You are on page 1of 81

University of Nevada, Reno

Structural Controls of the San Emidio Geothermal System,


Northwestern Nevada

A thesis submitted in partial fulfillment of the


requirements for the degree of Master of Science in Geology

by
Gregory T. Rhodes

Dr. James E. Faulds/Thesis Advisor

December, 2011

THE GRADUATE SCHOOL

We recommend that the thesis


prepared under our supervision by

   

entitled
    
   !
"
be accepted in partial fulfillment of the
requirements for the degree of
#$
% &


James E. Faulds, Ph.D., Advisor

Robert E. Karlin, Ph.D., Committee Member

Mae S. Gustin, Ph.D., Graduate School Representative

Marsha H. Read, Ph. D., Associate Dean, Graduate School

December, 2011

Abstract
Detailed geologic analyses integrated with geophysical and well data have
provided insight into the kinematics, stress state, and structural controls of the San
Emidio geothermal system. San Emidio lies within the Basin and Range province of
northwestern Nevada, ~100 km north-northeast of Reno.

The San Emidio area is

dominated by middle to late Miocene Pyramid sequence volcanic rocks and late Miocene
to recent sedimentary rocks, all overlying Mesozoic metasedimentary rocks. Currently, a
small geothermal power plant produces 3.6 MWe from a 152C reservoir at 520 m depth
at the south end of the active San Emidio fault system. Abundant hydrothermal alteration
along this fault zone, including acid-sulfate leaching and bleaching, native sulfur
deposits, and boiling groundwater at depths less than 100 m, suggest, however, that the
San Emidio geothermal resource extends several kilometers northward from the currently
producing well field. Thermal anomalies identified along this fault zone also extend
northward from the current production zone and intersect alteration around the relatively
young (Pleistocene/Pliocene) Wind Mountain epithermal mineral deposit. Structural and
lithologic similarities are established between this epithermal deposit and the modern
geothermal system.
Kinematic analysis, including slip and dilation tendency analysis, suggest that
north-northeast striking faults and fractures are favorably oriented for fluid flow under
west-northwest-directed extension.

The San Emidio geothermal system and Wind

Mountain epithermal deposit occupy intersections between favorably oriented northnortheast-striking normal faults and multiple closely-spaced north-striking normal faults.

ii

Additionally, these intersections correspond to right steps in hard-linked, north-striking,


normal fault zones along the west flank of the northern Lake Range. Favorably oriented
faults and fractures accompanied by increased fault and fracture density at fault
intersections likely produce the permeability necessary for deep fluid circulation within
the San Emidio system. Preliminary analysis of similar right-steps in mainly northstriking normal fault systems elsewhere within the Basin and Range province suggests
that structural characteristics may serve as a viable and complementary exploration tool
for discovering both epithermal mineral deposits and geothermal systems in extensional
tectonic settings.

iii

Acknowledgments
I want to thank my advisor, Dr. James Faulds, for providing me with the freedom,
guidance, and wisdom necessary to fully develop this work. I also want to thank Nick
Hinz for the assistance and guidance throughout this journey as well as his unmatched
enthusiasm for the study of the earth. Thank you to Dr. Mark Coolbaugh for helping to
characterize the alteration in my study area and greatly expanding my knowledge of
hydrothermal systems. I thank my committee members, Drs. Mae Gustin and Robert
Karlin, for their advice and patience. I also gratefully acknowledge the assistance of
Alan Ramelli, Irene Seelye, Inga Moeck, Mariana Eneva, John Bell, Satish
Pullammanappallil, and Joe Moore. I thank Bill Teplow of U.S. Geothermal, Inc. and Joe
Kizis for providing much of the existing data for this project. This work was partially
funded by the U.S. Department of Energy (Agreement Numbers DE-FG36-02ID14311
and DE-EE 0002847), Nevada Petroleum Society, and American Association of
Petroleum Geologists. I also want to thank my friends at UNR, Betsy Littlefield, Brad
Cantor, and Dr. Sean Long, for fruitful discussions on this work and life as a graduate
student. Finally, I must thank Mallory Principe and my family for their unwavering
support and belief in me.

iv

Table of Contents

1. Introduction

2. Geologic Setting
GEOTHERMAL SYSTEM

5
8

3. Stratigraphic Framework

4. Structural Framework
GEOMETRY
FAULT KINEMATICS
STRESS INVERSION
SLIP AND DILATION TENDENCY
INSAR DATA
SUBSURFACE CONSTRAINTS

13
13
17
18
20
21
23

5. Discussion
TIMING OF DEFORMATION
APPARENT STRUCTURAL CONTROLS
WIND MOUNTAIN EPITHERMAL MINERAL DEPOSIT
BASIN AND RANGE GEOTHERMAL SYSTEM ANALOGS

28
28
28
35
39

6. Conclusions and Implications

41

References

43

Appendix A: Description of Map Units

50

Appendix B: 40Ar/39Ar Geochronology

58

Appendix C: Seismic Reflection Profiles

62

Plate 1

73

List of Figures
1. Introduction
Figure 1.1 Identified structural settings for geothermal systems

Figure 1.2 Location of study area with major tectonic provinces

2. Geologic Setting
Figure 2.1 Geothermal systems in the Black Rock Desert

Figure 2.2 Aerial image of the study area

3. Stratigraphic Framework
Figure 3.1 Simplified geologic map

10

Figure 3.2 Stratigraphic column with geochronologic age spectra

11

4. Structural Framework
Figure 4.1 Stereographic projections of density plots for poles to layering 14
Figure 4.2 Fault map of the study area

15

Figure 4.3 Geologic cross sections

16

Figure 4.4 Stereographic projections of fault planes and slip vectors

18

Figure 4.5 Stereographic projections of stress axes

19

Figure 4.6 InSAR image showing surface deformation

22

Figure 4.7 Seismic reflection profile with interpretation

25

Figure 4.8 Seismic reflection profile with interpretation

26

Figure 4.9 Geophysical pattern

27

5. Discussion
Figure 5.1 Calculated dilational faults in the study area

30

Figure 5.2 Geophysical, temperature, and fault pattern

31

Figure 5.3 Seismic reflection and velocity patterns

32

Figure 5.4 Dilational fault segments and representative textures

33

Figure 5.5 Conceptual model of hydrothermal flow

35

vi

Figure 5.6 Wind Mountain epithermal textures

37

Figure 5.7 Dilational faults and textures at Wind Mountain

38

Figure 5.8 Basin and Range geothermal system analogs

40

1. Introduction
The continued growth of geothermal power generation as an alternative energy
source requires exploration techniques capable of discovering new geothermal resources
and enhancing existing field production. Most operating geothermal fields in Nevada
were initially recognized by hydrothermal surface expressions, such as hot springs.
However, identification of new blind (i.e., hidden) systems is needed for future
development of geothermal resources within the Basin and Range province, as ~35% of
the known geothermal systems with thermal groundwater temperatures 100C lack
surface expression (Coolbaugh et al., 2007) and as much as 75% of the total geothermal
resources in the region may be blind (Coolbaugh and Shevenell, 2004; Coolbaugh et al.,
2007). Many of these hydrothermal sites are spatially related to normal and oblique-slip
faults, suggesting that these faults contribute to and control much of the geothermal
activity in the Basin and Range. Although these fields are known to be structurally
controlled and fault related (Blackwell, 1983; Barton et al., 1995; Curewitz and Karson,
1997; Lowell and Rona, 2005; Faulds et al., 2004, 2006; Bell and Ramelli, 2007),
detailed analyses of the structural controls are lacking for most systems. By identifying
the most favorable conditions for geothermal activity and specific structural controls of
individual geothermal fields, exploration strategies can be refined for the Basin and
Range province and similar tectonic settings.
In an effort to characterize the structural controls of geothermal systems in the
Basin and Range, we have been conducting a comprehensive analysis of geothermal
systems through comparative and detailed studies of representative sites (e.g., Faulds et
al., 2003, 2005, 2006, 2010, 2011; Faulds and Melosh, 2008; Vice et al., 2007; Hinz et

