You are on page 1of 121

University of Nevada, Reno

Discovery and Analysis of a Blind Geothermal System in


Southeastern Gabbs Valley, Western Nevada

A thesis submitted in partial fulfillment of the


requirements for the degree of Master of Science in
Geology

by

Jason Westlee Craig

Dr. James E. Faulds/Thesis Advisor

December, 2018
THE GRADUATE SCHOOL

We recommend that the thesis


prepared under our supervision by

JASON W. CRAIG

Entitled

Discovery And Analysis Of A Blind Geothermal System,


Southeastern Gabbs Valley, Western Nevada

be accepted in partial fulfillment of the


requirements for the degree of

MASTER OF SCIENCE

James E. Faulds, Ph.D., Advisor

John N. Louie, Ph.D., Committee Member

Scott A. Mensing, Ph.D., Graduate School Representative

David W. Zeh, Ph.D., Dean, Graduate School

December, 2018
i

ABSTRACT

This study assessed the potential for high-temperature (≥130C) blind geothermal

systems in southeastern Gabbs Valley by employing a geothermal play fairway analysis

approach involving integration of geologic, geophysical, and geochemical datasets. Gabbs

Valley is a complex, tectonically active basin within the Great Basin on the boundary between

the transtensional central Walker Lane domain and extensional Basin and Range province. The

termination of the Petrified Springs fault, a major dextral fault of the central Walker Lane, in an

array of normal faults indicates that southeastern Gabbs Valley occupies a displacement transfer

zone, which is a favorable structural setting for geothermal activity. The displacement transfer

zone is a structurally complicated area with many faults of varying geometries, kinematics, and

ages. A substantial northwest-trending gravity high within the south-central part of southeastern

Gabbs Valley is produced from offsets along concealed northwest-striking dextral-normal faults

that intersect strands of north-northeast-to east-northeast-striking normal faults. This area of

tightly spaced faults and fault intersections is a particularly favorable structural setting within the

broader-scale displacement transfer zone. Multiple lines of direct and indirect evidence suggest

the presence of a relatively high-temperature (130C) blind geothermal system in this area,

including collocated intersecting gravity gradients, magnetic-low, low-resistivity, and 2-m

temperature anomaly. Potentially related, water samples from agricultural wells ~7 km northwest

of the 2-m anomaly yield geothermometers indicating subsurface fluid temperatures of 130-

140°C.

Six temperature-gradient holes were drilled to target the extent of the shallow-

temperature and geophysical anomalies to define the geothermal system in southeastern Gabbs

Valley. Two wells contained high temperatures exceeding boiling with bottom-hole temperatures
ii

of 114.5°C and 124.9°C, respectively, and the remaining wells displayed elevated to background

temperatures ranging from 79.2°C to 28.7°C. The observed temperature gradients for the two

hottest drill holes necessitate drill-hole intercepts of convecting hydrothermal fluids. The 2-m

survey is spatially and thermally consistent with the temperature-gradient holes and constrains

the core of upflow at a shallow depth interval (<250m) to an area <0.5-1 km west-southwest of

the hottest drill hole. The zone of inferred upwelling occupies the area corresponding to the

center of the magnetic-low.

This project has several significant implications. The newly discovered blind geothermal

system in southeastern Gabbs Valley may be capable of supporting a binary-type geothermal

power plant. This detailed study provides an initial validation of the play fairway approach to

geothermal exploration. It also demonstrates both the broad applicability of the play fairway

exploration strategy and the large untapped potential for commercial-grade blind geothermal

systems in many of the basins in the Great Basin region.


iii

ACKNOWLEDGMENTS

I am extremely grateful to my advisor Dr. James Faulds for giving me the opportunity to

work on this project, guiding me along the process, providing valuable criticism, and shaping me

into a better scientist and geologist. It has been an honor to learn from you. I am grateful to Dr.

Mark Coolbaugh for his experience and foresight on this project. Thanks to Nicholas Hinz for his

help and insight into all matters from geothermal to ArcGIS. I am indebted to Andrew Sadowski

for his teaching and wisdom in the early stages of this project. Thank you to Lisa Shevenell for

all the help troubleshooting numerous problems along the way. Thanks to Chris Sladek for his

patience and support with the 2-m survey. Thanks to Steve DeOreo for managing the drilling

program. Thank you to the crew at Zonge International, Inc. for completing the initial gravity

survey. I am grateful to the USGS Geophysics Unit of Menlo Park including Dr. Johnathan Glen,

Dr. Jared Peacock, Tait Earney, and William Schermerhorn for their help and collaboration.

Thank you to Dr. Drew Siler for the help conducting slip and dilatational tendency analysis. I

gratefully acknowledge Irene Seelye for all her cartographic help and endless ArcGIS

knowledge. Thank you to Emma McConville for the support and inspiration. Thank you to my

committee members Dr. John N. Louie and Dr. Scott A. Mensing. This work was funded by a

U.S. Department of Energy grant (DE-EE0006731) awarded to Faulds. Student research

scholarships from the Nevada Petroleum and Geothermal society and the American Association

of Petroleum Geologists also contributed to this work. Lastly, I want to thank my friends and

family for all their support over the past few years- it’s been a wild ride and I wouldn’t have

made it without you.


iv

TABLE OF CONTENTS

Abstract ............................................................................................................................................ i

Acknowledgments.......................................................................................................................... iii

Table of Contents ........................................................................................................................... iv

List of Tables ................................................................................................................................. vi

List of Figures ............................................................................................................................... vii

1.0 Introduction ................................................................................................................................1

2.0 Regional Tectonic Setting ........................................................................................................11

3.0 Geologic Setting of Gabbs Valley ...........................................................................................16

3.1 Stratigraphic Framework ......................................................................................................18

3.2 Structural Framework ...........................................................................................................23

3.2 Geothermal Framework........................................................................................................27

4.0 Methods and Results ................................................................................................................29

4.1 Geologic Investigation .........................................................................................................29

4.2 Geophysical Surveys ...........................................................................................................43

4.2.1 Gravity Survey ..........................................................................................................43

4.2.2 Magnetic Survey .......................................................................................................47

4.2.3 Magnetotelluric Survey.............................................................................................49

4.3 Cross Sections ......................................................................................................................51

4.4 Slip and Dilatational Tendency Analysis .............................................................................51


v

4.5 Fluid Geochemistry Investigations .......................................................................................54

4.6 Shallow Temperature Survey ...............................................................................................57

5.0 Discussion ................................................................................................................................59

5.1 Conceptual Structural Model ...............................................................................................59

5.2 Geothermal Prospects ...........................................................................................................65

5.3 Play Fairway Analysis ..........................................................................................................72

5.4 Temperature-Gradient Drilling.............................................................................................76

5.5 Blind Geothermal System in Southeastern Gabbs Valley ....................................................83

5.6 Broader Implications ............................................................................................................86

6.0 Conclusions ..............................................................................................................................89

References Cited ............................................................................................................................92

Appendix A: Detailed Geologic Unit Descriptions .....................................................................107

Plate 1...........................................................................................................................................110
vi

LIST OF TABLES
Table 1. Simplified unit descriptions of southeastern Gabbs Valley .............................................22

Table 2. Fluid geothermometry of Gabbs Valley ..........................................................................56


vii

LIST OF FIGURES

Figure 1.1. Map of high-temperature geothermal systems of the Great Basin region in the Basin

and Range province..........................................................................................................................4

Figure 1.2: Schematic diagrams of favorable structural settings .....................................................6

Figure 1.3. Geothermal play fairway map of west-central to eastern Nevada .................................9

Figure 1.4. Nevada play fairway modeling workflow ...................................................................10

Figure 2.1. Tectonic map of western North America. ..................................................................13

Figure 2.2. Historic earthquakes of central Nevada seismic belt in western Nevada ....................15

Figure 3.0.1. Regional overview map of west-central Nevada. .....................................................17

Figure 3.0.2. Area map for Gabbs Valley ......................................................................................18

Figure 3.1.1. Simplified stratigraphic column of southeastern Gabbs Valley study area ..............20

Figure 3.2.1 Simplified geologic map of Gabbs Valley ................................................................24

Figure 3.2.2. Quaternary faults of Gabbs Valley ...........................................................................26

Figure 3.3.1. Known geothermal systems of Gabbs Valley ..........................................................28

Figure 4.1.1. Simplified geologic map of the southeastern Gabbs Valley study area ...................30

Figure 4.1.2. Extent of LiDAR coverage in the southeastern Gabbs Valley study area ................32

Figure 4.1.3. Quaternary geologic map of the Petrified Springs fault, southeastern Gabbs Valley

........................................................................................................................................................34
viii

Figure 4.1.4. Normal fault splays of the southern Gabbs Valley fault zone ..................................36

Figure 4.1.5. Quaternary faults in southeastern Gabbs Valley with attributed rupture recency and

slip rate values................................................................................................................................37

Figure 4.1.6. Geologic map of Gabbs Mountain ...........................................................................40

Figure 4.1.7. Lower-hemisphere, equal-area stereographic projections of measured attitudes in

rocks ...............................................................................................................................................41

Figure 4.2.1. Residual isostatic gravity map of southeastern Gabbs Valley..................................46

Figure 4.2.2. Reduced to pole (RTP) magnetic map of the study area ..........................................48

Figure 4.2.3. MT resistivity depth profiles ....................................................................................50

Figure 4.4.1. Slip and dilatation tendency analyses for the southeastern Gabbs Valley fault

system ............................................................................................................................................53

Figure 4.5.1 Piper diagram of water samples from southeastern Gabbs Valley ............................55

Figure 4.5.2. GeoT multicomponent chemical equilibria model results........................................55

Figure 4.6.1. Shallow 2-m temperature survey in southeastern Gabbs Valley. ............................58

Figure 5.1.1. Southeastern Gabbs Valley displacement transfer zone ...........................................61

Figure 5.1.2. Geologic map and cross section of Gabbs Mountain with highlighted fault

intersections ...................................................................................................................................63

Figure 5.2.1. Favorable structural settings of southeastern Gabbs Valley.....................................66


ix

Figure 5.2.2. Collocated intersecting gravity gradients, magnetic low, low resistivity, 2-m

temperature anomaly, and favorable oriented faults for slip and dilation .....................................68

Figure 5.2.3. Geothermometor sample locations in relation to favorable structural settings and 2-

m temperature anomaly..................................................................................................................70

Figure 5.3.1. Phase I-III play fairway maps...................................................................................74

Figure 5.3.2. Phase I-III fairway error maps ..................................................................................75

Figure 5.4.1. Temperature gradient data and drill hole locations for the southeastern Gabbs

Valley blind geothermal system.....................................................................................................78

Figure 5.4.2. Spatial relationships of temperature and magnetic data ...........................................81

Figure 5.4.3. Phase I-III direct evidence maps ..............................................................................82


1

1. Introduction

Geothermal energy is a form of power production that utilizes the naturally

occurring heat stored in the Earth to produce constant energy with minimal carbon

footprint per megawatt generated. Geothermal is a baseload energy resource, defined as

the ability to produce energy continuously, year-round. Geothermal energy is available in

multiple geologic settings (e.g., volcanic, extensional, amagmatic) across the globe.

Further, geothermal development provides solutions to environmental, economic, and

social issues associated with increased energy needs for the growing population.

A primary limiting factor for development of geothermal resources for electrical

generation is discovering economic-grade hydrothermal systems. Three critical physical

properties must spatially coalesce to manifest geothermal activity: 1) a heat source (e.g.

magmatic body or high geothermal gradients in amagmatic settings), 2) reservoir to host

the fluids, and 3) fluids to transport heat stored at depth to the surface. Most of the

geothermal resources that have been developed into power plants were first identified

from surface manifestations of a hydrothermal system (e.g., hot springs, geysers,

fumaroles, mud pots, etc.) and are known as conventional geothermal systems. Recently,

exploration focus has shifted to discovering geothermal systems that are concealed in the

subsurface and lack obvious surface manifestations. These buried geothermal systems are

known as ‘hidden’ or ‘blind’ resources. Distinguishing surficial features associated with

hidden geothermal systems are subtle, including paleo-hydrothermal deposits (e.g.,

siliceous sinter, travertine, and tufa), rock alteration, and vegetation anomalies produced

from outflow of fluids and gasses. Paleo-hydrothermal deposits may have developed in
2

different hydrologic conditions and in some cases have been shown to overlie an active

geothermal system hidden in the subsurface (Coolbaugh et al., 2009). Blind geothermal

systems are contained entirely in the subsurface and lack any surface expression or paleo-

manifestations of hidden and conventional systems.

Estimates suggest a large undiscovered resource base of blind systems in the

Great Basin region, which could account for more than 3.2 times the number of currently

developed geothermal systems in the region (Coolbaugh and Shevenell, 2004; Coolbaugh

et al., 2006). Discovery and development of blind resources could exceed production of

previously known systems. This study is primarily motivated by the substantial potential

of undiscovered geothermal resources in the Great Basin region and the scientific

complexities associated with exploration for blind geothermal resources.

The primary purpose of this study was to define the structural setting of a

potential blind geothermal system in southeastern Gabbs Valley, Nevada, and to use

multiple exploration methods to characterize a prospective resource in this area. The

overarching goal of this work is to use play fairway methodology, outlined in the Nevada

play fairway project (Faulds et al., 2015a, 2016a, 2016b, 2017b) and discussed in detail in

subsequent sections, to discover new economic-grade, relatively high-temperature

(≥130°C) systems in the Great Basin region.

The study area is in southeastern Gabbs Valley, Nevada, in the west-central part

of the Basin and Range physiographic province (Figure 1.1). The study area spans the

boundary of two tectonic regimes; the transtensional Walker Lane domain and the

extensional Basin and Range province. The interaction between two different structural

domains within the study area enhances the structural complexity and thus increases the
3

potential of geothermal activity. Notably, this area had no previously known geothermal

activity or geothermal exploration.

This study involved collection and integration of multiple geologic, geophysical,

and geochemical datasets to characterize a potential blind geothermal system in

southeastern Gabbs Valley. Major tasks included: 1) analysis of stratigraphic relations in

the area, 2) delineating the geometry and kinematics of fault systems, 3) identifying

favorable structural settings for geothermal activity, 4) discovering and defining the

locations of thermal anomalies associated with zones of geothermal upwelling, 5)

selecting and surveying the most favorable exploration target(s) for drilling, and 6)

estimating the viability of the geothermal system for further development.

Overview of Great Basin Geothermal Systems

Geothermal activity in the Great Basin region of the western United States is

prolific (Figure 1.1, Coolbaugh et al., 2002), and the abundance of geothermal systems is

linked to anomalously high geothermal gradients (Blackwell, 1983; Blackwell et al.,

2010, 2011) produced from regional extension and resultant thinning of the crust (Faulds

et al., 2004). The Great Basin region is currently the most actively extending part of the

Basin and Range province and includes nearly all of Nevada, western Utah, southern

Idaho, southeastern Oregon, easternmost California, and northwestern-most Arizona

(Figure 1.1). Globally, most geothermal systems are associated with magmatism and

located in subduction, rift, or hotspot settings (Moeck et al., 2015; Curewitz and Karson,

1997). Volcanism within the Great Basin largely ceased between 10 and 3 Ma

(Christiansen and McKee, 1978), thus excluding magmatism as a source of heat for most
4

of the geothermal systems in the region. With the exception of a few magmatic-

influenced systems (e.g., Steamboat, Coso, and Roosevelt), most geothermal systems in

the Great Basin are amagmatic. Amagmatic geothermal systems typically derive heat

from high mantle heat flow to shallow levels of the crust (McKenzie, 1978; Henley and

Brown, 1985; Curewitz and Karson, 1997).

Figure 1.1. Map of high-temperature geothermal systems of the Great Basin region in the Basin and Range
province. The Great Basin region is outlined in red with major geothermal systems displayed as circles
(yellow <150˚C, red >150˚C, from Faulds et al., 2004). Magmatic systems are large pink circles.
Quaternary faults are black lines. The boundary of the Walker Lane-eastern California shear zone is gray
polygon with cross-hatch pattern. Study area is shown in the blue outline.
5

Permeability is commonly the most important limiting factor for geothermal

activity (Faulds et al., 2006, 2011). Generally, a local increase in permeability is

necessary to allow for the transmission and convection of fluids (Curewitz and Karson,

1997). Faulting is a mechanism capable of increasing the permeability of rocks through

propagation and dilation of interconnected fracture networks in the core and damage zone

of a fault (Caine et al., 1996). Specific fault geometries have been shown to increase

dilatation along fault systems, resulting in a conduit for fluid transport ( Curewitz and

Karson, 1997; Ferrill and Morris, 2003; Micklethwaite and Cox, 2004; Faulds et al.,

2006, 2010, 2011, 2013). Favorable structural settings in the Great Basin region include

discrete fault step-overs in normal faults, fault terminations, fault intersections,

accommodation zones (cf., Faulds and Varga, 1998), and transtensional displacement

transfer zones (Faulds et al., 2006, 2010, 2011, 2013). Because geothermal systems

within the Great Basin are primarily controlled by faults, the identification of favorable

structural settings provides a method for guiding geothermal exploration (Figure 1.2).

The combination of multiple structural settings at a single locality has been shown to

further increase the probability and extent of geothermal activity (Faulds et al., 2013;

Faulds and Hinz, 2015).


6

Figure 1.2: Schematic diagrams of favorable structural settings (modified from Faulds and Hinz, 2015).
Faults are shown in black. Normal faults are indicated by fault balls on the hanging walls, and motion along
strike-slip faults is shown by arrows. Red circles and ellipsoids indicate potential geothermal upwelling
zones. Structural settings in the Great Basin region associated with geothermal activity include: (A) major
normal fault, (B), fault bends in a range-front normal fault, (C) major normal fault terminating in horsetails
or fault splays, (D) hard or soft-linked major fault step-overs (relay ramps), (E) fault intersection between
two or more major faults, (F) accommodation zones between two sets of oppositely dipping normal faults,
(G) displacement transfer zones where a strike-slip fault terminates by transferring displacement to several
normal faults, and (H) transtensional pull-aparts formed by releasing steps in strike-slip faults. Favorable
structural settings for geothermal systems include C through E. Favorable structural settings observed in
southeastern Gabbs Valley are indicated with yellow stars.

All relatively high-temperature (≥130°C) amagmatic geothermal systems in the

Great Basin region are associated with Quaternary faults (Bell and Ramelli, 2007; 2009).

Therefore, it is essential to understand fault rupture history and attribute fault strands with

age of rupture. Faults with the most recent rupture event younger than ~750 ka are more

likely to host a geothermal system (Faulds et al., 2015a, 2017b). Over the lifespan of a

geothermal system, dissolved solutes in hydrothermal fluids precipitate along the flow

pathway and reduce flow rates within structural conduits. This is especially true for
7

geothermal systems with a boiling horizon exolving volatiles (Drummond and Ohmoto,

1985). Active fault systems with minimum mineral precipitation and open fractures in the

reservoir are ideal for producing the high flow rates needed for economic geothermal

operations.

Nevada Geothermal Play Fairway Project

The development of geothermal prospects requires significant upfront exploration

cost. Much of the initial investment involves the drilling program, which is inherently

risky with no guarantee of returns if the wells are unproductive. If a drill program is

successful, the project will typically pay off the initial investment in 4-8 years (Salmon et

al., 2011).

The play fairway methodology was first developed by the oil and gas industry

(Baker et al., 1986; Doust, 2003, 2010) by establishing a systematic approach of

integrating multiple datasets to narrow and rank drill targets to mitigate risk in

exploration programs. The U.S. Department of Energy initiated a program to encourage

the development of a play fairway methodology for geothermal exploration in Nevada

(Faulds et al., 2015a, 2015b, 2016a, 2016b), Utah (Wannamaker et al., 2015, 2017a,

2017b), Hawaii (Lautze et al., 2017), Idaho (Shervais et al., 2016), Washington (Forson

et al., 2016), and New Mexico (Nash and Bennett, 2015). The play fairway analysis is a

multidisciplinary approach, intended to minimize financial risk by utilizing relatively

low-cost geologic, geophysical, and geochemical methods to elucidate the quality of a

geothermal prospect for further investment.


8

The Nevada play fairway project is a three-phase study aimed at conducting one

of the first comprehensive regional-scale geothermal exploration programs. Phase 1 of

the project generated a regional geothermal potential map (Figure 1.3) covering 96,000
2
km of the Great Basin. The map spans west-central to eastern Nevada across a regional

strain gradient, a range of crystalline to carbonate basement lithologies, and differing

structural domains (Faulds et al., 2015a, 2015b). Nine geologic and geophysical

parameters were synthesized into the map (Figure 1.4) to show areas of high potential to

guide geothermal exploration. The nine parameters integrated into the geothermal

potential map include: 1) structural settings, 2) age of Quaternary faulting, 3) slip rates on

Quaternary faults, 4) regional-scale strain rates, 5) slip and dilation tendency on

Quaternary faults, 6) earthquake density, 7) gravity gradients, 8) temperatures at 3 km

depth, and 9) geochemistry from springs and wells. The first seven of these nine

parameters were distributed into subgroups to establish regional-scale permeability,

intermediate-scale permeability, and local-scale permeability layers for a robust

combined permeability model (Figure 1.4, Faulds et al., 2015a, 2015b). The combined

permeability model was merged with temperatures at 3 km depth to produce the fairway

or geothermal potential map (Figure 1.3). The geothermal potential map was then

iteratively combined with lines of direct evidence (e.g. well temperatures and

geothermometers) and a degree of exploration modifier to produce the additional

favorability maps (Faulds et al., 2015a, 2015b). The “fairway” identified areas with high

undiscovered geothermal potential and was subsequently used to guide exploration of

twenty-four of the highest-ranking prospects in Phase II (Faulds et al., 2016a, 2016b).

After evaluation of the twenty-four sites, five of the most favorable areas were down-
9

selected for detailed studies (Faulds et al., 2017a, 2017b). Among the down-selected sites

was southeastern Gabbs Valley. This study includes detail studies conducted in Phase II

and III of the Nevada play fairway project in southeastern Gabbs Valley.

