are observed and measured to organize and group the natural soil bodies into categories and classes. There are more than 20,000 classes for soil series in the US alone. Soil Orders Map of USA can be viewed at http://www.nrcs.usda.gov/wps/portal/nrcs/main/soils/survey/class/maps/ 20 20 2.4.1 Paleosols A paleosol is formed when a pre-existing soil is buried by deposition. This can be thought of as deposition rate exceeding soil formation rate. It does not take too much deposition to exceed soil formation rate. The fan areas downstream of the mouths of Sierran streams are an ideal place to preserve multiple paleosols because these are environments of very low erosion rates with essentially no deposition within interglacials versus high deposition rates during glacial periods during which time soils are buried. This results in a stack of multiple paleosols we see in San Joaquin Valley Quaternary stratigraphy. Typically, soils are characterized by zones (profile layers or horizons) created from physical, chemical, and biological processes exposed at the surface that are dictated by climate, vegetation, and time. However, when soils become buried, they tend to display distinct properties (Walker, 2005). These buried soils are classified as paleosols. Younger paleosols (prehistoric and historic time) tend to preserve both the A and B horizons whereas older paleosols tend to only preserve the B horizon with rare relict A horizons (mineral portion), and therefore, can be identified on a basis of its distinctive clay content which differs from that of the B horizon (Walker, 2005). The paleo A horizons are rare because fluvial deposition can be rather high energy and they are typically very loose and easily eroded, whereas the B horizon is much more indurated. Important diagnostics for a paleosol B horizon include color, textural variations, weathered minerals, and enrichment (dry climate) or depletion (wet climate) of carbonate content (Walker, 2005). Walker (2005) mentions some complexities when characterizing paleosols which are: (a) older buried paleosols are polygenetic, and reflect more than one phase of pedogenesis; (b) the buried soil may be affected by post-depositional diagenesis due to changing groundwater 21 21 conditions or differential compaction; and (c) when a paleosol is re-exposed to the surface, new pedogenic process can overprint the preexisting soil and in the case of welded soils (such as a duripan), a younger soil profile will develop above or in some cases merge with the relict soil. Generally, a paleosol can be thought of as a fossil soil. Paleosols for this thesis are identified by iron-silica hardpans (duripan) and root traces in the form of hollowed root tubes and root castings which will be discussed in more detail in the results section of this thesis. Clay-formation or - accumulation can occur from different processes and therefore is not used a key for identifying paleosols. However, buried root traces are absolute indicators of paleosols.

2.5 San Joaquin Valley Stratigraphic Nomenclature The San Joaquin Valley stratigraphic formation nomenclature is summarized in Lettis and Unruh (1991) report (Figure 7). These publications describe the units with type sections, which is the conventional way to name geological units. For instance, Piper et al. (1939) designated type sections for formations in the Area, Marchand and Allwardt (1981) defined most of the type sections for N.E. San Joaquin Valley, Lettis (1982) for West- Central San Joaquin Valley, and Bartow (1984) for the Southeastern San Joaquin Valley. The type sections of the Laguna, Turlock, Riverbank, Modesto, and post- Modesto Formations, described by Marchand and Allwardt (1981), are located within the northeastern San Joaquin Valley and are roughly 60 miles north of Fresno. In addition, these same formations are correlated to the units as far as the Kaweah River fan (Weissmann et al., 2005) as seen in Figure 8. According to Neuendorf, et al. (2005) a type section or stratotype is the originally described sequence of strata that constituted a named stratigraphic unit. In other words, type 22 22 sections are the standard stratigraphic unit that is used to compare to other portions of the same unit found in other areas. The type section serves as the standard site for naming formations and has to be published as a type section and can never be changed. For this reason, type section locations are included along with a brief description of geological formations. The geologic formation exposures for the Kaweah River fluvial fan can be seen on the geologic map in the results section of this report, however, the type sections are located 100 miles to the north. In general, alluvium within the valley is derived from surrounding mountain ranges. Sierra Nevada alluvium (mostly igneous and metamorphic detritus) fills the east-side of the San Joaquin Valley whereas the Coast Ranges’ alluvium (Marine sediment) deposits into its western half creating an east-west compositional variance across the valley. On the valley floor, lacustrine, marsh, and pond deposits interfinger into the alluvium along the mountain flanks. The Turlock Lake, Riverbank and Modesto Formations all contain lacustrine deposits (Marchand and Allwardt, 1981), which is now correlated in this thesis, to Croft and Gordon (1968) and Page and LeBlanc (1969) lacustrine layers such as the E- clay layer identified as the Corcoran Clay Member of the upper Turlock Lake and Tulare Formations.

2.6 Late Pliocene and Pleistocene Stratigraphy of the Eastern San Joaquin Since Marchand and Allwardt (1981) designated most of the type sections and are to date, the most descriptive, the following section on stratigraphy is mostly a summation of their report and is used to describe the Kaweah River fan in the Weissmann et al. (2005) report. Convention for describing geologic formations is in order of the oldest to the youngest. Kaweah River fan stratigraphy is generally defined at the surface, yet consequently, subsurface 23 23

Figure 7. The San Joaquin Valley stratigraphic nomenclature (from Lettis and Unruh, 1991). Note: Marchand and Allwardt (1981) will be used in this study.

24 24

Figure 8. East San Joaquin Valley geologic formations (from Weissmann et al., 2005). Note: Pleistocene geologic formations and Holocene deposits of the eastern San Joaquin Valley fluvial fans. This includes the study area located on the Kaweah River fan which shares a border with the Kings River fluvial fan both within the Tulare Basin. The fluvial fan boundaries are outlined in bold-black. 25 25 geology is not well-understood. The older stratigraphic units created during the Paleogene (66 - 23Ma) and Neogene (23 -2.588Ma) are the Ione, Valley Springs, Mehrten, and the Laguna Formations yet will not be described in detail in this thesis with exception to the Laguna. Laguna is briefly mentioned since it has been found in the Kings River fan (Weissmann et al. 2005) and the Kaweah River fan. This section is organized to focus on key descriptions, origins, ages, and soil series associations with each formation.

2.6.1 Laguna Formation The Laguna Formation underlies the North Merced Gravel and was being formed during a period when the Tulare Basin was still an open basin connecting to the Pacific Ocean between cities of Coalinga and Santa Cruz. The marine shelf at that time was between the modern cities of Fresno, and Coalinga, California (Bartow, 1991). East of the Marine shelf, alluvium filled the basin as a result of deposition from Sierra Nevada river drainages. The Laguna Formation has been estimated as being Pliocene and possibly, in part, Pleistocene in age (Piper et al., 1939; and Marchand and Allwardt 1981). Additionally, Marchand and Allwardt (1981) also state that the Laguna Formation thins rapidly toward the south and does not crop out between the Chowchilla and the San Joaquin Rivers. The Laguna Formation is divided into two units called the upper (younger) and lower units (older). The China Hat Member is included as a portion of the upper unit of the Laguna Formation and is similar to the North Merced Gravel except it is older and arkosic; this is what distinguishes the Laguna Formation from the locally derived North Merced Gravel (Marchand and Allwardt, 1981). Laguna Formation was being deposited during McGee and Glacier Point (Weissmann et al., 2002) Sierra Nevada glacial stages (Figure 5). 26 26 2.6.2 North Merced Gravel Type section for North Merced Gravel as described by Marchand and Allwardt, (1981) is located off the Merced County landfill access road where it is exposed overlying the lower unit of the Laguna Formation with an unconformable contact. It has been roughly estimated to be Pleistocene or Pliocene by Marchand and Allwardt (1981) and later correlated in age to the lower Tulare Formation as Pleistocene by Lettis (1982). This formation is the parent material of the overlying Redding soil series which has been classified as an Abruptic Durixeralfs, and does not crop out in the incised banks of the San Joaquin River or any river fan further to the south. North Merced Gravel deposits occurred during the McGee glacial stage (Lettis, 1982). This formation is missing in the Tulare Basin where the study-site for this thesis was conducted. Moreover, it has not been found in the Kings River fan by Weissmann et al. (2002 and 2005).

2.6.3 Turlock Lake Formation Turlock Lake Formation was first designated its type section by Davis and Hall (1959). However, because an unconformity and well developed buried soil were not included in the original description, Marchand and Allwardt (1981 p. 26) assigned a new type section area located on a series of road-cuts in Turlock Lake State Park, uphill from the Park Ranger Quarters (Turlock Lake State Park, Cooperstown 7.5-minute quadrangle, Stanislaus County). This formation is divided into two units called the lower and upper units similar to the Laguna. The upper unit includes two members called the Corcoran Clay and Friant Pumice. The Turlock Lake Formation has excellent marker horizons such as Brunhes- Matayama magnetic reversal (contact between the upper and lower units of the Turlock Lake Fm.), Corcoran Clay Member (well defined lacustrine bed throughout the Central Valley), Friant Pumice (distinct volcanic ash in 27 27 conformable contact on top of the Corcoran Clay), and paleosols (identified as duripan, pedogenic clay, or root traces found in the lower Turlock Lake type section). Furthermore, it can be recognized by its arkosic fine sand, silt, and clay at the base grading upward into coarse sand and occasional coarse pebbly (granitic, metamorphic, volcanic, and quartz-vein rocks) sand or gravel and lacks the medium gravels of the North Merced Gravel (Marchand and Allwardt, 1981). Turlock Lake Formation typically rests between the North Merced Gravel at its base and the Riverbank at its top. However, North Merced Gravels are not always found in contact with the lower Turlock Lake Formation; the bottom of the Turlock Lake Formation unconformably overlies the Miocene-Pliocene Mehrten (Davis and Hall, 1959) and Laguna (Weissmann et al., 2002). This formation can be easily distinguished from the underlying Laguna Formation, especially where North Merced Gravel is absent (Marhcand and Allwardt, 1981). Since Laguna and Turlock Lake Formations are similar, Marchand and Allwardt (1981, p.32) list criteria useful for differentiating the two formations as well as some key characteristics seen within the type section with respect to degree of alteration and location. The Turlock Lake and upper-Tulare Formation (stratigraphic nomenclature for West-Central San Joaquin Valley) correlate in age and share the Corcoran Clay Member (Lettis, 1982; Lettis and Unruh, 1991). In the Chowchilla area, thicknesses of the Turlock Lake Formation can range from 17-70 m in the upper unit and 33-150 m in the lower section (Marchand and Allwardt, 1981) and both increase in thickness as they approach the center of the San Joaquin River basin or valley topographic low. Several authors agree that paleosols can be observed within the sequence (Arkley, 1962; Janda, 1966; Janda and Croft, 1967; Marchand and Allwardt, 1981, and; Weissmann, 2002) since this can be used to help identify the unit. The Donner 28 28

Lake or Red Meadow Sierra Nevada glacial stages are related to the upper Turlock Lake Formation (including the Friant Pumice and Corcoran Clay Members) and the Sherwin or El Portal glacial stage to the lower unit (Weissmann et al., 2002). The Turlock Lake Formation recorded a geomagnetic reversal chron known as the Brunhes-Matuyama reversal (Davis et al., 1977; Marchand and Allwardt, 1981; Weissmann et al., 2002) which occurred 781,000 years ago (Gradstein et al., 2012). This magnetic reversal event is found near the contact between the upper and lower units of the Turlock Lake Formation beneath the Corcoran Clay Member as seen in Figure 5. The lower unit also recorded the Jaramillo subchron. The Corcoran Clay Member was created from the massive Pleistocene Lake that existed in the San Joaquin Valley (Frink and Kues, 1954) and was named the Corcoran Lake by Bartow (1991). Moreover, the Corcoran Clay Member is by far the best marker-bed within Quaternary San Joaquin Valley stratigraphy due to the continuous conformable contact with Bishop ash and its massive diatomaceous clay content (Lettis, 1982) that extends throughout most of the San Joaquin Valley subsurface. The upper unit of the Turlock Lake Formation has been estimated as being about 600,000 years by Marchand and Allwardt (1981) based on the Brunhes-normal magnetic polarity, vertebrate fauna (Janda, 1965,1966), uranium- series ages of two Montpelier soils (560,000±80,000 and 540,000±50,000 yr), and stratigraphic relations with K-Ar dating of the Friant Pumice Member (615,000±31,000 yr. by Janda, 1965,1966; Janda and Croft, 1968). Newer publications suggest the Brunhes-Matayama magnetic reversal (Gradstein et al., 2012) and Friant Pumice (Sarna-Wojcicki et al., 2000) to be older than previously reported by Marchand and Allwardt (1981). The Friant Pumice Member type section is found at a quarry (Outback Materials Inc.) located roughly 20 miles north-east of the city of Fresno on the 29 29 western edge of the Sierra Nevada where the foothills connect the San Joaquin Valley basin. The Friant Pumice Member has been correlated with Bishop Tuff and ash from the Long Valley Caldera which erupted roughly about 758,900 (±1.8 Ka) years ago (Sarna-Wojcicki et al., 2000). Newer data suggest the Friant Pumice Member has an average age of 750.0 ±4 Ka (Sarna-Wojcicki et al., 2000). Note that soil series were used by Marchand and Allwardt (1981) and Weissmann et al. (2005) to help create geologic maps and is used to construct a geologic map of the Kaweah River fan in this thesis. For this reasons, and to preserve originality of Marchand and Alwardts correlations of soil series to formations, the soil series of each formation is included for all formations. Typically, the Montpellier soil series formed in upper Turlock Lake Formation, whereas the lower Turlock lake soils series include the Rocklin, Whitney, Cometa, and the nonacid variety of Corning. In addition, the subgroup Mollic Haploxeralfs is commonly found in this formation (Marchand and Allwardt, 1981). An iron-silica duripan (Rocklin) can be found near the type section of the lower unit of the Turlock Lake Formation and includes a sequence of gravel and coarse sand that overlies finer sand, silt, and clay of possible lacustrine origin (Marchand and Allwardt, 1981). A disconformity is common where the youngest Turlock Lake and oldest Riverbank alluvium come in contact. Mima-mound microrelief (hog-wallow) characterizes those areas where duripan soils, that are typically associated with Riverbank erosional surfaces, bevel across Turlock Lake deposits (Marchand and Allwardt, 1981). Figure 9 shows a map from Preston (1981) and sketch from Means and Holmes (1900) that shows mima- mound microrelief of the Tulare Lake Basin before land was altered by farming. Duripan soils are good markers to help identify stratigraphy since they are distinct pedogenic subsurface horizons. Typically, duripans are found in Redding 30 30 soils of North Merced Gravel Formation and China Hat Member; the Rocklin soils of lower Turlock Lake Formation; the Alamo, Exeter, Madera, Madera, San Joaquin, and Yokohl soils of the Riverbank Formation; and finally, the Fresno soil series of the Modesto Formation (Marchand and Allwardt, 1981). In some cases, the duripans are relict soils that are paleosols that have been assimilated in younger soil horizons (Marchand and Allwardt, 1981). According to the Soil Survey Staff (2010), duripan is a silica-cemented subsurface horizon with or without auxiliary cementing agents (such as iron, magnesium, or calcite) and must satisfy 4 required characteristics or diagnostics. Duripan in the Kaweah River region is typified by the Riverbank Formation.

2.6.4 Riverbank Formation The Riverbank Formation consists of three (upper, middle, and lower) units primarily composed of arkosic sediment derived from Sierra Nevada detritus that typically form disconformities above the upper unit of the Turlock Lake Formation (Marchand and Allwardt, 1981). The three units are divided on a basis of superposition, soil type and alteration, and other geomorphic evidence such as the comparison of formation elevation relative to each other in the San Joaquin Valley (Marchand and Allwardt, 1981). Units tend to have abundant sand mixed with scattered pebble and gravel lenses, a few interbedded fine sands and silt, and are sometimes stratified due to local ponds during deposition (Marchand and Allwardt, 1981). Eolian sandy deposits can be found in this formation where Snelling and Ramona soils exist yet both the eolian and lacustrine deposits are rare (Marchand and Allwardt, 1981). The type section is located on the south bluff of the Stanislaus River in the town of Riverbank; however, all three units can be seen in superposition located north of the in the southeastern part of the 31 31

Figure 9. Duripan soils of the Tulare Basin. Note: A) Mimi-mounds microrelief map (modified from Preston, 1981) also called hog-wallows are typical of duripan soils. B) Sketch of hog-wallow mounds (from Means and Holmes, 1900). The Kings and Kaweah River fans cover sections of the Riverbank duripan. Duripan is used to identify the Riverbank Formation.