al., 2008). Thus far, normal faulting is the most apparent shared characteristic among
many geothermal fields. However, certain structural settings are more favorable for the
formation of individual geothermal systems. These settings include subvertical fluidflow conduits in highly permeable fault zones oriented approximately perpendicular to
the least principal stress direction (Barton et al., 1995; Faulds et al., 2006). Among faults
most optimally oriented for failure, the most favorable settings identified for geothermal
activity include: 1) discrete steps or relay ramps in steeply dipping normal fault zones, 2)
the horse-tailing terminations of major normal faults, and 3) fault intersections (Figure
1.1) (Curewitz and Karson, 1997; Faulds et al., 2006, 2011). These structural settings are

Figure 1.1 Identified favorable structural settings for geothermal systems in the Great Basin (from Faulds
et al., 2006).

commonly expressed topographically as: 1) major steps in range-fronts, 2) interbasinal


highs, 3) mountain ranges consisting of relatively low, discontinuous ridges, and 4)
lateral terminations of mountain ranges (Faulds et al., 2006).

In addition, most

geothermal systems have a spatial association with Holocene normal and oblique-slip
faults (Bell and Ramelli, 2007).
As part of this larger study, the San Emidio geothermal field, located ~100 km
north-northeast of Reno (Figure 1.2), provides an opportunity to examine the structural

Figure 1.2 Digital elevation model showing the study area, ~100 km north-northeast of Reno (outlined in
blue), and major tectonic provinces. The inset shows a National Agriculture Imagery Program (NAIP)
image of the study area.

controls of a commercially operative geothermal system with the potential for enhanced
exploitation. San Emidio lies in the actively extending Great Basin of northwestern
Nevada and occurs along a northerly striking normal fault system that bounds the
northern part of the Lake Range. The main goals of this study were to: 1) determine the
ages of post-Mesozoic stratigraphy, faulting, and alteration in the northern Lake Range
and San Emidio Desert, 2) provide a more complete understanding of the geometry and
kinematics of fault systems in the San Emidio area, 3) determine whether dilational fault
segments or multiple fault intersections control the geothermal system, 4) elucidate
features conducive to fluid flow, 5) examine the evolution of the San Emidio geothermal
system, and 6) refine exploration models in extensional settings based on our findings.
In order to elucidate the structural controls of the San Emidio geothermal field,
detailed analyses were applied to the eastern San Emidio Desert and adjacent northern
Lake Range. This study involved: 1) detailed geologic mapping of ~100 km2 at 1:24,000
scale to define stratigraphy and faults (Plate 1), 2) structural analysis to delineate the
geometry and kinematics of faults and estimate principal stress orientations in the area, 3)
petrographic analysis to define and correlate stratigraphic units and styles of
hydrothermal alteration, 4) analysis and integration of available seismic reflection,
gravity, magnetic, and well data to better define the subsurface geometry of fault zones,
5)

40

Ar/39Ar geochronology to assist in correlating strata, defining offsets on faults, and

constraining the timing of both deformation and alteration, and 6) GIS compilation of the
geologic map, including the extent of hydrothermal alteration, to facilitate interpretations
of the structural controls on geothermal activity.

2. Geologic Setting
The San Emidio Desert is located within the Basin and Range province of
northwestern Nevada directly south of the Black Rock Desert and ~100 km northnortheast of Reno. San Emidio lies in the southwestern part of a series of geothermal
systems (Figure 2.1) described by Faulds et al. (2006) as the Black Rock Desert

Figure 2.1 Shaded relief


image of northwestern Nevada
and surrounding areas showing
a series of geothermal systems
referred to as the Black Rock
Desert (BRD) geothermal belt
(modified from Faulds et al.,
2004).
Yellow circles are
geothermal
systems
with
o
temperatures of 100-160 C; red
circles are geothermal systems
o
with temperatures >160 C;
PLFZ-Pyramid Lake Fault
Zone.

geothermal belt, which is dominated by north-northeast-striking normal faults locally


linked by east-northeast-striking sinistral-normal faults. The San Emidio geothermal
system is expressed as a) abundant silicified sediments, primarily within the Wind
Mountain gold and silver deposit, and b) native sulfur, acid-sulfate alteration, and
subaqueous spring deposits (tufa) along a Holocene fault scarp at the eastern edge of the
San Emidio Desert (Figure 2.2). The San Emidio Desert is bounded to the west by the
Fox Range and to the east by the northern Lake Range.

Figure 2.2 NAIP image showing the study area outlined in blue, surficial hydrothermal alteration and the
Wind Mountain epithermal deposit in orange, and the geothermal power plant in black. Wells of interest to
this study are also shown. The San Emidio Desert is bounded to the west by the Fox Range and to the east
by the northern Lake Range.

The stratigraphy in the surrounding ranges consists of Mesozoic sedimentary,


volcanic, and intrusive units nonconformably overlain by Miocene volcanic and
sedimentary rocks (Bonham and Papke, 1969; Drakos, 2007; Vice, 2008). This study
focuses on the eastern half of the San Emidio Desert and adjacent northern Lake Range.
The Lake Range, ~60 km long and up to 12 km wide, trends approximately north to
north-northeast typical of other ranges in the region north of the Pyramid Lake fault zone.
The Pyramid Lake fault zone is a major northwest-striking right-lateral fault in the
northern Walker Lane (Figure 1.2) (Bell and Slemmons, 1979; Briggs and Wesnousky,
2004; Faulds et al., 2005a). Briggs and Wesnousky (2004) suggested that the Pyramid

Lake fault zone accommodates a major portion of contemporary right slip in the northern
Walker Lane. Based on the offset of tuff-filled Oligocene paleovalleys, the Pyramid Lake
fault zone appears to have accommodated ~10 km of dextral slip since ~9 to 3 Ma
(Faulds et al., 2005a).
The northern Walker Lane is a system of overlapping, left-stepping, northwest
striking dextral faults that accommodate ~15-25% of Pacific-North American plate
motion (Faulds et al., 2005a; Faulds and Henry, 2008). The northwest-striking dextral
faults of the northern Walker Lane, including the Pyramid Lake fault zone, terminate in
arrays of north-striking normal faults suggesting that northwest directed dextral shear
from the northern Walker Lane is transferred to the western Basin and Range and
manifested as northwest-directed extension (Faulds et al., 2004, 2005b, 2006; Drakos,
2007). For example, the Pyramid Lake fault zone dies out northward in the Pyramid
Lake area, where it merges with a system of primarily west-dipping normal faults.
Cumulative extension accommodated by this normal fault system is similar to the
magnitude of dextral shear on the Pyramid Lake fault (Drakos, 2007), which suggests
that the normal and strike-slip fault systems are kinematically related.
Geodetic studies of the northern Walker Lane indicate that the ~100 km wide
zone experiences a relatively high rate of northwest-directed dextral shear, approximately
10 mm/year (Bennett et al., 2003; Hammond and Thatcher, 2004; Hammond et al., 2009;
Kreemer et al., 2009).