Figure 1.3. Geothermal play fairway map of west-central to eastern Nevada. Warmer colors indicate areas
of higher potential; cool colors show areas of relatively low potential overlain on a DEM (digital elevation
model) hillshade (Faulds et al., 2017b). Fairway values are not normalized and are based on parameters and
relative weighting shown in Figure 1.4. Areas down selected for detailed studies are indicated by polygons
with black hachures. Black arrow points to the southeastern Gabbs Valley area. Cities are labeled on map
and major highways shown as black lines. Boundary of play fairway map is shown as red rectangle over the
state of Nevada for spatial reference in lower right corner of figure. Abbreviations for known geothermal
systems in the region: Br, Bradys; Bw, Beowawe; DP, Desert Peak; LA, Lee-Allen; MH, McGinness Hills;
SE, San Emidio; SL, Soda Lake; St, Stillwater; SW, Salt Wells; TM, Tungsten Mountain; WR, Wild Rose
or Don Campbell.
10

Figure 1.4. Nevada play fairway modeling workflow (Faulds et al., 2017b). Red numbers indicate relative
weights determined from weights of evidence. Black numbers indicate expert driven weights used in the
analysis. In all cases, the expert driven weights took into account the statistical analyses.
11

2. Regional Tectonic Setting

The Great Basin region is part of the Basin and Range province, which is a vast

Cenozoic extensional tectonic regime (Stewart, 1998a; Wernicke, 1992) across the

western United States and Mexico. In the strict sense, the Great Basin is a region of

western North America that has no hydrographic connection to the ocean. All

precipitation distributed into this watershed evaporates, infiltrates into the subsurface, or

is captured by topographic lows (e.g. lakes and basins). In this study, we define the

“Great Basin region” as a broad area (Figure 2.1) of active extensional to transtensional

tectonism that encompasses the Great Basin hydrographic area as well as neighboring

areas similar in tectonic setting. Although multiple tectonic events have affected this

region since the Paleozoic, Cenozoic episodes of extension are most relevant to this study

and therefore discussed herein.

The western Great Basin has experienced multiple extensional events since the

late Oligocene, including several pulses predating late Cenozoic Basin and Range

faulting and extension. For example, the Stillwater Range experienced two earlier

extensional events: 1) late Oligocene to early Miocene (initiation ca. 24-23 Ma) east-west

oriented extension, and 2) subsequent early to middle Miocene (initiated ca. 19-18 Ma)

relatively minor northeast-directed extension, as evidenced by northwest-striking dikes

(John, 1992, 1995; Hudson et al., 2000). Northwest-striking normal oblique-slip faults are

observed in the northern Wassuk Range, and the age of these faults is bracketed between

ca. 26 and 14 Ma (Dilles and Gans, 1995). In addition, Hardyman and Oldow (1991)

suggested a northeast-southwest oriented extension direction for the central Walker Lane

from late Oligocene to early Miocene (28 to 17 Ma). Initiation of northwest directed
12

Basin and Range extension is constrained to <15 Ma in the Yerington district (Dilles and

Gans, 1995, Surpless et al., 2002) and to 15-13.3 Ma in the southern Stillwater Range

(John, 1992). This extensional event has continued to present-day (Wesnousky et al.,

2005; Wesnousky et al., 2012) and has generated the characteristic north to northeast-

trending fault blocks of the region (Figure 2.1).

Initiation of dextral shear along the Walker Lane in the western Great Basin

approximately coincided with the northward migration of the Mendocino triple junction

that marked the termination of subduction on the plate margin and retreat of arc

volcanism to the northwest ( Faulds et al., 2005; Faulds and Henry, 2008). Geodetic

measurements show that ~20% percent of the dextral motion between the Pacific and

North American plates is currently accommodated by the eastern California shear zone

and Walker Lane (Figure 2.1, Hammond and Thatcher, 2005, 2007; Faulds and Henry,

2008; Hammond et al., 2009).

The eastern California shear zone and Walker Lane contain complex structural

domains of discontinuous northwest-striking right-lateral faults and areas of east-

northeast-striking left-lateral faults that accommodate clockwise vertical-axis rotation of

fault blocks (Stewart, 1988; Cashman and Fontaine, 2000; Faulds and Henry, 2008).

Estimates for total right-lateral offsets decrease northwestward from 40-110 km in the

eastern California shear zone and southern Walker Lane ( Beanland and Clark, 1982;

Dokka and Travis, 1983), to 40-78 km in the central Walker Lane (Ekren and Byers,

1984; Faulds and Henry, 2008), and 20-30 km in the northern Walker Lane (Faulds et al.,

2005; Faulds and Henry, 2008). In northeastern California, at the northernmost extent of

the Walker Lane, cumulative dextral offset is negligible (Faulds and Henry, 2008). As the
13

Walker Lane terminates northwestward, dextral shear is transferred to northwest directed

extension (Figure 2.1) across the northwestern Great Basin (Faulds et al., 2004, 2005),

with significant diffusion of dextral shear into extension within the central Nevada

seismic belt (Wesnousky et al., 2005; Wesnousky, 2005a; Faulds et al., 2005).

Figure 2.1. Tectonic map of western North America. Map shows boundaries of the Great Basin region in
red, structural domains of the Walker Lane-eastern California shear zone in blue (modified from Faulds and
Henry, 2008), and Quaternary faults as black lines (USGS, 2010) overlain on a geodetic strain map
(Kreemer et al,. 2012). Strain rates reflect the second invariant strain rate tensor (10-9/yr). Red arrow
indicates the general regional extension direction of N65ºW.
14

The central Nevada seismic belt is a north-northeast trending zone in the western

Great Basin (Figure 2.2) of enhanced seismicity, including several historical (<150

years), large magnitude (6.1-7.3Ms) earthquakes that extend from Monte Cristo Valley to

Pleasant Valley (Figure 2.2, Wesnousky et al., 2005). It is also characterized by high

strain rates relative to surrounding areas (Figure 2.1, Thatcher, 2003; Blewitt et al.,

2009). To the north and east of the central Nevada seismic belt, geodetic measurements

(Hammond and Thatcher, 2007) and paleoseismic studies (Wesnousky et al., 2005)

suggest lower strain rates throughout the rest of the Great Basin. Beyond the central

Nevada seismic belt and Walker Lane/eastern California shear zone, the highest rates of

strain and seismicity are along the margins of the Great Basin in the vicinity of the major

normal fault systems bounding the Sierra Nevada and Wasatch Range (Dokka and Travis,

1990; Hammond and Thatcher, 2005).

Near the boundary between the Walker Lane and Basin and Range province,

historic earthquakes illustrate the transition from dextral shear to extension. For example,

the 1932 Cedar Mountain earthquake ruptured on a series of discontinuous fault traces

extending for 75 km (Figure 2.2) from Monte Cristo Valley in the southeast, through

Stewart Valley, and extending northwest through Gabbs Valley (Bell et al., 1999). As the

fault rupture propagated north, the style of faulting changed from right-slip in Monte

Cristo Valley to mostly dip-slip normal displacement in Gabbs Valley. Additional

examples of this style of dextral to normal faulting includes the 1954 Fairview Peak and

Dixie Valley earthquakes (Figure 2.2, Caskey and Wesnousky, 1997). Dominantly right-

oblique slip ruptures propagated north along the eastern range front of Fairview Peak,
15

followed by dip-slip faulting four minutes later and ~25 km northwest along the Dixie

Valley fault, which then propagated 46 km to the north (Caskey et al., 1996). These

earthquakes suggest a transfer of dextral shear in the Walker Lane into northwest-directed

extension in the Great Basin.

Figure 2.2. Historic earthquakes of central Nevada seismic belt in western Nevada. Historic fault ruptures
are displayed in red (USGS Quaternary fault and fold database, 2010), and the boundary of the Walker
Lane is blue. Outline of map boundary (blue polygon), with respect to the state of Nevada (black polygon),
is shown for spatial reference in upper right hand corner. Study area boundary is shown as black polygon in
map. DV, Dixie Valley; FP, Fairview Peak; GV, Gabbs Valley; GVR, Gabbs Valley Range; MCV, Monte
Cristo Valley; PV, Pleasant Valley; SV, Stewart Valley.
16

3. Geologic Setting of Gabbs Valley

Gabbs Valley is a relatively large intermontane basin in west-central Nevada

(Figure 3.0.1) approximately 40 km south of highway 50 and 40 km northeast of the town

of Hawthorne. The valley, as a whole, trends northwest, with drainages flowing to a

hydrographic-low in the northwest portion of the basin. The study area encompasses

~400 km2 in the southeastern part of the basin, including the northern extent of the Gabbs

Valley Range, southernmost Monte Cristo Range, Gabbs Mountain, and the southern part

of Cobble Questa (Figure 3.0.2). The Gabbs Valley Range trends northwest and bounds

the southern edge of the valley and study area (Figure 3.0.2). Mystery Ridge is the

northernmost extent of the Gabbs Valley Range and marks the western margin of the

study area. Gabbs Mountain is a lone peak (1900 m) ~ 4 km north of the Gabbs Valley

Range on the southeast margin of the study area. The area north of the Gabbs Valley

Range between Gabbs Mountain and Mystery Ridge is a large gently north-sloping

piedmont composed of alluvial fans and channels sourced from the surrounding high

topography. The lowest elevation within the study area lies north of the piedmont and is

characterized by a relatively flat basin locally interrupted by active drainage channels.

Cobble Cuesta is a northerly trending ridge that rises out of the basin on the northern

margin of the study area. The north-northeast-trending Monte Cristo Range and Fissure

Ridge divide the Gabbs basin into two discrete western and eastern lobes (Figure 3.0.2).
17

Figure 3.0.1. Regional overview map of west-central Nevada. Map shown on DEM (digital elevation
model) topographic hillshade. Major roads are orange lines. Quaternary faults are black lines. Blue polygon
is the study area boundary.
18

Figure 3.0.2. Area map for Gabbs Valley. Major physiographic features are labeled. The town of Gabbs
and the Diamond A Ranch are also indicated for reference.

3.1. Stratigraphic Framework

In ascending order, the major stratigraphic units in the southeastern Gabbs Valley

study area include Paleozoic-Mesozoic sedimentary and metasedimentary rocks,

Cretaceous granodiorite, Oligocene-Miocene ash-flow tuffs and andesitic-mafic lavas,

middle to late Miocene sedimentary rocks, and Pliocene-Quaternary sediments and


19

surficial deposits (Figure 3.1.1). The composite thickness of the Tertiary volcanic section

in the Gabbs Valley and Gillis Ranges is estimated to be ~2 km (Ekren et al., 1980). The

local thickness of Miocene-Quaternary basin-fill sediments varies between ~100 m to 1.5

km based on gravity modeling (Earney et al., 2018) and prior stratigraphic studies

(Stewart et al., 1999).

Basement rocks in Gabbs Valley comprise Triassic and Permian sedimentary,

metasedimentary, and volcanic rocks locally intruded by Cretaceous to Jurassic

granodiorite (Muller and Ferguson, 1939; Ekren et al., 1980; Ekren and Byers, 1985a,

1985b, 1986a). Mesozoic sedimentary and metasedimentary rocks (^vc) are exposed in

the northwest portion of the study area at Mystery Ridge and Fissure Ridge area of the

southernmost Monte Cristo Range. Metasedimentary rocks (^vc) contain volcaniclastic

facies within the Luning Formation (Muller and Ferguson, 1939) and consist of gray,

thin-bedded, laminated siltstones, claystones, sandstones, and pebbly conglomerates

(Ekren and Byers, 1985a). Jurassic to Cretaceous granodiorite (KJgd) crops out in the

northwest portion of the study area in Fissure Ridge. Granodiorite from the western

extent of Gabbs Valley yielded K/Ar ages of 158±4 Ma and 155±7 Ma (Ekren and Byers,

1984). Granodiorite exposures within the study area are likely similar in age to these

nearby plutons. The age of this granodiorite coincides with emplacement of much of the

Sierra Nevada batholith (Evernden and Kistler, 1970).


20

Figure 3.1.1. Simplified stratigraphic column of southeastern Gabbs Valley study area. Dates shown are
provided for the tuff of Gabbs Valley (Tgv; Henry and John, 2013), Hu-Pwi Rhyodacite and lavas of
Mount Ferguson (Ekren and Byers, 1980), and Esmeralda Formation (Stewart et al., 1999).

Oligocene-Miocene ash-flow tuffs and andesitic-basaltic lavas nonconformably

overlie Paleozoic-Mesozoic basement. The three main divisions of volcanic rocks are, in

ascending order: 1) tuff of Gabbs Valley, 2) Hu-Pwi Rhyodacite, and 3) lavas of Mount

Ferguson (Ekren and Byers, 1980, 1985a, 1985b, 1986a). The tuff of Gabbs Valley is a

sequence of Oligocene ash-flow tuffs exposed in Fissure Ridge of the Monte Cristo

Range with a K/Ar date of 25.15 ±0.06 Ma (Henry and John, 2013). The tuff of Gabbs

Valley is disconformably overlain by younger volcanic and sedimentary rocks. The


21

lowermost volcanic unit in the Gabbs Valley Range, south of exposures of the tuff of

Gabbs Valley, is the lavas of Pointsettia Mine, which are a propylitically altered series of

lavas of intermediate composition. The Hu-Pwi Rhyodacite is a >1 km thick group of

rhyodacitic tuffs that overlie the lavas of Poinsettia Mine and have established K/Ar dates

of 22.6-23.3 Ma (Ekren et al., 1980). The Hu-Pwi Rhyodacite is unconformably overlain

by the lavas of Mount Ferguson. The omission of a tuff unit in the stratigraphic sequence

(~100 m thick Copper Mountain Tuff) between the upper Pointsettia Tuff Member of the

Hu-Pwi section and lavas of Mount Ferguson was interpreted to indicate substantial

erosion in this interval (Ekren and Byers., 1980). The lavas of Mount Ferguson include a

sequence of multiple volcanic flows with a wide range in composition from basalt to

rhyolite. Intermediate lavas comprise much of the rocks exposed in the Gabbs Valley

Range and Gabbs Mountain and range in composition from trachyandesite to quartz latite

(Ekren and Byers, 1980). K/Ar ages range from 22.5±0.6 Ma on basal lava to 15.0±0.5

Ma on the youngest flows (Ekren and Byers., 1980).

Miocene volcanic rocks are nonconformably overlain by Miocene fluviolacustrine

sedimentary rocks of the Esmeralda formation and are locally intruded by Miocene

rhyolite and basalt dikes. The strata of the Esmeralda formation have a minimum

thickness of 850 m and consist of siltstones, sandstones, and conglomerates that have

been constrained between 12.96 to 9.24 Ma using tepherochronology (Stewart et al.,

1999). Pliocene-Quaternary alluvial fans and fluviolacustrine sediments overlie all

bedrock units.
22

Table 1. Simplified unit descriptions of southeastern Gabbs Valley


Label Unit Description Age Thickness
Qp Playa deposits High albedo surface composed of fine-grained sand, silt, and clay. Holocene <100 m
Qe Eolian deposits Sand sheets generally <50 cm in thickness. Holocene <3 m
Qa Active stream Active and recently abandoned stream channel deposits. Holocene <20 m
channel deposits
Qay Young basin Basin-fill alluvium composed of sand, silt, and clays. Light color in imagery, Holocene <30 m
alluvium smooth surface, and not typically offset by fault scarps.
Qao Old basin Basin-fill alluvium composed of sand, silt, and clays. Displays moderately dark Holocene to <100 m
deposits tone in imagery, smooth surface, and contains fault scarps. Pleistocene
Qfy Young alluvial Light-mid gray tone, smooth textured unit of active stream channels and Holocene <100 m
fans abandoned anastomosing gullies in young fan surfaces.
Qfi Intermediate Composed of pebble to cobble sized clasts forming a desert pavement. Holocene to <100 m
alluvial fans Moderately dark tone and rough texture, interfluves are broad and flat. Pleistocene
Qfo Old alluvial fans Light toned, smooth fan surface composed of a well-developed desert pavement Middle to late <100 m
of pebble to cobble sized clasts. Pleistocene
Tcg Cobble Conglomerate of rounded to subrounded cobbles, boulders, and pebbles. Pliocene- latest < 350 m
conglomerate Miocene
Tb Basalt Dark gray aphanitic basalt containing sparse phenocrysts of olivine and Miocene <10 m
clinopyroxene.
Te Esmerelda Tannish brown to yellowish brown, thin to thickly bedded fluviolacustrine Miocene- 9.24 < 1,500 m
formation sandstone and siltstone, with ubiquitous beds of tephra. to 12.96 Ma
Trl Rhyolite of Purple light gray flow-laminated rhyolite. Miocene- <10 m
Gabbs Valley 19.2±0.7 Ma
Tmi Mafic intrusions Dark gray aphanitic andesite and latite with hornblende phenocrysts <4 cm. Miocene
Tlf Lavas of Mount Sequence of many volcanic flows with a wide range in composition from basalt Miocene- 15 to 1000-2000
Ferguson, to ash-flow tuff, with the majority of flows compositionally intermediate. 22.5 Ma m
undivided
Tlfh Hornblende- Black and dark gray latite containing ≤2 cm long hornblende-pyroxene Miocene- 15 to 0-300 m
pyroxene latite phenocrysts in an aphanitic silicic groundmass. 22.5 Ma
Tlfa Andesitic lavas Brownish-red andesite, lower unit is relatively massive, whereas the upper Miocene- 15 to 0-500 m
portion contains excellent flow-foliations. 22.5 Ma
Tlfbx Volcanic breccia Boulder to pebble sized clasts composed of distinct red/oxidized vesiculated Miocene- 15 to 0-150 m
andesitic-basalt lavas. 22.5 Ma
Tlfp Hypersthene Dark gray basalt-andesite lavas containing few small phenocrysts of plagioclase, Miocene- 15 to 0-50 m
andesite orthopyroxene, and clinopyroxene. 22.5 Ma
Tlfl Flow-laminated Cliff-forming brownish-purple dense flow-laminated rhyolite or silicic quartz- Miocene- 15 to 0-30 m
quartz-latite latite with phenocrysts of plagioclase and biotite. 22.5 Ma
Tlfb Biotite rhyolite White and slope-forming nonwelded ash-flow tuff with biotite phenocrysts in an Miocene- 15 to 0-20 m
ashy matrix. 22.5 Ma
Tlfq Hornblende Distinct light gray unit with large hornblende phenocrysts 1-4 cm long. Miocene- 15 to 100-200 m
quartz latite 22.5 Ma
Tml Mafic lavas Greenish-gray to gray hornblende andesite. Miocene 0-75 m
Tcm Tuff of Copper Multiple flow, simple cooling unit of silicic quartz latite and rhyolite tuff. Miocene 0-80 m
Mountain
Tp Hu-Pwi Tan rhyodacitic tuffs with abundance of ferromagnesian minerals and Oligo-Miocene- <1,000 m
rhyodacite characteristically deficient in quartz. 22.6-23.3 Ma
Tpc Unit C Simple cooling unit of rhyodacite tuff. Oligo-Miocene- 0-150 m
22.6-23.3 Ma
Tpb Unit B Simple cooling unit of rhyodacite tuff that is typically lighter color than Oligo-Miocene- 0-70 m
underlying and overlying cooling units. 22.6-23.3 Ma
Tpa Unit A Simple cooling unit of rhyodacite tuff. Oligo-Miocene- 0-400 m
22.6-23.3 Ma
Tlp Lavas of Gray-purple propylitically altered series of lavas of intermediate composition. Oligo-Miocene 0-400 m
Pointsettia Mine
Tgv Tuff of Gabbs Three major cooling units of highly differentiated rhyolite tuff. Oligocene- <400 m
Valley, undiv. 25.15 Ma
Tgv3 Unit 3 Simple cooling unit of red, densely welded tuff. 25.15 Ma <30 m
Tgv2 Unit 2 Compound cooling unit of alternating moderately and densely welded pinkish- Oligocene- 30-300 m
gray and red devitrified tuff. 25.15 Ma
Tgv1 Unit 1 Simple cooling unit of red to reddish-gray, densely welded tuff with a thick (<30 Oligocene- 30-65 m
m) black basal vitrophyre. 25.15 Ma
KJgd Granodiorite Fine-medium-grained granodiorite containing abundant hornblende, biotite, Cretaceous-
clinopyroxene, and only interstitial quartz. 155-158 Ma
^vc Metasedimentary Metasedimentary volcaniclastic rocks consisting of gray, thin-bedded, well- Triassic
rocks graded laminated siltstones, claystones, sandstones, and pebbly conglomerates.
23

3.2. Structural framework


Gabbs Valley is an active tectonic basin with a complex Quaternary fault network

that lies on the northern margin of the central Walker Lane at the transition to the Basin

and Range extensional province (Figure 3.2.1). Directly southwest of Gabbs Valley, the

central Walker Lane is dissected by a series of four northwest-striking, left-stepping, en-

echelon dextral Quaternary faults. These strike slip faults, from southwest to northeast,

include the Indian Head, Gumdrop Hills, Benton Springs, and Petrified Springs faults

(Figure 3.2.1). The northern portion of Gabbs Valley contains several north-northeast-

striking Quaternary normal fault zones, including the southern Gabbs Valley, Hot

Springs, Monte Cristo, Phillips Wash, and Paradise Range fault zones (Figure 3.2.1). The

study area is positioned at the northern termination of the Petrified Springs fault and

southern termination of normal fault zones in Gabbs Valley. It thus occupies a broad zone

of multiple fault intersections.

The Petrified Springs fault is a 45 km long northwest-striking dextral fault that cuts

across the southern part of the study area (Figure 3.2.1) and ends proximal to Mystery

Ridge. Ekren and Byers (1984) suggested total right-lateral offset of ~12-16 km on the

Petrified Springs fault on the basis of correlating tuffs of Mount Ferguson across the

fault. Angster (2018) estimated a late-Pleistocene slip rate of 0.6 ± 0.1 mm/yr for the

Petrified Springs fault based on the age of offset Quaternary alluvial surfaces.
24

Figure 3.2.1 Simplified geologic map of Gabbs Valley (modified from Stewart and Carlson, 1978).
Quaternary faults (USGS, 2010) shown in black; border of study area shown in blue; and approximate
boundary between the Walker Lane and Basin and Range tectonic domains shown in yellow. Geologic map
overlain on DEM hillshade. SGSV, southern Gabbs Valley fault zone.

The Gabbs Valley basin contains numerous Holocene and historic (<150 yr) ruptures

along strands of the Monte Cristo, Hot Springs, Phillips Wash, and southern Gabbs

Valley fault zones. The 1932 Cedar Mountain earthquake (Ms 7.2) produced surface
25

ruptures on north-northeast-striking faults with normal dip-slip on faults in the Monte

Cristo, Phillips Wash, and southern Gabbs Valley fault zones (Figure 3.2.2, Bell et al.,

1999). The 1954 Fairview Peak earthquake (Ms 7.2) primarily occurred to the north of

the study area, but faults ruptured within Gabbs Valley, including the northern extent of

the Hot Springs fault and segments of the Monte Cristo and Phillips Wash fault zones

(Figure 3.2.2, Caskey et al., 1996). Surface ruptures associated with the 1932 and 1954

earthquakes overlap in the northern part of Cobble Cuesta (Bell et al., 1999; Caskey et

al., 1996). Payne (2013) and Stewart (1999) interpreted Cobble Cuesta as an extensional

anticline within an accommodation zone formed between oppositely dipping range-front

normal faults, including the east-dipping Monte Cristo fault zone on the western side of

the valley and west-dipping Paradise Range fault on the east side of the valley (Figure

3.2.2).