32 32

Turlock Lake and the southwestern part of the Snelling (Davis and Hall, 1959; and Marchand and Allwardt, 1981). The duripan, similar to the Madera, normal variant, and Snelling weak variant, developed on the Riverbank sediments (Marchand and Allwardt 1981). Thicknesses of the units vary within the Riverbank (45-60 m) yet they tend to increase toward the major river channels and basin-ward (Marchand and Allwardt, 1981). Croft and Gordon (1968) reported a thin hardpan layer in excavations at depths of about 25 feet (7.6 m) below land surface in Visalia, which they suggested as the top of the older alluvium which has since been named Riverbank Formation. In general, the Riverbank Formation age ranges from 130,000 to 450,000 years (Marchand and Allwardt, 1981). More specifically, however, the three units of the Riverbank have been correlated with cold-water marine isotope by Marchand (1977) as stages 6 (127,000-190,000), 8 (247,000-276,000), and 10 (336,000-356,000) and/or stage 12 (425,000-457,000) estimates based on Hays et al., (1976). Rancholabrean vertebrate fauna, Illinoian or Wisconsinan in age, was found associated with the Riverbank (Janda, 1965, 1966). Riverbank deposits probably occurred during the Tahoe, Mono Basin or Casa Diablo Sierra Nevada glacial stages (Lettis, 1988; Weissmann et al., 2002). Lettis (1988) suggest the units of the Riverbank were deposited beginning at 130 Ka for the upper, 250 Ka for the middle, and 330 Ka for the lower unit. Marchand and Allwardt (1981) define the lower unit as everything between the middle unit of the Riverbank and youngest unit of the Turlock Lake Formations which include deposits from more than one aggradation cycle. The lower unit alters to strong variants of Snelling and Ramona soil and weak variants of Montpellier and Cometa soil series (Marchand and Allwardt, 1981). The lower unit is best preserved near its type section (Marchand and Allwardt, 1981). 33 33

The middle unit is generally comprised of a broad sandy-alluvial surface representative of a single aggradation unit that is commonly deposited over eroded Turlock Lake Formation (Marchand and Allwardt, 1981). In addition, it is reasonably well preserved along the flank of the Sierra Nevada, especially in the Chowchilla and basins (Marchand and Allwardt , 1981). The middle unit alters to normal variants of Snelling (Typic Haploxeralfs), San Joaquin (Abruptic Durixeralfs), Madera (Typic Durixeralfs) and Alamo soil series (Typic Duraquolls), in addition to strong variants of duripan within the San Joaquin, Madera, and Yokohl soil series (Marchand and Allwardt , 1981). These duripans horizons are thicker and more strongly cemented, oxidized, and clay-rich as compared with those formed on the upper unit (Marchand and Allwardt, 1981). The upper unit, like the middle, is representative of a single aggradation unit believed to be related to major rivers draining the glaciated interior of the Sierra Nevada (Marchand and Allwardt , 1981). Furthermore, the upper unit, like the other two units, consists of an upward coarsening sequence of alluvial silt and sand, with thin pebble or coble lenses near the top of the sequence (Marchand and Allwardt, 1981). Both the upper and middle units are separated by a disconformity in the form of a moderately developed duripan of the San Joaquin, Madera and Exeter soil series (Marchand and Allwardt, 1981). This can be seen near the south bluff of the San Joaquin River about one-km west of California Highway 99 (NE 1/4NE1/4 sec. 6, T. 13S., R. 19 E.; Herndon quadrangle) (Marchand and Allwardt , 1981). This unit alters to weak variants of Cometa, Madera, San Joaquin, Exeter, and a weak variant of Ramona soil series (Marchand and Allwardt, 1981). 34 34 2.6.5 Modesto Formation The Modesto Formation overlays the Riverbank Formation and is divided into two units, the younger upper unit which rarely exceeds 10 m in thickness and the older and thicker lower unit which measures 25 m or more (Marchand and Allwardt, 1981). Generally, the Modesto Formation is composed of arkosic sediments (virtually identical to the Laguna, Turlock Lake, and Riverbank Formations) and represents the last major series of aggradation, which is usually associated with the best soils for agriculture and foundations of most major cities in the eastern San Joaquin Valley (Marchand and Allwardt, 1981) including Visalia. The terraces and young fans of the Modesto Formation can be differentiated from the other formations by associations with lower elevations, moderate to slight degree of erosional modification, and a soil profile characterized by more weathering than Holocene deposits with exception to the upper Modesto eolian sand (Marchand and Allwardt, 1981). Thickness range from 15-40m with thickest portions found near river channels, furthermore, local facies changes within the formation render stratigraphic descriptions almost meaningless (Marchand and Allwardt, 1981). The revised type section for the Modesto Formation is located on a series of cuts exposed in a private road down south bluff of , just east of Mitchell Road Bridge, Stanislaus County (Marchand and Allwardt, 1981). The Modesto Formation is Wisconsin (includes Tahoe and Tioga) in age based on its youthful topography, geomorphic position, soils, radiocarbon dating, and disconformity between the lower Modesto and upper Riverbank Formations. The Formation must be older than 9,000 years B.P. based on radiocarbon dating on wood found in subsurface deposits (Janda and Croft, 1967; Croft, 1972) and about 14,000 and 42,000 years B.P. based on wood found in the upper member 35 35

(Marchand and Allwardt, 1981). The lower member has a minimum age of 29,407±2,027 years B.P. using open-system uranium-series age on bone; a medial age of 42,400±1,000 years B.P. based on radiocarbon age of wood buried in the Modesto flood basin deposits; and a maximum age of 95,000±45,000 years B.P. based on 12 Uranium-series dating tests on Greenfield soil (Marchand and Allwardt, 1981). In addition, the upper Member has a minimum age of 9,000 years B.P.; a medial age of 14,060±450 years B.P. (Ives et al., 1967; Croft, 1968) based on wood found at the top of lacustrine sediment; and has a maximum age 29,407±2,027 years based on its contact to the lower unit (Marchand and Allwardt, 1981). Since the Tioga Glaciation ended about 9,000 (Adam, 1967) or as early as 11,000 (Clark and Gillespie, 1997) years ago, this suggests that the closing phases of the Tioga Glaciation of the Sierra Nevada correlate to the upper Modesto (Janda and Croft, 1967; Marchand, 1977; Marchand and Allwardt, 1981). The Altonian (early and middle Wisconsinan) lower unit of the Modesto Formation is typically characterized by basin soils (Natric and Typic Haploxeralfs), pond soils (Natric Durixeralfs and Typic Natrixeralfs) and alter to Borden (Natric Hapoxeralf), Greenfield (Typic Halpoxeralfs), Fresno, Waukena, Traver and Dinuba soil series (Marchand and Allwardt, 1981). Typically these soils show minimal argillic horizons with Munsell soil color of 10YR to 7.5YR hues and their profiles are shallower, higher pH , are as red as those developed on Riverbank alluvium, and are absent of the silica-iron duripans (Marchand and Allwardt, 1981). However, the Fresno soil series contains calcium carbonate silica-cemented duripans with overlying argillic horizons, which developed under impeded drainage in early Modesto overbank and flood basin rim deposits (Marchand and Allwardt, 1981). The lower member also displays well-stratified silt and fine sand (commonly near its base) similar to the Laguna, Turlock Lake, 36 36 and Riverbank Formations, yet is more abundant in gravels near the foothills and above the lower Modesto fan apexes (Marchand and Allwardt, 1981). Furthermore, it can include a large amount of locally derived andesitic and metamorphic detritus (Marchand and Allwardt, 1981). The Modesto Formation is identified in the Kaweah River fan using soil series from Marchand and Allwardt (1981), using the boundaries of the older fan deposits devoid of duripan, using soil descriptions and maps, and comparing to the proximal (lower unit) and distal (upper unit) fan deposits from Weissmann et al. (2005). The upper member of the Modesto Formation is divided into four phases defined by terraces fans and stream channel deposits from major Sierra Nevada rivers (Marchand and Allwardt, 1981). They can be recognized by their topographic position between the lower post-Modesto (Holocene) terrace and flood-plain surfaces, and the higher terraces and fan surfaces of the lower member and the Riverbank Formation (Marchand and Allwardt, 1981). Both phases 1 (oldest-base) and 4 (youngest-top) form narrow benches along present rivers and are often closer in elevation to Holocene flood plain terraces than phases 2 and 3 (Marchand and Allwardt, 1981). Phase 1 typically alters to thick AC horizons in Pachappa, Oakdale, Hanford, and Chino soils whereas, phases 2 and 3 alter to Ap/AC/Clox/Cn profiles developed in Grangeville, Cajon, Wunjey, and Visalia soils, and phase 4 develops in Tujunga soil (Marchand and Allwardt, 1981). Delhi and Hilmar soil profiles (Typical Xeropsamment; Aerie Halaquept) are eolian sand deposits commonly, in large amounts, overlying alluvium of all four upper member terraces (Marchand and Allwardt, 1981). Terrace-fan surfaces may represent separate alluvial fills in which they could be designated as a subunits (Marchand and Allwardt, 1981). 37 37 2.7 Holocene Stratigraphy

2.7.1 Post-Modesto Formation The Post Modesto Formation is divided by age into four Holocene stages informally called Post-Modesto I (oldest), II, III, and IV (youngest) and are generally unweathered and thin alluvial fans which incise or cover the Modesto Formation (Marchand and Allwardt, 1981). These stages are recognized along most of the major rivers by topographic expression and position in addition to minor differences in soil profile development (Marchand and Allwardt, 1981). Typically Post-Modesto I underlie low terraces or high flood plains about 6-8 m above base flow levels along the Stanislaus, Tuolumne, Merced, and San Joaquin Rivers and stands slightly above the more extensive post-Modesto II terrace which can be seen along the lower Merced River (Marchand and Allwardt, 1981). Post- Modesto II terraces do not converge significantly downstream toward the modern stream profile and deposits include channel, point bar, levee, crevasse splay, interdistributary and flood basin alluvium about 3-5 m above river level) and usually measures close to pH 8 (Marchand and Allwardt, 1981). Post-Modesto III deposits include fluvial sand, silt, and clay associated with abandoned meander loops, arcuate levees (2-4 m above present river level) within the San Joaquin Valley and include modern flood plain which stands slightly above the modern channel-ways (Marchand and Allwardt, 1981). Post-Modesto IV is comprised of modern day deposits which include lacustrine, swamp, marsh and overbank flood deposits in addition to un-vegetated channel and point bar alluvium along all major streams (Marchand and Allwardt, 1981). Post Modesto Formation ranges from 9000 years B.P. to present day. Early Holocene Post-Modesto I (9000-4000 B.P.) contains charcoal and organic carbon both found in an A/Cox/Cn soil from the Yolo series which dated about 38 38

4,000 years B.P. by Shlemon and Begg (1972), and are found in the A horizon of an A/AC/Cox/Cn profile dating at 9,150±650 years B.P. (Marchand and Allwardt, 1981). Prehistoric post-Modesto II (4000-3000 years B.P.) deposits contained charcoal which measured 3,020±130 years B.P. using carbon-14 age dated and bone collagen yielding an age of 2,985±160-year B. P and 4,440±230-year B.P. using uranium-thorium age by University of California, Davis (Marchand and Allwardt, 1981). Historic Post-Modesto III has an unknown age but probably spans the last few hundred years whereas Post-Modesto stage IV represent modern day sediments (Marchand and Allwardt, 1981). This suggests that Post-Modesto III deposition occurred, in part, during Matthes glaciation, however, no evidence of glaciation is seen in San Joaquin Valley Holocene sediments as opposed to the older sediments. Soil development of post-Modesto I typically contain Cox horizon development toward an AC horizon, yet exposures are rare (Marchand and Allwardt, 1981). Local bodies of eolian sand, probably reworked from late Modesto eolian deposits, and in some cases related agricultural practices, can be found in both post-Modesto I and II terraces (Marchand and Allwardt, 1981). Post-Modesto II contain dark, clay-rich Temple and Merced soils in addition to Tujunga (Typic Xeropsamment) and Grangeville (Fluraquentic Haploxeroll) soils (Marchand and Allwardt, 1981). Major river alluvium altered to Foster (derived from arkosic alluvium along the major rivers), Honcut, and Yolo series along with the Columbia (Aquic Xerofluvent) on locally derived Coast Range-derived detritus (Marchand and Allwardt, 1981). Post-Modesto phase III soils developed in underdeveloped variants of the Columbia, Tujunga, Grangeville, Foster, Honcut, and Marguerite soil series (Marchand and Allwardt, 1981). Post-Modesto 39 39

IV modern deposits underlie presently active depositional surfaces and show neither A nor Cox soil horizons (Marchand and Allwardt, 1981).

2.8 San Joaquin Valley Hydrogeology

2.8.1 General Background The Central Valley (Figure 3), also known as the Great Valley, is one of the ten hydrological Provinces of California (Belitz et al., 2003; Burton and Belitz, 2008; Faunt et al., 2009a). It is bound by four mountain ranges: The Klamath Mountains and southern Cascades to the north, the Sierra Nevada to the east, and both Northern and Southern Coast Ranges to the west (Figure 2). The Central Valley is divided into two smaller valleys by the Sacramento River and its southern-most tributary, the south fork of the (Faunt et al., 2009a). These smaller valleys are called the Sacramento (north-side) and San Joaquin (south-side) Valleys. Structurally, the San Joaquin and Sacramento basins are separated by the buried Stockton arch that was uplifted during the early Paleogene along the Stockton fault (Bartow, 1991). Furthermore, the San Joaquin Valley is divided at the interface between the Kings River and Los Gatos Creek alluvial fans into two smaller basins known as the San Joaquin and Tulare basins (Faunt et al. 2009a). The California Department of Water Resources has further divided the San Joaquin Valley into groundwater basins and their subbasins to help manage groundwater. Descriptions and maps of the groundwater basins and subbasins are available at California Department of Water Resources website (www.water.ca.gov/). The Kaweah River fan covers most of the Kaweah subbasin (groundwater subbasin number 5-22.11) area which is within the Tulare Lake groundwater basin. This can cause confusion since the Tulare Lake basin (location of Tulare Lake bed) and Tulare Lake groundwater basin share the same 40 40 name. Tulare Lake groundwater basin is not used for this thesis since its region includes parts of the Sierra Nevada. For further inquiry into groundwater basins refer to the DWR website. For modeling purposes, Faunt et al. (2009a; 2010) divided the Central Valley into 9 spatial provinces based on general textural characteristics from select driller’s logs (Figure 10). This includes Redding, and Western and Eastern Sacramento spatial provinces for the Sacramento Valley, likewise, Tracy/Delta- Mendota Westside, Northern San Joaquin, Chowchilla-Madera, Kings, and Tulare/Kern spatial provinces for the San Joaquin Valley. Spatial provinces were produced from interpolated geostatistics based on soil textures of the subsurface which was used to estimate hydrologic properties. The Kaweah River has supplies sediment found in the Kaweah Subbasin of the Tulare/Kern spatial province and covers a large portion of its area (Tulare and Kings Counties).