East of the Walker Lane, approximately 3 mm/yr of west-

northwest to east-west directed extension is variably distributed across the northern Basin
and Range (Hammond and Thatcher, 2005). Within this transtensional tectonic setting
between northwest-directed dextral shear to the west and west-northwest to east-west

extension to the east, the extension direction in the San Emidio area is oriented
approximately N71W (Ichinose et al., 2003; Kreemer et al., 2006, 2009).

GEOTHERMAL SYSTEM
The San Emidio geothermal field contains two apparent, potentially related
thermal anomalies (Fig. 2.2): 1) a southern anomaly and associated small power plant;
and 2) a northern anomaly near Wind Mountain. The southern system was initially
exploited for power generation in 1988 using a 141C resource at 152 m depth.
Subsequent drilling resulted in the discovery of 152C geothermal fluids at 520 m depth.
Currently, a small power plant produces 3.6 MWe from this reservoir. The original wells
at 152 m depth are currently utilized as injection wells (Fig. 2.2).

Abundant

hydrothermally altered rocks and native sulfur deposits occur in the vicinity of the current
production wells. Temperature reversals and a silica cap in the production and injection
wells suggest that shallow geothermal outflow occupies a depth zone from ~35-100 m.
Shallow temperatures (~100 m depth) decrease away from the current production wells.
Geothermal upwelling at a depth greater than 520 m has yet to be discovered. The
altered rocks and native sulfur deposits associated with the southern anomaly extend ~4
km northward presumably along a narrow fault zone. The northern-most expression of
this narrow altered zone merges with the northern thermal anomaly ~0.75 km west of
Wind Mountain (Fig. 2.2).

Three former mine-water supply wells in this northern

anomaly have a maximum temperature of 116C at a depth of 100 m. An important


question is whether the northern and southern thermal anomalies are part of one larger

system associated with a single zone of upwelling or whether they represent two discrete
geothermal systems.

3. Stratigraphic Framework
The eastern San Emidio Desert and adjacent northern Lake Range are composed
of Mesozoic metasedimentary basement, Tertiary volcanic and sedimentary rocks, and
Quaternary alluvium, lacustrine sediments, and spring deposits (Figures 3.1 & 3.2)
(Appendices A and B).

The basement in the Lake Range consists of Mesozoic

metasedimentary rocks assigned to the Triassic-Jurassic Nightingale Sequence (Bonham


and Papke, 1969). This sequence, exposed along the western side of the Lake Range,
consists of a thick section of metamorphosed and folded low-grade, argillaceous phyllite,
with some slate and schist, intercalated carbonate and quartzite horizons, and felsic
intrusions (Moore, 1979; Wood, 1990).
A thick section (~1.5 km) of Tertiary volcanic and sedimentary rocks
nonconformably overlies the Mesozoic basement. It can be separated into an isolated late
Oligocene dacite flow, the middle Miocene Pyramid sequence (Bonham and Papke,
1969; Drakos, 2007), and a late Miocene section of

mainly sedimentary rocks

temporally correlative with the Truckee and Coal Valley Formations (Moore, 1979). The
basal dacite flow crops out in only the south-central part of the map area. Hornblende
from this flow yielded an 40Ar/39Ar age of 24.10.4 Ma (Figure 3.2).

10

Figure 3.1 Simplified geologic map of the northern Lake Range and eastern San Emidio Desert showing
locations of dated samples, production and injection wells, and seismic reflection lines. See Figure 3.2 for
unit descriptions. Bar and balls shown on downthrown sides of normal faults.

11

Figure 3.2 Generalized stratigraphic column of the northern Lake Range showing relative unit
thicknesses, approximate degree of tilting, and 40Ar/39Ar age spectra and ideograms for dated samples (A
through E). Lithologic units: TrJn-Nightingale metasedimentary rocks; KTr-rhyolite dikes; Td-dacite
(sample E); Tpts-tuffaceous volcaniclastic rocks; Tpb-basalt; Tpp-porphyritic basaltic andesite (sample B);
Tpbl-sparsely porphyritic basaltic andesite lower; Tpb-sparsely porphyritic basaltic andesite (sample C);
Tpts-tuffaceous volcaniclastic rocks upper; Tpd- dacite (sample B); Tcla-clay-rich sedimentary rocks; Tssedimentary rocks; Tss-silicified sedimentary rocks and siliceous sinter; Tr-rhyolitic tuff (sample A); QTaLate Tertiary-Quaternary basin-fill deposits undivided; Qa-Quaternary deposits.

The Pyramid sequence dominates the northern Lake Range and primarily consists
of basaltic andesite flows and flow breccias and lesser non-welded tuffs and tuffaceous
sedimentary rocks. The lower portion of the Pyramid sequence is mainly a volcaniclastic
assemblage (e.g., tuffaceous sandstone, non-welded pumiceous ash-flow tuff, and
conglomerate) hosting three distinct, semi-continuous, mostly aphanitic basaltic andesite
lenses, generally overlying the Mesozoic basement. These rocks locally pinch out against
paleo-topographic highs in the Mesozoic basement. Above the sedimentary units, the

12

Pyramid sequence chiefly consists of sparsely porphyritic, olivine basalt and basaltic
andesite lavas with subordinate intercalated sedimentary units. Recent

40

Ar/39Ar

geochronology performed throughout the Lake Range and nearby areas has yielded ages
ranging from 13-16 Ma for the Pyramid sequence (Henry et al., 2004; Drakos, 2007;
Vice, 2008).
In the northernmost part of the Lake Range, the Pyramid sequence is overlain by a
younger sedimentary section that may temporally correlate with the upper Miocene
Truckee Formation (Bonham and Papke, 1969; Moore, 1979; Wood, 1990). This section
consists of conglomerate, sandstone, clay-rich siltstone, and a lens of non-welded
rhyolitic ash-flow tuff, which has yielded an age of 4.80.9 Ma from feldspar (Figure
3.2).
The Wind Mountain deposit consists of highly silicified strata (with intercalated
sinter horizons) and lesser argillic alteration of the upper Miocene/Pliocene sedimentary
section. Hydrothermal alteration mineralization within this area includes chalcedony,
opal, siliceous sinter, quartz, calcite, montmorillonite, kaolinite, illite, and lesser sulfur
and gypsum.