The bedrock structure consists of moderately tilted fault blocks with variable amounts

of Quaternary faulting that overprints preexisting structures. The Gabbs Valley Range

consists of northwest-trending, northeast-tilted (45-65º) fault blocks bound by early to

middle Miocene northwest-striking normal faults (Ekren and Byers; 1985a, 1985b,

1986a) that parallel the younger dextral faults in the central Walker Lane. Gabbs

Mountain has the same structural fabric as the Gabbs Valley Range, but many of the

older northwest-striking normal faults have been reactivated as dextral faults. Northwest-

striking normal faults observed in the Gabbs Valley Range and Gabbs Mountain likely

pre-date the onset of strike slip faulting and are consistent with a northeast-southwest

oriented extension direction in the early to middle Miocene (ca. <19-10 Ma) (Hardyman

and Oldow, 1991). These normal faults were well oriented to later accommodate strike-
26

slip motion when dextral-shear initiated in the Walker Lane in the late Miocene (ca. 11-9

Ma) (Hardyman and Oldow, 1991). In contrast to Gabbs Mountain and the Gabbs Valley

Range on the border of the Walker Lane, the Monte Cristo Range is the southernmost

extent of Basin and Range extension within the valley and is a north-trending, gently

east-tilted (25º) horst block bound by active normal faults on either side. Within the

Monte Cristo Range are a series of both north-northeast-striking normal and northwest-

striking oblique-slip faults. The onset of major extension along north-northeast-striking

normal faults in the region is constrained to 17-15 Ma (Dilles and Gans, 1995, Surpless et

al., 2002).

Figure 3.2.2. Quaternary faults of Gabbs Valley. Mapped quaternary faults (USGS, 2010) color coded
according to rupture age overlain on DEM hillshade. SGFZ, southern Gabbs Valley fault zone.
27

3.3 Geothermal Framework

The Gabbs Valley area has several known geothermal systems, including one hot

spring, two previously documented 2-m temperature anomalies, and an operational

geothermal power plant (Figure 3.3.1). Rawhide Hot Springs is a series of pools located

on the west side of Fissure Ridge in the western arm of the basin. The springs have a

flow rate of 0.8593 m3/min, and geothermometry suggest subsurface fluid temperatures

of 130-140ºC (Kratt et al., 2008). One 2-m thermal anomaly lies on the western edge of

the Monte Cristo Range in western Gabbs Valley (Figure 3.3.1, Kratt et al., 2008). The

second 2-m thermal anomaly is located east of Fissure Ridge in the eastern basin (Figure

3.3.1, Payne et al., 2011, 2013). These thermal anomalies have been interpreted as

thermal signals from upwelling zones of blind geothermal systems (Kratt et al., 2008;

Payne, 2013). The Don Campbell power plant, owned by Ormat Nevada Inc., lies in the

southwestern portion of the basin (Figure 3.3.1). The power plant is a binary system that

taps a ~130°C reservoir and has a 16 MW (net) output (capacity data at

http://nbmg.unr.edu/Geothermal). The discovery of this blind geothermal system was

initiated from the identification of a low magnetic anomaly collocated with hot wells

drilled for agricultural purposes in addition to high-temperatures in drill holes from a

mineral exploration program (Orenstein et al., 2015). A 2-m survey constrained the

location of the thermal anomaly (Kratt et al., 2010), and subsequent geothermal drilling

programs later characterized the system as an economic-grade resource (Orenstein et al.,

2015).

Prior to this study, no previous geothermal exploration had been conducted in the

southeastern portion of the basin. The following sections will address the exploration
28

methods used to identify and characterize a previously unknown blind geothermal system

in this area.

Figure 3.3.1. Known geothermal systems of Gabbs Valley. Don A. Campbell geothermal power plant (pink
dot), Rawhide Hot Springs (yellow ellipse), and previously documented 2-m temperature anomalies (red
ellipses) are shown. The location of the Diamond A Ranch is also indicated for reference.
29

4. Methods and Results

Methods employed to achieve geologic and geothermal exploration objectives

include: 1) geologic investigations, 2) geophysical surveys consisting of gravity,

magnetic, and magnetotelluric data, 3) cross-section construction, 4) slip and dilation

tendency analyses, 5) fluid geochemistry investigations, and 6) 2-m temperature survey.

Geologic mapping was used to define the geometry and kinematics of the local fault

systems, relative ages of faults, and identification of favorable structural settings.

Geophysical surveys were used to identify concealed faults, areas of hydrothermal

alteration, and conductive anomalies associated with geothermal activity. Detailed cross

sections were used to conceptualize geologic structure and contribute to the conceptual

model of possible geothermal systems in the area. Slip and dilation tendency analyses

were preformed to indicate faults that are more favorably oriented to transmit

hydrothermal fluids within the present stress field. Fluid geochemistry was used to

evaluate temperatures of potential geothermal reservoirs at depth. A 2-m temperature

survey was deployed to detect areas of shallow anomalous temperatures likely sourced

from geothermal upwellings.

4.1 Geologic Investigations

Geologic investigations were carried out across a study area encompassing ~400

km2 in southeastern Gabbs Valley and adjacent mountain ranges. Geologic investigations

included geologic mapping of Quaternary and bedrock units, analysis of Quaternary

faults, and structural analysis of bedrock units.


30

Detailed geologic mapping of approximately 180 km2 was completed at 1:24,000

scale. Reconnaissance scale mapping at ~1:48,000 was completed on the remaining ~220

km2. The map area includes Gabbs Mountain, southeastern Gabbs Valley, southern

Cobble Cuesta, and the northern extent of the Gabbs Valley Range. Previous mapping by

Ekren and Byers (1985a, 1985b, 1986a, 1986b), Stewart (1999), and Payne (2013) was

incorporated to generate a composite geologic map of the study area (Figure 4.1.1 and

Plate 1).

Figure 4.1.1. Simplified geologic map of the southeastern Gabbs Valley study area. See Table 1 and
Appendix A for unit descriptions and Plate 1 for detailed geologic map.
31

Field mapping was completed on 1:24,000 scale color stereo air photos and aided

by 160 km2 of LiDAR data. LiDAR coverage (Figure 4.1.2) was sourced from a previous

DOE-funded geothermal project (Payne, 2013), which covered the northern part of the

study area at 5.1 pulse per m2 (ppm) resolution, and an NSF-funded tectonics research

project (Angster, 2018), which covered a 1.5 km by 20 km strip along the Petrified

Springs fault at ≥8 ppm. LiDAR data were collected from aircraft-flown, laser point-

cloud measurements that produce extremely accurate topographic models. Topographic

models from LiDAR data were converted into hillshades to distinguish geologic features,

especially Quaternary fault scarps, which are commonly not identifiable in the field or on

air photos. Air photos were scanned and digitized utilizing Vr software (by Cardinal

Systems) that ortho-rectifies and georeferences the photos. Vr software generated a 3-D

topographic model of each stereo-pair upon which faults, contacts, and other features

were attributed. The Vr files were exported into ArcGIS 10.3 for assimilation into a

geodatabase and geologic map. Cross-sections were constructed from the geologic map

and geophysical data to display key geologic relationships and structures at depth in the

field area.

Because most relatively high temperature geothermal systems are associated with

Quaternary faults (Bell and Ramelli, 2007), Quaternary mapping was completed

throughout the entire study area in southeastern Gabbs Valley. Distinguishing alluvial

surfaces was based on observations indicative of the relative age of units, such as the

degree of weathering, stratigraphic succession, and cross-cutting relationships. Older

alluvial surfaces generally have greater amounts of incision, wider channels, rougher
32

texture, and lighter tone from deposition of fine-grained eolian-sourced sediments. In

contrast, younger surfaces commonly display little incision, smaller closely spaced

channels, smooth texture, and darker tone. Development of desert pavement (e.g., Wells

et al., 1995) and soils (Birkeland, 1999) is a function of increasing surface age and is an

important tool for surface correlation and relative age determination of alluvial surfaces.

Additionally, the presence of desert varnish on clasts can be another indicator of surface

age.

Figure 4.1.2. Extent of LiDAR coverage in the southeastern Gabbs Valley study area. LiDAR data shown
as gray hillshade with boundary of coverage indicated by red polygon. Base map is DEM hillshade.
33

Quaternary faults were identified by scarps in alluvial surfaces and linear

landforms (e.g. drainages). Greater scarp heights relative to increasing age of surfaces

signify a component of dip-slip displacement across a fault surface. Normal faults

commonly rupture in relatively sinuous traces produced from the expression of the

dipping surface across topography, changes in dip angle, or lateral fault steps. Strike-slip

faulting is characterized by linear fault traces, laterally offset drainages, up-hill facing

scarps, pop-up structures, sag-ponds, and shutter ridges. Uphill facing scarps, offset

drainages, and shutter ridges are all produced from the lateral motion across a strike-slip

fault. Pop-up structures and sag ponds result from variations in the strike of the strike-slip

fault, producing localized zones of transpression and transtension (Cunningham and

Mann, 2007).

Detailed Quaternary mapping was completed along the trace of the Petrified Springs

fault (Figure 4.1.3) in the southern part of the study area. LiDAR data and field

observations were employed to accurately map the trace of the fault, identify alluvial

surfaces of similar ages across the fault trace, estimate scarp heights, and distinguish

geomorphic features indicative of fault kinematics. In the vicinity of this fault,

Quaternary alluvial units generally overlie the Esmeralda Formation (Tes) on Quaternary-

Tertiary pediment surfaces (Plate 1). The primary feature of the Petrified Springs fault is

a sublinear northwest-trending, steeply inclined northeast-facing scarp cutting across

alluvial fans. The youngest alluvial fans and active channels (Qfy) are incised into the

scarp with no apparent offset across the trace. The height of the scarp increases relative to

older units, which display greater development of soil, desert pavement, desert varnish,

dissection by streams and gullies, and smoothing of surfaces compared to younger units.
34

The scarp heights observed in progressively older units range from approximately 2 m in

Qfi, 22 m in Qfo, and ≤50 m in Qfo2. These observations indicate a component of dip-

slip across the fault. Additionally, a 10 m2 travertine deposit was identified on the fault

scarp overlying all Quaternary deposits (Figure 4.1.3).

Figure 4.1.3. Quaternary geologic map of the Petrified Springs fault, southeastern Gabbs Valley. A.
Detailed Quaternary map with faults shown as black lines with annotations pointing to features along the
fault trace. Travertine deposit shown as small red polygon. B. Inset showing study area boundary with
extent of Quaternary map boundary (A) in red for spatial reference.

Geomorphic features along the trace of the fault indicate right-lateral

displacement in addition to the dip-slip component (Figure 4.1.3). Numerous active


35

streams along the fault trace are deflected in a dextral sense. One deflection in the north-

central part of the detailed map area displays shutter-ridge morphology and a series of

uphill facing scarps/knobs adjacent to the scarp on the downthrown side. A left-step of

the fault trace in the central part of the detailed map area has produced a “pop-up”

structure. Within this transpressional ridge, the otherwise gently northeast sloping Qfo

surface is tilted to a sub-horizontal attitude, and scarp height locally increases. Beheaded

and abandoned channels, in addition to a closed depression, are observed in the

southeastern portion of the detailed map area. An abandoned channel in the Qfi unit

displays over 90 m of right-lateral offset based on correlating surfaces bounding the

active and abandoned channels. Relative ages and distinguishing characteristics of

alluvial units established along the Petrified Springs fault were used as reference for

reconnaissance-scale mapping of surficial deposits throughout the study area by utilizing

field observations, available LiDAR data, and aerial imagery.

LiDAR data were also used to map the northern part of the study area at the

southern extent of Cobble Cuesta in the vicinity of the southern Gabbs Valley fault zone

(Figure 4.1.4). A series of previously undocumented fault scarps were identified in this

area, adding to some faults identified by Payne (2013) (Figure 4.1.4). Faults in this area

include multiple splays in a zone up to ~5 km wide, display vertical displacements ≤4 m,

range in length from 0.1-3 km, and cut basin sediment deposits (Qay and Qai) and

bedrock (Kcg, Te, and Tlf). Faults strike north-northeast and dip both east and west, with

the west-facing faults displaying greater vertical displacements and longer fault traces.

Traces of the normal faults are largely not observed in younger units (Qa, Qay, and Qfy)
36

bordering the fault zone, indicating possible ruptures <750 ka (Ramelli personal

communication, 2017).

Figure 4.1.4. Normal fault splays of the southern Gabbs Valley fault zone. A. Mapped faults are displayed
as black lines with direction of displacement indicated by fault ball. Faults are overlain on LiDAR
topographic hillshade. B. Inset showing study area boundary with extent of map (A) in red for spatial
reference.

Quaternary faults were attributed with age of most recent rupture (Figure 4.1.5A)

and estimated slip rate (Figure 4.1.5B). Fault age was assessed based on the youngest

offset unit (Ramelli personal communication, 2018). A slip rate of 0.6 ± 0.1 mm/yr was

inferred for the Petrified Springs fault based on work by Angster (2018). Significant
37

Figure 4.1.5. Quaternary faults in southeastern Gabbs Valley with attributed rupture recency and slip rate
values. A. Fault age (ka) shown as colored faults with warmer colors indicating more recent ruptures. B.
Fault slip rate (mm/yr) displayed as colored faults with warmer colors indicating higher slip rates.
38

Quaternary faults with no existing slip rate data were assigned slip rates between 0.001

and 0.01 mm/yr based on their location, size of scarp, age of offset surface (Ramelli,

personal communication, 2018), and previous analyses (Bell et al., 1999; Payne, 2013).

Geologic mapping and structural analysis were also conducted in bedrock units

surrounding southeastern Gabbs Valley to better characterize the structural and

stratigraphic framework of the area, particularly within the subsurface of Gabbs Valley.

The bedrock structure of the area is vital for identifying favorable structural settings that

may host blind geothermal systems. Bedding, flow-foliation, and compaction foliation

measurements were recorded to help constrain the attitude of bedrock units and tilt

orientation of fault blocks. A total of 110 attitudes were integrated into the geologic map

(Plate 1) from new mapping efforts and previous work (Ekren and Byers (1985a, 1985b,

1986a, 1986b; Stewart, 1999; Payne, 2013). Faults were identified by offset of

stratigraphic units, scarps, and fault surfaces. The attitudes of fault surfaces were

measured in addition to any kinematic indicators recording sense of slip. Kinematic

indicators include slickenlines, rough facets, and Riedel shears (e.g., Angelier et al.,

1985). Relatively few fault surfaces with useable kinematic indicators were observed in

this study.

Gabbs Mountain is well situated adjacent to the Petrified Springs fault to allow

for detailed analysis of bedrock structure affected by deformation at the transition from

strike-slip faulting of the Walker Lane to normal faulting of the Basin and Range.

Furthermore, bedrock structure in the Gabbs Mountain area is likely analogous to that

beneath southeastern Gabbs Valley and may therefore elucidate concealed structures in
39

the basin that control geothermal activity. Miocene lavas of Mount Ferguson comprise

the entirety of rock types exposed at Gabbs Mountain. Fault blocks typically strike west-

northwest and dip 25-65° to the northeast (Figure 4.1.6 and Figure 4.1.7A). The Miocene

volcanic strata in this area are cut by three distinct sets of faults: 1) northeast-striking

normal faults accommodating down-to-the-northwest offset; 2) northwest-striking strike-

slip or oblique-slip faults; and 3) northwest-striking normal faults accommodating down-

to-the-southwest offset (Figure 4.1.6). The northeast-striking normal faults accommodate

offsets ranging from 10-150 m. Volcanic flows of unit Tlfh on the summit and southern

extent of Gabbs Mountain have been offset right-laterally along steeply dipping,

northwest-striking dextral-normal faults. Unit Tlfh overlies in angular unconformity other

volcanic units of the area, exhibits a shallow dip (0-5°) in comparison to the moderate

dips (25-50°) of underlying units, and is interpreted to be the youngest unit at Gabbs

Mountain. Measured right-lateral separations of the lower contact of unit Tlfh on the

summit of Gabbs Mountain range from 20-60 m. Northwest-striking faults are observed

south of the summit and along the southwestern-most ridges, with a few outliers in the

northern ridges (Figure 4.1.6). Fault scarps range in height from 0.5-3 m and cut both

bedrock and commonly Quaternary colluvium and young alluvial fans (Qfy). These faults

are interpreted to have ruptured <750 ka based on offsets in colluvium (Ramelli personal

communication, 2018). The final style of faulting observed in this area involves

northwest-striking normal faults accommodating down-to-the-southwest displacement of

stratigraphic units. These normal faults lack recent scarps characteristic of the other two

sets of faults in the area, do not truncate the capping volcanic unit at the summit of Gabbs

Mountain (Tlfh), are distinguished by apparent southwest dip-slip offsets ranging from
40

100-500 m, and are interpreted to be the oldest faulting sequence in this area. These faults

are likely early to middle Miocene and may correlate with faults of similar geometry and

kinematics in the Stillwater Range (e.g., John, 1992; Hudson et al., 2000) and northern

Wassuk Range (Dilles and Gans, 1995).

Figure 4.1.6. Geologic map of Gabbs Mountain. A. Geologic map of Gabbs Mountain showing faults as
black lines, attitudes, and contacts. B. Inset showing location of map within study area in red.
41

Figure 4.1.7. Lower-hemisphere, equal-area stereographic projections of measured attitudes in rocks.


Density contour plots for poles to foliations (black dots), using a contour interval of 2 % per 1 % area. N is
number of measurements. Locations shown in Figure 3.0.2 and Plate 1. A. Poles to flow-foliations in rocks
comprising the lavas of Mount Ferguson (Tlf) in Gabbs Mountain and Gabbs Valley Range. B. Poles to
joints in Cretaceous intrusive rocks (KJgd) of Fissure Ridge. Poles to bedding (C) and joint (D) planes in
metasedimentary rocks comprising the Luning formation (^vc) at Mystery Ridge. Poles to bedding (E) and
joint (F) planes in metasedimentary rocks comprising the Luning formation (^vc) at Fissure Ridge.
Stereonet plots were generated using Stereonet 10.1.1. software (Allmendinger et al., 2012; Cardozo and
Allmendinger, 2013).

Within the Gabbs Valley Range, southwest of the Petrified Springs fault, a series

of fault blocks bound by northwest-striking, southwest-dipping normal faults are

composed of Tertiary volcanic rocks and are tilted 40-60º northeast. A major northwest-

striking normal fault bounds Mystery Ridge on the southwest and juxtaposes Tertiary

volcanic rocks on the south against Mesozoic metasedimentary rocks on the north (Plate
42

1). The Quaternary northwest-striking normal and dextral/normal faults observed in

Gabbs Valley and Gabbs Mountain are generally not observed in the Gabbs Valley

Range. An exception to the lack of strike-slip faulting in the Gabbs Valley Range

includes a lone northwest-striking strike-slip fault observed in volcanic rocks near the

fault contact with Mesozoic basement south of the summit of Mystery Ridge (Plate 1).

The Mesozoic basement of the area is a potential reservoir for geothermal systems

in southeastern Gabbs Valley, and thus understanding the orientation of joint and fracture

sets in these basement rocks is useful for conceptualizing permeable pathways that may

influence transmission of hydrothermal fluids in the area. Reconnaissance-scale mapping

and detailed fracture and joint analysis were conducted in Mesozoic basement exposures

at both Mystery Ridge and Fissure Ridge. Mystery Ridge exposes Triassic and Jurassic

metasedimentary basement rocks. Bedding in metasedimentary rocks generally strikes

northeast and dips moderately (20-40°) to the southeast (Figure 4.1.7C). Three sets of

joints were noted (Figure 4.1.7D): 1) north-northeast-striking, moderately (30-60°)

southeast-dipping, 2) northwest-striking, steeply (50-80°) southwest-dipping, and 3)

northwest-striking, steeply dipping (70-90°) to the northeast and southwest. Additional

analyses are needed to determine if the first two sets of presumed joints are associated

with foliation planes in these rocks.

Fissure Ridge consists of Triassic-Jurassic metasedimentary rocks of the Luning

Formation intruded by a Cretaceous granodiorite pluton. Intrusive rocks display a wide

range of fracture orientations, with at least two or more sets observed (Figure 4.1.7B): 1)

steeply dipping to vertical, east-striking joints; and 2) north-striking, moderately east-

dipping fractures. Metasedimentary rocks (^vc) at Fissure Ridge are folded into a gentle
43

syncline (Figure 4.1.7E). These rocks are also highly fractured but have no dominant

orientation of joints (Figure 4.1.7F). The most notable fracture set strikes northeast and

dips steeply. Exposures of metasedimentary basement at Mystery Ridge vary from those

at Fissure Ridge in several ways: 1) strata consistently dip to the east-southeast at

Mystery Ridge in comparison to the gentle fold indicated by west to north-northeast dips

at Fissure Ridge; and 2) foliation and joint sets are much more consistent and better

developed at Mystery Ridge versus those at Fissure Ridge.

4.2 Geophysical Surveys

Geophysical surveys were employed to characterize basin architecture, identify

concealed faults, and detect features associated with geothermal activity. Potential field

surveys included measuring gravity and magnetic fields. An MT survey measured

variations in subsurface resistivity. Mapping of intrabasinal structures involved a

multidisciplinary approach that integrated geologic mapping with these geophysical data,

particularly the gravity surveys. Geophysical surveys, methods, and data interpretation

included in this study area are also discussed in Earney et al. (2018) and Peacock et al.

(2018).

4.2.1 Gravity Survey

Detailed gravity surveys were employed to constrain basin architecture and

identify major concealed faults. Gravity anomalies caused by faulting have been

documented at several geothermal systems in the Great Basin (Glen et al., 2008, 2018;

Schwering et al., 2018). Identification and structural characterization of gravity

anomalies can be used to assess potential targets for blind geothermal systems within
44

basins.

Detailed gravity surveys in southeastern Gabbs Valley were completed by Zonge

International, Inc. and the U.S. Geological Survey. Zonge completed the initial gravity

survey in this study, collecting 274 stations during December, 2016 and March, 2017.

The U.S. Geological Survey subsequently collected 480 gravity stations in November,

2017 (Earney et al., 2018). Geodetic-grade GPS equipment was used to determine station

locations and elevations. The observed gravity is produced from both the position

(elevation and latitude) of the station and the variation of subsurface density. To remove

local effects of a measurement, a series of reductions were completed, including latitude

(Blakley, 1984), free air, Bouguer, curvature (LaFehr, 1991), and terrain corrections. A

density of 2.67 g/cc was used for reductions (Earney et al., 2018). Corrections were

conducted in Geosoft Montaj software.