2.8.2 Controls of Groundwater Flow According to Page and LeBlanc (1969) the groundwater occurrence and flow is not controlled by faulted units, especially in the deep subsurface of the San Joaquin Valley; however, two nearly vertical joint sets and one nearly horizontal set outline a network of large rectangular blocks found in the basement complex that crop out along the mountain front and presumably underlie the buried shelf which permits the basement complex to yield water. With exception to joint sets within basement rock, the main controls of groundwater in the San Joaquin Valley are the prominent Quaternary confining beds. The confining beds present in the San Joaquin Valley trough consist of lenses of greenish-gray, gray and blue

(GLEY: 5G, 5GY, 10G, 5B, 5BG) silt, clayey silt and subordinant silty clay as a

41 41

Figure 10. Spatial provinces map (modified from Faunt et al., 2009). Note: Figure based on soil texture map of the Central Valley textural distribution based on selected driller’s logs from ground water basins and subbasins. They are grouped into nine regions or spatial provinces to a depth of 1200 feet. The Kaweah Subbasin is highlighted in red. 42 42 result of flood, lacustrine and paludal deposition composed of both Coast Range and Sierra Nevada bedrock lithology (Lettis, 1982 p 142). The E-clay, also known as Corcoran Clay Member of the upper (Figure 11- A) Turlock lake Formation is the thickest and most extensive of all the lacustrine and marsh deposits (Page and LeBlanc, 1969) and is modeled as layers 4 and 5 as an aquitard (Figure 11-B) by the state hydrogeologists (Faunt et al., 2010). The locations of the cross-sections lines are sketched on Figure 10. The Tulare Lake basin has twice the amount of lacustrine and marsh deposit as seen in the rest of the San Joaquin Valley and is generally shaped like a deep, somewhat cylindrical plug from which several irregularly shaped discs, or tongues, radiate at irregular intervals (Croft and Gordon, 1968; Page and LeBlanc, 1969). These prominent clay layers are informally, from oldest to youngest, the F, E, D, C, B, and A clays for the Tulare Basin, however the E, C, and A clays are present through the rest of the San Joaquin Valley and Fresno area (Croft and Gordon, 1968; Page and LeBlanc, 1969; Page, 1986). Both the C and A clays are not found north of the city of Mendota yet exist in the southern extent of the San Joaquin Valley and near the San Joaquin Basin axial trough (Page, 1986). Furthermore, beneath the Tulare Lake bed within the Tulare Basin, the E-clay boundary separates the upper semi confined zone from the lower confined zone that together extends to maximum depths at about 3000 feet. These distinct clays are confining beds that are the boundaries between aquifer layers seen in Figure 11B and will be discussed in more detail in the confining beds section. 43 43

Figure 11. Cross sections of the San Joaquin Valley. Note: A) Generalized cross section modified from Lettis (1982) of the northern San Joaquin Valley showing Pleistocene formations. Corcoran Clay was formed from a massive Pleistocene lake and is the Corcoran Clay Member of the upper Turlock Lake Formation. B) Generalized hydrogeologic cross section showing groundwater flow system of the Central Valley of California which is modified from Faunt et al. (2010). Corcoran clay is an aquitard labeled as Layers 4 and 5 and also called the E-clay. The confining beds are denoted as the A,B,C,D,E,F,G, H, I, and J clays and are boundaries between aquifer Layers. See Figure 10 for cross-section line sketch. 44 44 2.8.3 The Confining Beds The confining beds (i.e. J, I, H, G, F, E, D, C, B, and A clays) in Figure 11B, are boundaries between aquifers in the San Joaquin Valley yet they have not been correlated to stratigraphic formations with exception to the E (Corcoran Clay Member) and A (Modesto Formation) clays. The E, C, and A clays exist in both the Tulare and San Joaquin Basins and are Quaternary in age (Page, 1986) which suggests the both B and D tongues found in the Tulare Basin are of the same age. The extent of these confining beds in the Tulare Basin has been mapped by Burton and Belitz (2008) seen in Figure 12. Generally, the confining beds extend from the Tulare Lake basin and axial trough of the San Joaquin Valley and interfinger into the fluvial fans deposits near the base of the mountain ranges. The F clay has a subsurface area similar to the historical Tulare Lake bed and interfingers into the reduced deposits of the older alluvium (Croft and Gordon, 1968). Furthermore, this confining bed has been found to be diatomaceous (Page and LeBlanc, 1969). The F clay resides in the Pleistocene reduced older alluvium described by Croft and Gordon (1968) and lies beneath the Corcoran Clay Member of the upper Turlock Lake and Tulare Formations. Page and LeBlanc (1969) estimated the F clay to be about 265 square miles beneath the western part of Fresno, and is reported as the largest aquiclude in that area. This suggests that the F clay possibly rests within the lower Turlock Lake formation, and therefore would correlate to Sherwin or El Portal Sierra Nevada Glacial stages. The E clay (Corcoran Clay Member) is a deposit from the Pleistocene Corcoran Lake that covered most of the San Joaquin Valley. The E clay is the thickest and most prevalent confining bed within the San Joaquin Valley and its great extent was first recognized by Frink and Kues (1954). It can be recognized by the dark greenish blue-gray, silty, diatomaceous clay that sits about 60-120 feet 45 45

Figure 12. Lacustrine clay-lenses regional extent map (from Burton and Belitz, 2008). Note: Both the B and D clay tongues share the same spatial extent as the E-clay (Corcoran Clay Member) lacustrine lens within the Tulare Basin (Croft and Gordon , 1968). These are the confining beds that bound San Joaquin Valley aquifers.

46 46 above the F clay and has an area roughly 1,000 square miles Croft and Gordon (1968). The D clay sits roughly 60 to 190 feet above the E clay and is mostly found in the Tulare Basin and therefore considered a tongue (Croft and Gordon, 1968) of the cylindrical clay plug beneath the Tulare Lake bed (Page and LeBlanc, 1969). The C clay deposit extends into both the Tulare and San Joaquin Basins and is found about 140 to 210 feet above the D clay, roughly 250 feet below land surface (Croft and Gordon, 1968). The B clay is a clay tongue of the Tulare Basin that is about 140 feet below land surface and extends roughly the boundaries of the D clay (Croft and Gordon, 1968). The D, C, and, B clays are likely interfingered into and part of the lower, middle, and upper sections of the Riverbank formation yet they have not been examined close enough to be certain. The D clay therefore would relate to Casa Diablo glacial stage and both the C and D clays to the Tahoe or Mono basin glacial stages respectively. Furthermore, all three clays are interfingered into the oxidized older alluvium and are all included in Layer 3 aquifer based on their depths and descriptions. The A clay exists roughly 40 to 50 feet below land surface that underlies the younger alluvium (Croft and Gordon, 1968). Younger alluvium is above the A clay (Croft and Gordon, 1968). Like the E and C clays, it extends throughout most of San Joaquin Valley, especially to the southern portions (Croft and Gordon, 1968; Page and LeBlanc, 1969; Page, 1986). The A clay was deposited during Wisconsin glaciation (Janda and Croft, 1967; Page, 1986). Based on Radio carbon dates of wood collected in the A clay yielded an age ranging from 26,780 ± 600 (USGS, 1965; Page, 1986) and 9,040 ± 300 years (Croft, 1972; Page, 1986). Croft’s age date for the A clay is unlikely due to previous radio carbon dates and lack of deformation along the San Andreas (Lettis, 1982; Page, 1986). The A clay 47 47 is found within the Modesto formation and therefore was deposited during the Hilgard, Tioga, or Tenaya glacial stages. Like the D, C, and B clays, the A clay interfingers into the oxidized older alluvium (Croft and Gordon, 1968). The Modesto formation is probably the Layer 2 aquifer whereas Layer 3 is Riverbank and upper Turlock Lake Formations. Holocene deposits likely make up large portions of the Layer 1 aquifer.

2.8.4 San Joaquin Valley Aquifers Stratigraphic formations have been found to contain aquifers. For instance, the Mehrten Formation yields large quantities of water in the San Joaquin Basin (Page, 1986). The Turlock Formation, as a whole, is an important aquifer in eastern Stanislaus County and its sand and gravels can yield up to two thousand gallons per minute of good quality water (Davis and Hall, 1959). The Riverbank Formation hosts the principal aquifer for Riverbank, CA area and produces water of good quality (Davis and Hall, 1959). The Modesto Formation is an important aquifer that yields large quantities of water (up to 2,000 gallons per minute), is characterized by local variations in permeability (such as lacustrine versus alluvium) due to the lenticular nature of sediments, and is generally found to produce good quality water (Davis and Hall, 1959). Tulare Basin aquifers were described in detail by Croft and Gordon (1968), from top to bottom, as unconfined (units that contain the Water table) above the A-clay, semiconfined (affected by pumping and loading and adjusts to water table) between the A and C clays, and the confined (artesian wells) aquifers beneath the E clay. Page and LeBlanc (1969) separated the aquifers into five bodies of water based on electric logs that can be applied to the San Joaquin Basin and differ from the Tulare Basin slightly. For instance, San Joaquin Basin’s aquifers are from top 48 48 to bottom: (a) the unconfined water body which is the major aquifer for the Fresno area as of 1969, (b) the shallow water body, (c) the confined water body below the A-clay (d) the confined water body below the C-clay, and (e) the confined water body below the E-clay. Generally, hydraulic head drops with increased depth that suggests the clays allow for slow downward passage of groundwater (Page and LeBlanc, 1969). The three confined aquifers below the A clay can be identified by differences in hydraulic head in nearby wells (Croft and Gordon, 1968; Page and LeBlanc, 1969). In the Fresno area, the depth to the E clay is 460 feet, to the C clay about 240 feet, and to the A clay about 60 feet (Page and LeBlanc, 1969; Mitten, 1984). Croft and Gordon (1968) suggest the oxidized older alluvium as highly permeable and is a major aquifer for Tulare County. Furthermore, DWR (2004) confirms that older alluvium contains the major aquifer for the Kaweah Subbasin. The oxidized older alluvium rests above the reduced older alluvium (Croft and Gordon, 1968, p.15). The older alluvium is Pleistocene and therefore correlates to the Pleistocene Formations for the San Joaquin Valley. Table 1 shows an attempt to correlate the aquifers of Faunt et al. (2010) to the stratigraphic formations (Marchand and Allwardt, 1981), Sierra Nevada glacial cycles (Fullerton 1986; Lettis, 1988; Clark and Gillespie, 1997; Weissmann et al., 2002), confining beds (Croft and Gordon, 1968; Page and LeBlanc, 1969; Croft, 1972; Page, 1986), and former Tulare Basin subsurface descriptions (Croft and Gordon, 1968). The unconfined aquifer (zone of fluctuating Water table) contains both the younger and oxidized older alluvium and flood plain deposits (Croft and Gordon, 1968). The semiconfined aquifer is composed of continental deposits and older alluvium (Croft and Gordon, 1968). The confined aquifer consists of reduced continental

Table 1. Summary and Correlations from Geologic Background. Note: The G, H,I, and J confining beds are hypothetical and are created for this report based on Faunt et al. (2009; 2010) boundaries between layers and Croft and Gordon (1968) descriptions. The asterisk (*) is representative of authors who suggest the location of A Clay in the upper Modesto Formation which are: USGS, 1965; Croft and Gordon, 1968; Croft, 1972; Marchand and Allwardt, 1981; Page, 1986)

Sierra Nevada Glacial Marine Stratigraphic Formations Hanford-Visalia Stages Pre-Holocene Tulare Basin Tulare Basin Isotope Approximate Age of the Eastern-Central Area Water (Fullerton, 1986; Terminology Confining Beds Confining Beds Aquifers stage Thousands of San Joaquin Valley Quality Type Section Authors: Lettis, 1988; and (Croft and Gordon, of River Fans the Tulare Basin (Faunt et al., (Lettis, Years (Marchand and (Croft and Weissmann et al., 1968) 2009 and 2010) 1988) Allwardt, 1981) Gordon, 1968) 2002)

Matthes (Clark and 10 Ka or Holocene Post-Modesto Formation Younger Alluvium and Gillespie, 1997) 1 (Lettis, 1988) or Holocene flood-basin deposits None Younger Alluvium Layer 1 Good Marchand and Allwardt, 1981

(8.2 to 14) Ka Modesto Formation Older Alluvium A-Clay top of U- Hilgard or Tioga 2 (Lettis, 1988) (Upper) (Oxidized) None Modesto Fm. (*) Layer 1 and/or 2 Good Marchand and Allwardt (1981)

A-Clay top of L- (20 to 70) Ka (Lettis, Modesto Formation Older Alluvium Modesto Fm. Tenaya or Tahoe 4 1988) (Lower) (Oxidized) Rare Duripan (Lettis, 1988) Layer 2 Good Marchand and Allwardt (1981)

Riverbank Formation Older Alluvium Tahoe or Mono Basin 6 130 Ka (Lettis, 1988) (Upper) (Oxidized) Duripan B-Clay? Layer 3 Good Marchand and Allwardt (1981)

C-Clay base M- Riverbank Formation Older Alluvium Indurated Riverbank Fm. Tahoe or Mono Basin 8 250 Ka (Lettis, 1988) (Middle) (Oxidized) Duripan (Lettis, 1988) Layer 3 Good Marchand and Allwardt (1981)

10 and/or Riverbank Formation Older Alluvium Davis and Hall (1959) and Casa Diablo 12 330 Ka (Lettis, 1988) (Lower) (Oxidized) Duripan D-Clay? Layer 3 Good Marchand and Allwardt (1981)

49

Donner Lake or Red (570 to >615) Ka Upper-Turlock Lake Older Alluvium Medow 16 and/or18 (Lettis, 1988) Formation (UTLF) (Oxidized) Paleosols D-clay? Layer 3 Good Marchand and Allwardt (1981) 750.0 ±4 Ka (Ave. from Sarna- Confomable Donner Lake or Red Wojcicki et al., UTLF (Friant Pumice Older Alluvium Friant Pumice contact above the E- Aquitard Top - Meadow 17 and/or18 2000) Member) (Reduced) Member Clay Layer 4 N/A Janda (1966) <781 Ka Bruhnes/Matayama Donner Lake or Red (Gradstein et al., UTLF (Corcoran Clay Older Alluvium Interfingers with Aquitard-Layers 4 Frink and Kues (1954); Lettis, Meadow 17 and/or18 2012) Member) (Reduced) Paleosols = E-Clay and 5 N/A 1982)

(778 to 900) Ka Lower-Turlock Lake Older Alluvium Paleosols and F (base of L6) and G Sherwin or El Portal 20 (Lettis, 1988) Formation (Reduced) Duripan (base of L7) Layers 6 and 7 Good Marchand and Allwardt (1981) Pleistocene or Pliocene (Marchand and ? 21? Allwardt, 1981) North Merced Gravel Not in area Duripan not in area not in area not in area Marchand and Allwardt (1981)

(2,000 to 2,500) Ka Older Alluvium, H (base of L8) Gales/Piper et al., (1939); or Late Pliocene lacustrine, and marsh and/or I (base of Janda 1965&1966; Marchand McGee or Glacier Point (Lettis, 1988) Upper-Laguna Formation deposits China Hat duripan L9)? Layers 8 and/or 9? Good and Allwardt (1981)

(2,000 to 2,500) Ka Older Alluvium, Gales/Piper et al., (1939); or Late Pliocene lacustrine, and marsh H (base of L8) Janda 1965&1966; Marchand McGee or Glacier Point (Lettis, 1988) Lower-Laguna Formation deposits None found and/or I (base of L9) Layers 8 and/or 9? Good and Allwardt (1981) Miocene to late Pliocene Gales/Piper et al., (1939); Davis (Marchand and Continental Deposits I (base of L9) and/or Layers 9 and/or and Hall (1959); Marchand and Allwardt, 1981) Mehrten Formation (Oxidized) paloesols J (base of L10)? 10? (NaHCO3-rich) Allwardt (1981) to Miocene (Marchand and Continental Deposits Davis and Hall (1959); Allwardt, 1981) Valley Springs Formation (Reduced) paleosol J-Clay? L-? (NaHCO3-rich) Marchand and Allwardt (1981)

Eocene (Marchand Allen (1929), Gales/Piper et al., and Allwardt, 1981) Ione Formation Marine Rocks L-? Poor (1939), and Gillam (1974).

Basement Complex Varies Basement Complex (Igneous-Meta rocks) Joints and faults Poor 50

51 deposits below and between the lacustrine clays within the Tulare Lake basin region (Croft and Gordon, 1968). Faunt et al. (2009; 2010) used textural descriptions from driller’s to model the aquifers of the entire Central Valley of California which resulted in the generalized hydrogeologic cross section seen in Figure 11-B which is an east–west line across the historical Tulare Lake bed of the Tulare Basin. The aquifers are designated as layers one through ten with Corcoran Clay aquitard designated as both layers 4 and 5. Typically, the prominent clay layers define the boundaries between the aquifers and can be correlated through until the clays pinch out in the alluvium. The A clay is the boundary between layers 1 and 2.