As a result of the silicification, these strata are much more resistant than

the typical late Miocene sedimentary rocks and thus form a broad ridge along the
northwest flank of the Lake Range (Wood, 1990). Additionally, organic material, such as
reeds, and travertine-like structures have been preserved by silicification in the Wind
Mountain deposit.
Quaternary sediments onlap older strata along the western margin of the Lake
Range. These include alluvial fan deposits of varying ages, recent playa sediments, and
Pleistocene Lake Lahontan silt, sand, and tufa.

Isolated tufa mounds, abundant

13

hydrothermally altered rocks, silicified Lake Lahontan sand, and cinnabar and native
sulfur deposits crop out along the eastern margin of the San Emidio Desert and mark the
zone of geothermal activity. This narrow zone of alteration has been extensively leached
and bleached by acid-sulfate fluids and includes kaolinite, gypsum, siliceous sinter, and
opal.

4. Structural Framework
GEOMETRY
The Lake Range consists of a well exposed series of mainly east-tilted (~20-30),
northerly trending fault blocks (Figure 4.1) bounded by moderately to steeply westdipping, en echelon normal faults (Drakos, 2007). The west side of the northern Lake
Range is bounded by a series of northerly striking, steeply west-dipping normal faults,
which expose relatively resistant Mesozoic and Tertiary strata (Figure 4.2 & 4.3). The
eastern-most of these normal faults is the range-bounding fault, here referred to as the
northern Lake Range fault. The east side of the northern Lake Range is an east-dipping,
dip-slope (~30) formed by resistant Pyramid sequence lavas overlain by a thin veneer of
Quaternary alluvium.
The most apparent exception to the northerly striking, west-dipping normal faults
in the northern Lake Range is a north-northeast to east-northeast-striking (~N20-80E)
sinistral-normal oblique-slip fault, here referred to informally as the Wind Mountain stepover fault zone (Figure 4.2A). This fault provides hard-linkage between two underlapping, en echelon strands of the northern Lake Range fault, resulting in a right-step in

14

the range-front. The faulted step-over in the range-front fault zone within this area is
analogous to a breached relay ramp (e.g., Larsen, 1988; Peacock and Sanderson, 1994).

Figure 4.1 Lower-hemisphere, equal-area stereographic projections of density plots for poles to bedding,
volcanic layering, and metamorphic foliation. A. Late Miocene/Pliocene sedimentary units (Ts); B. middle
Miocene flow-foliated sparsely porphyritic basaltic andesite (Tpb); C. Middle Miocene tuffaceous
volcaniclastic units (Tpts); D. foliation in Nightingale metasedimentary rocks (TrJn). n=number of
measurements. Density contour interval equals 2% of the data per 1% area (e.g. for C: contours =
2,4,6,8,10,12,14,16,18,20%).

15

Figure 4.2 Shaded relief map of the eastern San Emidio Desert and northern Lake Range showing
mapped faults in black; major faults are shown in blue; SEF-San Emidio fault; NLRF-northern Lake
Range fault. Inset A NAIP image with mapped faults showing orientations of sliplines; WMSF-Wind
Mountain step-over fault just southeast of the Wind Mountain mine. Inset B NAIP image with mapped
faults showing the Empire fault (EF), San Emidio fault (SEF), geothermal power plant (black box), and
wells.

16

Figure 4.3 Geologic cross sections A-A (top) and B-B (bottom) based on detailed geologic mapping,
seismic reflection and velocity data, and well logs. The sections reveal east-tilted strata cut by westdipping normal faults in the eastern San Emidio Desert and northern Lake Range. See Plate 1 and Figure
3.2 for explanation of units.

The hanging wall of this right-step contains both the upper Miocene/Pliocene
sedimentary section and the Wind Mountain epithermal mineral deposit. The northern
boundary of the epithermal deposit is also a northeast-striking fault of lesser strike-length
than that to the south. North of the right-step and Wind Mountain step-over fault,
normal, down-to-the-west displacement on the main range-bounding fault decreases
northward but is accommodated by multiple normal fault strands to the west in the
vicinity of Wind Mountain. Range-front fault displacement continues to decline farther

17

north resulting in the northern Lake Range fault terminating in a north-plunging anticline
(Fig. 3.1).
Another important exception to the northerly striking, west-dipping normal faults
is a north-northeast-striking, west-dipping normal fault, here referred to informally as the
Empire fault, in the vicinity of the current production and injection wells for the
geothermal power plant (Figure 4.2B). The hanging wall to the intersection between this
north-northeast-striking fault and multiple north-striking normal fault strands hosts the
current geothermal production wells and is marked by abundant hydrothermal alteration
and silicified alluvium. The bulk of the mapped surficial hydrothermal alteration lies
north of the Empire fault along a north-trending, west-dipping, curvilinear Holocene
scarp, here referred to as the San Emidio fault, which lies approximately 1 km west of the
range-front strand of the northern Lake Range fault (Fig. 4.2).

Though the north-

northeast-striking fault and the majority of the north-striking faults lack well-exposed
surficial expressions in the vicinity of the geothermal production area, geophysical and
well data confirm both their presence and the relatively high fault and fracture density of
the area, as evidenced by gravity gradients, offset seismic reflectors, and lost circulation
zones within the wells.

FAULT KINEMATICS
In order to analyze the kinematics and characterize fault populations, the attitudes
of exposed faults were measured along with the orientations of observed kinematic
indicators, such as slickenlines, rough facets, and Riedel shears (Figure 4.4A) (e.g.,
Angelier et al., 1985; Gauthier and Angelier, 1985; Petit, 1987). Analysis of stratigraphic

18

separation and kinematic data indicate a west-northwest-trending slip direction on most


faults (Figure 4.4B), approximately parallel to the regional extension direction inferred
from geodetic data and available earthquake focal mechanism solutions (Ichinose et al.,
2003; Kreemer et al., 2006, 2009). However, a subpopulation of faults has a westsouthwest-trending slip direction (Figure 4.4C).

Figure 4.4 Lower-hemisphere, stereographic projections of fault planes (great circles) and slip vectors
(arrows), as deduced from kinematic indicators (e.g., striae, Riedel shears, and rough facets). A. All fault
measurements. B. Faults indicating west-northwest slip direction. C. Faults indicating west-southwest slip
direction. D. Representative fault surface with slickenlines indicating dip-slip motion. N=number of
measurements.

STRESS INVERSION
The kinematic data were used to estimate principal stress orientations and stress
ratios (Figure 4.5). Principal shortening and extension axes were determined for each

19

fault to identify kinematic consistencies and heterogeneities within the slip data (e.g.,
Marrett and Allmendinger, 1990). Heterogeneous fault sets, potentially indicating stress
fields of different ages, were separated into homogeneous data sets using the PTB

Figure 4.5 - Lower-hemisphere, stereographic projections of P-compressional(blue dots) and T-shear(red


dots) axes for calculated linked Bingham strain axes derived from kinematic indicators of measured faults.
Average principal stress orientations: 1-P axis; 2-Intermediate axis; 3-T axis. A. Fault population
indicating a west-northwest-trending least principal stress (T-axis). B. Fault population suggesting a westsouthwest-trending least principal stress. N=number of measurements.