The complete Bouguer Anomaly (CBA), isostatic gravity, and maximum

horizontal derivative (MHG) were completed. The complete Bouguer anomaly was

completed by combining the simple Bouguer anomaly with terrain corrections. Finally,

the regional isostatic gravity field was removed from the Bouguer anomaly using

established methods (Blakley and Simpson, 1986; Athens et al., 2014) to produce

isostatic gravity and residual isostatic gravity maps of the area. Isostatic gravity

anomalies are caused by lateral density variations in the middle to upper crust and can be

produced by variations in subsurface lithologies, sedimentary basins, and faults (Blakely

and Simpson, 1986; Athens et al., 2014). Modeling was completed by Zonge

International Inc. and the U.S. Geological Survey, with the USGS integrating the prior

survey by Zonge into the final gravity map (Figure 4.2.1).


45

The residual isostatic gravity values range from -4.7 mGal to 5.8 mGal and define a

series of relatively prominent northeast-trending gravity lows and highs, punctuated by

north to northwest-trending anomalies. Fissure Ridge and County Line Ridge are two

north-northeast-trending bedrock exposures that match the observed gravity highs sharing

the same orientation. Within the basin is a northeast oriented alignment of several gravity

lows, with the most significant low located northwest of County Line Ridge (point A in

Figure 4.2.1). An intrabasinal northeast-trending gravity high (point B in Figure 4.2.1)

occurs west and adjacent to the previously mentioned northeast-trending low. Volcanic

rocks exposed at the northwestern margin of the intrabasinal gravity high at the

southwestern edge of Cobble Cuesta suggest the presence of a basement high in this area.

Mystery Ridge and the Gabbs Valley Range are two northwest-trending bedrock

exposures associated with northwest-trending gravity highs. Gabbs Mountain shows

gravity highs in locations with bedrock exposure and also displays two north- and

northwest-trending gravity highs projecting into the basin. The westernmost gravity high

(point C in Figure 4.2.1) is bound by substantial gradients (~2.6 to -0.4 mGals) on its

northwest, southwest, and northeast margins. East of Gabbs Mountain is a gravity low

(point D in Figure 2.1.1) connected to a small isolated basin in the southeastern-most part

of the study area. Between Fissure Ridge and Mystery Ridge is a northwest-trending

gravity low (point E in Figure 4.2.1) associated with a small basin connecting the western

and eastern lobes of Gabbs Valley. West of Mystery Ridge and Fissure Ridge is another

north-northeast-trending gravity low (point F in Figure 4.2.1) that continues west of the

study area into western Gabbs Valley.


46

Figure 4.2.1. Residual isostatic gravity map of southeastern Gabbs Valley. Mapped faults shown as black
lines. Colored gravity plot overlain on DEM hillshade with warmer colors indicating relatively high gravity
values (mGals). Gravity stations are black dots. Gravity contours are shown as dark lines and with 0.5
mGal intervals. Gravity data collected and processed by the USGS and Zonge International, Inc.

Synthesis of geologic mapping efforts with the gravity data was used for

identifying and mapping concealed fault zones in the basin of southeastern Gabbs Valley

(Figure 4.2.1). Major gravity gradients in the area are likely produced by the
47

juxtaposition of stratigraphic units of differing density across faults, particularly where

low-density basin-fill sediments are in contact with higher-density volcanic and basement

lithologies. Generally, substantial gravity gradients within the basin were interpreted as

concealed faults if consistent with the established structural framework of the area as

identified in the geologic mapping. In many cases, major faults observed in the

surrounding ranges projected directly into major gravity gradients within the basin, thus

supporting the interpretation of many of these gradients as faults. Structural

interpretations of gravity data are described in detail in the Discussion, but the results of

these interpretations are incorporated into subsequent sections on cross sections and slip

and dilation tendency analyses.

4.2.2 Magnetic Survey

A ground magnetic survey was conducted to identify subsurface faults and areas

of hydrothermal alteration. Lateral variations in the magnetization of rocks produce

magnetic anomalies. The variations in magnetization may result from juxtaposition of

different rock types across faults and differences in the magnetic polarity and intensity of

volcanic rocks. Magnetic lows can be associated with geothermal activity due to the

alteration of rocks (e.g., Glen et al., 2008, 2018; ; Orenstein et al., 2015; Schwering et al.,

2018).

A total of 300 line-km of high resolution ground magnetic data were collected by

the U.S. Geological Survey on foot and by use of ATV-mounted systems using

Geometrics G858 and G859 cesium vapor magnetometers with integrated Global

Positioning Systems (GPS). Corrections were applied for variations in the local magnetic
48

field. Magnetic susceptibility values were measured from outcrop locations and hand

samples using a ZH instruments SM30 meter (resolution 1×10-7 SI units) for all major

Figure 4.2.2. Reduced to pole (RTP) magnetic map of the study area. The magnetic survey is overlain on
DEM hillshade. Warmer colors indicate relatively higher magnetic values (nT) than cooler colors. Letters
indicate location points referenced in text. Magnetic data collected and processed by USGS.

bedrock units in the area (Earney et al., 2018). Remnant magnetization properties

(magnitude, inclination, and declination) were obtained through previous paleomagnetic

surveys in the region (Carlson, 2018). The magnetic susceptibility and remnant
49

magnetization data were utilized for modeling of data acquired in the magnetic survey.

Southeastern Gabbs Valley contains several distinct magnetic anomalies.

Magnetic values range from -1319.3 nT to -565.3 nT. Magnetic highs are observed within

or proximal to volcanic bedrock exposures of Gabbs Mountain (points A and B in Figure

4.2.2). Magnetic highs also mark the Gabbs Valley Range in areas with volcanic strata,

intermediate-mafic intrusions, and basaltic lava flows (points C, D, and E in Figure

4.2.2). A magnetic low corresponds to Mesozoic metasedimentary rocks of Mystery

Ridge (point F in Figure 4.2.2). Moreover, an anomalous 2 km2 elliptically-shaped

magnetic low is observed in the south-central part of the basin to the west of Gabbs

Mountain and north of the Gabbs Valley Range (point G in Figure 4.2.2). This magnetic

low is surrounded by positive magnetic gradients generally ranging from -1121.0 nT to -

953.8 nT.

4.2.3 Magnetotelluric Survey

Magnetotelluric (MT) surveying is a passive geophysical method that measures

the electrical response of the Earth to natural time-varying magnetic fields (Vozoff, 1991)

and is very useful for imaging geothermal activity in the subsurface. The frequency

dependent ratio of the induced electric field and the inducing magnetic field, the

impedance tensor, provides both directional and depth information about subsurface

electrical resistivity structure. Inverting the impedance tensor in three-dimensions

provides a spatial representation of subsurface resistivity (Vozoff, 1991). Fluids, thermal

anomalies, and clay mineral assemblages associated with geothermal alteration can be

imaged in the subsurface as low-resistivity (conductive) anomalies (Ussher, 2000;


50

Cumming, 2009; Wannamaker et al., 2011, 2013, 2017). Several geothermal systems in

the Great Basin have documented conductive MT anomalies (Wannamaker et al., 2011,

2013).

Figure 4.2.3. MT resistivity depth profiles. Warmer colors indicate areas with higher measured
conductivity values (Ohm-m). Study area boundary shown in inset at top-center with extent of the MT
survey displayed as red polygon. MT station locations are black dots. A. Depth slice at 108-m. B. 199-m
depth slice. C. 392-m depth slice. D. 642-m depth slice. MT data collected and processed by USGS.

A total of 24 MT stations were collected by the U.S. Geological Survey (Figure

4.2.3) using ZEN data loggers, ANT-4 induction coils, and Borin Ag-AgCl

electrodes. Data were recorded for 20 hours and processed using the bounded influence
51

robust processing code developed by Chave and Thompson (2004), where synchronous

stations were used as remote references. The data were inverted using ModEM (Kelbert

et al., 2014) on NASA's supercomputer Pleiades.

The MT data indicate several low-resistivity anomalies in the study area (Figure

4.2.3). The largest low-resistivity anomaly has a semi-ellipsoidal geometry that is 10 km2,

trends northeast, and ranges from 100-1000 m in depth. This low-resistivity anomaly is

located in the southeastern part of the basin and is collocated with conspicuous gravity

gradients and the 2 km2 magnetic-low. Because the coverage and number of stations were

limited due to land status, there is some uncertainty in MT results. Methods and results

are further discussed in Peacock et al. (2018).

4.3 Cross Sections

Five detailed cross sections were constructed based on the integration of geologic

mapping with the geophysical surveys (Plate 1). The thickness of basin-fill sediments

(i.e., depth to top of the Tertiary volcanic section) was constrained by gravity forward

modeling completed by the U.S. Geological Survey (Earney, Schermerhorn, and Glen

personal communication, 2018). Lack of well and seismic reflection data within the study

area results in some uncertainty in the depth to volcanic rocks.

4.4 Slip and Dilation Tendency Analyses

Faults that are optimally oriented for slip and dilation are commonly the most

active conduits for fluid transmission (Ferrill et al., 1999; Ferrill and Morris, 2003). Slip

and dilation tendency analyses are useful methods for identifying faults with high
52

potential for transmitting geothermal fluids. Slip-tendency analysis assesses the potential

for faults to rupture in a known or inferred stress state for the area. Slip tendency (Ts) is a

function of the frictional resistance on a fault plane and the ratio of shear stress () to

normal stress (n) acting on that plane (Morris et al., 1996). This relationship is

represented by the equation:

Ts=/n

The value of slip tendency (Ts) is only dependent on the stress field and orientation of the

surface (Ferrill et al., 1999).

Dilation tendency analysis of fault surfaces can be completed on mapped faults

similar to slip tendency analysis based on a known stress field. Fracture dilation is mainly

controlled by the normal stress, which is in turn related to the acting fluid pressure and

the tectonic and lithostatic stresses. Dilation tendency (Td) for a fault or fracture surface

is defined as:

Td=(1-n)/(1-3)

Where 1 is the maximum principal stress, n is normal stress on the fault or fracture

plane, and 3 is the least principal stress.

Slip and dilation tendency analyses were conducted for the mapped fault system

in the study area (Figure 4.4.1) using a compilation of stress magnitudes from the western

Great Basin (Siler, personal communication 2018). Calculations used a minimum

principal stress direction (Shmin) oriented 124. Several assumptions were made for the

calculation, notably that all normal faults dip 60, and all strike-slip faults are vertical.

Slip and dilation tendency results showed that north-northeast-striking normal faults have
53

the highest slip and dilation tendency values, with a maximum along normal faults

striking N32E. Northwest-striking oblique-slip and strike-slip faults have high slip

tendency yet low dilation tendency. East-northeast-striking faults display moderate to low

slip and dilation tendency values.

Figure 4.4.1. Slip and dilatation tendency analyses for the southeastern Gabbs Valley fault system. Faults
are colored based on slip and dilation tendency values, with warmer colors indicating high slip and dilation
tendency.
54

4.5 Fluid Geochemistry Investigations

Water chemistry, water temperatures, and fluid geothermometry were evaluated

for 20 locations in this study. Out of the 20 water analyses, 11 are historical water

samples, 5 are new samples gleaned from gray literature (Payne, 2013; Delwich personal

communication, 2017), and 4 new samples were collected for this study at the Diamond

A Ranch in the northwest part of the study area (Figure 3.0.2). It is worth noting that

anomalously warm wells (32ºC) were identified in the study area at the Diamond A

Ranch. This was the first indication of a potential blind geothermal system in the area and

prompted more detailed reconnaissance and surveys.

Established methods were used for geothermometer calculations from cation and

silica concentrations in water samples (Fournier and Truesdell, 1973; Fournier, 1977a,

1977b, 1983; Fournier and Potter, 1979, 1982; Giggenbach, 1984; 1988). Samples with

charge balances ± 10% were used (Shevenell personal communciation, 2018). The

available geochemical data are predominantly from wells and springs in Gabbs Valley

and the surrounding area, with the four new analyses sourced from irrigation wells at the

Diamond-A Ranch. Several geothermometers for the Diamond A Ranch wells indicate

moderate to high subsurface fluid temperatures (Table 2). These geothermometers

included: 1) quartz conductive yielding temperatures between 123-136ºC, 2) quartz

adiabatic between 120-131ºC, and 3) several Na/K geothermometers with indicated

temperatures between 170-233ºC. A Piper diagram with plotted water samples show that

most samples are in the chloride-sulfide-bicarbonate facies (e.g., Schwartz and Zhang,

2003) (Figure 4.5.1). A GeoT multicomponent chemical equilibria model (Spycher et al.,

2016) was completed for the water samples collected at the Diamond A Ranch (Figure
55

4.5.2) and yielded equilibration temperatures of ~130-140ºC (Spycher personal

communication, 2017).

Figure 4.5.1 Piper diagram of water samples from southeastern Gabbs Valley. Water samples plotted using
different shapes and colors. Sample abbreviations and chemistry listed in Table 2. In the upper diamond
plot of the Piper diagram, water facies are characterized by Back (1961) and shown in blue, purple, green,
and yellow.

Figure 4.5.2. GeoT multicomponent chemical equilibria model results (Spycher, personal communication,
May 2017). The modeled minerals attain equilibria under the modeled conditions for Diamond A Ranch
Well 1 (22.4°C) at 133 ± 5°C, suggesting last temperature of equilibration near 133°C.
56

Table 2. Fluid geothermometry of Gabbs Valley


57

4.6 Shallow Temperature Survey

A 2-meter temperature survey was deployed to detect thermal anomalies associated with

zones of geothermal upwelling. Previous 2-m temperature surveys have been successful in

locating blind, relatively high-temperature geothermal systems that were subsequently developed

into geothermal power plants, including the Don A. Campbell plant in Gabbs Valley (Kratt et al.,

2010, Orenstein et al., 2014). Established methods were followed for the surveys in the study

area (Sladek et al., 2007; Coolbaugh et al., 2007; Sladek et al., 2009; Sladek et al., 2012; Sladek

and Coolbaugh, 2013). The shallow temperature survey was completed in three discrete time

intervals, with the initial collection of 61 stations in August, 2016, followed by two campaigns

totaling 65 stations in December, 2016. A total of 126 stations were collected in the study area

(Figure 4.6.1).

Data across the temporally-spaced surveys were influenced by seasonally driven changes

in ground temperatures. Base station measurements were therefore used as means to adjust the

surveys to a similar temperature threshold for identifying anomalous temperature zones.

Following seasonal correction, the average background temperature was calculated to be

14.94°C, and deviation from average background (DAB) was calculated for each station utilizing

methods outlined in Sladek and Coolbaugh (2013). Areas displaying DAB greater than 3°C are

considered statistically significant temperature anomalies (Sladek and Coolbaugh, 2013).

Normalized 2-m temperatures range from 13.3°C to 21.8°C (Figure 4.6.1). The 2-m

survey established a 7 km2 area with elevated temperatures in the southeastern part of the valley

(Figure 4.6.1) with the hottest station deviating from average background by 5.36°C, which is

well above the statistically significant temperature threshold. The thermal anomaly is elliptically

shaped with a northerly oriented long axis and highest measured temperatures in the north-
58

central part of the anomaly. Temperatures systematically decrease on all sides of the anomaly.

Background temperatures were detected throughout the rest of the basin with the exception of

one moderately anomalous warm point directly east of Mystery Ridge.

Figure 4.6.1. Shallow 2-m temperature survey in southeastern Gabbs Valley. Stations are shown as colored circles.
Colors correspond to measured temperatures and are attributed from colder temperatures displayed as blue to
warmest measurements as red.
59

5. Discussion

The purpose of this study was to evaluate the potential for high-temperature (≥130C) blind

geothermal system(s) in southeastern Gabbs Valley. The methods employed a quantitative

approach to exploration termed the geothermal play fairway analysis whereby geologic,

geophysical, and geochemical datasets were integrated to individually rank prospects to guide

assessment and potential future development. The sections below present a conceptual model of

the area based on bedrock structure and basin architecture that was used to identify favorable

structural settings. This model permitted selection of temperature-gradient holes, the results of

which, in turn, provided a means by which to evaluate the Nevada play fairway methodology.

The geothermal implications and direct evidence for the discovery of a high-temperature

geothermal system are thus presented.

5.1 Conceptual Structural Model

The geologic mapping, detailed gravity surveys, relatively high regional strain rates

(Kreemer et al., 2012), and numerous Quaternary faults with relatively high slip-rates (Angster,

2018; Payne, 2013; Bell et al., 1999) demonstrate that Gabbs Valley is a complex, tectonically

active basin. It is positioned at the southern extent of the central Nevada seismic belt. The

heightened seismicity and strain rates within this region probably result from the transfer of

dextral shear from the Walker Lane into northwest directed extension (Wesnousky et al., 2005;

Faulds et al., 2005; Faulds and Henry, 2008). The relatively abrupt transition from northwest-

striking faults and fault blocks in the southernmost part of the study area to the north-northeast-

trending structural grain to the north epitomizes this transfer of dextral shear into northwest-

directed extension. Integration of previous geologic mapping (Ekren and Byer, 1985a, 1985b,
60

1986a, 1986b; Stewart, 1999; Payne, 2013) and new mapping paired with detailed gravity data

provide insights into the structural framework of the area. The geologic map and cross-sections

from this study (Plate 1) illustrate the spatial-temporal relationships and kinematics of fault

systems in the area.

Analysis of the Quaternary fault network shows how active deformation and strain is

distributed in southeastern Gabbs Valley. The Petrified Springs fault is the most significant

Quaternary fault with a late Pleistocene slip rate of 0.6 mm/yr (Angster, 2018). It has a distinct

linear northwest-trending fault trace with strike-slip geomorphic features and vertical

displacements (50 m) indicating oblique-dextral displacement. The surficial trace of the

Petrified Springs fault dies out in the northwestern portion of the study area near Mystery Ridge,

indicating that the fault is losing total displacement along strike to the northwest and terminating

in or near the northwestern edge of the study area (Plate 1). North-northeast-striking Quaternary

normal faults observed in the northern portion of the area between Fissure Ridge and County

Line Ridge, through the southern extent of Cobble Cuesta (Plate 1 and cross-section D-D’), are

inferred to project south-southwest through the basin to intersect the Petrified Springs fault. The

apparent northwestward termination of the Petrified Springs fault suggests that dextral shear

strain is transferred to northwest-directed extension on the multiple splays of north-northeast-

striking normal faults. The termination of a major strike-slip fault of the central Walker Lane in

an array of normal faults implies that southeastern Gabbs Valley contains a displacement transfer

zone (Figure 5.1.1), which is one of the key favorable structural settings for hosting geothermal

systems in the region (Faulds et al., 2011; Faulds and Hinz, 2015).
61

Figure 5.1.1. Southeastern Gabbs Valley displacement transfer zone. Structural domains of the area are shaded to
illustrate the spatial distributions of faulting: Walker Lane dextral shear in red, displacement transfer zone in orange,
and Basin and Range extension shown in yellow. The trace of the Petrified Springs fault is shown as a red line.
Favorable structural settings shown as orange-red dashed polygons.

The area in and surrounding the displacement transfer zone is structurally complicated

with many faults of varying geometries, kinematics, and ages. The three main styles of faults in

this area include: 1) late Miocene-Quaternary northwest-striking oblique-slip and dextral faults,
62

2) late Miocene-Quaternary north-northeast-striking normal faults, and 3) late Oligocene to

middle Miocene northwest-striking normal faults. The late Oligocene to middle Miocene

northwest-striking normal faults are the primary set of faults observed in the Gabbs Valley

Range, signifying that this fault block has been minimally affected by more recent Walker Lane

deformation and is currently translating as a relatively coherent block between two major dextral

faults of the central Walker Lane, the Petrified Springs and Benton Springs faults. The relative

lack of Quaternary faulting in the Gabbs Valley Range indicates that the transfer of dextral shear

strain is primarily distributed on the northwest side of the Petrified Springs fault.

On the northwest side of the Petrified Springs fault, Gabbs Mountain contains a large

exposure of Tertiary bedrock that illustrates the complex fault structure within the displacement

transfer zone. Gabbs Mountain displays all three sets of faults observed in the study area in a

dense array of faulting. The late Oligocene to middle Miocene northwest-striking normal faults

cut volcanic stratigraphy throughout Gabbs Mountain, similar to the Gabbs Valley Range.

However, multiple younger faults overprint the northeast-tilted fault blocks at Gabbs Mountain.

The type, density, and interaction of recent fault sets at Gabbs Mountain varies based on the

general location and distance from the Petrified Springs fault (Figure 5.1.2, Plate 1 and cross-

section C-C’). Northwest-striking dextral and oblique-slip faults dominate the area south of the

summit of Gabbs Mountain proximal to the Petrified Springs fault. In contrast, the ridges north

and northwest of the summit exhibit primarily north-northeast-striking normal faults. The

overlap of normal and dextral/oblique-slip fault zones are best observed west of the summit in an

area with numerous intersections between these two sets of faults (Figure 5.1.2).
63

Figure 5.1.2. Geologic map and cross section of Gabbs Mountain with highlighted fault intersections. A. Geologic
map with major fault intersections between north-northeast-striking normal faults and northwest-striking oblique-
slip faults are circled in red. Location of C-C’ shown on map. B. Inset showing location of geologic map as red
rectangle relative to study area. C. Portion of cross section C-C’. Fault intersection circled in red.
64

As evident in the residual isostatic gravity (Figure 4.2.1) and geologic mapping data

(Plate 1), southeast Gabbs Valley is a complex composite basin composed of several sub-basins

that are controlled by the tectonic domains of the area. The northern portion of study area,

extending from western Gabbs Valley to Stewart Valley (Plate 1, cross section D-D’), is

composed of a series of north-northeast-trending half-grabens, grabens, and horsts. Fissure Ridge

and County Line Ridge are both horst blocks, bound by oppositely dipping normal faults. The

basin between these two horst blocks is primarily composed of a series of east-tilted half-grabens

generated by displacement along west-northwest-dipping, north-northeast-striking normal faults

that are locally punctuated by secondary, east-dipping antithetic faults (Plate 1, cross section D-

D’). The east-tilting of these sub-basins is inferred from the east tilts observed in both Fissure

and County Line Ridge, as well as the dominant set of west-dipping normal faults. The most

substantial sub-basin in this area is located northwest of County Line Ridge (point A in Figure

4.2.1) and is produced from normal displacement along east-northeast-striking, north-northwest-

dipping faults. West of this sub-basin is a northeast-trending intrabasinal high (point B in Figure

4.2.1) that may be a horst or half-graben and could be a concealed expression of the southern

extent of the Cobble Cuesta anticline. Between the intrabasinal high and Fissure Ridge is another

graben and sub-basin (Figure 4.2.1) bound on either side by oppositely-dipping normal faults.