2.8.5 Groundwater Flow Directions The direction of groundwater flow under natural pre-development conditions moved roughly southwestward from the mountains toward the valley trough. However, pumping depressions has caused local changes in directions of groundwater movement. Conversely, groundwater mounds are suggestive of high infiltration. This is evident on groundwater contour maps of the San Joaquin Valley produced by DWR. Under the influence of current water demands, the shallow water body flows northwest into the unconfined water body and then moves toward the center of a depression in the northwest with some motion downward through the A, C, and E clays where pressure head is lowered by pumping (Page and LeBlanc, 1969). Generally, in the Tulare Basin, groundwater migrates in a direction towards the Tulare Lake bed from the unconfined and semiconfined aquifers into the confined aquifers that exist between prominent lacustrine clays layers (Croft and Gordon, 1968). 52 2.8.6 Water Supply Tulare County, and therefore, the Kaweah Subbasin receive water supplies from precipitation (snow-melt from the Sierras and rain in the valley), streamflow, imports (controlled by irrigation districts), and groundwater (Croft and Gordon 1968). The principal source of recharge to the unconfined and semiconfined aquifers is seepage from rivers, streams, irrigation canals and ditches (Croft and Gordon, 1968). Other sources of include artificial recharge ponds and direct precipitation which contribute minor quantities of water to the aquifers (Croft and Gordon, 1968). Kaweah River is the main source of water for the Kaweah Subbasin yet additional water resources can be attained form the Friant-Kern Canal (Croft and Gordon 1968). The unconfined aquifer above the A clay is recharged from local precipitation and seepage from surface water channels, whereas the sub-surface region between the A and C clays, which become confined as distance to Tulare Lake bed decreases, is recharged from unconfined aquifer at the northeast and east portions of the Tulare Basin. Furthermore, the confined aquifer beneath the E clay receives most of its water from aquifers above, and from again, the northeast and east portions of the Tulare Basin. Since many wells are perforated between the confined aquifers this allows for artificial recharge from the units above (Croft and Gordon, 1968) and, therefore, can carry pollution such as nitrates to deeper aquifers. Page and LeBlanc (1969) suggest the best sites for recharge are in areas beneath river channels due to the high permeability of sedimentary deposits and the ease of distributing large quantities of water through the channels. Typically, the river channels become losing streams downstream until they reach the clay deposits of the terminal basin. Incidentally, Page and LeBlanc (1969) also suggest 53 the lower reaches the river channels probably overlie the A clay which will restrict downward movement of recharge.

2.8.7 Water Quality Hanford and Visalia groundwater quality appears to follow the water quality of the surface water (Croft and Gordon, 1968). Generally, the dissolved solids content of the groundwater in the Fresno area (San Joaquin Basin) rarely exceeds 600 mg/l (milligrams per liter) although at depths ranging from about 700 to 3,000 feet, it can exceed 2,000 mg/l (Page and LeBlanc, 1969) later updated to depths ranging from about 600 to 3,000 feet (Page, 1973; Mitten, 1984). The groundwater dissolved solids begin to exceed 2,000 mg/L at 3000 feet for the Central Valley model (Faunt et al. 2009a) which agrees with Croft and Gordon’s (1968) water quality for the Tulare Basin. The salinity below the E clay is unsuitable for most crops except very highly salt tolerant crops on a gypsiferous soil in the San Joaquin Basin (Page and LeBlanc, 1969).

2.8.8 Kaweah Subbasin Water Table Seasonal fluctuation of the water table occurs in the unconfined aquifer which range in depths of about 30-115 feet below the surface. Hydrographs from well water heights suggest a general trend of high-water level in the winter and early spring and low-water level in the late summer and early autumn (Page and LeBlanc, 1969). Historical records show a trend of depletion of the water reserves in the upper most aquifer. During the early 1900’s the water tables of the alluvial fans of the Kaweah, Tule, and Kings Rivers were typically 20 feet below the surface (Croft and Gordon, 1968). It is evident from Figure 13 that the uppermost aquifer is being depleted of its water resources over the last 100 years. Figure 13 combines precipitation (inches) from NOAA, depth to water table (feet below 54 surface) from groundwater equal elevation levels maps produced by the California Department of Water Resources (DWR), Kaweah River discharge data in acre-feet (DWR), California historical hydrological events from Paulson et al. (1991 ) and are plotted to their dates. The results show water table depth fluctuations related to high volumes of Kaweah River Discharge which suggest the Kaweah River is the main source of recharge for its fan. Precipitation is not effective therefore never reaches the aquifer. Droughts lower the water table and last for years whereas flood events happen in shorter yearly intervals yet they are somewhat effective on raising the water table. Flooding events are much less of a problem within the San Joaquin Valley since the construction of dams along all major rivers. was completed in 1965; however, massive precipitation above 950 feet elevation and within 24 hours produced more water than the dam could hold causing the last flood event of 1966 within the Kaweah River region (Dean, 1971). If the Terminus Dam had not have been built just the previous year, the flooding would have been catastrophic.

2.8.9 Transmissivity The historic data of the aquifers in the SJ Valley are very valuable, but more descriptive than quantitative so far. In order to help determine how much water is in the layers of aquifers and what kind of pumping discharges are available in a particular area, transmissivity (T) of every layer must be determined. Most estimates for hydraulic conductivity use pumping tests that result in a horizontal hydraulic conductivity estimate. Transmissivity, however, requires lab tests and requires a measure of vertical hydraulic conductivity. The Central Valley hydraulic properties govern the transmission and storage of groundwater and are represented by hydraulic conductivity and thickness (Faunt et al., 2009b).

Figure 13. Local surface water effects on groundwater. Note: It is apparent that ground water recharge lags in time behind both precipitation and Kaweah River Discharge. Furthermore, this shows depletion of well-water height for the City of Visalia for the last 100 years conventionally measured during the spring.

55

56

For the purpose of the model, it uses hydraulic conductivity values based on texture for the model (Faunt et al., 2009b). According to Faunt et al. (2009a), the model for the Central Valley aquifer hydraulic conductivities and transmissivities were estimated on a basis of distribution of drillers and geophysical logs and sedimentary texture on account of limited stratigraphic control and measurements of hydraulic properties of the aquifer systems. The hydraulic conductivity for the San Joaquin Valley model are based on laboratory values from Bertoldi et al. (1991) seen in Table 2. Select values that will be compared to the results from the permeameters in this thesis.

Table 2. Hydraulic Properties of Central Valley Soils (from Bertoldi et al., 1991). Note: Laboratory results for the Central Valley, CA from Bertoldi et al. (1991). Vertical hydraulic conductivities of Sand (3.5 m/d), Sand-silt-clay (0.0061 m/d), and Silty clay (0.00003 m/d) are compared with results from the permeameters in Figure 26.

3. METHODOLOGY

3.1 General Description of Work Laboratory measurements from core cuttings are a deterministic method for measuring hydrologic properties in addition to being the best way to identify stratigraphy when outcrops are not available. However, the costs for drilling are expensive and require a large budget. Pumping tests can be used to measure horizontal hydraulic conductivity, yet in order to measure vertical hydraulic conductivity, which is a variable for transmissivity estimates, tests on actual material is needed. Core cuttings were a requirement to make this thesis possible. After the core was washed and documented, the cuttings were reduced in size and slipped into clear plastic liners so it could be tested using the permeameters. This required the creation of a sampling frame capable of reducing core to tightly fit into the liners at either 0.5 or 1.0 foot intervals. Sampling was divided based changes in lithology such as soil texture, cementing, evidence of pedogenesis, and consistency. Once the material was inserted into liners, their hydraulic conductivity was measured using both the falling head and constant head permeameters. Measurements for permeameter samples ranged from several hours (sandy material) to over a week (fined grained samples) per sample depending on texture. Identifying the stratigraphic formations in core was done by associating the descriptions from the type sections. The two main characteristics used to identify the formations were soil differentiation and volcanic ash. Soil differentiation could be identified by the accumulations of carbon, clays, silica and/or root traces. Volcanic ash was identified and compared with samples from known volcaniclastic deposits under microscope and photographed for proof. Identifying 58 the stratigraphic formations was used to relate the proper hydrogeological aquifer for the Central Valley model summarized in the interpretive correlations from Table 1. This gave the thickness variable needed for transmissivity estimates.

3.2 Core Drilling Technicon Engineering Services Inc. was contracted to drill one vertical hole to a depth of 225 feet (68.5 m) using a Central Mining Equipment truck mounted auger drill (CME-75) which recovered 2.25 in. diameter core (Figure 14). The target depth was a large compact clay body reported by Thompson (1892, p14) at 212 feet (64.6 m) and was based on an 8-inch bore hole for city water that was drilled at the intersection of Court and Main Streets, Visalia which is just under a mile away and same elevation as this drill site. Consequently, due to difficult drill conditions at depths below 75 feet, the total depth of the hole was 132 feet (40.2 m) with an overall recovery measured at 67 percent. Subsequently, most losses in recovery occurred in the unconsolidated sands.

Figure 14. Core drilling operation. Note: Mining Equipment truck mounted auger drill (CME-75) is operated by Technicon Engineering Services Inc. Core was cut to 2.25 inches in diameter, drilled to 132 feet, and preserved in core-boxes. 59

The borehole was drilled in the city of Visalia, California, which is situated on the Kaweah River fan. The drill site is located at WGS 84 decimal degree coordinates N 36.33967 (36° 20’ 22.81”) and W 119.27986 (119° 16’ 47.49”) with an elevation of 330 feet (Figure 1). Once core was recovered, it was washed, core- logged, and photographed (Figure 15). The core then was transported in core boxes to CSU, Fresno Hydrology Lab for sample preparation for permeameters. The drill site is approximately 275 feet west of N. Ben Maddox Way in an abandoned lot. The property is accessible directly from N. Ben Maddox Way and is bound by the cross-streets E. Houston Ave. to the north and E. Douglas Ave. to the south. This work took place on undeveloped land that historically was used for cultivation of walnuts.

Figure 15 Core recovered from drilling. Note: This is a photo-log of the washed core before it was altered by sampling. The wood blocks show the depth in foot below the land surface. Geotechnical measurements and core description were documented before material was sampled for permeameter tests. 60

Figure 15. Photo log (cont.) 61

Figure 15. Photo log (cont.) 62

Figure 15. Photo log (cont.) 63

Figure 15. Photo log (cont.) 64

Figure 15. Photo log (cont.)

3.3 Technique for Sampling Core A stainless steel 2 inch core sampler (manufactured from AMS) was used to reduce the core diameter and insert the samples into plastic liners later to be measured for saturated hydraulic conductivity. This was done by constructing a PVC-pipe frame that stabilized the AMS sampler so core can be inserted intact (Figure 16). Core was held stable by a 3 inch PVC pipe that reduced core diameter from 2.25 inches to the inside diameter of two-inch PVC pipe that was again reduced and slipped into the AMS sampler at the base of the frame. This technique was used on fine grained material due to the strength of the clay-pan and lack of cementation. Hydration of the fine grained core was necessary to render the material to a state that samples could be easily inserted into the liners. This was done in 3 inch PVC reservoir (Figure 17). To keep core intact during hydration, sectioned pipe (bottom left image of figure 17) was made porous and placed under the core in the reservoir. This porous sectioned pipe also served to 65 carry and insert the hydrated core into the sampling frame. The porous sectioned pipe held the samples of core in the reservoir filled with water as a sprayer was used to hydrate the top of the sample. When the sample was sufficiently hydrated, it was pushed into the frame with a 2 inch diameter wood dowel which can be hit with a hammer once material is in the 2 inch PVC neck. This is done until all material has filled the AMS sampler and liners at the base. Plastic liners now filled with material were recovered and capped for later analysis. With greater depth, samples became more difficult to both hydrate and insert into liners. Unconsolidated sand could be packed directly into plastic liners without using the sampling frame. Consequently, the duripan was too hard to sample, however, based on Soil Survey Staff (2010) soils characteristics, it is impermeable.

3.4 Constant Head Permeameter Saturated hydraulic conductivity (Ks) quantifies the ease with which water can move through pore spaces or fractures. It is dependent on porosity (pore space matrix) of a soil medium yet can be measured independently. Note that saturated hydrologic conductivity and permeability are not the same yet are directly proportional since Ks=k*pg/u, where k is intrinsic permeability of the material, p and u are respectively fluid density and viscosity. Ks was measured using two different devices depending on the texture of samples. Course grained (sandy) material was measured using a constant-head permeameter (Figure 18) apparatus designed and constructed by Dr. Zhi Wang, Department of Earth and Environmental Sciences, CSU, Fresno Hydrology Lab (Figure 19) based on Klute and Dirksen’s (1986) design and Equation 1 seen below:

VL Ks  Ath (1) 66

Figure 16. PVC Frame for AMS stainless steel sampler. Note: AMS sampler frame was built to reduce core diameter and insert sample into plastic liners for permeameter tests. It can be used for sediment or soil. 67

Figure 17. PVC Reservoir. Saturation process of core. Note: Core was moistened until soft enough to be inserted into AMS sampler frame in Figure 15.

68

Figure 18. Constant head permeameter. Note: Transparent constant head permeameter (The Tank) designed by Dr. Zhi Wang, Department of Earth and Environmental Sciences, CSU, Fresno.

Figure 19. The tank type constant-head permeameter. Note: Constant head permeameter used to measure several samples at once yet is limited to coarse grained material.

69

Equation 1 is a variation of Darcy’s equation that allows laboratory measurement from coarse grained soil-core samples. The length (L) of the sample in the liner and cross-sectional area (A = πr2 = πd2/4) can be measured before sample is saturated. The length of sample in the liners varies whereas the cross- sectional area is constant. Samples are saturated when the hydraulic head reaches a maximum height and stops rising. The difference in head (Δh) from the water level in the tank and water column inside the transparent hydraulic head extender is measured when stabilized. Coarse grained material will stabilize (become saturated) fast due to high permeability; conversely fine grained materials take longer periods of time to stabilize. Once the sample is saturated, a small siphon was inserted and stabilized to a constant flow. The stabilized siphon outlet volume

(V) was collected, timed (t), and measured and later used to estimate Ks for all samples abundant in coarse grained material. This design allows for multiple samples to be tested at once and is used to saturate the soil before being tested. The units of measure were milliliters or cubic centimeters for volume, centimeters for lengths, and minutes for time resulting in Ks of centimeters per minute. Good contact between the sample material and liner is necessary otherwise measured Ks would produce invalid results that are greater than true values. This can happen when pressurized water creates a sinuous flow-path along the side of the liner which is called channeling. Channeling can be observed in colorless plastic liners. For this reason, any samples displaying channeling were removed from the data-set. Channeling was more common in the falling head permeameter apparatus due to greater hydraulic pressure and will be discussed in the next section. 70 3.5 Falling Head Permeameter The schematic for the falling head apparatus (Figure 20) was provided by

Dr. Zhi Wang. Saturated hydraulic conductivity (Ks) of the fine grained material was measured using a falling head permeameter (Figure 21B) constructed by Dr. Wang at CSU, Fresno. However, due to slow conductivities of samples, a second apparatus was constructed by the author to increase measuring time (Fig. 21A photo of your device). This allowed for as many as six samples to be tested at once. These permeameters are modified versions of Johnson et al. (2005) and were designed by Dr. Wang, Department of Earth and Environmental Sciences, CSU, Fresno. Falling head permeameters are properly suited to measure fine grained mediums such as silt- or clay-rich materials and uses a derivation of Darcy’s equation seen in equation 2 from Johnson et al. (2005):

2 d t L ho Ks  2 ln( ) (2) d c t h

Figure 20. Falling head permeameter design. Note: Diagram of falling head permeameter. Image provided by Dr. Zhi Wang, Department of Earth and Environmental Sciences, CSU, Fresno. 71

Figure 21. Falling head permeameters. Note: Falling head permeameters labeled A and B are designed by Dr Zhi Wang, Department of Earth and Environmental Sciences, CSU, Fresno and were used to measure hydraulic conductivity of fine grained material.