(P=Pressure T=Tension B=Intermediate) method (Sperner et al., 1993), Hoeppner


diagrams (Meschede, 1994), and fluctuation histograms (Etchecopar et al., 1981).
Statistical analysis of the calculated versus measured strain axes and stress ratios suggest
that the decomposed homogeneous data sets represent single stress regimes (e.g., Yamaji,
2000). This method of stress inversion was performed with the computer programs
TectonicsFP 1.5 of Reiter and Acs (1999) and MIM (Yamaji et al., 2005).
Analysis of the local stress field in the San Emidio area reveals an extensional
stress regime with a west-northwest-trending (azimuth of ~299o) least principal stress.
However, heterogeneity within the fault slip data indicates a secondary and less

20

statistically substantiated extensional stress regime with a west-southwest-trending


(~250o) least principal stress.

SLIP AND DILATION TENDENCY


By analyzing the behavior of faults within the current stress field, the tendency of
individual fault orientations to slip or dilate was determined for the San Emidio area (e.g.,
Morris et al., 1996, Moeck et al., 2009). The slip tendency is defined by the ratio of shear
stress to the calculated normal stress on a particular fault segment and indicates whether
or not that segment is critically stressed (i.e., likely to slip). The likelihood that certain
fault orientations will slip under the current stress field provides a better understanding of
fault reactivation potential during any adjustments in fluid pressure associated with
production and/or injection within the geothermal field (Moeck et al., 2009). The slip
tendency may also provide an indication of the permeability potential of faults that have
been favorably oriented for slip throughout the duration of a certain stress regime (Barton
et al., 1995; Caine et al., 1996; Curewitz and Karson, 1997). Those favorably oriented
and critically stressed faults with a higher slip tendency could potentially have a more
protracted slip history and thus possibly a better developed fracture network with
sustained and renewed permeability. In the San Emidio area, the faults with the highest
slip tendency have a north-northeast (~17-36) strike and bi-modal 65 dip.
Dilation tendency analysis indicates which fault and/or fracture orientations have
the potential to dilate under the current stress field and act as conduits for geothermal
fluids (Gudmundsson, 2002; Moeck et al., 2009, 2010). This technique, along with
analysis of subsurface data revealing fault dip variations (i.e., refracted or steepened fault

21

segments in more competent units), allow for the identification of potentially dilatant
zones at depth (Ferrill and Morris, 2003). In the San Emidio area, the faults and/or
fractures with the highest dilation tendency have a north-northeast (~18-42) strike and
bi-modal 88 dip.

INSAR DATA
This study incorporated satellite radar interferometry (InSAR) analysis performed
by Mariana Eneva of Imageair, Inc.

This technique utilized satellite radar scenes

collected from 1992 to 2001 to measure millimeter-scale ground surface deformation of


permanent radar signal reflectors, such as boulders or buildings. This technique was
intended to identify uplift or subsidence related to tectonic processes or geothermal
production and injection activities. This analysis identified a zone of subsidence that
centers around the current production wells and extends to the north along the San
Emidio fault zone (Eneva et al., 2011) (Figure 4.6). The analysis also suggests that
mapped faults are influencing hydrologic properties and groundwater flow in the eastern
San Emidio Desert.

22

Figure 4.6 InSAR results showing surface deformation from 1992 to 2001 with seismic reflection lines,
mapped faults, wells, and the San Emidio fault zone (SEF) (modified from Eneva et al., 2011). Rates are
mapped by color and shown in mm/year. Yellow to red colors indicate negative movement away from the
satellite (subsidence) and green to blue colors indicate positive movement.

23

SUBSURFACE CONSTRAINTS
This study incorporated available geophysical and well data, including seismic
reflection, gravity, magnetic, and shallow temperature surveys. These data, combined
with the detailed geologic mapping, allowed for the production of detailed cross sections.
These data sets confirm the presence of the mapped north-striking, west-dipping normal
faults cutting east-dipping strata in the San Emidio area and revealed additional subparallel, west-dipping normal faults.
As part of this study, ~32 km of active source seismic data were acquired in the
eastern San Emidio Desert (Optim Inc., 2011). Ten profiles, each approximately 3048 m
(10,000 feet) in length, had 49 shot points per line with a geophone spacing of 17 m (55
feet) and a shot point spacing of 67 m (220 feet). Three-component (3C) geophones were
used to record data at each station. Shots were generated using a vibroseis truck, and data
were recorded into each geophone three times, including an initial shaking in the vertical
direction followed by shaking in both the horizontal inline (along the geophone array)
and horizontal cross-line (perpendicular to the geophone array) directions.

Three

component shot sources included two P-wave vibroseis trucks operating in series to
generate energy in the vertical direction and two shear-wave vibroseis trucks operating in
the inline and cross-line directions. Data were acquired for 6 seconds at a 2 millisecond
sample rate. First arrival times picked from raw shot gathers were used to produce
velocity models using SeisOpt 2D processing software. First arrival information
from both P-wave and SH-wave data was employed to derive P-wave and S-wave
velocities under each line. This allowed tomographic estimation of compressional (Vp)
and shear (Vs) wave velocities for the stratigraphic section. The P-wave velocity model

24

was then put through a Kirchhoff pre-stack depth migration to derive images of faults and
fractures. The pre-stack migration algorithm incorporated the velocity models for
accurate calculation of travel times down to and up from every point within the reflection
data volume. It produced images by summing the value of seismograms within the data
volume at discrete points of time based on travel-time calculations through the velocity
model. Because the pre-stack migration is free of assumptions about dip of bedding and
structure, it generated images that reveal the true-depth location and geometry of variably
oriented features.
For interpretive purposes, these depth-migrated seismic reflection profiles along
with velocity models were incorporated with available well data into OpendTect, open
source seismic interpretation software from dGB Earth Sciences.

Profiles were

interpreted by analyzing prominent reflectors, identifying offset reflectors, and applying


detailed geologic map, well data, seismic velocity, and gravity survey information
(Figures 4.7 and 4.8) (Appendix C). Faults were interpreted on the basis of offset
reflection packages, changes in velocity, and/or some steeply dipping reflectors within
the profiles. Many faults imaged relatively well. However, major stratigraphic units
were generally poorly imaged and more difficult to interpret. Problems in imaging the
strata probably resulted from attenuation of seismic waves within the relatively thick
volcanic section of the Pyramid sequence as well as problems in velocity stacking due to
the lower velocity sedimentary package underlying the higher velocity basalts within the
Pyramid sequence. As a result of these problems, interpretations of stratigraphic units in
the subsurface of the San Emidio basin should be considered with caution, although well

25

data did provide important constraints in some areas. Interpreted profiles were directly
imported into illustration software to facilitate construction of detailed cross sections.

Figure 4.7 A. Seismic reflection profile (2-2) with color velocity model overlay. B. Interpreted seismic
reflection profile showing faults, relative motion, and stratigraphic horizons. Lithologic units: TrJnNightingale metasedimentary rocks; Tptsu-tuffaceous sedimentary units undivided; Tpb-Pyramid
sequence volcanic rocks; QTa-undivided basin-fill deposits.