The basin architecture of the southern portion of southeastern Gabbs Valley contrasts

with that in the north, as northwest-trending sub-basins dominate and are primarily controlled by

transtensional faulting. Basins in this area are generally located between dextral-oblique faults,

and changes in fault geometry, in addition to interaction with normal faults, has influenced sub-

basin development. For example, an inferred right-step of the Petrified Springs fault between

Mystery Ridge and Fissure Ridge transfers dextral shear strain across a series of north-northeast-
65

striking normal faults. Displacement along these normal faults generated a relatively shallow

northwest-trending basin that connects the two larger basins of western and eastern Gabbs Valley

(Figure 5.2.1). On the western and eastern margins of this right-step are two pull-apart zones that

concentrate strain along the bounding normal faults and have produced two discrete deeper sub-

basins (Figure 5.2.1, points D and E). The area between the shallow northwest-trending

transtensional sub-basin within the right-step and the deep northeast-trending extensional sub-

basin northwest of County Line Ridge is a transition zone, where the basin architecture gradually

changes in orientation and structure between the Walker Lane and Basin and Range province.

The structural framework of Gabbs Mountain is probably a good analogue for concealed

structures in the basin that may control geothermal activity. Within the basin directly west of

Gabbs Mountain is a substantial northwest-trending gravity high produced from offsets along

concealed faults. Tightly spaced gravity stations at 100-250 m within this area allow for high-

resolution control on subsurface fault locations. This northwest-trending gravity high ends

abruptly to the northwest along a northeast-trending gravity gradient in the central part of the

basin. This gravity high is probably bound by northwest-striking dextral-normal faults, linked to

major faults in Gabbs Mountain along its southeastern and northeastern margins (Plate 1 and

cross-section B-B’), and northwest-dipping normal faults projecting from Cobble Questa and

County Line Ridge on its northwest margin (cross-section E-E’). The northwest-striking oblique-

slip faults are interpreted to intersect splays of northeast- to east-northeast-striking normal faults

within a dense array of fault intersections near the northwest margin of the gravity high.

5.2 Geothermal Prospects

The geologic and geophysical data indicate at least six individual, favorable structural
66

settings for geothermal activity within the broader displacement transfer zone in southeastern

Gabbs Valley (Figure 5.1.1). These include fault intersections of varying complexity, apparent

pull-aparts and releasing bends along the Petrified Springs dextral fault, and possible

terminations of normal fault zones (Figure 5.2.1). The multitude of favorable settings within one

moderately sized basin attests to both the structural complexity and vast geothermal potential of

the Great Basin region.

Figure 5.2.1. Favorable structural settings of southeastern Gabbs Valley. Residual isostatic gravity map of the area
with warmer colors indicating higher relative gravity values. Brown lines are 0.5 mGal gravity contours. Favorable
structural settings shown as red dashed polygons with letters corresponding to locations referenced in the text.
67

The area of tightly spaced faults and fault intersections in the south-central part of

southeastern Gabbs Valley between the north-northeast-striking southern Gabbs Valley fault

zone and northwest-striking oblique-slip faults emanating from Gabbs Mountain define a

particularly favorable structural setting (point A in figure 5.2.1). Multiple lines of direct and

indirect evidence suggest a relatively high-temperature (130C) blind geothermal system in this

area, including collocated intersecting gravity gradients, magnetic-low, low-resistivity, and 2-m

temperature anomaly (Figure 5.2.2). The intersecting gravity gradients, in particular, suggest a

complex fault intersection (Figure 5.2.2A). Numerous north-northeast striking faults intersecting

northwest-striking faults, as mapped from gravity data, are favorably oriented to both slip and

dilate in the current stress field (Figure 5.2.2D). The 2 km2 magnetic-low (Figure 5.2.2B) is

likely produced from hydrothermal alteration of magnetic minerals in the volcanic stratigraphy.

The ~10 km2 low-resistivity anomaly (Figure 5.2.2C) is inferred to result from argillic alteration

in the system. The geometry of the resistivity anomaly is elliptically shaped and mostly

distributed between depths of 250-750 m in volcanic rocks near the contact with basin-fill

sediments. The shape of the low-resistivity anomaly is consistent with the model of an argillic

lithocap straddling the geothermal system, similar to that documented in other geothermal

systems (Pellerin et al., 1996; Newman et al., 2008; Munoz, 2014). The shallow temperature

survey defined a discrete 7 km2 anomaly (Figure 5.2.2) within this area with temperatures

elevated as much as 5.36°C above average background, which is well above the statistically

significant temperature threshold used to classify shallow thermal anomalies as potentially linked

to zones of hydrothermal upwelling (Sladeck and Coolbaugh, 2013). The 2-m temperature data

illustrate the surficial extent of the thermal anomaly. Anomalous 2-m temperatures
68

Figure 5.2.2. Collocated intersecting gravity gradients, magnetic-low, low-resistivity, 2-m temperature anomaly,
and favorable oriented faults for slip and dilation. Faults are shown as black lines (except in D). 2-m temperature
stations are shown as colored circles with warmer colors indicating higher measured temperatures. Favorable
structural settings shown as red dashed polygons. All maps are overlain on DEM hillshade. A. Residual isostatic
gravity map of the area with warmer colors indicating higher relative gravity values. Brown lines are 0.5 mGal
gravity contours. B. Reduced to pole (RTP) magnetic map of study area. Warmer colors indicate relatively higher
magnetic values (nT) than cooler colors. C. 199 m MT resistivity depth profile. Warmer colors indicate areas with
lower measured resistivity values (Ohm-m). D. Slip and dilatation tendency analyses. Faults are colored based on
slip and dilation tendency values with warmer colors indicating high slip and dilation tendency. E. Map showing
area of datasets as red polygon.
69

systematically decline to the north and eventually drop to background temperatures in the basin

at the inferred near-surface elevation of the water table (Figure 5.4.2). Thus, the 2-m survey

appears to detect the thermal signature produced from a zone of hydrothermal upflow/outflow of

a concealed geothermal system and essentially maps the flow path to the north into the central

part of the basin, where the thermal signature is masked by cool groundwater.

Interestingly, water samples collected from agricultural wells at the Diamond A Ranch

~7 km northwest of the 2-m anomaly (Figure 5.2.3) were anomalously warm (32°C) and

produced geothermometers indicating subsurface fluid temperatures of 130-140°C (Table 2 and

Figure 4.5.2). However, due to the proximity of several other favorable structural settings (points

B, D, and F in Figure 5.2.3), including the 2-m thermal anomaly documented by Payne (2013)

near Cobble Cuesta (point F in Figure 5.2.3), the warm water and promising geothermometers

from these wells are not clearly related to the geothermal system encountered by the 2-m survey

in the south-central part of the study area. Thus, geothermometry from the water samples at

Diamond A Ranch is not considered a definitive temperature estimate for the inferred geothermal

system in the south-central part of the basin. It is notable, however, that the geothermal systems

in western Gabbs Valley, including Rawhide Hot Springs and the system powering the Don A.

Campbell geothermal plant, are not likely sources based on fluid chemistry, structural barriers to

flow, and location down hydrologic gradient from the ranch. Thus, geothermometry results

indicate the existence of a high-temperature (≥130°C) geothermal system in eastern Gabbs

Valley.
70

Figure 5.2.3. Geothermometor sample locations in relation to favorable structural settings and 2-m temperature
anomaly. Temperature data overlain on residual isostatic gravity map of the area with 0.5 mGal contours. Wells at
the Diamond A Ranch, where water samples were collected and used for geothermometry, are shown as yellow
squares. 2-m temperature stations are shown as colored circles with warmer colors indicating higher measured
temperatures. Favorable structural settings shown as red dashed polygons.

Several other favorable structural settings were identified within the study area (Figure

5.2.3). Two additional fault intersections were noted: 1) near the eastern edge of Mystery Ridge,

where a northeast-striking normal fault inferred from gravity data intersects the Petrified Springs
71

fault within a small releasing bend along the Petrified Springs fault (point B in Figure 5.2.3), and

2) along the northwestern side of County Line Ridge, where multiple north-striking faults

intersect east-northeast-striking faults, as evidenced by mapped faults, including Quaternary

scarps (e.g., 1932 Cedar Mountain earthquake), and gravity gradients (point C, Figure 5.2.3). On

the west side of the map area, the inferred right-step in the Petrified Springs fault between

Mystery and Fissure Ridges (Figure 5.2.3) displays two discrete gravity lows on either side of a

gravity saddle that may correspond to transtensional pull apart zones, where faulting and

extension are concentrated (points D and E in Figure 5.2.3). This inferred right-step implies that

the Petrified Springs fault does not completely terminate within the study area and that some

dextral strain extends northwestward into western Gabbs Valley (Plate 1).

The final favorable structural setting is located in the northern part of the study area

directly west of Cobble Cuesta (point F in Figure 5.2.3), where an anomalous gravity high within

the basin lies adjacent to a previously documented 2.5 km2 thermal anomaly possibly associated

with a zone of geothermal upwelling (Payne, 2013). Payne (2013) proposed three potential

structural settings for this inferred upwelling: 1) fault termination of the Philips Wash normal

fault zone (e.g., ruptures associated with the 1954 Fairview Peak earthquake, Figure 3.2.1), 2)

nested north-northeast-striking normal faults, or 3) fault intersection of a concealed northwest-

striking strike-slip fault and north-striking normal faults. New gravity data collected in this study

show a west-northwest-trending gravity gradient in this area, which supports either the concealed

fault intersection setting or hybrid setting involving the southward terminating Phillips Wash

fault zone amid multiple intersections with the concealed northwest-striking fault. Further work

is needed to test these hypotheses.


72

5.3 Play Fairway Analysis

The Nevada play fairway methodology is a probabilistic estimate for finding a relatively

high-temperature geothermal system (130ºC) based on integrating geologic, geophysical,

geochemical, and temperature data sets to show areas that are most favorable for further

exploration efforts. All datasets acquired in the detailed study have been synthesized to perform

a play fairway statistical analysis to quantitatively estimate targets in the area most favorable for

temperature-gradient drilling and further exploration. The general methodologies for producing

regional predictive maps in the initial phase of the project (Faulds et al., 2015b) were followed in

building detailed predictive maps for southeastern Gabbs Valley at the completion of this study.

Two main sets of predictive maps were generated: 1) play fairway maps (Figure 5.3.1), and 2)

play fairway error maps (Figure 5.3.2).

Methodologies were modified from the initial to final phases of the play fairway project

to include new geologic mapping, geophysical data, geochemical data, and temperature data.

The permeability model for the area was improved by attributing Quaternary faults within the

area with slip and dilation tendency values, slip rates, and estimated most recent rupture age.

Additional modifications included incorporation of: 1) a structural settings quality factor used to

model the strength or quality of individual structural settings; and 2) 2-m temperature anomalies

utilizing established methods of assessing degrees above background and potential errors (e.g.,

Sladek and Coolbaugh, 2013). A probability of occurrence of a 130°C geothermal system was

assigned to the 2-m temperature anomaly as follows: a DAB of <2°C = 0 probability, 2-3°C =

0.15 probability, 3-4°C = 0.25 probability, 4-5°C=0.40 probability, and 5-6°C = 0.45 probability.

New data acquired in this study significantly enhanced the play fairway analysis from the

initial phase (Figure 5.3.1A) to Phase III (Figure 5.3.1C). The hybrid structural setting of the
73

densely spaced faults and fault intersections within the displacement transfer zone (point A in

Figure 5.3.1C) in the south-central part of the study area is the area most likely to host a

geothermal system based on this analysis. The favorable settings are ranked and scored in the

following order: 1) zone of multiple fault intersections in the south-central part of the area (point

A in Figure 5.3.1C) with a score of 7.2, 2) the eastern pull-apart zone of the Petrified Springs

fault adjoining the Diamond A Ranch (point D) with a score of 5.44, 3) the western pull-apart

zone of the Petrified Springs fault (point E) and fault intersection near the southern edge of

Mystery Ridge (point B) tied with a score of 4.8, 4) multiple fault intersections (point C)

bordering the northwest edge of County Line Ridge with a score of 2.34, and 5) the fault

termination or fault intersection near Cobble Questa (point F) with a score of 1.335. This

analysis indicates two structural settings with scores greater than 5, which signifies potential for

these settings to host a ≥130ºC blind geothermal system.

Although the play fairway scores have similar overall values in both the initial and final

phases of the project, the final play fairway analysis has defined locations of higher favorability

in much greater detail (Figure 5.3.1). Figure 5.3.1D illustrates the difference between the initial

and final phases whereby newly identified areas of permeability are shown in red compared to

areas that have a reduced permeability score in blue. Fairway error analysis maps (Figure 5.3.2)

display areas that have a decrease in error as positive numbers, whereas areas that have an

increase in error correspond to negative numbers. Error analysis results indicate a decrease in

error throughout the study area from the initial to final phases of the play fairway analysis, which

is attributed to the input of new data. The primary conclusion from this analysis is that the final

fairway map has reduced the exploration risk and improved targeting of blind geothermal

prospects in southeastern Gabbs Valley.


74

Figure 5.3.1. Phase I-III play fairway maps. Play fairway raster images overlain on DEM hillshade. A. Phase I
fairway map with initial structural settings and Quaternary faults (USGS, 2010). B. Phase II fairway map with newly
mapped faults and revised structural settings. Extent of phase II play fairway map covers the prior map boundary. C.
Phase III play fairway map with updated fault map and revised structural settings displaying local permeability
score. D. Difference between phase III and phase I play fairway maps. Blue colors indicate areas that have lower
fairway favorability scores after phase III analysis, and red colors are areas that have received higher fairway
favorability scores after phase III analysis.
75

Figure 5.3.2. Phase I-III fairway error maps. Error maps are overlain on DEM hillshade and show locations of
Quaternary faults and revised structural settings. A. Phase I fairway error map. Quaternary faults mapped by USGS
(2010). B. Phase II fairway error map. C. Phase III fairway error map.
76

5.4 Temperature-Gradient Drilling

On the basis of the high play fairway score and collocated 2-m temperature anomaly,

intersecting gravity gradients, magnetic-low, and low-resistivity, the complex fault intersection

within the south-central part of the study area (point A on Figure 5.2.3) was chosen for

temperature-gradient drilling in this study. Specific drill targets were selected to span and

straddle the collocated features to better define the length and breadth of the potential geothermal

system.

Six temperature-gradient holes were drilled by the U.S. Geological Survey research

drilling unit from late May to late June 2018 to a depth of ~152 m (500 ft) each. A depth of 152

m was chosen based on regional observations (Benoit personal communication, 2018) suggesting

that this is a sufficient depth interval to detect concealed geothermal activity. The budget only

allowed for six drill holes. A sealed narrow conductive casing was placed in the wellbore and

cemented in place with geothermal grout. The sealed casing was filled with water and allowed to

equilibrate with the surrounding rock temperature, and subsequent surveys tracked temperature

equilibration. Five temperature surveys were completed to track temperature equilibration in the

holes from June 9, 2018 to September 28, 2018 to ensure equilibrated temperature-gradient

profiles of the holes. Surveys involved lowering a thermocouple on a wire-line down the sealed

casing and measuring temperatures in 10 m intervals. All surveys were found to have an

instrument drift/error less than 2ºC.

The six temperature-gradient holes were distributed roughly across the extent of the

shallow-temperature and geophysical anomalies, including the full north-south extent of the

anomalous area and partially across the east-west extent (Figure 5.3.1B). Wells 31-2 and 52-2

contained high temperatures exceeding boiling with bottom-hole temperatures of 114.5°C and
77

124.9°C, respectively. The remaining wells displayed moderate to background temperatures

ranging from 79.2°C to 28.7°C. The drill data indicate that the thermal anomaly has a ~2 km

north-south length and a ~1 km east-west breadth. Drill cuttings are homogeneous across all

wells and are composed of basin-fill sediments of sand, silt, and gravels primarily composed of

Tertiary volcanic clasts sourced from the surrounding mountains. Tertiary and Mesozoic bedrock

was not intersected in any of these drill holes.

Analysis of the temperature-gradient drilling data yields insights into the distribution and

source of subsurface thermal anomalies and implications for the interpretation of other datasets.

Five of the six wells yielded temperature-gradients significantly higher than the average for the

Great Basin region. Holes 65-33 and 87-3 display linear trends of temperature increasing with

depth (Figure 5.3.1A) consistent with a geothermal gradient produced solely by conductive

heating. The calculated temperature gradient for hole 65-33 is 56.7ºC/km and considered within

the range of the average Great Basin geothermal gradient of 25-70ºC/km (Blackwell and

Richards, 2004; Coolbaugh, 2005). Hole 87-3 has a geothermal gradient of 190.8ºC/km, which is

highly elevated. Holes 85-3 and 41-1 exhibit even more highly elevated geothermal gradients

(Figure 5.3.1A). The calculated gradient for 85-3 is 400ºC/km, which is considered an extreme

gradient and indicative of a localized heat-source proximal to the hole. Similar to holes 65-33

and 87-3, 85-3 has a linear conduction temperature gradient, whereas hole 41-1 has a gently

curved temperature/depth gradient. The non-linear temperature/depth gradient of 41-1 indicates

that this hole is not solely influenced by conduction and possibly received some secondary

thermal input between 30-100 m from convective fluids. Holes that do not exhibit conduction

were not chosen for temperature gradient calculations based on the uncertainty in estimating the

most representative gradient per hole.


78

Figure 5.4.1. Temperature-gradient data and drill-hole locations for the southeastern Gabbs Valley blind geothermal
system. Temperature-gradients were measured following temperature equilibration of the holes from the latest
survey taken on September 28th, 2018. A. Temperature vs. depth graphs with well names and interpretations. For
wells displaying conduction, the calculated linear gradient is extended to 1 km to yield a °C/km estimation, which is
listed below the well name. B. Location of temperature-gradient holes (colored triangles) with respect to the
favorable structural setting (red dashed ellipse) and faults (black lines). Warmer colored triangles indicate higher
bottom-hole temperatures at 152 m.
79

The temperature-gradient profiles for holes 31-2 and 52-2 differ significantly from the

profiles produced from the other four drilled holes (Figure 5.3.1A). These differences include:

1) significantly higher geothermal gradients and bottom-hole temperatures, 2) concordant

temperature reversals in both holes between 60-80 m depth, and 3) relatively isothermal

temperatures from 90 m to the bottom of the holes at 152 m. The observed temperature-gradients

for 31-2 and 52-2 necessitate drill-hole intercepts of convecting hydrothermal fluids to produce

the distinct temperature reversals and isothermal anomalies, which cannot be produced purely

from a conductive heat source. The matching isothermal interval between 60-80 m depth

strongly indicates the presence of hydrothermal fluids, which is likely a zone of lateral outflow in

the basin sediments. The 90-150 m isothermal interval also indicates hydrothermal fluids, which

may be in a zone of upwelling or lateral outflow. The relatively shallow depths of these holes do

not capture the full extent of the deeper isothermal thermal anomaly, and thus several questions

remain. Notably, temperature-gradient programs targeting established geothermal systems in

nearby areas, including Dixie Valley (Blackwell et al., 2000), Salt Wells (Edmiston and Benoit,

1985; Hinz et al., 2014), Hawthorne (Blake et al., 2017), and western Gabbs Valley (Orenstein et

al., 2015), display similar temperatuer/depth relationships to the blind geothermal system in

southeastern Gabbs Valley. Nonetheless, deeper drilling is needed to extend the temperature

profiles to resolve whether these holes intersected a zone of geothermal upflow (temperature

increases and/or remains constant with depth) or lateral outflow (decrease in temperature with

depth).

It is noteworthy that the 2-m survey captured a broad area with elevated temperatures that

are spatially and thermally consistent with the temperature-gradient holes (Figure 5.4.2). For

example, the hottest 2-m stations are clustered ~700-900 m to the north and down hydrologic
80

gradient from the hottest temperature-gradient hole (Figure 5.4.2). Thus, the core of upflow at a

shallow depth interval (<250m) is likely <0.5-1 km west-southwest of the hottest drill hole, as

evidenced by: 1) tight spatial correlation in three dimensions between the highest temperature

gradient holes and 2-m stations tracking upflow/outflow, 2) intersecting gravity gradients

indicating a favorable structural setting, and 3) discrete magnetic-low and low-resistivity

anomaly.

The direct evidence map for temperatures indicative of a blind geothermal system in

southeastern Gabbs Valley has been significantly improved in this study with the addition of new

geothermometry, 2-m data, and temperature-gradient hole data. Figure 5.4.3A shows that there

was no indication of a blind geothermal system in the study area in the initial phase of the play

fairway analysis. Subsequent detailed studies obtained 130-140ºC geothermometry at the

Diamond A Ranch and identified the 2-m thermal anomaly in Phase II (Figure 5.4.3B).

Temperature-gradient drill data acquired in the final phase is another line of evidence for a

geothermal system and correlates with the 2-m data (Figure 5.4.3C). Figure 5.4.3D shows the

difference between direct evidence in the initial and final phases.


81

Figure 5.4.2. Spatial relationships of temperature and magnetic data. Temperature-gradient holes are shown as
colored triangles with warmer colors indicating higher temperatures at 152 m. The shallow temperature survey is
shown as colored circles with warmer colors indicating higher temperatures at 2 m. Magnetic survey shown as
colored basemap with warmer colors indicating higher relative magnetic values (nT). Favorable structural setting is
shown as red dashed polygon, which also approximately encompasses the extent of the thermal anomaly. Location
of deep potential well targets indicated by red circular bulls-eyes. Yellow circle indicates the inferred core area of
shallow (<250 m) hydrothermal upwelling.
82

Figure 5.4.3. Phase I-III direct evidence maps. A. Phase I direct evidence map. B. Phase II direct evidence map. C.
Phase III direct evidence map. D. Map showing difference in direct evidence between phase I and phase III. Note
that direct evidence near Cobble Cuesta is data from Payne (2013) that was added to the fairway map following
Phase I and not new data acquired in this study.
83

5.5 Blind Geothermal System in Southeastern Gabbs Valley

All available datasets indicate that relatively high-temperature hydrothermal fluids are

upwelling in the south-central part of southeastern Gabbs Valley in the area of densely spaced

faults and fault intersections collocated with the thermal anomaly (as defined by both

temperature-gradient holes and 2-m temperature data), intersecting gravity gradients, magnetic-

low, and low-resistivity (Figure 5.2.2). Fluids likely upflow along north-northeast striking

normal faults proximal to and along fault intersections. North-northeast-striking normal faults

have the highest tendency to slip and dilate in the current stress field, and fracture density is

probably greater near the fault intersections. The Mesozoic basement likely hosts a geothermal

reservoir, and joint and fracture analysis of basement exposures in the study area documented

pervasive fracture density likely increasing overall permeability of these lithologies. Primary

fracture orientations in the basement strike north-northeast, parallel to major normal faults in the

area, and are similarly favored for slip and dilation in the current stress field. Hydrothermal

fluids presumably convect from basement into the overlying volcanic and sedimentary rocks.