72

Both the inner diameters of the sample-liners (dc) and supply tubes (dt) are constant (see Fig. 19), however, the lengths (L) of core samples vary and must be measured individually after saturation. Samples were saturated from the base to the top to insure complete saturation of material being tested. The permeameter base is filled with water before the material-filled liner is inserted to insure water to material contact without air. This is important since the transmission of water is being measured as opposed to air. In addition, this helps to lubricate the rubber holder and allows the liner to slide into the base with ease. Once sample is tightened in place, a highly permeable filter is fitted into the top of the liner and filled with water so that the head height within the liner is constant and at maximum. Cap and upper filter is then secured to hold material in place so that it does not get pushed out or cause channeling once hydraulic pressure is increased by filling the supply tube. Once cap is in place, the supply tube can be filled until the meniscus is at a maximum desired graduated measure (h0), which is when measuring time is initiated. The supply tube is then loosely capped to limit evaporation yet not tight enough to cause a vacuum as water level drops. Samples can take up to several hours to a few weeks to measure depending on the amount of silt and clay. Typically, the finer grained the material, the longer the waiting period. All water should migrate through the material and out the drain tube at the top of the cap. This insures that all water being measured has traveled through the medium. When the water level in the supply tube has dropped sufficiently, the total time (t) is measured and ending water level (h) in the supply tube is recorded.

The units of measure were minutes and centimeters resulting in a Ks of centimeters per minute. 73 3.6 Transmissivity Transmissivity (T) describes the ease water can move through a unit (Schwartz and Zhang, 2003). The quantitative determination of transmissivity (T) is hydraulic conductivity (K) multiplied by an aquifer thickness (H). In other words, when using Darcy's law, the available discharge Q = KHW* dh/dl = TW*dh/dl (where K is conductivity, H is aquifer thickness, W is aquifer width or arc length, and dh/dl is the groundwater pressure gradient could be the ratio of drawdown and well influence radius), thus the available discharge Q = T at W=1 (unit width) and dh/dl = 1 (unit gradient). The hydraulic conductivities were measured using the permeameters described above. Geologic formations thicknesses are based on the core log descriptions and are then correlated to the aquifer layers in Table 1. Once hydraulic conductivity and aquifer thicknesses were measured, the transmissivity was calculated by multiplying unit thickness by its average saturated hydraulic conductivity using Equation 3 taken from Schwartz and Zhang (2003):

T  KH (3)

3.7 Microscopy Friant Pumice Member of the upper Turlock Lake Formation was identified using Microscopy techniques. Sample preparations included removing moisture from the sample using a Fisher Scientific Isotemp (Model 725-F) oven, loosening grain to grain contacts by crushing with a roller, and finally sieving material to very fine sand and silt sized using a USA standard Test Sieve, ASTM E-11 in a Valoras shaker. Typically, volcanic glass shards range in sizes from 0.002-2.0 mm (Soil Survey Staff, 2010). Once grain-sizes of material were separated a microscope was used to identify glass shards which are characteristic of volcanic 74 ashes. This was done with three different samples collected by the author: (a) Lava Creek ash from Tecopa Canyon Bad Lands near China Ranch, Mojave Desert, California; (b) Friant Pumice Member from the type section collected from Outback Materials Friant quarry, CA; and (c) Friant Pumice Member of the Kaweah River fan drill core from this thesis. Since Kaweah River ash has not been found until this thesis, samples from well-known volcanic ash beds were collected to ‘train the eye’ and show evidence that Kaweah sediments do indeed have volcanic ash.

3.8 Mastersizer 3000 Particle Analyzer A Mastersizer 3000 was used to measure the particle size distribution of sediments. This yielded 52 samples that were sieved to include all material except gravels (2 mm). The material was mixed with deionized water and shaken for an hour on a shaker box and then measured using the large volume wet sample dispersion unit or Hydro LV. This unit of the Mastersizer 3000 has a detection range of 10 nanometers to as large as 2.5 millimeters. Ultrasound was used during testing to lower aggregation of fine particles. Data was saved and used to construct histograms for the particle size distributions of samples.

3.9 GIS Digital Data Geographic information system (GIS) software was used to create maps for figures 1 and 2 in addition to the geologic map in the results section. The geologic map was created from shapefiles of soils maps downloaded from NRCS website (http://www.nrcs.usda.gov/wps/portal/ nrcs/site/national/home/). Surface stratigraphy was associated with soils map and manuscripts (Stephens, 1982; Arroues and Anderson, 1986; and Wasner and Arroues, 2003). The geologic map was compared and calibrated to the color 75 codes of Weissman et al. (2005) San Joaquin Valley map which differs only slightly for formations. In contrast, most of the lithologic units were identified for this thesis using soil survey descriptions and taxonomy, and locations and distributions that were related to type section descriptions . The Riverbank was identified by soil descriptions with duripan and then was subdivided based on whether the duripan was strongly indurated (middle and lower Riverbank) or not (upper Riverbank). The indurations level is described by Marchand and Allwardt (1981). Modesto Formation was identified using soil survey maps as old fan deposits that did not have duripan since duripan is not found in its type section. The remaining soil series were then designated as Holocene. Upland soils were not included since they are not important to this study. Soil series correlations by Marchand and Allwardt (1981), soil taxonomy (especially Orders and Subgroups), 5 foot interval contour map made from DEM maps, river locations, and Weissmann et al., 2005 map were used to help construct an accurate map. The watersheds and rivers are modified shapefiles from California Department of Water Resources website (http://www.water.ca.gov/). Lake, reservoir, swamp, marsh, and California coastline shapefiles and all Digital Elevation Files (DEM) are modified files from the USGS website (http://viewer.nationalmap.gov/viewer/). City boundaries and California state highway 99 are modified shapefiles from files from the Caltrans website (http://www.dot.ca.gov/hq/tsip/gis/datalibrary/index.php). Latitude-Longitude, cities, and interstate shapefiles are modified from Chang (2012).

4. RESULTS

4.1 Geologic Formations of the Kaweah River Fan The geologic map (Figure 22) of the Kaweah River fan shows the extent of stratigraphic formations. These are in order of oldest to youngest: Middle and upper Riverbank Fm., lower and upper Modesto Fm. and Holocene deposits. Late Pleistocene Vertisols ( such as Porterville or Seville soil series) are also mapped since they have been associated in age to the middle Riverbank by Marchand and Allwardt (1981) who also noted that they are derived from andesitic and metavolcanic parent sources. For this reason they are mapped separately. Soils characterized by duripan are mapped as Riverbank. Distinction between Upper and middle Riverbank was based on Marchand and Allwardt (1981) descriptions that suggest the middle Riverbank is more acidic and indurated than the upper Riverbanks. Similarly, the lower and upper Modesto Formations were identified on a basis of previous work along with soil information from the soils survey manuscripts and topography. Holocene deposits were identified using soil survey descriptions and topography. The results of soil series names correlated to stratigraphic nomenclature is seen in Table 3. These data were used to produce the geologic map using digital files from NRCS website. Holocene deposits are the most recent and blanket over the Modesto Formation. Holocene deposits show no evidence of glacial cycles in the sedimentary record yet both the upper (Hilgard or Tioga) and lower (Tenaya or Tahoe) Modesto units are related to the glacial cycles. The extent the upper fan is seen deposed below the Holocene deposits. Riverbank Formation (Tahoe, Mono

Basin, or Casa Diablo), again related to glacial cycles is exposed near the edges of

77

Figure 22. Geologic map of the Kaweah River fan region. Note: Geologic map of Late Pleistocene formations was created from soil survey digital data and reports. Orange-brown is Riverbank which is the oldest unit. Modesto is represented as green and is younger than the Riverbank. Holocene or the youngest units are yellow. Soil data was obtained from Department of Agriculture, Natural Resources Conservation Service and the manuscripts of Arroues and Anderson (1986), Arroues (2006), Stephens (1982), and Wasner and Arroues (2003).

Table 3. Correlations Between Soil Series and Geologic Formations. Note: Many soil series correlations were already done by Marchand and Allwardt (1981). When soil series correlations did not exist from previous literature, soil orders and subgroups, topography, soil survey descriptions were used to help identify geologic formations. These were used to generate the geologic map from digital data from NRCS website.

78

Table 3 (cont.)

79

Table 3 (cont.)

80

Table 3 (cont.)

81

82 the Sierra Nevada foothills and continues below the Modesto Formation. Only the upper and middle units are exposed along the Kaweah River fan. The Turlock Lake Formation does not crop out anywhere on the map area.

4.2 Stratigraphy of the Kaweah River Fan Figure 23 is a stratigraphic column of the bore hole sediments recovered from the Kaweah River fan. Recovery of core was 67 percent of the 132 feet (40.2 m) drilled. Recovery represented the percentage of core brought to the surface versus actually drilled. The missing portions of the core, not recovered, are indicated with an X in the stratigraphic column. Most losses in recovery were seen in sandy sediments since the loose sediments fail to latch to the recovery- tube. Conversely, fine grained material consisted of higher recoveries. The soil series have been included along with formation names and are interpreted based on location and age. These soils were also compared to survey descriptions and/or compared with its subgroup such as duripan. Figure13 shows the local surface water effects on groundwater in addition to illustrating that Layer 3 aquifer is where most of the water in wells are found locally. Munsell soil color was done on both dry (d) and wet (w) material in direct sunlight and after core was washed to insure the color of the material was measured as opposed to drill mud color. Generally, soil colors ranged in hues of 10Yr with moderate chroma and values that change with respect to moisture content. Typically, wet samples are brown to dark brown whereas dry material is very pale brown. The iron oxides (FeOx) and organic carbon (Organics) contents heavily influence the colors of the soils. Iron oxides such as hematite, jarosite, and goethite, typically change Munsell values to darker varieties of brown or varieties of reddish brown (5YR). Organic carbon is black and will darken or 83

Figure 23. Stratigraphic Column of the Kaweah River, CA. Note: The prominent aquifer at this locality is Layer-3 which includes the upper Turlock Lake and Riverbank Formations. 84 lower the Munsell values of any soil color. Carbon-rich soils are the darkest and are localized on the topsoil.

The abundance of calcite (CaCO3) was identified using 10% hydrochloric acid on both wet and dry core. Both Nord and Tagus soils are described as having disseminated calcite (Wasner and Arroues, 2003). Calcite abundances were high near the topsoil, and subsequently, no reactions with the acid were observed below a depth of 22 feet (6.7m), just above the first sandy layer. The duripan had sparse calcite stringers a few millimeters in width that appeared to infill cracks yet no reaction from HCl were seen within its matrix. According to Wasner and Arroues, (2003) soils survey map, drilling began in Nord Soil Series which is a Cumulic Haploxerolls (Mollosols) which classifies it as Holocene based on organic carbon age. This soil series also is described as having an Ab layer horizon or buried soil horizon believed to be Tagus soils which is upper Modesto. The descriptions of the B soil horizons of Tagus soils show similar Munsell colors to the Nord soils. Both are characterized by disseminated calcite, and Tagus soil is exposed at the surface nearby. For this reason, Tagus or upper Modesto is at the Ab horizon at 4 feet (1.2 m) of depth. Consequently, soil surveys typically characterize the soils to depths of 72 inches or less and, therefore, are limited to the topsoil. Conventionally, the contacts of the stratigraphic formations have been separated on a basis of disconformities typically seen as soil differentiation. At a depth of 9 feet (2.7 m) another buried soil (Abk) is interpreted to be the contact between the upper and lower Modesto Formation. Duripan was identified by its abnormal hardness both dry and wet and assigned to Flamen soils series of the middle Riverbank Formation which crops out just north of the drill-site. Duripan is typically at 25 feet (7.6 m) depth in the Visalia area according to Croft and 85

Gordon, (1968). These buried duripans are typically preserved B soil horizons of the Riverbank Formation (Marchand and Allwardt, 1981). The middle Riverbank on the geologic map appears to be much more extensive than the other two sections and in lower elevations. According to Marchand and Allwardt (1981) the lower Riverbank is the least extensive of the three. For these reasons the duripan was interpreted as Flamen soils of the middle Riverbank whereas the upper and lower sections appear to be missing. The Friant Pumice Member of the upper Turlock Lake Formation is found at a depth of 79 feet (24.0 m) below Visalia. This was verified through microscope (Figure 24A, B, and C) by identifying glass shards characteristic of acidic volcanoclastic material. When the pumice was recovered at the surface, it appeared as a light gray (w-10YR/7/1) fine-sandy, fat-clay that was both elastic and sticky. Two known volcanic ash samples were used for a microscopic comparison to the Kaweah River ash so there was no doubt that volcanic glass shards existed at this locality. These were Friant Pumice Member samples (from the laminated beds below the coarse pumice-pebbles) from the type section described by Janda (1966), and Lava Creek ash collected near China Ranch, Mojavi Desert, CA. The Friant pumice Member (Figure 24B) is the same ash found in the core sample (Figure 24C) and is related to a Long Valley Caldera near , CA. Lava Creek ash (Figure 24A) is related to Yellowstone Caldera and has not been reported in the San Joaquin Valley sediments. Both calderas are composed of acid volcaniclastic material and therefore can be used to compare for glass shards. The Kaweah River ash is believed to be Friant Pumice Member based on the work of several authors (Davis et al., 1977; Janda 1965;1966; Marchand and Allwardt, 1981; Izett, 1988; and Sarna-Wojcicki et al., 2000) who reported Bishop ash in the San Joaquin Valley. 86

A. Lava Creek volcanic ash related to Yellowstone caldera deposits in CA.

B. Friant Pumice Member type section related to Long Valley caldera, CA.

C. Friant Pumice Member from this study related to Long Valley caldera, CA.

Figure 24. Microscopy of volcanic ash. Note: A. Lava Creek ash sampled from the Badlands of near China Ranch, Mojavi Desert, CA. Grains are sieved to very fine sand sized and magnified. B. Type section-Friant Pumice Member of the upper Turlock Formation sampled with permission from Outback Materials Friant quarry, near the city of Friant, California. C. Friant Pumice Member (very fine volcanic sandstone) collected from the subsurface of the Kaweah River fan, Visalia, CA.