Additionally, analysis of gravity, magnetic, and well data reveal zones of


hydrothermal alteration along closely spaced fault zones to the north, south-southwest,

26

and beneath the current geothermal production zone. The identified intersection of the
north-northeast-striking Empire fault with the San Emidio fault appears as a prominent
north- to north-northeast-trending dogleg in the gravity, magnetic, and shallow
temperature surveys (Figure 4.9 A-C).

Figure 4.8 A. Seismic reflection profile (7-7) with color velocity model overlay and production wells.
B. Interpreted seismic reflection profile showing faults, relative motion, and stratigraphic horizons.
Lithologic units: TrJn-Nightingale metasedimentary rocks; Tptsu-tuffaceous sedimentary units undivided;
Tpb-Pyramid sequence volcanic rocks; QTa-undivided basin-fill deposits.

The geothermal production zone is also identified in the seismic reflection


profiles and associated velocity model as a well-defined low velocity zone, which also

27

corresponds to the interpreted closely spaced faults and the highest identified thermal
anomaly.

Conversely, a slightly higher velocity zone with prominent reflectors at

shallower levels is observed in several seismic reflection profiles (5-9) along the San
Emidio fault to the north and along the Empire fault to the south-southwest of the current
production zone, respectively. This may delineate a silicified zone along the San Emidio
fault, which is compatible with several other features along this fault, including a)
extensive surface alteration, b) a thermal anomaly, c) well logs indicating a silicified zone
at ~35-100 m depth, and d) a slight gravity high (or plateau) associated with increased
density due to silicification in basin-fill sediments (e.g., Locke et al., 1999; Wood and
Thomas, 2002). The San Emidio fault also corresponds to a zone of low magnetic
intensity, which may indicate that hydrothermal fluids have oxidized magnetite within
volcanic units along the fault zone (Henkel and Guzman, 1977; Grant, 1985).

Figure 4.9 Shaded relief map of the study area showing mapped faults in black. These contours are
shown without contour intervals to simply illustrate the pronounced dogleg at the intersection of the San
Emidio (SEF) and Empire faults (EF). A. General gravity survey pattern in green. B. General ground
magnetic survey pattern shown in purple. C. Shallow temperature survey pattern shown in orange. All
indicate a distinct north to north-northeast-trending dogleg at the intersection of the San Emidio (SEF) and
Empire faults (EF). The current production wells occupy this fault intersection.

28

5. Discussion
TIMING OF DEFORMATION
40

Ar/39Ar geochronologic analysis and structural data in this study reveal that

volcanic and sedimentary strata of the Pyramid sequence were deposited ~14-16 Ma and
have concordant tilts of ~30 east. This suggests that major extension postdates ~14 Ma.
This is compatible with previous studies that indicated extension began ~12 Ma in
northwestern Nevada (e.g., Duffield and Mckee, 1986; Henry and Perkins, 2001; Trexler
et al., 2000; Colgan et al., 2006). Also, the ~10 east tilt of the late Tertiary sedimentary
section (Ts), as compared to the ~30o tilt of the middle Miocene Pyramid sequence,
indicates that significant extension occurred between ~14 and 4.8 Ma. The tilting and
faulting of these early Pliocene sedimentary rocks and abundant Quaternary fault scarps
further demonstrate that extension continued after ~4.8 Ma and into the Holocene.

STRUCTURAL CONTROLS ON GEOTHERMAL ACTIVITY


The aim of this study was to evaluate which structures permit the deep circulation
and ascent of geothermally heated meteoric fluids and why. Because previous work has
indicated that many geothermal systems occur in highly permeable, critically stressed,
and/or dilatant fault zones oriented approximately perpendicular to the least principal
stress direction (e.g., Barton et al., 1995; Gudmundsson, 2002; Sheridan and Hickman,
2004; Faulds et al., 2006; Moeck et al., 2009, 2010), detailed geologic mapping,
structural analysis of faults, and slip and dilation tendency analyses were integrated with
geophysical and well data to provide a better understanding of the kinematics, stress
state, and structural controls of the San Emidio geothermal field.

29

The kinematic data, along with available geodetic data (Kreemer et al., 2006,
2009) and earthquake focal mechanism solutions (Ichinose et al., 2003), have revealed a
west-northwest-trending extension direction and least principal stress. A subpopulation
of faults suggests west-southwest-trending extension and may reflect an earlier period of
west-southwest-oriented extension that affected the northern Basin and Range prior to
~10-7 Ma (Zoback et al., 1981; Bellier and Zoback, 1995). Slip and dilation tendency
analysis using the west-northwest orientation indicates that moderately to steeply dipping
fault segments that strike north-northeast have the highest tendency to slip and dilate
(Figure 5.1). It is important to note that the predominant north-striking normal fault
zones in the San Emidio area are not completely orthogonal to the current westnorthwest-trending extension direction.

Moreover, both the current geothermal

production zone and the Wind Mountain epithermal deposit occur at discontinuities in the
north-striking normal fault zones that contain north-northeast-striking fault segments.
A significant shared structural characteristic between the two apparent loci of
geothermal/hydrothermal activity is the intersection of a dilational north-northeaststriking normal fault segment with multiple closely spaced roughly north-striking normal
faults. At the current production zone, this fault intersection is manifested as abundant
hydrothermal alteration, north- to north-northeast-trending gravity, magnetic, and shallow
temperature anomalies, and economically viable geothermal temperatures and flow rates
(Figure 5.2 A, B). For example, the shallow temperature survey reveals a distinct

30

Figure 5.1 Shaded relief topographic map showing mapped faults. Faults with the highest calculated
dilational tendency are shown in red. Bar and balls are shown on downthrown sides of faults.

temperature anomaly that marks this fault intersection. On seismic reflection profiles,
this fault intersection appears as a zone of poor reflectors and low velocity, which

31

corresponds to intervals of lost circulation in available well data (Figure 5.3). These data
suggest that this fault intersection is an area of high fracture density, increased
permeability, hydrothermal alteration, and hydrothermal flow. This zone of increased
hydrothermal flow is also compatible with calculations indicating that the northnortheast-striking Empire fault is more prone to slip and dilation.

Figure 5.2 NAIP image showing mapped faults with calculated dilational fault segments shown in red.
A. Gravity survey contours (contour interval=0.5 mgal) shown in green and mapped surficial hydrothermal
alteration outlined in purple; the gravity high along the San Emidio fault (SEF) deflects to the southsouthwest at the intersection with the north-northeast-striking Empire fault (EF). B. General ground
magnetic survey contours shown in purple and shallow temperature contours (contour interval=22C) in
orange; both the magnetic low and shallow temperature contours along the SEF deflect to the southsouthwest at the intersection with the EF. Production wells are labeled in B with the highest measured
temperature in C.