Multiple geologic and geophysical collocated features suggest that the Tertiary volcanic and

sedimentary rocks also host a geothermal reservoir. Temperature-gradient hole and 2-m survey

data constrain the zone of shallow upwelling (Figure 5.4.2), and the location of the magnetic-low

may best define the spatial extent of the geothermal system in volcanic rocks. The zone of

inferred upwelling in volcanic rocks occupies the area corresponding to the center of the

magnetic-low, which is ~750 m west of the hottest temperature-gradient hole (52-2). The fine-

scale resolution (<200 m) constraining the area with hydrothermal upflow is yet to be

ascertained. It is worth reiterating that this is a blind geothermal system, whereby fluids do not

daylight. Hydrothermal fluids are instead inhibited by a clay cap and/or captured by highly-
84

permeable alluvium and directed down hydrologic gradient to the north-northwest into lower

parts of the basin.

Based on synthesis of all available data sets and generation of a conceptual model of the

geothermal system, deeper drill targets are recommended to delineate the zone of upflow and

characterize the system. This is a natural progression following relatively shallow temperature-

gradient drilling in assessing a geothermal system. Figure 5.4.2 shows locations of suggested

drill targets. Four wells placed near the highest temperature-gradient hole, centered on the

magnetic low and drilled to progressively increasing depths from east to west (500-1000 m) are

positioned to intersect northwest-striking faults that may be channelizing hydrothermal fluids

proximal to intersections with northwest-striking faults. The increasing depths for the wells from

east to west is aimed at targeting the northeast-striking, northwest-dipping normal faults at

progressively lower depths within the geothermal system.

The geothermal system supplying the Don A. Campbell power plant ~15 km to the

northwest of the study area is an analogous system that shares many of the same geologic,

geochemical, and geophysical characteristics as the new geothermal discovery in southeastern

Gabbs Valley. Similar to southeastern Gabbs Valley, the Don A. Campbell geothermal system is

blind with no surface expression of geothermal activity. A reconnaissance 2-m survey initially

documented a 6.5 km2 temperature anomaly at that site (Kratt et al., 2008). The geophysical

expression of the Don Campbell system shares the same patterns observed in southeastern Gabbs

Valley with a conspicuous magnetic-low anomaly and northeast-trending gravity gradients. The

Don A. Campbell system occupies a hybrid structural setting, including fault intersections

between possible splays of the Benton Springs dextral fault and north-northeast-striking normal

faults in an apparent displacement transfer zone. The power plant taps a ~130°C reservoir
85

(Orenstein et al., 2015). Geothermometers for the Diamond A Ranch show similar patterns and

slightly lower values in comparison to water samples sourced from production wells at the power

plant. Well temperatures at ~150 m are ~130°C for six of the eight wells at the power plant.

Below 150 m, these six wells are isothermal down to terminal depths, which vary from 500-1200

m (Orenstein et al., 2015). The 150 m temperature gradient profiles of the wells at the power

plant share similar isothermal temperature/depth relationships, associated with zones of

hydrothermal upwelling, as the system identified in southeastern Gabbs Valley.

The newly discovered blind geothermal system in southeastern Gabbs Valley is an

excellent candidate for further development for electrical energy production. Temperature-

gradient hole measurements and potentially related geothermometry suggest that the system has

temperatures ≥130ºC. If so, the system would probably be capable of supporting a binary-type

geothermal power plant. Further exploration is needed, however, to test if the area has sufficient

reservoir temperatures and permeability over a broad enough area to maintain flow rates needed

to sustain a power plant. Drilling deeper temperature-gradient holes or full sized well bores is

recommended (Figure 5.4.2). Completion of these wells is necessary to assess the exact location

of hydrothermal upwelling, gauge the permeability of the reservoir, and directly sample water

from the system for improved geothermometry. Seismic reflection surveys would also be helpful

in imaging concealed faults and more accurately defining the basin architecture.

Other favorable structural settings identified in this study (Figure 5.2.1) have potential for

hosting a blind geothermal system, and further work is needed to delineate the quality of these

prospects. Reconnaissance geologic mapping, gravity surveys, and shallow temperature surveys

employed in this study have covered most of the southeastern part of the Gabbs Valley basin.

The majority of the 2-m survey in the lower parts of the basin encountered background
86

temperatures likely produced from cool groundwater. Typical 2-m temperatures in these areas do

not rule out the potential for blind geothermal systems, as thermal anomalies are possibly

suppressed by cool shallow ground water. The favorable structural setting directly east of

Mystery Ridge (Figure 5.2.1, point B) contains one moderately anomalous 2-m station that could

signal some geothermal activity. Collection of new magnetic and MT data in areas lacking

coverage would be helpful in determining whether any of the other favorable structural settings

are promising prospects. In settings displaying anomalous geophysical characteristics, such as a

magnetic-low or low-resistivity anomalies, temperature-gradient hole and/or GeoProbe drilling

(e.g., Zehner et al., 2012) would be recommended. The thermal anomaly identified by Payne

(2013) near Cobble Cuesta (Figure 3.3.1) is an excellent candidate for future exploration in

southeastern Gabbs Valley.

5.6 Broader Implications

The play fairway analysis for southeastern Gabbs Valley yielded significant differences

between the initial and final phases of the project (Figure 5.3.1A and 5.3.1C). The location of the

study area in a large late Cenozoic basin necessitated collection of new geophysical data to refine

and/or identify intrabasinal, favorable structural settings. Future exploration subsequently

focused on the most promising setting that displayed the highest structural complexity collocated

with other features indicative of geothermal potential (Figure 5.2.2). A primary difference

between the initial and final phases of this project is that the initial regional analysis recognized a

relatively broad, favorable structural setting in the southeastern part of Gabbs Valley (Figure 1.3

and Figure 5.3.1). As is typical in any exploration program, it is difficult to discern the detailed

characteristics of a particular area to select the most favorable targets for drilling in the early,

more regional stages. Detailed geological, geochemical, and geophysical investigations were
87

thus carried out in subsequent phases to apply the play fairway analysis at a finer scale, refine

exploration targets, and ultimately select temperature-gradient drilling sites within the basin. The

positive results epitomize the importance of detailed studies to refine exploration targets.

Considering that nearly half of the Great Basin region is covered by basins, this also

demonstrates the broad applicability of such detailed studies, as well as the large untapped

potential for commercial-grade blind geothermal systems in many of these basins. The discovery

of an apparent blind geothermal system in southeastern Gabbs Valley also provides an initial

validation of the Nevada play fairway methodology.

This methodology of accurately targeting blind geothermal systems through play fairway

analysis can be applied to future exploration programs in the Great Basin, particularly to

development in broad Neogene basins. Evolution of future methods may encompass targeting

geothermal systems on the scale of entire basins, which contain several individual prospects,

instead of a single system-scale analysis. System-scale analysis involving development of one

geothermal system at a time in isolated localities has been the standard since the 1980’s. An

exploration program targeting multiple geothermal systems at a time would yield several

advantages, such as lower long-term exploration costs (e.g., geophysical survey and drill rig

mobilization), reduced infrastructure expense (e.g., transmission lines), and a possibly more

robust investment model of diversified risk and higher total energy capacity projects. Conducting

exploration programs in this fashion would be similar to how oil fields are developed and

maximized (e.g., Merrill and Sternbach, 2017). Results from this study indicate that southeastern

Gabbs Valley has at least one blind geothermal system worthy of further investment and possibly

several other systems of similar quality. There are numerous basins within the Great Basin that

may hide multiple blind systems and are prime for this type of analysis. Locating and harnessing
88

large amounts of blind conventional geothermal resources in this fashion may change how the

geothermal industry operates in relation to other renewable energy industries in the coming

decades. Large capacity renewable infrastructure is expected to be built to transition away from

fossil-fuel energy resources. Proving that geothermal can compete with large wind and solar

projects will be important to capitalize on during this transition period. Play fairway analysis can

clearly play a crucial role in the continued advancement in the development of geothermal

energy.
89

4. Conclusions

This study has assessed the potential for high-temperature (≥130C) blind geothermal

system(s) in southeastern Gabbs Valley by utilizing a quantitative play fairway exploration

approach integrating geologic, geophysical, and geochemical datasets to individually rank

prospects to guide future development. Gabbs Valley is a complex, tectonically active basin

within the Great Basin on the boundary of the transtensional central Walker Lane and

extensional Basin and Range province. The relatively abrupt transition from northwest-striking

faults and fault blocks in the southernmost part of the study area to the north-northeast-trending

structural grain to the north reflects a transfer of dextral shear into northwest-directed extension.

The termination of the Petrified Springs fault, a major strike-slip fault of the central Walker

Lane, into an array of normal faults indicates that a broad displacement transfer zone occupies

the area.

The geologic and geophysical data indicate at least six individual, favorable structural

settings for geothermal activity within the broader displacement transfer zone in southeastern

Gabbs Valley. These include fault intersections of varying complexity, apparent pull-aparts and

releasing bends along the Petrified Springs dextral fault, and possible terminations of normal

fault zones. The area of tightly spaced faults and fault intersections in the south-central part of

southeastern Gabbs Valley is a particularly favorable structural setting. Multiple lines of direct

and indirect evidence suggest a relatively high-temperature (130C) blind geothermal system in

this area, including collocated intersecting gravity gradients, magnetic-low, low-resistivity, and

2-m temperature anomaly. Water samples collected from agricultural wells ~7 km northwest of

the 2-m anomaly produced geothermometers indicating subsurface fluid temperatures of 130-

140°C. However, geothermal fluids contributing to these wells are not clearly related to the blind
90

geothermal system in southeastern Gabbs Valley.

Six temperature-gradient holes were drilled to target the extent of the shallow-

temperature and geophysical anomalies to define the geothermal system in southeastern Gabbs

Valley. Two wells contained high temperatures exceeding boiling, with bottom-hole

temperatures of 114.5°C and 124.9°C, respectively. The remaining wells displaying elevated to

background temperatures ranging from 79.2°C to 28.7°C. The observed temperature gradients

for the two hottest drill holes necessitate drill-hole intercepts of convecting hydrothermal fluids.

Deeper drilling is needed to extend the temperature profiles to resolve whether these holes

intersected a zone of geothermal upflow or lateral outflow. The 2-m survey is spatially and

thermally consistent with the temperature-gradient holes, and integration of these datasets

constrains the core of upflow at a shallow depth interval (<250m) to an area <0.5-1 km west-

southwest of the hottest drill hole. Fluids likely upflow along north-northeast striking normal

faults proximal to and along fault intersections. The core of upwelling probably occupies the area

corresponding to the center of magnetic-low, which is ~750 m west of the hottest temperature-

gradient hole.

The newly discovered blind geothermal system in southeastern Gabbs Valley is an

excellent candidate for development of a geothermal power plant. Temperature-gradient hole

measurements and potentially related geothermometry suggest that the system in southeastern

Gabbs Valley has temperatures ≥130ºC and is likely capable of supporting a binary-type

geothermal power plant. Further exploration is needed to test if the area has sufficient reservoir

temperatures and permeability over a broad enough area to maintain flow rates needed to sustain

a power plant.
91

The discovery of an apparent blind geothermal system in southeastern Gabbs Valley

provides an initial validation of the play fairway methodology. This detailed study demonstrates

the broad applicability of this exploration strategy and the large untapped potential for

commercial-grade blind geothermal systems in many of the basins in the Great Basin region.

Evolution of future methods may encompass targeting multiple geothermal systems on the scale

of entire basins. There are numerous basins within the Great Basin that may hide multiple blind

systems and are prime for this type of analysis.


92

References Cited

Allmendinger, R. W., Cardozo, N., and Fisher, D., 2012, Structural geology algorithms: Vectors
and tensors in structural geology: New York, Cambridge University Press, 304 p.

Angelier, J., Colletta, B., and Anderson, R.E., 1985, Neogene paleostress changes in the Basin and
Range: A case study at Hoover Dam, Nevada-Arizona: Geological Society of
America Bulletin, v. 96, p. 347–361, doi: 10.1130/0016-
7606(1985)96<347:NPCITB>2.0.CO;2.

Angelier, J., 1994, Fault slip analysis and paleostress field reconstruction, in Hancock, P.L., ed.,
Continental Deformation: Terrytown, New York, Pergamon Press, p. 53-100.

Angster, S.J., 2018, Contributions to the neotectonics of the central and northern Walker Lane,
[Ph.D. dissertation]: University of Nevada, Reno, 172 p.

Arnórsson, S., 1983, Chemical equilibria in Icelandic geothermal systems – Implications for
chemical geothermometry investigations: Geothermics, v. 12, p. 119-128.

Athens, N.D., Glen, J.M., Morin, R.L., and Klemperer, S.L., 2011, ATV magnetometer systems for
efficient ground magnetic surveying: The Leading Edge, v. 30, p. 394-398.

Athens, N.D., Ponce, D.A., Jayko, A.S., Miller, M., McEvoy, B., Marcaida, M., Mangan, M.T.,
Wilkinson, S.K., McClain, J.S., Chuchel, B.A., and Denton K.M, 2014, Magnetic and
gravity studies of Mono Lake, east-central California: U.S. Geological Survey Open File
Report 2014-1043, 14 p.

Atwater, T., and Molnar, P., 1973, Relative motion of the Pacific and North American plates
deduced from seafloor spreading in the Atlantic, Indian and south Pacific Oceans, in
Kovach, R.L., and Nur, A., eds, Proceedings of the Conference on Tectonic Problems of the
San Andreas fault system: Stanford University Publication Geological Sciences, v. 13, p.
136-148.

Atwater, T., and Stock, J., 1998, Pacific – North American plate tectonics of the Neogene
southwestern United States – An update: International Geology Reviews, v. 40, p. 375-402.

Back, W., 1961, Techniques for mapping of hydrochemical facies: U.S. Geological Survey
Professional Paper 424, 382 p.

Baker, R.A., Gehman, H. M., James, W. R., and White, D. A., 1986, Geologic field number and oil
and gas plays, in Rice, D.D., ed., Oil and gas assessment- Methods and applications:
American Association of Petroleum Geologists Studies in Geology, v. 21, p. 25-31.

Barton, C.A., Hickman, S., Morin, R., Zoback, M.D., Finkbeiner, T., Sassj, J., and Benoit, D.,
1997, Fracture permeability and its relationship to in-situ stress in the Dixie Valley,
93

Nevada: Proceedings, 22nd Workshop on Geothermal Reservoir Engineering, Stanford


University, Stanford, California, p. 147-152.

Beanland, S., and Clark, M.M., 1982, The Owens Valley fault zone, Eastern California, and surface
faulting associated with the 1872 Earthquake: U.S. Geological Survey Bulletin 1982, 38 p.

Bell, J.W., DePolo, C.M., Ramelli, A.R., Sarna-Wojcicki, A.M., and Meyer, C.E., 1999, Surface
faulting and paleoseismic history of the 1932 Cedar Mountain earthquake area, west-central
Nevada, and implications for modern tectonics of the Walker Lane: Geological Society of
America Bulletin, v. 111, p. 791–807, doi: 10.1130/0016-
7606(1999)111<0791:SFAPHO>2.3.CO;2.

Bell, J., and Ramelli, A., 2007, Active faults and neotectonics at geothermal sites in the western
Basin and Range: Preliminary results: Geothermal Resources Council Transactions, v. 31,
p. 375–378.

Bell, J.W., and Ramelli, A.R., 2009, Active fault controls at high-temperature geothermal sites:
Prospecting for new faults: Geothermal Resources Council Transactions, v. 33, p. 425–429.

Birkeland, P.W., 1999, Soils and geomorphology: Oxford University Press, 430 p., doi:
10.1016/0016-7061(86)90046-7.

Blackwell, D.D., 1983, Heat Flow in the northern Basin and Range Province, in The role of heat in
the development of energy and mineral resources in the northern Basin and Range province:
Geothermal Resources Council Special Report 13, p. 81–92.

Blackwell, D., Stepp, P., and Richards, M., 2010, Comparison and discussion of the 6 km
temperature maps of the western U.S. prepared by the SMU Geothermal Lab and the USGS:
Geothermal Resources Council Transactions, v. 34, p. 515–520.

Blackwell, D.D., Richards, M., Frone, Z., and Batir, J., 2011, Temperature-at-depth maps for the
conterminous U.S. and geothermal resource estimates: Geothermal Resources Council
Transactions, v. 35, p. 1545-1550.

Blake, K., and Davatzes, N.C., 2012, Borehole image log and statistical analysis of FOH-3D,
Fallon Naval Air Station, NV: Proceedings, 37th Workshop on Geothermal Reservoir
Engineering, Stanford University, Stanford, California, 14 p.

Blakley, R.J., 1995, Potential theory in gravity and magnetic applications: Cambridge, New York,
Cambridge University Press, 135 p.

Blakely, R.J., and Simpson, R.W., 1986, Approximating edges of source bodies from magnetic or
gravity anomalies: Geophysics, v. 51, p. 1494-1498.

Blewitt, G., Hammond, W.C., and Kreemer, C., 2009, Geodetic observation of contemporary
deformation in the northern Walker Lane: 1. Semipermanent GPS strategy, in Oldow, J.S.,
94

Cashman, P.H., Late Cenozoic structure and evolution of the Great Basin- Sierra Nevada
transition: Boulder, Colorado, Geological Society of America Special Paper 447, p. 1-15.

Boden, D.R., 2017, Geologic fundamentals of geothermal energy: Boca Raton, Florida, CRC Press,
398 p.

Brundy, M., and Zoback, M., 1999, Drilling-induced tensile wall-fractures: Implications for
determination of in-situ stress orientation and magnitude: International Journal of Rock
Mechanics and Mining Sciences, v. 36, p. 191-215.

Caine, J.S., Evans, J.P., and Forster, C.B., 1996, Fault zone architecture and permeability structure:
Geology, v. 24, p. 1025–1028, doi: 10.1130/0091-7613(1996)024<1025

Cardozo, N., and Allmendinger, R. W., 2013, Spherical projections with OSXStereonet: Computers
and Geosciences, v. 51, p. 193 – 205, doi:10.1016/j.cageo.2012.07.021.

Cashman, P.H., and Fontaine, S.A., 2000, Strain partitioning in the northern Walker Lane, western
Nevada and northeastern California: Tectonophysics, v. 326, p. 111–130, doi:
10.1016/S0040-1951(00)00149-9.

Caskey, S.J., Wesnousky, S.G., Zhang, P., and Slemmons, D.B., 1996, Surface faulting of the 1954
Fairview Peak (MS 7.2) and Dixie Valley (MS 6.8) earthquakes, central Nevada:
Seismological Society of America Bulletin, v. 86, p. 761–787.

Caskey, S.J., and Wesnousky, S.G., 1997, Static stress changes and earthquake triggering during
the 1954 Fairview Peak and Dixie Valley earthquakes, central Nevada: Seismological
Society of America Bulletin, v. 87, p. 521–527.

Chave, A.D., Thomson, D.J., 2004, Bounded influence magnetotelluric response function
estimation: Geophysical Journal International, v. 157, p. 988-1006.

Christiansen, R.L., and McKee, E.H., 1978, Late Cenozoic volcanic and tectonic evolution of the
Great Basin and Columbia Intermontane regions, in Smith, R.B., Eaton, G.P., eds.,
Cenozoic Tectonics and Regional Geophysics of the Western Cordillera: Geological Society
of America Memoir 152, p. 283–312.

Colgan, J.P., Howard, K.A., Fleck, R.J., Wooden, J.L., 2010, Rapid middle Miocene extension and
unroofing of the southern Ruby Mountains, Nevada: Tectonics, v. 29, p. 38.

Coney, P.J., and Harms, T.A., 1984, Cordillerian metamorphic core complexes: Cenozoic
extensional relics of Mesozoic compression: Geology, v. 12, p. 550-554.

Coolbaugh, M.F., Taranik, J. V, Raines, G., Shevenell, A., Sawatzky, D., Bedell, R., and Minor,
T.B., 2002, A geothermal GIS for Nevada: Defining regional controls and favorable
95

exploration terrains for extensional geothermal systems: Geothermal Resources Council


Transactions, v. 26, p. 485–490.

Coolbaugh, M.F, and A. Shevenell, L., 2004, A method for estimating undiscovered geothermal
resources in Nevada and the Great Basin: Geothermal Resources Council Transactions, v.
28, p. 13-18.

Coolbaugh, M.F., Raines, G.L., Zehner, R.E., Shevenell, L., and Williams, C.F., 2006, Prediction
and discovery of new geothermal resources in the Great Basin: Multiple evidence of a large
undiscovered resource base: Geothermal Resources Council Transactions, v. 30, p. 867–
874.

Coolbaugh, M., Sladek, C., Faulds, J., Zehner, R., and Oppliger, G., 2007, Use of rapid temperature
measurements at a 2-meter depth to augment deeper temperature gradient drilling:
Proceedings, 32nd Workshop on Geothermal Reservoir Engineering, Stanford University,
Stanford, California, SGP-TR-183, p. 109-116.

Coolbaugh, M., Lechler, P., Sladek, C., and Kratt, C., 2009, Carbonate tufa columns as exploration
guides for geothermal systems in the Great Basin: Geothermal Resources Council
Transactions, v. 33, p. 461–466.

Coolbaugh, M., Sladek, C., Zehner, R., and Kratt, C., 2014, Shallow temperature surveys for
geothermal exploration in the Great Basin, USA, and estimation of shallow aquifer heat
loss: Geothermal Resources Council Transactions, v. 38, p. 115-122.

Cumming, W., 2009, Geothermal resource conceptual models using surface exploration data:
Proceedings: 34th Workshop on Geothermal Reservoir Engineering, Stanford University,
Stanford, California, 6 p.

Cunningham, W.D., and Mann, P., 2007, Tectonics of strike-slip restraining and releasing bends:
London, England, Geological Society of Special Publications 290, p. 1-12.

Curewitz, D., and Karson, J.A., 1997, Structural settings of hydrothermal outflow: Fracture
permeability maintained by fault propagation and interaction: Journal of Volcanology and
Geothermal Research, v. 79, p. 149–168, doi: 10.1016/S0377-0273(97)00027-9.