87

Paleosols were used by Weissmann and Fogg, (1999) to differentiate formation contacts within the Kings River fan that connects the Kaweah River fan to the north. These paleosols are relatively continuous within the fan stata (Weissmann and Fogg, 1999). The Kaweah River fan paleosols are characterized by root traces (Figures 25 A and B). Root traces are used to identify contacts between stratigraphic formations and are either in the form of root tubes (RT) or root castings (RC). Therefore, paloesols with root traces divide both units of Turlock Lake Formation in addition to the Laguna Formation. Layer 3 aquifer was assigned to the formations based on literature review which is summed up in Table 1. Both Layers 1 and 2 are too thin and high in fine grained material to constitute as an aquifer. More than likely, this is due to being high on the fan. Apparently, these upper layers vary in thickness. However, as Marchand and Allwardt, (1981) pointed out, the units get thicker as distance is decreased from the major depressions of the valley. Recall that Layers 4 and 5 are the aquitard known as the Corcoran Clay Member which was not found in the bore hole. However, the Friant Pumice Member forms a conformable contact with the Corcoran Clay and marks where the aquitard would be if the lens extended this high into the fan. The extent of this massive clay lens is seen in Figures 11B and 12. Furthermore, Figure 12 shows how Layer 3aquifer bounds Layers 4 and 5 aquitard in the terminal basin and bounds the lower Turlock Lake paleosol (top of Layer 6 aquifer) in the fluvial fans which it typically identified using magnetostratigraphy. Magnetostratigraphy is the best way to find the bottom of Layer 3 when the Corcoran Clay is not present or the top of Layer 6 aquifer. Due to time constrains, magnetostratigraphy was not measured therefore, the well- developed paleosol was used as a marker for the lower contact instead. Deeper aquifer intercepts are undetermined at this time. 88

A. Root tubes (RT)

B. Root casts (RC)

Figure 25. Root traces (paleosols) seen in Kaweah River fan. Note: A) Root tubes with iron oxide stained rinds. B) Root casts found at greater depths. 89 4.3 Hydraulic Conductivities of the Kaweah River Fan Once the stratigraphy was identified the hydraulic conductivities (Ks) and related transmissivities could be compared and estimated. The results of the saturated hydraulic conductivities (raw data are in Appendix) are plotted against the stratigraphic formations as seen in Figure 26. Lab values from Bertoldi et al. (1991) in Table 2 and were plotted as constants. These were the K-values for sand, sand-silt-clay (or sandy silt), and silty clay (or clayey silt). Permeameter measurements of the Kaweah River fan sediments are plotted with respect to depth with the same intervals as the simplified stratigraphic log. Note that the constant head permeameter was used for course grained material which yielded values much greater than Bertoldi et al. (1991) sand-sized averages. This suggest that the constant head permeameter measures rates greater than actual amounts and/or the small amount of fine gravels and coarse sands increased K substantially. Regardless, coarse grained material has a high permeability and porosity and therefore can hold high volume of water. Fine grained material was measured using the falling head permeameters which fell below silty-clay average K from Bertoldi et al. (1991). The results from Figure 26 clearly show that the paleosols do indeed act as aquitards or as Weissmann and Fogg (1999) state, constitute important confining units due to their lateral continuity and high clay content. Furthermore, the Friant Pumice Member acts as a confining bed. These data show that the uppermost aquifer for the Kaweah River fan is hosted in the Riverbank and upper Turlock Lake Formation sediments which together comprise Layer 3 aquifer seen in the Central Valley hydrogeological model. The depths for this unconfined aquifer rang from roughly 14 to 90 feet (4.3-27.4). Average hydraulic conductivities of the stratigraphic formation are used to estimate transmissivities. Also note that the duripan at 14-15.5 feet (4.3-4.7m) was too hard to sample for K-measurements. 90

Figure 26. Hydraulic conductivity of formations and Layer 3 aquifer (blue). Note: The K-values for sand (green), sand-silt-clay (yellow), and silty clay (red) are from Table 2. All values between the sand-silt-clay and silty-clay were measured using the constant head permeameter to add quantitative data to these confining beds. Duripan was too indurated to test for K-values (4.4-4.7 m), however, K-values were low in material below the duripan due to increased clay. 91 4.4 Malvern Mastersizer 3000 Particle Size Analyses of Impermeable Sections Texture describes the amounts of sand, silt, and clay fractions of the soil or sediment. Wentworth size class was used to sort material sizes and classes. The Mastersizer 3000 was used to measure these size fractions of the confining units of the Kaweah River fan. These values are best viewed as particle distributions which can be seen in the 52 histograms in Figure 27. The sample code is designed to have the first number representative of sample count (1 to 52 samples) with the lowest number, “1” representing the shallowest depth. The next value followed by a hyphen represents actual depth in feet. For example, sample “14-34” represents the 14th sample from the surface whereas the 34 represent the depth in feet. Constant values of each size class were plotted against the particle distributions to help clarify. Clay averaged 10.7 percent for all samples yet was as high as 24 percent in sample 14-34 which also had the highest amount of silt at 75 percent. Generally, for all samples, the most common size class was silt with exception to the Friant Pumice which is abundant in fine sand. The reduced hydraulic conductivity is probably from the hydrophobic properties common in volcanic ashes. The silt averaged 64 percent for all samples. Only 3 samples had moderate amounts of fine sand which are samples 11-16, 12-16, and 18-79 (F. pumice). Again the Friant Pumice had the greatest abundance of medium and coarse sand. Otherwise, medium and coarse sands are in very low quantities overall.

92

Figure 27. Histograms from particle size analysis using Mastersizer 3000. Note: Clay and silt boundary is blue, the silt and fine sand boundary is green, the fine sand and medium sand boundary is yellow, the medium sand and course sand boundary is orange, and the course sand and gravel line is red. These 52 samples were selected based on hydraulic conductivities less than or equal to Bertoldi et al. (1991) sand-silt-clay values. 93

Figure 27 (cont.)

94

Figure 27 (cont.)

95

Figure 27 (cont.)

96

Figure 27 (cont.)

97

Figure 27 (cont.)

98

Figure 27 (cont.)

4.5 Transmissivity of Kaweah River Fan Stratigraphic Units Results for stratigraphic unit transmissivities for the Kaweah River fan are shown in Table 4. This was estimated by multiplying unit (formation or aquifer layer) thickness by the average vertical hydraulic conductivities measured with the permeameters. The results show that the upper Turlock Lake Formation has the biggest transmissivities measured at 2719 square feet per day (253m2/d). Based on unit correlations, both the upper Turlock Lake and Riverbank Formations encompass Layer 3 aquifer which together has a transmissivity of 2492 square feet per day (232 m2/d). This suggests that Layer 3 aquifer is the uppermost aquifer in the Kaweah River fan and Kaweah subbasin. The other units have relatively poor transmissivities and act as confining beds. 99 Table 4. Results of Hydraulic Conductivity (K) and Transmissivity (T). Formation Depth Ave-K Ave-K T T or Aquifer (ft) H (ft) H (m) (ft/d) (m/d) (ft2/d) (m2/d) Holocene 0-4 4 1.2 0.012 0.004 0.048 0.004 Upper Modesto 4-9 5 1.5 0.022 0.007 0.112 0.010 Lower Modesto 9-14 5 1.5 0.204 0.067 1.020 0.103 Middle Riverbank 14-28 14 4.3 0.190 0.058 2.657 0.247 Upper Turlock Lake 28-89 61 18.6 44.574 13.586 2719.0 252.60 Layer 3 Aquifer 14-89 75 22.9 33.231 10.129 2492.3 231.54 Lower Turlock Lake 89-114 25 7.6 1.101 0.336 27.533 2.558

5. DISCUSSION

5.1 Recharging Layer 3 Aquifer in the Kaweah River Fan Since it is already understood in the geologic background that aquifers are being depleted of water resources with increased time and that the river channels are the arteries for recharging aquifers, this suggests that the river channels should be used as the main source of recharge. Consequently, when the water reaches the Corcoran Clay boundary (Figure 12), recharge from the river channels will be reduced from the abundance of clay in the terminal basin. It should also be understood from the background section that the aquifers of the terminal basin (Tulare Lake basin), including the confined aquifer below the E-clay, are recharged from the river fans. This concludes that by recharging the upper river fan aquifers, indirectly, the Tulare Lake basin aquifers below the clay layers will be recharged. Moreover, the current irrigation-for-farming techniques do not recharge the aquifer especially where duripan exists. This is a common misconception from local farmers. Duripan should be removed from any recharge basins. Furthermore, low hydraulic conductivities at the surface further restrict vertical transmission of water. This exemplifies that recharge of the Layer 3 aquifer relies on water from the Kaweah River fan natural streams and would indirectly recharge the aquifers below the clay layers in the Tulare Lake basin. Note the Tulare Basin includes both the Tulare Lake basin and related river fans.

5.2 One Borehole Can Improve Knowledge of an Aquifer on a Sub-Regional Scale The San Joaquin Valley is reasonably well studied for both the surface and subsurface geology. Since abundant data is available, including near this study 101 area, one well described borehole can verify the stratigraphy at depth. This is a similar technique used for exploration of base and precious metals yet instead of looking for ore bodies, hydrogeologist look for aquifers. Since the borehole from the Kaweah River fan shows strong similarities to the Kings River fan previously studied by Weissmann and Fogg (1999) and Weissmann et al. (2002 and 2005) spatial data can be inferred especially when it has been determined from core cuttings. The geologic map shows the spatial distribution formation outcrops which were derived from soil survey maps. Paleosols that develop in basin soils show a stable environment or periods of depositional hiatus and are therefore assumed to be extensive through the river fans (Marchand and Allwardt, 1981; Lettis, 1988; Weissmann et al., 2005). Furthermore, these unconformities extend throughout the fans and are likely, below the lacustrine and marsh deposits which exist along the axial trough of the San Joaquin Valley’s closed basin. Since paleosols (including the duripan) are used as stratigraphic markers, are characterized by low hydraulic conductivity (confining beds), are found on a regional scale in Sierra Nevada river fans, and are used to define the confining beds within this thesis, this demonstrates that they can be modeled on a regional scale. More boreholes would help constrain the overall thickness of the aquifers in these fans and help create a three dimensional model, however drilling core are very expensive and requires a large budget. Since aquifer and stratigraphic formations are correlated, transmissivity measurements from this thesis can help with water management since it defined the confining beds of the uppermost aquifer. 102 5.3 Confining Beds of the River Fans Versus Terminal Basins The clay layer mentioned by Croft and Gordon (1968) does not exist everywhere in the river fans, indicating that confining beds in the fluvial fans above the lake and swamp deposits are instead paleosols. Since soil formation occurs during interglacial periods, and conversely, glacial periods are the wettest periods (Lettis, 1988), this suggests that lake levels rise to their maximum elevations during glacial periods which suggests the lacustrine clays defined by Croft and Gordon (1968) were probably deposited during glacial peaks. Advancement of glaciers is associated with high-water stages in Pleistocene lakes such as Lakes Bonneville and Lahontan (Blackwelder, 1931) and likely the lacustrine and marsh deposits of the San Joaquin Valley as well. Conversely, the paleosols form during interglacial periods and therefore would extend deep into the basin until they reached minimum lake levels such as the Tulare Lake bed.

5.4 Undifferentiated Confining Beds of the Terminal Basin The cross-section seen in Figure 11-B from Faunt et al. (2010) shows the aquifer layers with the confining beds bounding them. The cross-section shows four more confining beds between Layers 10, 9, 8, and 7 with Layer 6 aquifer above the F-clay. This suggest that there are as many as four more confining beds below the F clay previously reported by Croft and Gordon (1968). In this thesis, these additional confining bed will be designated the J, I, H and G clays respectively, and are hypothetical based on the textual geostatistical interpolation by Faunt et al. (2010). These clays are more than likely within the Laguna and lower Turlock Lake Formations which then relate to the McGee or Glacier Point (Lettis, 1988; Weissmann et al., 2002) Sierra Nevada glacial cycles, and therefore, correlate to the older alluvium, lacustrine, and marsh deposits described by Croft 103 and Gordon (1968). The J Clay is the deepest and oldest confining bed and rests below the deepest fresh water aquifer (Layer 10). The confining beds J, I, H, and G and their associated aquifers, Layers 10, 9, 8, 7 and 6 aquifers are currently not well understood. This relationship can be seen in Table 1.

5.5 The Importance of Transmissivity The transmissivity can be used to estimate how much drawdown have occurred in the valley due to extended, excessive pumping for agriculture and residential uses, causing many wells to go dry. In other words, T helps determine the change in hydraulic head in an aquifer or well due to pumping. However, the critical T distribution is not available in the existing maps and databases. This thesis brings new data available for the local and similar areas. This will help to manage groundwater reserves since it relates directly to the rate at which the water table in wells is affected by pumping.

SUMMARY AND CONCLUSIONS

This thesis was completed to provide a correlation between stratigraphic nomenclature and hydrogeologic units of the eastern San Joaquin Valley of California’s Central Valley. Furthermore, it completes a detailed map of the Kaweah River fan which was undifferentiated in Wiessmann et al. (2005). This is the first study to define Kaweah River fan stratigraphy. In order to measure the saturated hydraulic conductivities of fan material using permeameters, an inexpensive way to sample core cuttings was devised. This research was designed to characterize the Kaweah River fan aquifers. Consequently, due to poor drilling conditions, only information on the uppermost aquifer could be attained. Regardless however, the information was complete by identifying material characterized with high transmissivities bounded by two confining beds. This identified the uppermost aquifer for the Kaweah River fan. Literature had suggested that confining beds between aquifers were related to lacustrine and swamp deposits, but certain elements of the picture were missing. Paleosols within the fluvial fans of the Sierra Nevada define the confining beds whereas the lacustrine and swamp deposits are within the terminal basin of the Tulare Lake bed and axial trough of the San Joaquin Valley. Furthermore, these facies are formed during different paleoclimates and therefore are likely to overlap as opposed to connecting as a continuous layer. Stratigraphy of the Kaweah River fan includes all the stratigraphic formations except the upper and lower Riverbank Formations. Riverbank Formation was identified by duripan. The Upper Turlock Lake formation was identified by the Friant Pumice Member. Other contacts were identified by translocated clays and root traces. The upper contact of the Laguna Formation is 105 believed to be located at a depth of 114 feet (34.7 m); however, the lower contact was never intercepted. The Laguna Formation is very likely the host for a deeper second aquifer yet is not characterized at this time. Hydraulic conductivities of the Kaweah River Fan stratigraphic formations showed that Holocene, Modesto, and Riverbank Formations impede the flow of water into deeper substrate. Duripan can be inferred as halting or slowing the flow of water based on its characterization. In addition, fine grained material below this mineral layer reduced hydraulic conductivities substantially. The paleosols and the Friant Pumice Member all acted to slow the passage of water and are considered confining beds. Conversely, the sandy material defined the aquifer and had the highest hydraulic conductivities as expected. Thus, the confining beds of the western Sierra Nevada fluvial fans are controlled mostly by paleosols (including duripan horizons) and the Friant Pumice Member of the upper Turlock Lake Formation. The Friant Pumice Member is in conformable contact with the top of the Corcoran Clay Member (E-Clay or Layer 4-5 aquitard) throughout the San Joaquin Valley in both fluvial fans and the lacustrine and marsh deposits of the basin floor; therefore, it can be used to relate the two different confining bed systems. Transmissivity calculations of the Kaweah River fan stratigraphic units show that the upper Turlock Lake Formation yielded the best results in comparison to other formations. However, Layer 3 aquifer encompasses both the upper Turlock Lake and Riverbank Formations which define the uppermost aquifer (unconfined) in the region. Conversely, the overlying units showed low values due to low average hydraulic conductivities and relatively thin beds. The lower Turlock Lake Formation transmissivity values are not nearly as great as the upper section, yet show a moderate T-value. Magnetostratigraphy of this contact 106 would have verified the stratigraphic call and should be followed up on in the future in order to insure that this sandy unit is not part of the upper section of the Turlock Lake. Over the last one hundred years, the Kaweah Subbasin shows clear evidence of a fluctuating water table and depletion of the overall water reserves within Layer 3 aquifer. The water table appears to be affected by , droughts, and most importantly the volume of water discharged from the Kaweah River (Figure 13). The valley precipitation is a non-effective source of recharge however precipitation in the Sierra Nevada can produce high volumes of water and results in high river discharge. Recharge of the aquifer is mostly affected by river seepage. In conclusion, the core recovered from the drill site helped to identify subsurface stratigraphic geologic formations, was used to measure hydraulic conductivity and estimate their transmissivity, which resulted in identification of the uppermost aquifer of the Kaweah River fan which can be correlated to Layer 3 aquifer. The upper most aquifer is hosted mostly in the upper Turlock Lake Formation due to the abundance of sands and the overall thickness of the formation. This can be seen by the estimated transmissivities of the formations. Furthermore, the research revealed that there are some key differences in confining beds of the fluvial fans and the terminal basin of the San Joaquin Valley. Thus core samples were taken depths of 132 feet (40.2 m) for visualization and measurements but more importantly to consolidate the model about the Kaweah River fan region that was missing in the historical data.

REFERENCES

REFERENCES

Adam, D. P., 1967, Late-Pleistocene and Recent palynology in the central Sierra Nevada, California, in Gushing, E. J., and Wright, H. E., Jr., eds., Quaternary paleoecology: New Haven, Yale Univ. Press, p. 275-301.

Allen, V. T., 1929, The lone Formation of California: California Univ. Dept. Geol. Sci. Bull., v. 18, p. 347-448.

Argus, D.F., and Gordon, R.G., 1991, Current Sierra Nevada–North America motion from very long baseline interferometry: Implications for the kinematics of the western United States: Geology, v. 19, p. 1085–1088, doi:10.1130/0091-7613(1991)019<1085:CSNNAM>2.3.CO;2.

Arkley, R. J., 1962, The geology, geomorphology, and soils of the San Joaquin Valley in the vicinity of the Merced River, California in Geologic guide to the Merced Canyon and Yosemite Valley: Calif. Div. Mines and Geology Bull. 182, p. 25-31.