At the Wind Mountain epithermal deposit, the intersection of the dilational


~north-northeast-striking normal fault with closely spaced north-striking normal faults is
manifested as high fracture and fault density, pervasive silica flooding, sheeted veins of
rhombohedral calcite, silicified fault zones exhibiting hydrothermal brecciation, and
open-space growth textures including sheeted veins of rhombohedral calcite, banded

32

Figure 5.3 Seismic reflection profile 7-7 with colored velocity overlay, interpreted W-dipping faults and
E-dipping strata in black, low velocity zone exhibiting poor reflector quality outlined in purple, and lost
circulation zones encountered during drilling of the production wells outlined in light blue. Lithologic
units: TrJn-Nightingale metasedimentary rocks; Tptsu-tuffaceous sedimentary units undivided; TpbPyramid sequence volcanic rocks; QTa-undivided basin-fill deposits.

amorphous silica gel drip structures, and silica molds and pseudomorphs of platy calcite
(Figure 5.4). The majority of these features appear concentrated in the north-northeast-to
northeast- striking portion of the mainly east-northeast-striking Wind Mountain step-over
fault. This part of the Wind Mountain step-over fault zone appears to have hosted
abundant and prolonged hydrothermal flow.

This silicified fault zone appears as a

topographically elevated arcuate fin, which serves as a link between the north-striking
northern Lake Range fault and the east-northeast-striking Wind Mountain step-over fault.
Calculations also indicate that this north-northeast- to northeast-striking fault segment has
a high slip and dilation tendency under the current stress regime (Figure 5.4).

33

Figure 5.4 A. NAIP image showing mapped faults with dilational fault segments in red; WMSF-Wind
Mountain step-over fault; SEF-San Emidio fault; NLRF-northern Lake Range fault. Letters on image A
correspond to locations of observed representative textures shown in B through G. B. Wind Mountain
open-pit mine showing highly-fractured and pervasively silicified sedimentary units. C. Sheeted calcite
veins at inferred dilational fault segment. D. Silicified fault surface with preserved fault striae. E.
Silicified fault breccia. F. Silica-gel drip structures indicative of open-space growth. G. Chalcedony mold
of rhombohedral calcite.

In summary, at both the current geothermal production zone and the Wind
Mountain mineral deposit, the correspondence of significant hydrothermal alteration with
predicted high slip and dilation tendency fault segments indicate that the structural

34

complexities at these two locales control fluid flow. Specifically, it appears that the
favorably oriented north-northeast- to northeast-striking normal fault segments act as the
main conduits for hydrothermal ascent at the current geothermal production zone and
Wind Mountain (Figure 5.5). Away from these apparent upflow zones, hydrothermal
flow is locally channeled along closely spaced north-striking fault segments, as evidenced
by surficial hydrothermal alteration, InSAR data suggesting some subsidence associated
with geothermal production, and anomalous shallow temperatures along the San Emidio
fault zone. This fault zone may be more permeable due to interfering fracture networks
between closely spaced north-striking normal faults (e.g., Fairley and Hinds, 2004). Near
the Wind Mountain epithermal deposit, this flow path is indicated by a decreasing
volume of silicified hydrothermal breccias and sheeted calcite veining northward away
from an apparent upflow zone at the Wind Mountain step-over fault (Figure 5.4).
The protracted record of hydrothermal flow at Wind Mountain and the close
spatial relationship with the active modern geothermal system suggest that this area has
long-term favorability for geothermal production. The northern thermal anomaly appears
to currently reside to the west of Wind Mountain. This may indicate a westward
migration of the locus of hydrothermal activity through time, possibly resulting from
sealing of older fluid flow conduits by mineral precipitation. It is possible that the
southern anomaly may represent lateral fluid flow from the Wind Mountain hydrothermal
upwelling zone, as abundant evidence for extensive lateral flow is observed along the San
Emidio fault zone (e.g., mineral alteration and subsidence). Alternatively, the southern
anomaly may represent the current and singular upwelling zone with the northern
anomaly resulting from fault-bounded outflow from the south. Additional delineation of

35

the thermal anomalies is necessary to definitively conclude whether two separate systems
exist at San Emidio.

Figure 5.5 Mapped faults at the geothermal production zone with calculated dilational fault segments in
red; large red arrows indicate inferred hydrothermal upwelling zone; smaller red arrows indicate inferred
zone of lateral hydrothermal flow; bar and balls shown on downthrown sides of normal faults; EF, Empire
fault; SEF, San Emidio fault.

WIND MOUNTAIN EPITHERMAL MINERAL DEPOSIT


Though the two inferred loci of hydrothermal activity discussed above exhibit
strong evidence for fault-controlled fluid flow, other features may have contributed to
hydrothermal flow at the Wind Mountain epithermal mineral deposit.

However,

36

determining which structures aided in the pervasive mineral-bearing, silica-rich


hydrothermal flooding of the mainly shallow water sedimentary rocks of Wind Mountain
requires knowledge of the age of the mineralization and stress field active during
formation of the deposit. Prior work in the San Emidio area and Lake Range estimated
that the sedimentary rocks at Wind Mountain were late Miocene/Pliocene (Bonham and
Papke, 1969; Moore, 1979; Wood, 1990).

40

Ar/39Ar geochronologic analysis in this study

reveals that a sedimentary section just east of and presumably correlative with the Wind
Mountain deposit is about 4.8 0.9 Ma (Figures 3.1 and 3.2), confirming previous
estimates. The sedimentary rocks of Wind Mountain are interpreted to be tectonically
down-thrown from their original position stratigraphically above the dated sedimentary
section and thus are younger than 4.8 0.9 Ma. However, intercalated sinter horizons,
geyserite, and silicified reed beds (some in growth-position) within the deposit indicate
that some, albeit less robust, hydrothermal activity in a hot spring environment was
ongoing during deposition of the sedimentary section (Figure 5.6). Without exposure of
the oldest appearance of this hot spring environment, the timing of the onset of activity
cannot be determined.
By relatively dating the Wind Mountain sedimentary section, a maximum age of
4.8 0.9 Ma can be placed on at least the pervasive silica flooding preserved in the
deposit. The modern west-northwest-trending least principal stress direction has been
acting on the northwestern Basin and Range since ~10-7 Ma (Zoback et al., 1981; Bellier
and Zoback, 1995). By determining that the Wind Mountain epithermal deposit formed
under the current stress conditions, it can be concluded that moderately to steeply dipping
fault segments and fractures within the deposit that strike north-northeast have the highest

37

tendency to slip and/or dilate. These favorably oriented structures may have acted as
fluid conduits during formation of the deposit. Relations observed in the deposit that
support this conclusion include a strong spatial correlation between calculated dilational
fault segments and hydrothermal breccias, sheeted calcite veins, open-space and boiling
textures, and even former open-pit mining operations (Figure 5.7).

The pervasive

silicification observed at Wind Mountain suggests that it has hosted prolonged and
significant hydrothermal flow.

Figure 5.6 Photos of textures observed in the Wind Mountain epithermal mineral deposit: A. Fractured,
silicified sinter horizons. B. Silicified silica pellets interpreted as geyserite. C and D. Silicified reeds.

38

Figure 5.7 - A. NAIP image showing mapped faults with calculated dilational fault segments in red and the
Wind Mountain open pits outlined in gray. Letters on image A correspond to photos of representative
textures observed at those locations. B. Hydrothermal breccia. C. Sheeted calcite and chalcedony veins.
D. Chalcedony forming pseudomorphs after platy calcite in sheeted chalcedony vein (platy calcite can
indicate boiling). E. Sheeted rhombohedral calcite veins in fault zone in Wind Mountain open-pit mine.