Dallmeyer, R.D., Snoke, A.W., and McKee, E.H., 1986, The Mesozoic-Cenozoic tectonothermal
evolution of the Ruby Mountains, East Humboldt Range, Nevada: A cordilleran
metamorphic core complex: Tectonics, v. 5, p. 931-954.

DeAngelo, J., Shervais, J.W., Glen, J.M., Nielson, D., Garg, S., Dobson, P., Gasperikova, E.,
Sonnenthal, E., Visser, C., and Liberty, L.M., 2016, Geothermal play fairway analysis of the
Snake River Plain: Phase II: Geothermal Research Council Transactions, v. 41, p. 2328-
2345.
96

Dilles, J.H., and Gans, P.B., 1995, The chronology of Cenozoic volcanism and deformation in the
Yerington area, western Basin and Range and Walker Lane: Geological Society of America
Bulletin, v. 107, p. 474-486.

Dokka, R.K., and Travis, C.J., 1983, Late Cenozoic strike‐slip faulting in the Mojave Desert,
California: Geology, v. 11, p. 305–308.

Dokka, R.K., and Travis, C.J., 1990, Role of the eastern California shear zone in accommodating
Pacific‐North American plate motion: Geophysical Research Letters, v. 17, p. 1323–1326.

Drummond, S.E., and Ohmoto, H., 1985, Chemical evolution and mineral deposition in boiling
hydrothermal systems: Economic Geology, v. 80, p. 126–147, doi:
10.2113/gsecongeo.80.1.126.

Doust, H., 2003, Placing petroleum systems and plays in their basin-history context: A means to
assist in the identification of new opportunities: First Break, v. 21, p. 73-83.

Doust, H., 2010, The exploration play: what do we mean by it?: American Association of
Petroleum Geologist Bulletin, v. 94, p. 1657-1672.

Ekren, E. B., and Byers, F.M., Jr., 1976, Ash-flow fissure vent in west-central Nevada: Geology, v.
4, p. 247-251.

Ekren, E.B., Byers, F.M., Jr., Hardyman, R.F., Marvin, R.F., and Silberman, M.L., 1980,
Stratigraphy, preliminary petrology, and some structural features of Tertiary volcanic rocks
in the Gabbs Valley and Gillis Ranges, Mineral County, Nevada: U.S. Geological Survey
Bulletin 1464, p. 1-54.

Ekren, E.B., and Byers, F.M., Jr., 1984, The Gabbs Valley Range – A well-exposed segment of the
Walker Lane in west-central Nevada, in Lintz, J., Jr., ed., Western geologic excursions, v. 4:
Reno, Nevada: Geologic Society of America Guidebook, Annual Meeting, p. 203-215.

Ekren, E. B., and Byers, F.M., Jr., 1985a, Geologic map of the Gabbs Mountain, Mount Ferguson,
Luning, and Sunrise Fault quadrangles, Mineral and Nye counties, Nevada: U.S. Geological
Survey Map I-1577, scale 1:48,000, 1 sheet.

Ekren, E. B., and Byers, F.M., Jr., 1985b, Geologic map of the Win Wan Flat, Kinkaid NW,
Kinkaid, and Indian Head Peak quadrangles, Mineral County, Nevada: U.S. Geological
Survey Map I-1578, scale 1:48,000, 1 sheet.

Ekren, E. B., and Byers, F.M., Jr., 1986a, Geologic map of the Mount Annie NE, Mount Annie,
Ramsey Spring, and Mount Annie SE quadrangles, Mineral and Nye counties, Nevada: U.S.
Geological Survey Map I-1579, scale 1:48,000, 1 sheet.

Ekren, E. B., and Byers, F.M., Jr., 1986b, Geologic map of the Murphys Well, Pilot Cone, Copper
Mountain, and Poinsettia Spring quadrangles, Mineral County, Nevada: U.S. Geological
Survey Map I-1576, scale 1:48,000, 1 sheet.
97

Evernden, J., and Kistler, R.W., 1970, Chronology of emplacement of Mesozoic batholithic
complexes in California and western Nevada: U.S. Geological Survey Professional Paper
623, 42 p.

Faulds, J.E., and Varga, R.J., 1998, The role of accommodation zones and transfer zones in the
regional segmentation of extended terranes, in Faulds, J.E., and Stewart, J.H., eds.,
Accommodation Zones and Transfer Zones: The Regional Segmentation of the Basin and
Range Province: Geological Society of America Special Paper 323. p. 1-46.

Faulds, J.E., Coolbaugh, M.F., Blewitt, G., and Henry, C.D., 2004, Why is Nevada in hot water?
Structural controls and tectonic model of geothermal systems in the northwestern Great
Basin: Geothermal Resources Council Transactions, v. 28, p. 649–654.

Faulds, J.E., Henry, C.D., and Hinz, N.H., 2005, Kinematics of the northern Walker Lane: An
incipient transform fault along the Pacific-North American plate boundary: Geology, v. 33,
p. 505–508.

Faulds, J.E., Coolbaugh, M.F., Vice, G.S., and Edwards, M.L., 2006, Characterizing structural
controls of geothermal fields in the northwestern Great Basin: A progress report:
Geothermal Resources Council Transactions, v. 30, p. 69–76.

Faulds, J.E., and Henry, C.D., 2008, Tectonic influences on the spatial and temporal evolution of
the Walker Lane: an incipient transform fault along the evolving Pacific – North American
plate boundary, in Spencer, J.E., and Titley, S.R., eds, Ores and orogenesis: Circum Pacific
tectonics, geologic evolution, and ore deposits: Arizona Geological Society Digest, p. 437–
470.

Faulds, J.E., Coolbaugh, M.F., Bouchot, V., Moeck, I., and Oguz, K., 2010, Characterizing
structural controls of geothermal reservoirs in the Great Basin, USA, and western Turkey:
Developing successful exploration strategies in extended terranes: Proceedings of the World
Geothermal Congress, Bali, Indonesia, 25-29 April, 10 p.

Faulds, J.E., Hinz, N.H., Coolbaugh, M.F., Cashman, P.H., Kratt, C., Dering, G., Edwards, J.,
Mayhew, B., and Mclachlan, H., 2011, Assessment of favorable structural settings of
geothermal systems in the Great Basin, western USA: Geothermal Resources Council
Transactions, v. 35, p. 777–783.

Faulds. J.E., Hinz, N., Kreemer, C., and Coolbaugh, M., 2012, Regional patterns of geothermal
activity in the Great Basin region, western USA: Correlation with strain rates and
distribution of geothermal fields: Geothermal Resources Council Transactions, v. 36, p.
897-902.

Faulds, J.E., Hinz, N.H., Dering, G.M., and Siler, D.L., 2013, The hybrid model — The most
accommodating structural setting for geothermal power generation in the Great Basin,
western USA: Geothermal Resources Council Transactions, v. 37, p. 3-10.
98

Faulds, J.E., Hinz, N.H., Coolbaugh, M.F., Shevenell, L.A., Siler, D.L., dePolo, C.M., Hammond,
W.C., Kreemer, C., Oppliger, G., Wannamaker, P., Queen, J.H., and Visser, C., 2015a,
Integrated geologic and geophysical approach for establishing geothermal play fairways and
discovering blind geothermal systems in the Great Basin region, western USA: Final
submitted report to the Department of Energy (DE-EE0006731), 106 p.

Faulds, J.E., Hinz, N.H., Coolbaugh, M.F., Shevenell, L.A., Siler, D.L., dePolo, C.M., Hammond,
W.C., Kreemer, C., Oppliger, G., Wannamaker, P.E., Queen, J.H., and Visser, C.F., 2015b,
Integrated geologic and geophysical approach for establishing geothermal play fairways and
discovering blind geothermal systems in the Great Basin region, western USA: A progress
report: Geothermal Resources Council Transactions, v. 39, p. 691-700.

Faulds, J.E., and Hinz, N.H., 2015, Favorable tectonic and structural settings of geothermal systems
in the Great Basin region, western USA: Proxies for discovering blind geothermal systems:
World Geothermal Congress 2015, Melbourne, Australia, 6 p.

Faulds, J. E., Hinz, N.H., Coolbaugh, M. F., Shevenell, L. A., and Siler D. L, 2016a, The
Nevada play fairway project — Phase II: Initial search for new viable geothermal systems
in the Great Basin region, western USA: Geothermal Resources Council Transactions, v.
40, p. 535-540.

Faulds, J.E., Hinz, N.H., Coolbaugh, M.F., Siler, D.L., Shevenell, L.A., Queen, J.H., dePolo, C.M.,
Hammond, W.C., and Kreemer, C., 2016b, Discovering geothermal systems in the Great
Basin region: an integrated geologic, geochemical, and geophysical approach for
establishing geothermal play fairways: Proceedings, 41st Workshop on Geothermal
Reservoir Engineering, Stanford University, Stanford, California, 15 p.

Faulds, J. E., Hinz, N.H., Coolbaugh, M. F., Shevenell, L. A., Sadowski, A.J., Shevenell, L.A.,
McConville, E., Craig, J., Sladek, C., and Siler D. L, 2017a, Progress report on the Nevada
play fairway project: Integrated geological, geochemical, and geophysical analyses of
possible new geothermal systems in the Great Basin region: Proceedings, 42nd Workshop
on Geothermal Reservoir Engineering, Stanford University, Stanford, California, 11 p.

Faulds, J. E., Hinz, N.H., Coolbaugh, M. F., Shevenell, L. A., Sadowski, A.J., Shevenell, L.A.,
McConville, E., Craig, J., Sladek, C., and Siler D. L, 2017b, Discovering Blind Geothermal
Systems in the Great Basin Region: An Integrated Geologic and Geophysical Approach for
Establishing Geothermal Play Fairways: Final report submitted to the Department of Energy
(DE-EE0006731), 37 p.

Ferrill, D.A., and Morris, A.P., 2003, Dilatational normal faults: Journal of Structural Geology, v.
25, p. 183–196.

Ferrill, D.A., Winterle, J., Wittmeyer, G., Sims, D., Colton, S., Armstrong, A., and Morris, A.P.,
1999, Stressed rock strains groundwater at Yucca Mountain, Nevada: Geologic Society
Today, v. 9, p. 1–8.
99

Forson, C., Czajkowski, J.L., Norman, D.K., Swyer, M.W., Cladouhos, T.T., and Davatzes, N.,
2016, Summary of phase 1 and plans for phase 2 of the Washington state geothermal play-
fairway analysis, Geothermal Resources Council Transactions, v. 40, p. 541-550.

Fosdick, J.C., and Colgan, J.P., 2008, Miocene extension in the East Range, Nevada: A two-stage
history of normal faulting in the northern basin and range: Geological Society of America
Bulletin, v. 120, p. 1198–1213, doi: 10.1130/B26201.1.

Fournier, R.O., and Truesdell, A.H., 1973, An empirical NaKCa geothermometer for natural
waters: Geochemica et Cosmochimica Acta, v. 37, p. 1255-1275.

Fournier, R.O., 1977a, Chemical geothermometers and mixing models for geothermal systems:
Geothermics, v. 5, p. 41–50.

Fournier, R.O., 1977b, The solubility of amorphous silica in water at high temperatures and
pressures: The American Mineralogist, v. 62, p. 1052-1056.

Fournier, R.O., and Potter II, R.W., 1979, Magnesium correction to the NAKCa chemical
geothermometer: Geochemica et Cosmochimica Acta, v. 43, p.1543-1550.

Fournier, R.O., and Potter II, R.W., 1982, Revised and expanded silica (quartz) geothermometer:
Geothermal Resources Council Bulletin, v. 11, p. 1-32.

Fournier, R.O., 1983, A method of calculating quartz solubilities in aqueous sodium chloride
solutions: Geochimica et Cosmochimica Acta, v. 47, p.579-586.

Giggenbach, W.F., 1984, Mass transfer in hydrothermal alteration systems – a conceptual


approach: Geochemica et Cosmochimica Acta, v. 48, p. 2693-2711.

Giggenbach, W.F., 1988, Geothermal Solute Equilibria. Derivation of N-K-Mg-Ca Geoindicators:


Geochemica et Cosmochimica Acta, v. 52, p. 2749-2765.

Glen, J.M., Egger. A.E., Ponce, D.A., 2008, Structures controlling geothermal circulation identified
through gravity and magnetic transects, Surprise Valley, California, Northwestern Great
Basin: Geothermal Resources Council Transactions, v. 32, p. 279-283.

Glen, J.M.G., Liberty, L., Peacock, J., Gasperikova, E., Earney, T., Schermerhorn, W., Siler, D.,
Shervais, J., Dobson, P., 2018, A geophysical characterization of the structural framework
of the Camas Prairie geothermal system, southcentral Idaho: Geothermal Resources Council
Transactions, v. 42, p. 466-481.

Hammond, W.C., and Thatcher, W., 2005, Northwest Basin and Range tectonic deformation
observed with the Global Positioning System, 1999-2003: Journal of Geophysical Research,
v. 110, B10405.
100

Hammond, W.C., and Thatcher, W., 2007, Crustal deformation across the Sierra Nevada, northern
Walker Lane, Basin and Range transition, western United States measured with GPS, 2000-
2004: Journal of Geophysical Research, v. 112, B05411.

Hammond, W.C., Kreemer, C., and Blewitt, G., 2009, Geodetic constraints on contemporary
deformation in the northern Walker Lane: 3. Central Nevada seismic belt postseismic
relaxation, in Oldow, J.S., Cashman, P.H., eds., Late Cenozoic Structure and Evolution of
the Great Basin Sierra Nevada Transition: Geological Society of America Special Paper
447, p. 33–54.

Hardyman, R. F., and Oldow, J. S., 1991, Tertiary tectonic framework and Cenozoic history of the
central Walker Lane, Nevada, in Raines, G. L., Lisle, R. E., Schafer, R. W., and Wilkinson,
W. H., eds., Geology and ore deposits of the Great Basin: Reno, Nevada, Geological
Society of Nevada Symposium Proceedings, v. 1, p. 279-301.

Henley, R.W., Brown, K.L., 1985, A practical guide to the thermodynamics of geothermal fluids
and hydrothermal ore deposits, in Berger, B.R., Bethke, P.M., ed., Geology and
geochemistry of epithermal systems: Reviews of Economic Geology, v. 2, p. 25–43.

Henry, C.D., Mcgrew, A.J., Colgan, J.P., Snoke, A.W., and Brueseke, M.E., 2011, Timing,
distribution, amount, and style of Cenozoic extension in the northern Great Basin: Geologic
Society of America Field Guide, v. 21, 40 p., doi: 10.1130/2011.0021(02).

Henry, C.D., and John, D.A., 2013, Magmatism, ash-flow tuffs, and calderas of the
ignimbrite flareup in the western Nevada volcanic field, Great Basin, USA: Geosphere, v. 9,
p. 951–1008, doi: 10.1130/GES00867.1.

Hickman, S.H., and Davatzes, N.C., 2010, In-situ stress and fracture characterization for planning
of an EGS stimulations in the Desert Peak geothermal field: 35th Workshop on Geothermal
Reservoir Engineering, Stanford University, Stanford, California, 14 p.

Hinz, N.H., Faulds, J.E., Siler, D.L., Tobin, B., Blake, K., Tiedeman, A., Sabin, A., Blankenship,
D., Kennedy, M., Rhodes, G., Nordquist, J., Hickman, S., Glen, J., Williams, C., Robertson-
Tait, A., Calvin, W., 2017, Stratigraphic and structural framework of the proposed Fallon
Forge Site, Nevada: Proceedings, 41st Workshop on Geothermal Reservoir Engineering,
Stanford University, Stanford, California, 12 p.

Hudson, M,R., John, D.A., Conrad, J.E., McKee, E.H., 2000, Style and age of late Oligocene-early
Miocene deformation in the southern Stillwater range, west central Nevada:
Paleomagnetism, geochronology, and field relations: Journal of Geophysical Research, v.
105, p. 929-954.

Humphreys, E.D., 1995, Post-Laramide removal of the Farallon Slab, western United States:
Geology, v. 23, p. 987-990.
101

John, D.A., 1992, Late Cenozoic volcanotectonic evolution of the southern Stillwater Range, west-
central Nevada, in Craig, S.D., eds., Structure, Tectonics, and Mineralization of the Walker
Lane: Geologic Society of Nevada, p. 64-92.

John, D.A., 1995, Tilted middle Tertiary ash-flow calderas and subjacent granitic plutons, southern
Stillwater Range, Nevada: Cross sections of an Oligocene igneous center: Geologic Society
of America Bulletin, v. 107, p. 180-200.

Kelbert, A., Meqbel, N., Egbert, G.D., Kush,T., 2014, ModEM: A modular system for inversion of
electromagnetic geophysical data: Computers and Geoscience, v. 66, p. 40-53.

Kratt, C., Coolbaugh, M., Sladek, C., Zehner, R., Penfield, R., and Delwiche, B., 2008, A new gold
pan for the west: Discovering blind geothermal systems with shallow temperature surveys:
Geothermal Resources Council Transactions, v. 32, p. 153–158.

Kratt, C., Coolbaugh, M., Peppin, B., and Sladek, C., 2009, Identification of a new blind
geothermal system with hyperspectral remote sensing and shallow temperature
measurements at Columbus Salt Marsh, Esmeralda County, Nevada: Geothermal Resources
Council Transactions, v. 33, p. 481–485.

Kratt, C., Sladek, C., and Coolbaugh, M., 2010, Boom and bust with the latest 2m temperature
surveys: Dead Horse Wells, Hawthorne Army Depot, Terraced Hills, and other areas in
Nevada: Geothermal Resources Council Transactions, v. 34, p. 567–574.

LaFehr, T.R., 1991, An exact solution for the gravity curvature (Bullard B) correction: Geophysics,
v. 56, p. 1179-1184.

Lautze, N., Thomas, D., Hinz, N., Apuzen-Ito, G., Frazer, N., and Waller, D., 2017, Play fairway
analysis of geothermal resources across the State of Hawaii: 1. Geological, geophysical, and
geochemical datasets: Geothermics, v. 70, p. 376-392.

MacCready, T., Snoke, A.W., Wright, J.E., and Howard, K.A., 1997, Mid-crustal flow during
Tertiary extension in the Ruby Mountains core complex, Nevada: Geologic Society of
America Bulletin, v. 109, p. 1576-1594.

McGrew, A.J., Peters, M.T., and Wright, J.E., 2000, Thermobarometric constraints on the
tectonothermal evolution of the East Humboldt Range metamorphic core complex, Nevada:
Geologic Society of America Bulletin, v. 112, p. 45-60.

Marrett, R., and Allmendinger, R.W., 1990, Kinematic analysis of fault-slip data: Journal of
Structural Geology, v. 12, p. 973–986.

Micklethwaite, S., and Cox, S.F., 2004, Fault-segment rupture, aftershock-zone fluid flow, and
mineralization: Geology, v. 32, p. 813–816.
102

Moeck, I.S., Beardsmore, G., Harvey, C.C., 2015, Cataloging worldwide geothermal systems by
geothermal play type: World Geothermal Congress Proceedings, Melbourne, Australia, 9 p.

Morris, A., Ferrill, D.A., and Henderson, D.B., 1996, Slip-tendency analysis and fault reactivation:
Geology, v. 24, p. 275–278.

Muller, S.W.M., and Ferguson, H.G., 1939, Mesozoic stratigraphy of the Hawthorne and Tonopah
quadrangles, Nevada: Geological Society of America Bulletin, v. 50, p. 1573–1624, doi:
10.1130/GSAB-50-1573.

Nash, G.D., and Bennett, C.R., 2015, Adaptation of a petroleum exploration tool to geothermal
exploration: Preliminary play fairway model of Tularosa Basin, New Mexico, and Texas:
Geothermal Resources Council Transactions, v. 39, p. 743-749.

Oldow, J.S., 1981, Structure and stratigraphy of the Luning allochthon and the kinematics of
allochthon emplacement, Pilot Mountains, west-central Nevada: Geological Society of
America Bulletin, v. 92, p. 888–911.

Oldow, J.S., 1984, Evolution of a late Mesozoic back-arc fold and thrust belt, northwestern Great
Basin, U.S.A, in Carlson, R.L., and Kobayashi, K., eds., Geodynamics of back-arc regions:
Tectonophysics, v. 102, p. 245-274.

Orenstein, R., Delwiche, B., and Lovekin, J., 2014, The Don A. Campbell geothermal project –
development of a low-temperature resource: Geothermal Resources Council Transactions,
v. 38, p. 91-88.

Payne, J.F, Bell, J.W, Calvin, W., and Spinks, K., 2011, Active fault structure and potential high
temperature geothermal systems: LiDAR analysis of the Gabbs Valley, Nevada, fault
system: Geothermal Resources Council Transactions, v. 35, p. 961–966.

Payne, J.F., 2013, Characterization of a blind geothermal prospect through LiDAR analysis and
shallow temperature survey, Gabbs Valley, Nye and Mineral Co., NV [MS. thesis]:
University of Nevada Reno, 113 p.

Peacock, J. R., Glen, J., Ritzinger, B., Earney, T., Schermerhorn, W., Siler, D., and Anderson, M.,
2018, Geophysical Imaging Geothermal Systems Spanning Various Geologic Settings,
Geothermal Resources Council Transactions, v. 42, p. 514-523.

Peiffer, L., Wanner, C., Spycher, N., Sonnenthal, E.L., Kennedy, B.M., and Iovenitti, J., 2014,
Optimized multicomponent vs. classical geothermometry: Insights from modeling studies at
the Dixie Valley geothermal area: Geothermics, v. 51, p. 154–169, doi:
10.1016/j.geothermics.2013.12.002.

Powell, T., and Cumming, W., 2010, Spreadsheets for geothermal water and gas geochemistry:
35th Workshop on Geothermal Reservoir Engineering Stanford University, Stanford,
California, February 1-3, 2010. SGP-TR-188.
103

Salmon, J.P., Meurice, J., Wobus, N., Stern, F., and Duaime, M., 2012, Guidebook to Geothermal
Power Finance: National Renewable Energy Laboratory, p. 1-61.

Shervais, J.W., Glen, J.M., Nielson, D., Garg, S., Dobson, P., Gasperikova, E., Sonnenthal, E.,
Visser, C., Liberty, L.M., Deangelo, J., Siler, D., and Evans, J.P., 2016, Geothermal play
fairway analysis of the Snake River Plain: Phase 1: Proceedings, 41st Workshop on
Geothermal Reservoir Engineering, Stanford University, Stanford, California, 7 p.