Arroues K.D. and Anderson C.H., 1986, Soils Survey of Kings County, United States Department of Agriculture, Natural Resources Conservation Service http://www.nrcs.usda.gov/wps/portal/nrcs/site/national/home/

Arroues K.D., 2006, Soils Survey of Fresno County, California, Western Part, United States Department of Agriculture, Natural Resources Conservation Service http://www.nrcs.usda.gov/wps/portal/nrcs/site/national/home/

Atwater, B.F., Adam, D.P., Bradbury, J.P., ,Forester, R.M., Mark, R.K., Lettis, W.R., Fisher, G.R., Gobalet, K.W. and Robinson, S.W, 1986, A fan dam for Tulare Lake, California, and implications for the Wisconsin glacial history of the Sierra Nevada Geological Society of America Bulletin 1986; 97, no. 1;97-109 Available at http://gsabulletin.gsapubs.org

Atwater, T., 1970, Implications of plate tectonics for the Cenozoic tectonic evolution of western North America: Geological Society of America Bulletin, v. 81, p. 3513–3536. doi:10.1130/0016- 7606(1970)81[3513:IOPTFT]2.0.CO;2

Atwater, T., and Stock, J., 1998, Pacific-North America plate tectonics of the Neogene southwestern United States: An update: International Geology Review, v. 40, p. 375–402. doi:10.1080/00206819809465216 109 109 Baldwin, M., Kellogg, C.E., and Thorp, J., 1938, Soil Classification: U.S. Dept. Agr. Ybk.: 979-1001. Illus

Bartow, J. A., McDougall, K.A., 1984, Tertiary stratigraphy of the southeastern San Joaquin Valley, California USGS Bulletin: 1529-J

Bartow, J.A., 1991, The Cenozoic evolution of the San Joaquin Valley, California, U.S. Geol. Surv. Prof. Paper 1501

Belitz,K., Dubrovsky, N.M., Burow, K.R., Jurgens, B., and Johnson, T., 2003, Framework for a groundwater quality monitoring and assessment program for California: U.S. Geological Survey Water-Resources Investigations Report 03-4166, 78 p.

Bertoldi, G.L., Johnston, R.H., and Evenson,K.D., 1991, Groundwater in the Central Valley, California; A summary report: U.S.G.S. Professional Paper 1401-A 44p.

Birkeland, P.W. and Janda, R.J., 1971, Clay Mineralogy of Soils Developed from Quaternary Deposits of the Eastern Sierra Nevada, California Geological Society of America Bulletin 1971;82, no. 9;2495-2514

Blackwelder, E. B., 1931, Pleistocene glaciation in the Sierra Nevada and Basin Ranges: Geol. Soc. America Bull., v. 42, p. 865-922.

Brown, R.D., 1990, The San Andreas Fault system California; Quaternary deformation; USGS Professional Papers 1515, p. 61-80

Burton, C.A., and Belitz, Kenneth, 2008, Groundwater quality data in the southeast San Joaquin Valley, 2005– 2006—Results from the California GAMA Program: U.S. Geological Survey Data Series 351, 103 p. Available at http://pubs.usgs.gov/ds/351/

Chang, K., 2012, Introduction to geographic information systems. 6th ed

Clark, D.H. and Gillespie A.R., 1997, Timing and significance of late-glaciation and Holocene cirque glaciation in the Sierra Nevada, California, Quaternary International, Vols 38/39, pp. 21-38, 1997. Copyright © 1996 INQUAJElsevier Science Ltd Printed in Great Britain. All rights reserved.

Croft, M.G., and Gordon, G.V., 1968, Geology, hydrology, and quality of water in the Hanford-Visalia area, San Joaquin Valley, California: U.S. Geol. Survey open-file rept. 110 110 Croft, M.G., 1972, Subsurface geology of the late Tertiary and Quaternary water- bearing deposits of the southern part of the San Joaquin Valley, California: U.S. Geol. Survey Water-Supply Paper 1999-H, 29 p.

Davis, S. N., and Hall, F. R., 1959, Water quality of eastern Stanislaus and northern Merced Counties, California: Stanford University Publications, Geological Sciences, v. 6, no. 1, 112 p.

Davis, P., Smith, J., Kukla, G.J., and Opdyke, N., 1977, Paleomagnetic study at a nuclear power plant site near Bakersfield, California: Quaternary Research, v7, p. 380-397

Dean, W.W., 1971, Floods of December 1966 in the Kern-Kaweah Area, Kern and Tulare Counties, California, Geological survey water-supply paper 1870-C

Division of Water Resources, California, 2004, San Joaquin Valley Groundwater Basin Kaweah Subbasin, Tulare Lake Hydrologic Region California’s Groundwater San Joaquin Valley Groundwater Basin Bulletin 118, www.water.ca.gov/

Faunt, C.C., R.T. Hanson , and K. Belitz, and, 2009a, Introduction, overview of hydrogeology, and textural model of California’s Central Valley; Chapter A, Groundwater availability of the Central Valley Aquifer, California, USGS groundwater resource program.

Faunt, C.C., R.T. Hanson, K. Belitz, W., Schmid, S.P. Predmore, D.L., Rewis, and K. McPherson, 2009b, Chapter C, Numerical model of the hydrologic landscape and groundwater flow in California’s Central Valley; Groundwater availability of the Central Valley Aquifer, California, USGS groundwater resource program.

Faunt, C.C., K. Belitz, and R.T. Hanson, 2010, Development of a three- dimensional model of sedimentary texture in valley-fill deposits of Central Valley , California , USA: p. 625–649, doi:10.1007/s10040-009-0539-7

Frink, J. H. and Kues, H. A., 1954, Corcoran clay - A Pleistocene lacustrine deposit in San Joaquin Valley, California: Am. Assoc. Petroleum Geologists Bull., v. 38, no. 11, p. 2357-2371.

Fullerton, D.S., 1986, Chronology and correlation of glacial deposits in the Sierra Nevada, California, in Sibrava, V., Bowen, D.Q., Richmond, G.M., (eds), Quaternary Glaciations in the Northern Hemisphere: Quaternary Science Reviews, v. 5m p. 161-169, Pergamon Press. 111 111 Gales, H. S., Piper, A. M., and Thomas, H. E., 1939, Geology of the Mokelumne area, California, in Piper, A. M., Gale, H. S., Thomas, H. E., and Robinson, T. W., Geology and groundwater hydrology of the Mokelumne area, California: U.S. Geol. Survey Water-Supply Paper 780, p. 14-100.

Gillam, M. L., 1974, Contact relations of the lone and Valley Springs Formations in the Buena Vista area, Amador County, California: Stanford Univ., Stanford, Calif., M. S. thesis, 180 p.

Gradstein, F., Ogg, J., Schmitz, M., and Ogg, G., 2012, eds. 2012. Amsterdam, NLD: Elsevier Science & Technology, 2012. ProQuest ebrary. Web. 5 December 2014. Copyright © 2012. Elsevier Science & Technology

Harland, W.B., Armstrong, R.L., Cox, A.V., Craig, L.E., and Smith, D.G., 1990, A Geologic Time Scale 1989: Cambridge, U.K., Cambridge University Press, 263 p.

Hays, J. D., Imbrie, J., and Shackleton, N. J., 1976, Variations in the earth's orbit: Pacemaker of the ice ages: Science, v. 194, no. 4270, p. 1121-1132.

Hinds, N.E.A., 1952, Evolution of the California landscape: California Div. Mines Bull. 158

Ives, P. C., Levin, Betsy, Oman, C. L., and Rubin, Meyer, 1967, U.S. Geological Survey radiocarbon dates IX: Radiocarbon, v. 9, p. 514.

Janda, R. J., 1965, Quaternary alluvium near Friant, California: Internat. Assoc. Quaternary Research Guidebook for Field Conference I, northern and California, p. 128-133.

-1966, Pleistocene history and hydrology of the upper San Joaquin River, California: California Univ., Berkeley, Ph.D. thesis, 425 p.

Johnson, D. O., Arriaga F. J., and Lowery B., 2005, Automation of a falling head permeameter for rapid determination of hydraulic conductivity of multiple samples, published in Soil Sci. Soc. Am. J. 69:828–833 (2005). doi:10.2136/sssaj2004.0014N

Klute, A. and C. Dirksen, 1986, Hydraulic conductivity and Diffusivity: Laboratory methods. In A. Klute (editor), Methods of Soil Analysis: Part 1. Physical and Mineralogical Methods, second edition, Madison, Wisconsin, USA, 687-732. 112 112 Lettis, W. R., 1982, Late Cenozoic stratigraphy and structure of the western margin of the central San Joaquin Valley, California: U.S. Geological Survey Open-File Report 82-526, 202 p.

Lettis, W.R., 1988, Quaternary geology of the Northern San Joaquin Valley, in Graham, S.A., ed., Studies of the Geology of the San Joaquin Basin: SEPM, Pacific Section, v. 60, p. 333–351.

Lettis, W.R. and Unruh, J.R., 1991, Quaternary geology of the Great Valley, California: Quaternary Nonglacial geology: Conterminous US: Geological Society of America, Geology of North America, v 2, p. 164-176

Lynn, W.C., Ahrens, R.J., and A.L. Smith, 2002, Soil Science Society of America, Series: 7 Soil Mineralogy with Environmental Applications. Soil minerals, their geographic distribution, and soil taxonomy, Ch. 23: p. 693.

Marchand, D. E., and Allwardt, A., 1981, Late Cenozoic stratigraphic units, northeastern San Joaquin Valley, California: U.S. Geological Survey Bulletin 1470

Means, T. H., Holmes, J. G., 1900, Soil survey around Fresno California: Division of soils, U.S. Department of Agriculture p. 333-384, http://soils.usda.gov/survey/online_surveys/california/

Neuendorf, K.K.E., Mehl, J.P., and Jackson, J.A., 2005, Glossary of geology, 5th ed., American Geological Institute

Page and LeBlanc, 1969, Geology, Hydrology, and Water Quality in the Fresno Area: U.S. Geological Survey Open-File Report

Page, R. W., 1973, Base of fresh water (approximately 3,000 micromhos) in the San Joaquin Valley, California: U.S. Geological Survey Hydrologic Investigations Atlas HA-489.

-1986, Geology of the fresh groundwater basin of the Central Valley, California, with texture maps and sections: U.S. Geological Survey Professional Paper 1401-C, 54 p.

Paulson, R.W., Chase,E. B. Roberts, R.S., and Moody, D.W., 1991, National Water Summary 1988-89 Hydrologic Events and Floods and Droughts United States Geological Survey Water-Supply Paper 2375

Preston, W.L., 1981, Vanishing Landscapes: Land and life in the Tulare Lake Basin 113 113 Raub, W., Brown, C.S., and Post, A., 2006, Inventory of Glaciers in the Sierra Nevada California: U.S. Geological Survey Open-File Report 2006-1239, 232p., online only

Sarna-Wojcicki, A.M., Pringle, M.S., Wijbrans, J., 2000, New 40Ar/39Ar age of the Bishop Tuff from multiple sites and sediment rate calibration for the Matuyama-Brunhes boundary Andrei M. Journal of Geophysical Research, v. 105, NO. B9, p. 21,431–21,443

Schaetzl, R., and Anderson, S., 2009, Soils Genesis and geomorphology: Cambridge University Press (Edu)

Shlemon, R. J., and Begg, E. L., 1972, A Holocene soil-landscape chronology, southwestern Sacramento Valley, California, in Adams, W.P., and Helleiner, F. M., eds., International geography: Internat. Geog. Cong., 22d, Montreal, Univ. Toronto Press, p. 277-279.

Schwartz, F.W., and Zhang, H., 2003, Fundamentals of groundwater (edu) Ch. 4.2 transmissive and storage properties of aquifers. John Wiley and sons

Soil Survey Staff, 2010, Keys to Soil Taxonomy, 11th ed, U.S.D.A.-N.R.C.S Agric. Handbook, printed in U.S.A.

Stephens, F.G., 1982, Soils Survey of Tulare County, California, Central Part, United States Department of Agriculture, Natural Resources Conservation Service http://www.nrcs.usda.gov/wps/portal/nrcs/site/national/home/

Thompson, H., Thomas, 1892, Official historical atlas Map of Tulare County

Unruh, J.R., 1991, The uplift of the Sierra Nevada and implications for late Cenozoic epeirogeny in the western Cordillera: Geological Society of America Bulletin, v. 103, p. 1395–1404,

U.S. Geological Survey, 1965, Geological research survey 1965: U.S. Geological Survey professional Paper 525-A, 376 p

Wahrhaftig, C., and, Birman, J.H., 1965, The Quaternary of the Pacific Mountain System in California. In: Wright, H.E,. Jr., and Frey, D.G. (eds) The Quaternary of the United States, A Review Volume for the VII Congress of the International Association for Quaternary Research. Princeton University Press, Princeton, NJ, 299–340. 114 114 Wakabayashi, J., and Sawyer, T.L., 2001, Stream incision, tectonics, uplift, and evolution of topography of the Sierra Nevada, California: The Journal of Geology, v. 109, p. 539–562,

Wakabayashi, J., 2013, Paleochannels, stream incision, erosion, topographic evolution, and alternative explanations of paleoaltimetry, Sierra Nevada, California: Geosphere

Walker, M.J.C., 2005, Palaeosols (eds): Quaternary dating methods (Chichester: John Wiley) ch 7.5, p. 225-235

Wasner, A.R., and Arroues K.D., 2003, Soils Survey of Tulare County, California, Western Part, United States Department of Agriculture, Natural Resources Conservation Service http://www.nrcs.usda.gov/wps/portal/nrcs/site/national/home/

Weissmann, G.S. and Fogg, G.E. 1999, Multi-scale alluvial fan heterogeneity modeled with transition probability geostatistics in a sequence stratigraphic framework: Journal of Hydrology, 226, 48–65.

Weissmann, G.S., Mount, J.F., and Fogg, G.E. 2002, Glacially driven cycles in accumulation space and sequence stratigraphy of a stream-dominated alluvial fan, San Joaquin Valley, California: U.S.A. Journal of Sedimentary Research, 72, 270–281.

Weissmann, G.S., Bennett, G.L., and Lansdale A.L., 2005, Factors controlling sequence development on Quaternary fluvial fans, San Joaquin Basin, California, USA, HARVEY, A.M.,MATHER, A.E. & STOKES, M. (eds) 2005. From Alluvial Fans: Geomorphology, Sedimentology, Dynamics. Geological Society, London, Special Publications, 251, 169–186. 0305– 8719/05/$15 © The Geological Society of London 2005.

APPENDIX: HYDRAULIC CONDUCTIVITY AND TRANSMISSIVITY DATA

116

Depth Depth Interval Interval Ks Ks Ks Actual-T Actual-T (ft) (m) (ft) (m) (cm/min) (m/day) (ft/day) (ft2/day) (m2/day) 0.5 0.2 0.5 0.15 0.00012 0.00171 0.00561 0.00280 0.00026 1 0.3 0.5 0.15 0.00007 0.00100 0.00330 0.00165 0.00015 1.5 0.5 0.5 0.15 0.00006 0.00087 0.00285 0.00143 0.00013 2 0.6 0.5 0.15 0.00003 0.00048 0.00159 0.00079 0.00007 3 0.9 1 0.30 0.00023 0.00336 0.01102 0.01102 0.00102 4 1.2 1 0.30 0.00100 0.01445 0.04742 0.04742 0.00441 4.5 1.4 0.5 0.15 0.00009 0.00131 0.00429 0.00215 0.00020 5 1.5 0.5 0.15 0.00006 0.00088 0.00287 0.00144 0.00013 5.5 1.7 0.5 0.15 0.00007 0.00102 0.00333 0.00167 0.00015 6 1.8 0.5 0.15 0.00004 0.00059 0.00194 0.00097 0.00009 7 2.1 1 0.30 0.00073 0.01048 0.03440 0.03440 0.00320 7.4 2.3 0.4 0.12 0.00015 0.00218 0.00714 0.00286 0.00027 9.5 2.9 2.1 0.64 0.00216 0.03116 0.10222 0.21467 0.01994 10 3.0 0.5 0.15 0.01040 0.14972 0.49120 0.24560 0.02282 10.2 3.1 0.2 0.06 0.00079 0.01144 0.03752 0.00750 0.00070 11 3.4 0.8 0.24 0.00764 0.11004 0.36103 0.28882 0.02683 12 3.7 1 0.30 0.00388 0.05587 0.18332 0.18332 0.01703 13 4.0 1 0.30 0.00364 0.05247 0.17215 0.17215 0.01599 14 4.3 1 0.30 0.00172 0.02477 0.08125 0.08125 0.00755 14.5 4.4 0.5 0.15 0.01823 0.26254 0.86136 0.43068 0.04001 16 4.9 1.5 0.46 0.00002 0.00034 0.00111 0.00167 0.00015 16.5 5.0 0.5 0.15 0.00014 0.00195 0.00640 0.00320 0.00030 17 5.2 0.5 0.15 0.00034 0.00484 0.01589 0.00795 0.00074 17.7 5.4 0.7 0.21 0.00260 0.03739 0.12268 0.08587 0.00798 18.2 5.5 0.5 0.15 0.01641 0.23628 0.77519 0.38759 0.03601 18.7 5.7 0.5 0.15 0.00818 0.11785 0.38666 0.19333 0.01796 19.3 5.9 0.6 0.18 0.00082 0.01182 0.03877 0.02326 0.00216 19.8 6.0 0.5 0.15 0.00055 0.00798 0.02618 0.01309 0.00122 20.3 6.2 0.5 0.15 0.00529 0.07621 0.25002 0.12501 0.01161 20.8 6.3 0.5 0.15 0.00130 0.01874 0.06148 0.03074 0.00286 21.8 6.6 1 0.30 0.00531 0.07643 0.25076 0.25076 0.02330 22.3 6.8 0.5 0.15 0.00195 0.02809 0.09216 0.04608 0.00428 22.8 6.9 0.5 0.15 0.00270 0.03893 0.12772 0.06386 0.00593 23.3 7.1 0.5 0.15 0.00340 0.04901 0.16081 0.08040 0.00747 23.8 7.3 1.5 0.46 0.00060 0.00863 0.02830 0.04246 0.00394 24.5 7.5 0.7 0.21 0.00262 0.03775 0.12385 0.08669 0.00805 25 7.6 0.5 0.15 0.00203 0.02923 0.09591 0.04796 0.00446 25.5 7.8 0.5 0.15 0.00406 0.05844 0.19174 0.09587 0.00891 26 7.9 0.5 0.15 0.00472 0.06803 0.22321 0.11160 0.01037 117 117 26.5 8.1 0.5 0.15 0.00206 0.02973 0.09753 0.04876 0.00453 27.5 8.4 1 0.30 0.00398 0.05736 0.18819 0.18819 0.01748 28 8.5 0.5 0.15 0.00506 0.07280 0.23884 0.11942 0.01109 29 8.8 1 0.30 0.00229 0.03303 0.10837 0.10837 0.01007 29.5 9.0 0.5 0.15 0.00207 0.02976 0.09764 0.04882 0.00454 30 9.1 0.5 0.15 0.00189 0.02728 0.08949 0.04475 0.00416 31 9.4 1 0.30 0.00875 0.04559 0.14957 0.14957 0.01390 31.5 9.6 0.5 0.15 0.00069 0.00987 0.03237 0.01619 0.00150 32 9.8 0.5 0.15 0.00171 0.02457 0.08062 0.04031 0.00374 32.5 9.9 0.5 0.15 0.00187 0.02695 0.08843 0.04421 0.00411 33 10.1 0.5 0.15 0.00083 0.01195 0.03921 0.01961 0.00182 34 10.4 1 0.30 0.00025 0.00357 0.01171 0.01171 0.00109 34.5 10.5 0.5 0.15 0.01860 0.34997 1.14820 0.57410 0.05334 35.5 10.8 1 0.30 0.14951 1.06260 3.48623 3.48623 0.32388 36.5 11.1 1 0.30 0.10244 1.47509 4.83955 4.83955 0.44961 37 11.3 0.5 0.15 0.00003 0.00045 0.00148 0.00074 0.00007 37.5 11.4 0.5 0.15 0.00004 0.00058 0.00190 0.00095 0.00009 38.3 11.7 0.8 0.24 0.00005 0.00076 0.00251 0.00200 0.00019 39 11.9 0.7 0.21 0.06374 0.33601 1.10239 0.77168 0.07169 39.5 12.0 0.5 0.15 0.51686 7.44283 24.41875 12.20937 1.13429 40 12.2 0.5 0.15 1.18428 17.05369 55.95044 27.97522 2.59898 40.5 12.3 0.5 0.15 1.93863 27.91629 91.58888 45.79444 4.25444 41 12.5 0.5 0.15 3.13336 45.12043 148.03290 74.01645 6.87635 41.5 12.6 0.5 0.15 2.25441 32.46353 106.50764 53.25382 4.94744 42 12.8 0.5 0.15 3.78776 54.54373 178.94923 89.47462 8.31246 42.5 13.0 0.5 0.15 2.12117 30.54481 100.21263 50.10632 4.65503 43 13.1 0.5 0.15 2.45902 35.40982 116.17396 58.08698 5.39646 43.5 13.3 0.5 0.15 3.42788 49.36144 161.94697 80.97349 7.52268 44 13.4 0.5 0.15 2.72108 39.18362 128.55517 64.27758 5.97158 45.4 13.8 1.4 0.43 2.00602 28.88662 94.77238 132.68134 12.32650 45.6 13.9 0.2 0.06 0.00205 0.02953 0.09689 0.01938 0.00180 46 14.0 0.4 0.12 0.02901 0.41778 1.37067 0.54827 0.05094 49 14.9 3 0.91 0.03200 0.46079 1.51178 4.53535 0.42135 49.5 15.1 0.5 0.15 0.62467 8.99524 29.51193 14.75597 1.37087 50 15.2 0.5 0.15 1.77449 25.55260 83.83399 41.91700 3.89422 51 15.5 1 0.30 1.86102 26.79866 87.92212 87.92212 8.16823 52 15.8 1 0.30 1.82829 26.32739 86.37597 86.37597 8.02459 53 16.2 1 0.30 6.91854 54.54373 178.94923 178.94923 16.62493 54 16.5 1 0.30 0.84286 12.13724 39.82033 39.82033 3.69943 55 16.8 1 0.30 3.78239 54.46636 178.69540 178.69540 16.60135 56 17.1 1 0.30 3.01648 43.43732 142.51089 142.51089 13.23970 57.3 17.5 1.3 0.40 3.33022 47.95515 157.33318 204.53313 19.00175 118 118 58 17.7 0.7 0.21 0.00101 0.01455 0.04775 0.03342 0.00311 58.5 17.8 0.5 0.15 0.13607 0.07588 0.24894 0.12447 0.01156 59 18.0 0.5 0.15 0.00718 0.10338 0.33919 0.16959 0.01576 63.9 19.5 4.9 1.49 0.22858 3.29150 10.79889 52.91458 4.91593 64.2 19.6 0.3 0.09 0.14212 1.54034 5.05361 1.51608 0.14085 65 19.8 0.8 0.24 1.34681 19.39408 63.62887 50.90309 4.72905 65.5 20.0 0.5 0.15 1.59938 23.03105 75.56118 37.78059 3.50993 66 20.1 0.5 0.15 1.73311 24.95672 81.87900 40.93950 3.80340 66.5 20.3 0.5 0.15 2.78093 40.04544 131.38267 65.69133 6.10292 67 20.4 0.5 0.15 3.16267 45.54243 149.41742 74.70871 6.94067 67.5 20.6 0.5 0.15 1.14258 16.45318 53.98025 26.99012 2.50746 68.8 21.0 1.3 0.40 0.14562 0.17409 0.57118 0.74253 0.06898 69.6 21.2 0.8 0.24 0.00338 0.04865 0.15963 0.12770 0.01186 76 23.2 6.4 1.95 0.12208 1.23684 4.05788 25.97042 2.41273 79 24.1 3 0.91 1.42710 20.55026 67.42212 202.26636 18.79116 79.5 24.2 0.5 0.15 0.00006 0.00085 0.00278 0.00139 0.00013 80 24.4 0.5 0.15 0.36736 5.28997 17.35554 8.67777 0.80619 81 24.7 1 0.30 1.15648 16.65333 54.63689 54.63689 5.07593 82 25.0 1 0.30 0.63592 9.15726 30.04351 30.04351 2.79113 83 25.3 1 0.30 0.22734 0.06218 0.20399 0.20399 0.01895 84 25.6 1 0.30 0.46078 6.63530 21.76936 21.76936 2.02244 84.5 25.8 0.5 0.15 0.03370 0.48532 1.59227 0.79614 0.07396 85 25.9 0.5 0.15 0.58262 0.02412 0.07915 0.03957 0.00368 86 26.2 1 0.30 0.20762 2.98974 9.80887 9.80887 0.91127 86.5 26.4 0.5 0.15 0.00016 0.00237 0.00777 0.00389 0.00036 87.2 26.6 0.7 0.21 0.00021 0.00301 0.00988 0.00691 0.00064 88 26.8 0.8 0.24 0.07576 0.01465 0.04807 0.03845 0.00357 89 27.1 1 0.30 0.01809 0.01019 0.03344 0.03344 0.00311 89.5 27.3 0.5 0.15 0.00087 0.01246 0.04088 0.02044 0.00190 91 27.7 1.5 0.46 0.00088 0.01266 0.04155 0.06232 0.00579 91.5 27.9 0.5 0.15 0.00734 0.10574 0.34692 0.17346 0.01611 92 28.0 0.5 0.15 0.00425 0.06122 0.20085 0.10042 0.00933 92.5 28.2 0.5 0.15 0.00089 0.01277 0.04189 0.02094 0.00195 93 28.3 0.5 0.15 0.00211 0.03040 0.09975 0.04987 0.00463 93.5 28.5 0.5 0.15 0.00022 0.00317 0.01041 0.00520 0.00048 94 28.7 0.5 0.15 0.00063 0.00904 0.02964 0.01482 0.00138 94.5 28.8 0.5 0.15 0.00069 0.00991 0.03250 0.01625 0.00151 95 29.0 0.5 0.15 0.00071 0.01029 0.03375 0.01688 0.00157 95.5 29.1 0.5 0.15 0.00005 0.00077 0.00254 0.00127 0.00012 96 29.3 0.5 0.15 0.00039 0.00561 0.01839 0.00920 0.00085 96.5 29.4 0.5 0.15 0.00054 0.00784 0.02571 0.01286 0.00119 97 29.6 0.5 0.15 0.00318 0.04582 0.15034 0.07517 0.00698 119 119 97.5 29.7 0.5 0.15 0.00010 0.00148 0.00484 0.00242 0.00022 98 29.9 0.5 0.15 0.00014 0.00206 0.00675 0.00338 0.00031 98.5 30.0 0.5 0.15 0.00045 0.00642 0.02105 0.01053 0.00098 99 30.2 0.5 0.15 0.00184 0.02655 0.08712 0.04356 0.00405 100 30.5 1 0.30 0.00003 0.00044 0.00146 0.00146 0.00014 101 30.8 1 0.30 0.00225 0.03233 0.10608 0.10608 0.00986 103 31.4 2 0.61 0.05440 0.78334 2.57002 5.14003 0.47752 104 31.7 1 0.30 0.02883 0.41522 1.36227 1.36227 0.12656 111 33.8 7 2.13 0.20487 2.95015 9.67895 67.75268 6.29443 113 34.4 2 0.61 0.26432 3.80618 12.48747 24.97493 2.32025 114.2 34.8 1.2 0.37 0.00279 0.04016 0.13177 0.15813 0.01469 114.8 35.0 0.6 0.18 0.00003 0.00041 0.00135 0.00081 0.00008 115.5 35.2 0.7 0.21 0.00012 0.00171 0.00562 0.00393 0.00037 116 35.4 0.5 0.15 0.00016 0.00234 0.00767 0.00384 0.00036 116.5 35.5 0.5 0.15 0.00010 0.00144 0.00471 0.00235 0.00022 117 35.7 0.5 0.15 0.00086 0.01241 0.04071 0.02036 0.00189 117.5 35.8 0.5 0.15 0.00025 0.00356 0.01167 0.00584 0.00054 118 36.0 0.5 0.15 0.00032 0.00467 0.01531 0.00765 0.00071 118.5 36.1 0.5 0.15 0.00006 0.00087 0.00285 0.00142 0.00013 119 36.3 0.5 0.15 0.00035 0.00504 0.01652 0.00826 0.00077 119.5 36.4 0.5 0.15 0.00013 0.00186 0.00609 0.00304 0.00028 120 36.6 0.5 0.15 0.00009 0.00127 0.00416 0.00208 0.00019 120.5 36.7 0.5 0.15 0.00003 0.00049 0.00159 0.00080 0.00007 121 36.9 0.5 0.15 0.00010 0.00146 0.00480 0.00240 0.00022 121.5 37.0 0.5 0.15 0.00009 0.00127 0.00416 0.00208 0.00019 122 37.2 0.5 0.15 0.00026 0.00381 0.01250 0.00625 0.00058 122.5 37.3 0.5 0.15 0.00002 0.00026 0.00086 0.00043 0.00004 123 37.5 0.5 0.15 0.00006 0.00093 0.00304 0.00152 0.00014 123.5 37.6 0.5 0.15 0.00002 0.00024 0.00080 0.00040 0.00004 124 37.8 0.5 0.15 0.00007 0.00096 0.00316 0.00158 0.00015 124.5 37.9 0.5 0.15 0.00002 0.00028 0.00092 0.00046 0.00004 125 38.1 0.5 0.15 0.00352 0.05070 0.16634 0.08317 0.00773 125.5 38.3 0.5 0.15 0.00006 0.00083 0.00271 0.00136 0.00013 126 38.4 0.5 0.15 0.00009 0.00134 0.00441 0.00221 0.00020 127 38.7 1 0.30 0.00003 0.00045 0.00148 0.00148 0.00014 127.5 38.9 0.5 0.15 0.00009 0.00126 0.00413 0.00207 0.00019 128 39.0 0.5 0.15 0.00006 0.00084 0.00275 0.00138 0.00013 128.5 39.2 0.5 0.15 0.00003 0.00046 0.00152 0.00076 0.00007 129 39.3 0.5 0.15 0.00067 0.00959 0.03146 0.01573 0.00146 132 40.2 3 0.91 0.05319 0.76590 2.51278 7.53835 0.70034

Fresno State

Non-Exclusive Distribution License (to archive your thesis/dissertation electronically via the library’s eCollections database)

By submitting this license, you (the author or copyright holder) grant to Fresno State Digital Scholar the non-exclusive right to reproduce, translate (as defined in the next paragraph), and/or distribute your submission (including the abstract) worldwide in print and electronic format and in any medium, including but not limited to audio or video.

You agree that Fresno State may, without changing the content, translate the submission to any medium or format for the purpose of preservation.

You also agree that the submission is your original work, and that you have the right to grant the rights contained in this license. You also represent that your submission does not, to the best of your knowledge, infringe upon anyone’s copyright.

If the submission reproduces material for which you do not hold copyright and that would not be considered fair use outside the copyright law, you represent that you have obtained the unrestricted permission of the copyright owner to grant Fresno State the rights required by this license, and that such third-party material is clearly identified and acknowledged within the text or content of the submission.

If the submission is based upon work that has been sponsored or supported by an agency or organization other than Fresno State, you represent that you have fulfilled any right of review or other obligations required by such contract or agreement.

Fresno State will clearly identify your name as the author or owner of the submission and will not make any alteration, other than as allowed by this license, to your submission. By typing your name and date in the fields below, you indicate your agreement to the terms of this distribution license.

Embargo options (fill box with an X).

Make my thesis or dissertation available to eCollections immediately upon

submission.

X Embargo my thesis or dissertation for a period of 2 years from date of graduation.

Embargo my thesis or dissertation for a period of 5 years from date of graduation.

Dustin C. White

Type full name as it appears on submission

April 21, 2006

Date