Spycher, N, L. Peiffer, S. Finsterle and E. Sonnenthal, 2016, GeoT user’s guide: A computer
program for multicomponent geothermometry and geochemical speciation, version 2.1:
LBNL Report, Rev. 1, June 6, 2016, 42 p.

Proffett, J.M., 1977, Cenozoic geology of the Yerington district, Nevada, and implications for the
nature and origin of Basin and Range faulting: Geological Society of America Bulletin, v.
88, p. 247–266.

Schwering, P.C., Karlin, R.E., 2012, Structural interpretation and modeling of the Dixie Meadows
geothermal prospect using gravity and magnetic data: Geothermal Resources Council
Transactions, v. 36, p. 53-58.

Shevenell, L., and DeRocher, T., 2005, Evaluation of chemical geothermometers for calculating
reservoir temperatures at Nevada geothermal power plants: Geothermal Resources Council
Transactions, v. 29, p. 303–308.

Shevenell, L.A., and Coolbaugh, M.F., 2011, A new method of evaluation of chemical
geothermometers for calculating reservoir temperatures from thermal springs in Nevada:
Geothermal Resources Council Transactions, v. 35, p. 657–661.

Schwartz, F.W., and Zhang, H., 2003, Fundamentals of Ground Water: New York, New York, John
Wiley and Sons, Inc., 583 p.

Sladek, C., Coolbaugh, M.F., and Zehner, R.E., 2007, Development of 2-meter soil temperature
probes and results of temperature survey conducted at Desert Peak, Nevada, USA:
Geothermal Resources Council Transactions, v. 31, p. 363-368.

Sladek, C., Coolbaugh, M.F., and Kratt, C., 2009, Improvements in shallow (two-meter)
temperature measurements and data interpretation: Geothermal Resources Council
Transactions, v. 33, p. 535-541.

Sladek, C., Coolbaugh, M.F., Penfield, R., Skord, J., and Williamson, L., 2012, The influences of
thermal diffusivity and weather on shallow (2-meter) temperature measurements:
Geothermal Resources Council Transactions, v. 36, p. 793-798.
104

Sladek, C., and Coolbaugh, M.F., 2013, Development of online map of 2 meter temperatures and
methods for normalizing 2 meter temperature data for comparison analysis: Geothermal
Resources Council Transactions, v. 37, p. 333-336.

Snoke, A.W., McGrew, A.J., Valasek, P.A., and Smithson, S.B., 1990, A crustal cross-section for a
terrain of superimposed shortening and extension: Ruby Mountains-East Humboldt Range
metamorphic core complex, Nevada, in Salisbury, M.H., and Fountain, D.M., eds., Exposed
Cross-Section of the Continental Crust: Kluwer Academic Publishers, v. 317, p. 103-135.

Stewart, J.H., and Carlson, J.E., 1978, Geologic Map of Nevada: U.S. Geological Survey and
Nevada Bureau of Mines and Geology, scale 1:500,000 (not part of any formal series,
printed and distributed by the U.S. Geological Survey, G75163, reprinted, 1981, G81386).

Stewart, J.H., 1978, Basin-range structure in western North America: A review, in Smith, R.B.,
Eaton, G.P., eds., Cenozoic Tectonics and Regional Geophysics of the Western Cordillera:
Geologic Society of America Memoirs, v. 152, p. 1-31.

Stewart, J.H., 1988, Tectonics of the Walker Lane belt, western Great Basin: Mesozoic and
Cenozoic deformation in a zone of shear, in Ernst, W.G., ed., Metamorphism and crustal
evolution of the western United States: Englewood Cliffs, New Jersey, Prentice Hall, p.
681-713.

Stewart, J.H., 1998, Regional characteristics, tilt domains, and extensional history of the later
Cenozoic Basin and Range Province, western North America, in Faulds J.E., and Stewart,
J.H., eds., Accommodation Zones and Transfer Zones: The Regional Segmentation of the
Basin and Range Province: Geological Society of America Special Paper 323. p. 47–74,
doi: 10.1130/0-8137-2323-X.47.

Stewart, J.H., Sarna-Wojcicki, A., Meyer, C.E., and Elmira, W., 1999, Stratigraphy,
tephrochronology, and structural setting of Miocene sedimentary rocks in the Cobble Cuesta
area, west-central Nevada: U.S. Geological Survey Open File Report 99-352, 21 p.

Sullivan, W.A., and Snoke, A.W., 2007, Comparative anatomy of core-complex development in the
northeastern Great Basin, U.S.A: Rocky Mountain Geology, v. 42, 29 p.

Telford, W.M., Geldart, L.P., and Sheriff, R.E., 1990, Applied Geophysics: Cambridge, New York,
Cambridge University Press, 792 p.

Thatcher, W., 2003, GPS Constraints on the kinematics of continental deformation: International
Geology Review, v. 45, p. 191–212, doi: 10.2747/0020-6814.45.3.191.

Trexler, J.H., Cashman, P.H., Henry, C.D., Muntean, T., Schwartz, K., TenBrink, A., Faulds, J.E.,
Perkins, M., and Kelly, T., 2000, Neogene basins in western Nevada document the tectonic
history of the Sierra Nevada – Basin and Range transition zone for the last 12 Ma, in
Lageson, D.R., Peters, S.G., and Lahren, M.M., eds., Great Basin and Sierra Nevada:
Geologic Society of America Field Guide 2, p. 97-116.
105

Truesdell, A.H., and Fournier, R.O., 1976, Conditions in the deeper parts of the hot spring systems
of Yellowstone National Park, Wyoming: U.S. Geological Survey Open File Report 76-428,
29 p.

U.S. Geological Survey, 2006, Quaternary fault and fold database for the United States:
http://earthquake.usgs.gov/hazards/qfaults/ (accessed January 2017).

Ussher, G., 2000, Understanding the resistivities observed in geothermal systems: Proceedings
World Geothermal Congress 2000, Kyushu-Tohoku, Japan, May 28-June 10, 6 p.

Vozoff, K., 1991, The Magnetotelluric Method, in M.N., Nabighian, ed., Electromagnetic Methods
in Applied Geophysics: Society of Exploration Geophysics, p. 641–712.

Wallace, R.E., 1979, Strain pattern represented by scarps formed during the earthquakes of October
2, 1915, Pleasant Valley, Nevada: Tectonophysics, v. 52, p. 599.

Wannamaker, P.E., Caldwell, T.G., Jiracek, G.R., Maris, V., Hill, G.J., Ogawa, Y., Bibby, H.M.,
Bennie, S.L., and Heise, W., 2009, Fluid and deformation regime of an advancing
subduction system at Marlborough, New Zealand: Nature, v. 460, p. 733–736, doi:
10.1038/nature08204.

Wannamaker, P.E., Maris, V., Hasterok, D.P., and Doerner, W.M., 2011, Crustal Scale Resistivity
Structure, Magmatic-Hydrothermal Connections, and Thermal Regionalization of the Great
Basin Tectonic/Geothermal Correlations with Deep Resistivity: Geothermal Resources
Council Transactions, v. 35, p. 1787–1790.

Wannamaker, P.E., Maris, V., Sainsbury, J., and Iovenitti, J., 2013, Intersecting Fault Trends and
Crustal-Scale Fluid Pathways Below the Dixie Valley Geothermal Area, Nevada, Inferred
From 3D Magnetotelluric Surveying: Proceedings, 38th Workshop on Geothermal Reservoir
Engineering, Stanford University, Stanford, California, v. 38, p. 1303–1311.

Wannamaker, P.E., Moore, J.N., Pankow, K.L., Simmons, S.D., Nash, G.D., Maris, V., Batchelor,
C., and Hardwick, C.L., 2015, Play fairway analysis of the Eastern Great Basin extensional
regime, Utah: Preliminary implications: Geothermal Resources Council Transactions, v. 39,
p. 793-802.

Wannamaker, P.E., Pankow, K.L., Moore, J.N., Nash, G.D., Maris ,V., Simmons, S.F., Hardwish,
C.L., Trow, A., and Allis, R., 2017a, Phase II activities in play fairway analysis for
structurally controlled geothermal systems in the Eastern Great Basin extensional regime,
Utah: 42nd Workshop of Geothermal Reservoir Engineering, Stanford University, Stanford,
California, SGP-TR-212, p. 1-12.

Wannamaker. P.E., Moore, J.N., Pankow, K.L., Simmons, S.F., Nash, G.D., Maris, V., Trow, A.,
and Hardwick, C.L., 2017b, Phase II of play fairway analysis for the eastern Great Basin
106

extensional regime, Utah: Status of indications: Geothermal Resources Council


Transactions, v. 41, p. 2368-2382.

Wells, S.G., McFadden, L.D., Poths, J., and Olinger, C.T., 1995, Cosmogenic 3He surface-
exposure dating of stone pavements. Implications for landscape evolution in deserts:
Geology, v. 23, p. 613–616.

Wernicke, B., 1992, Cenozoic extensional tectonics of the US Cordillera, in Burchfiel, B.C.,
Lipman, P.W., and Zoback, M.L., eds., The Cordilleran Orogen: conterminous U.S:
Boulder, Geologic Society of America, The Geology of North America, v. G-3, p. 553-581.

Wesnousky, S.G., 2005a, Active faulting in the Walker Lane: Tectonics, v. 24, p. 1–35, doi:
10.1029/2004TC001645.

Wesnousky, S.G., 2005b, The San Andreas and Walker Lane fault systems, western North
America: Transpression, transtension, cumulative slip and the structural evolution of a
major transform plate boundary: Journal of Structural Geology, v. 27, p. 1505–1512.

Wesnousky, S.G., Barron, A.D., Briggs, R.W., Caskey, S.J., Kumar, S., and Owen, L., 2005,
Paleoseismic transect across the northern Great Basin: Journal of Geophysical Research, v.
110, p. 1–25, doi: 10.1029/2004JB003283.

Wesnousky, S.G., Bormann, J.M., Kreemer, C., Hammond, W.C., and Brune, J.N., 2012,
Neotectonics, geodesy, and seismic hazard in the northern Walker Lane of western North
America: Thirty kilometers of crustal shear and no strike slip?: Earth and Planetary Science
Letters, v. 329–330, p. 133–140, doi: 10.1016/j.epsl.2012.02.018.

Wyld, Sandra. J., 2002, Structural Evolution of a Mesozoic backarc fold-and-thrust best in the U.S.
Cordilleran: New evidence from northern Nevada: Geological Society of America Bulletin:
v. 114, p. 1452-1468.

Zehner, R.E., Tullar, K.N., and Rutledge, E., 2012, Effectiveness of 2-meter and geoprobe shallow
temperature surveys in early stage geothermal exploration, Geothermal Resources Council
Transactions, v. 36, p. 835-841.

Zoback, M.L., 1989, State of stress and modern deformation of the northern Basin and Range
province: Journal of Geophysical Research, v. 94, p. 7105-7128.

Zoback, M.D., Barton, C.A., Brundy, M., Castillo, D.A., Finkbeiner, T., Grollimund, B.R., Moos,
D.B., Peska, P., Ward, C.D., and Wiprut, D.J., 2003, Determination of stress orientation and
magnitude in deep wells: International Journal of Rock Mechanics and Mining Sciences, v.
40, p. 1049-1076.
107

APPENDIX A: Complete lithologic unit descriptions of southeastern Gabbs Valley area for
PLATE 1
Label Unit Name Description Age Thickness
Qp Playa deposits High albedo surface composed of fine-grained sand, silt, Holocene <100 m
and clay with localized evaporite deposits in western
portion of basin.
Qe Eolian Sand sheets generally <50 cm in thickness. Holocene <3 m
deposits
Qa Active stream Active and recently abandoned stream channels. Holocene <20 m
channel
deposits
Qay Young Basin-fill alluvium composed of sand, silt, and clays. Light Holocene <30 m
alluvium and color in imagery, smooth surface, and not typically offset
basin deposits by fault scarps.
Qao Old basin Basin-fill alluvium composed of sand, silt, and clays. Holocene to <100 m
deposits Displays moderately dark tone in imagery, smooth surface, middle
and contains fault scarps. Pleistocene
Qfy Young Fans displaying light-mid gray tone and smooth texture Holocene <100 m
alluvial fans with active stream channels and abandoned anastomosing
gullies. Interfluves are round, ~1-2 m high, and form
debris flow levees lacking desert pavement. Interfluves are
composed of pebble to cobble sized clasts and lesser
boulders (<1 m). Fluves display dendritic pattern, are
spaced tightly, contain fine sediment, and range from
active stream channels with vertical levees to abandoned
gullies. Unit incises Petrified Springs fault scarp.
Qfi Intermediate Fans with moderately dark tone and rough texture. Holocene to late <100 m
alluvial fans Interfluves are broad, relatively flat, and composed of Pleistocene
pebble to cobble sized clasts forming a desert pavement.
Fluves contain fine sand and silt deposits with well-
rounded channel margins and are mostly abandoned stream
gullies. Surface is offset by a 2-4 m scarp that displays
apparent dextral offset (<90 m) at the Petrified Springs
fault.
Qfo Old alluvial Fan surface displaying light tone and smooth texture. Middle to late <100 m
fans Contains broad flat interfluves that form linear ridges. Pleistocene
Interfluves are deeply incised by fluves (<6 m). Fluves are
spaced far apart and bound by gently rounded margins.
Surface is composed of a well-developed desert pavement
of pebble to cobble sized clasts that are almost all covered
in desert varnish.
QTf Fan alluvium, Undivided fan alluvium. In cross sections only. Holocene to late <1000 m
undivided Pleistocene
Tcg Cobble Rounded to subrounded cobbles, boulders, and pebbles Pliocene- latest < 200 m
conglomerate composed of intermediate lavas with lesser clasts of tuff of Miocene (Ekren
Gabbs Valley and Mesozoic granodiorite and sedimentary and Byers, 1986a;
rocks (Ekren and Byers, 1986a). Payne, 2013)
Tb Basalt Dark gray aphanitic basalt containing sparse phenocrysts Miocene <10 m
of olivine and clinopyroxene (Ekren and Byers, 1986a).
Te Esmerelda Tannish brown to yellowish brown, thin to thickly bedded Miocene- 9.24 to < 1,500 m
formation lacustrine sandstone and siltstone, fluvial and deltaic 12.96 Ma
sandstone, and conglomerate, with ubiquitous beds of (Stewart, 1999)
tephra (Ekren and Byers, 1986a).
Tes Loess-like Light toned, pale-brown to orange-brown, cross-bedded Miocene <200 m
sandstone fine sand and silt deposits. Readily erodes and incised by
channels in dendritic pattern. Derived from re-worked
quartz latite ash (Ekren and Byers, 1985a).
Trl Rhyolite of Purple to light gray flow-laminated rhyolite. Miocene- <10 m
Gabbs Valley 19.2±0.7 Ma
108

(Ekren and Byers,


1980)
Tmi Mafic Two rock types: the younger rock is dark gray, aphanitic, Miocene
intrusions and contains oxidized phenocrysts of hb (as long as 4mm)
in a pilotaxitic-trachytic groundmass of plagioclase and
pyroxene microlites and brown glass; this rock forms a
large volcanic neck that intrudes lavas of Poinsettia mine
area. Along the southern border, several small masses
occur that are probably older than the large volcanic neck;
these exhibit both extrusive and intrusive contacts with
adjacent tuff; most of these contain both hornblende and
clinopyroxene and have seriate plagioclase feldspar; these
masses possibly were feeders for some of the lavas of
Mount Ferguson (Ekren and Byers, 1986a).
Tlf Lavas of Sequence of many volcanic flows with a wide range in Miocene- ages 1000-
Mount composition from basalt to ash-flow tuff, with the majority range from 2000 m
Ferguson, of flows compositionally intermediate. 15.0±0.5 to
undivided 22.5±0.6 Ma
(Ekren and Byers,
1980)
Tlfh Hornblende- Black and dark gray latite containing >2 cm long Miocene- 0-300 m
pyroxene hornblende-pyroxene phenocrysts in an aphanitic silicic 15.0±0.5 to
latite groundmass. 22.5±0.6 Ma
Tlfa Andesitic Brownish-red, forms cliffs, and contains phenochrysts (30- Miocene- 0-500 m
lavas 40% total), including plagioclase (60-70%, 1-5 mm long), 15.0±0.5 to
and orthopyroxene (30-40%, 0.5-3mm). The lower portion 22.5±0.6 Ma
of the andesite unit is relatively massive, displaying no
paleohorizontal indicators, whereas the upper portion
contains excellent 1-20 cm thick flow-foliations.

Tlfbx Volcanic Boulder to pebble sized clasts composed of distinct Miocene- 0-150 m
breccia red/oxidized vesiculated andesitic-basalt lavas. 15.0±0.5 to
22.5±0.6 Ma
Tlfp Hypersthene Dark gray andesite composed of few small phenocrysts of Miocene- 0-50 m
andesite plagioclase, orthopyroxene, and clinopyroxene. 15.0±0.5 to
22.5±0.6 Ma
Tlfl Flow- Cliff-forming brownish-purple dense flow-laminated Miocene- 0-30 m
laminated rhyolite or silicic quartz-latite with phenocrysts of 15.0±0.5 to
quartz-latite plagioclase and biotite. 22.5±0.6 Ma
Tlfb Biotite White and slope-forming nonwelded ash-flow tuff or Miocene- 0-20 m
rhyolite tuffaceous sediment with biotite phenocrysts in an ashy 15.0±0.5 to
matrix. 22.5±0.6 Ma
Tlfq Hornblende Light gray unit with large hornblende phenocrysts 1-4 cm Miocene- 100-200
quartz-latite long. 15.0±0.5 to m
22.5±0.6 Ma
Tml Mafic lavas Mapped to include two distinctive lithologies: the most Miocene 0-75 m
common type is a greenish-gray hornblende andesite and
the other is a gray hornblende andesite (Ekren and Byers,
1986a).
Tcm Tuff of Multiple flow; simple cooling unit of silic quartz latite and Miocene 0-80 m
Copper rhyolite tuff; light reddish gray on fresh surface,
Mountain weathering brownish gray and brown. The most
distinguishing characteristic of this tuff is abundant
accessory sphene, as many as 16 grains per thin section.
Phenocrysts 28-44% total composition: quartz, 3-25%;
alkaline feldspar, 20-30%; plagioclase feldspar, 35-50%;
biotite, 6-9%; hornblende, 1-4%; clinopyroxene, trace-3%;
olivine, 1-2% (Ekren and Byers, 1986a)
Tp Hu-Pwi Tuffs are characteristically deficient in quartz, contain an Oligo-Miocene- <1,000 m
rhyodacite abundance of ferromagnesian minerals, and display perlitic dated between
Pointsettia gray pumice fragments as long as 20 cm. Composed of 30- 22.6-23.3 Ma
109

Tuff Member, 50% phenocrysts: plagioclase (75-85 % total), biotite (6-13 using K-Ar
undivided %), hornblende (trace), clinopyroxene (2-9 %), and methods (Ekren
orthopyroxene (trace). The four units contain essentially and Byers, 1980)
the same phenocryst assemblage and show the same
variations in color, grading from light gray, where
nonwelded or weakly welded, to light brown and brownish
or reddish gray where moderately to densely welded
(Ekren and Byers, 1985a, 1986a).
Tpc Unit C Simple cooling unit of rhyodacite tuff. Oligo-Miocene- 0-150 m
22.6-23.3 Ma
Tpb Unit B Simple cooling unit of rhyodacite tuff that is typically Olio-Miocene- 0-70 m
lighter color than underlying and overlying cooling units; 22.6-23.3 Ma
contains zones in which pumice is pale brown and tuff
matrix is light gray.
Tpa Unit A Simple cooling unit of rhyodacite tuff. Oligo-Miocene- 0-400 m
22.6-23.3 Ma
Tlp Lavas of Gray-purple propylitically altered series of lavas of Oligo-Miocene 0-400 m
Pointsettia intermediate composition. The unit contains 20-30%
Mine phenocrysts: plagioclase feldspar, hornblende,
clinopyroxene, and orthopyroxene that are set in a silicic
groundmass (Ekren and Byers, 1986a).
Tgv Tuff of Gabbs Three major cooling units of highly differentiated rhyolite Oligocene- dated <400 m
Valley, tuff. at 25.15 ±0.06
undivided Ma (Henry and
John., 2013)
Tgv3 Unit 3 Tgv3 is a simple cooling unit of red, densely welded tuff as Oligocene- 25.15 <30 m
much as 30 m thick. Tgv3 contains 8%phenocrysts: quartz, ±0.06 Ma
28%; alkali feldspar, 41%; plagioclase feldspar, 28%;
biotite, 1.5%; opaque minerals, 1.5% (Ekren and Byers,
1986a).

Tgv2 Unit 2 Tgv2 is a compound cooling unit of alternating moderately Oligocene- 25.15 30-300 m
and densely welded pinkish-gray and red devitrified tuff ±0.06 Ma
having a white, partly welded zone locally at base, unit has
numerous lithic fragments of rhyolite and intermediate lava
throughout; it contains 20-35% phenocrysts: quartz, 25%;
alkali feldspar, 39%; plagioclase feldspar, 31%; biotite,
<1% (Ekren and Byers, 1986a).
Tgv1 Unit 1 Tgv1is a simple cooling unit of red to reddish-gray, Oligocene- 25.15 30-65 m
densely welded tuff characterized by abundant lithic ±0.06 Ma
fragments of rhyolite and intermediate lava and a thick
(<30 m) black basal vitrophere. Tgv1 is composed of 4-
11% phenocrysts: quartz, 9-34%; alkalai feldspar, 32-46%;
plagioclase feldspar, 19-42%; biotite, trace; hornblende, 0-
5%; opaque minerals, trace (Ekren and Byers, 1986a).
KJgd Granodiorite Fine-medium-grained and contains abundant hornblende, Cretaceous- dated
biotite, clinopyroxene, and only interstitial quartz (Ekren at 158±4 Ma and
and Byers, 1985a). The relative proportions of phenocrysts 155±7 Ma (Ekren
in KJgd are: quartz, 5-14%; alkali feldspar (perthite and and Byers, 1984)
cloudy orthoclase), 15-30%; plagioclase feldspar (in some
rocks extensively saussuritized), 34-48%; hornblende, 50-
30%; biotite 10-14%; clinopyroxene, 4-18% (Ekren and
Byers, 1985a).
^vc Metasediment Metasedimentary rocks (^vc) consisting of volcaniclastic Triassic
ary facies within the Luning Formation (Muller and Ferguson,
volcaniclastic 1939). Consists of gray, thin-bedded, laminated siltstones,
rocks claystones, sandstones, and pebbly conglomerates (Ekren
and Byers, 1985a). Sandstone and conglomerate layers
display clasts that are moderate to well-rounded and well-
to poorly-sorted.
110

You might also like