BASIN AND RANGE GEOTHERMAL SYSTEM ANALOGS


The structural setting and stress conditions at San Emidio are not unique. The
mainly north-striking, right-stepping fault pattern, locally linked by east-northeast-

39

striking sinistral-normal fault step-overs, is fairly common in the northwestern Basin and
Range (Faulds et al., 2006). Additionally, close spatial relationships between young (<7
Ma) epithermal mineralization and active geothermal systems occur at numerous
localities in the northern Basin and Range, including Florida Canyon/Humboldt House,
Jersey Valley, Colado, Hycroft, and Tipton Ranch (Coolbaugh et al., 2005; Faulds et al.,
2005b).

Preliminary analysis of each of these sites reveals some structural and

hydrothermal similarities with San Emidio.


Mainly north- to north-northeast-striking, right-stepping normal fault zones with
northeast- to east-northeast-striking fault linkages are present at each of these sites
(Figure 5.8). At Florida Canyon and Hycroft, the right steps appear as simply fault-strike
discontinuities along a single fault. However, at San Emidio, Jersey Valley, Colado, and
Tipton Ranch, the right steps are clearly breaches in relay ramps between two rightstepping, normal faults. Each of these sites also possesses Quaternary fault scarps that
typically gradually transition from north- to northeast-striking in the vicinity of the
geothermal system. Further, the orientation of the Quaternary scarps generally parallels
that of the thermal anomaly and hydrothermal mineralization in each of these systems.
The locus of geothermal activity along such fault segments may in part result from their
orientation perpendicular to the current west-northwest-trending least principal stress
direction.

The increased fracture density within such favorably oriented steps may

generate relatively high permeability along subvertical conduits and thus allow for
sustained flow of hydrothermal fluids.

40

Figure 5.8 A. Tectonic setting map of Nevada showing San Emidio (SE) and hydrothermal system
analogs (B-F), which are all being evaluated for geothermal power generation potential; CNSB central
Nevada seismic belt. B. Florida Canyon/Humboldt House in right-step in normal fault zone. C. Jersey
Valley in right-step between two overlapping normal fault zones. D. Colado in right-step formed by broad
step-over fault between two overlapping normal fault zones. E. Hycroft epithermal gold and silver deposit
in right-step in easternmost normal fault zone. F. Tipton Ranch in right-step formed by step-over fault
between two overlapping normal fault zones. Hydrothermal alteration (outlined in green) and modern
geothermal activity (outlined in red) in figures B-F are concentrated at fault bends/intersections.

41

In addition, the northeast- to east-northeast-trending steps may link north-striking


normal fault zones that originally developed during ~east-west extension prior to the
current west-northwest-oriented extension direction (i.e., > 10 Ma). After the extension
direction rotated to west-northwest at ~10-7 Ma, northeast-striking structures within fault
step-overs became more favorably oriented for hydrothermal flow. Step-overs or relay
ramps are the most common structural setting for geothermal systems within the Basin
and Range (Faulds et al., 2011). However, moving east across the Basin and Range and
roughly beginning at the central Nevada seismic belt (CNSB), the extension direction
becomes more easterly (Hammond and Thatcher, 2004) and thus, northeast-striking stepovers may not be as favorably oriented for fluid flow. One hypothesis that remains to be
tested is whether the predominant orientation of step-overs that host geothermal systems
progressively changes across the Basin and Range in conjunction with the change in
extension direction.

6. Conclusions and Implications


Detailed geologic analyses integrated with geophysical and well data have
provided insight into the kinematics, stress state, and structural controls of the San
Emidio geothermal system. The San Emidio area consists of several gently to moderately
east-tilted fault blocks bounded by west-dipping normal faults and consisting primarily of
middle Miocene to Pliocene volcanic and sedimentary strata that nonconformably overlie
Mesozoic metasedimentary rocks. Kinematic data from faults, as well as earthquake
focal mechanisms and geodetic data, indicate a west-northwest-trending extension
direction. The west-northwest-trending extension direction results in moderately to

42

steeply dipping, north-northeast-striking faults and fractures with the highest tendency to
slip and dilate.
The current geothermal production zone and the Wind Mountain epithermal
mineral deposit occur at the intersection of favorably oriented north-northeast striking
normal faults with multiple closely-spaced north-striking normal faults.

Favorably

oriented north-northeast striking normal faults in right-steps are inferred as the major
conduits for hydrothermal upwelling. The Wind Mountain epithermal mineral deposit is
no older than 4.8 0.9 Ma and formed under west-northwest-directed extension.
Structures contributing to hydrothermal flow within this deposit may include northnortheast-oriented faults and fractures.
The structural settings at San Emidio and Wind Mountain, specifically right-steps
in north-striking normal fault systems, are not unique. The similarities between this
young epithermal mineral deposit and the associated modern geothermal system suggest
that similar exploration strategies could aid in identifying both geothermal and epithermal
mineral systems that are younger than ~7-10 Ma within the northwestern Basin and
Range. The identification of similar structural settings for several other geothermal
systems in the region indicates that structural characteristics may serve as a viable
exploration tool for discovering geothermal systems in extensional tectonic settings. A
simple review of fault systems and step-overs in extensional settings with a known
extension direction could provide an initial group of potential targets for further analysis.
Additionally, identification of favorable conduits for fluid flow in certain structural
settings could aid in enhancing the production of existing fields in similar settings.

43

References
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.
Barton, C.A., Zoback, D.M., Moos, D., 1995, Fluid flow along potentially active faults
in crystalline rock: Geology, v. 23, no. 8, p. 683-686.
Bell, E.J., and Slemmons, D.B., 1979, Recent crustal movements in the central Sierra
Nevada-Walker Lane region of California-Nevada: Part II, the Pyramid Lake
right-slip fault zone segment of the Walker Lane: Tectonophysics, v. 52, p. 571583.
Bell, J.W., and Ramelli, A.R., 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.
Bellier. O., and Zoback, M.L., 1995, Recent state of stress change in the Walker Lane
zone, western Basin and Range province, United States: Tectonics, v. 14, no. 3, p.
564-593.
Bennett, R.A., Wernicke, B.P., Niemi, N.A., Friedrich, A.M., and Davis, J.L., 2003,
Contemporary strain rates in the northern Basin and Range province from GPS
data: Tectonics, v. 22, no. 2, p. 1-31.
Blackwell, D.D., 1983, Heat flow in the northern Basin and Range province: Geothermal
Resources Council Special Report 13, p. 81-93.
Bonham, H.F., and Papke, K.G., 1969, Geology and mineral deposits of Washoe and
Storey Counties, Nevada: Nevada Bureau of Mines and Geology Bulletin 70,
scale 1:250,000, 140 p.
Briggs, R.W., and Wesnousky, S.G., 2004, Late Pleistocene fault slip rate, earthquake
recurrence, and recency of slip along the Pyramid Lake fault zone, northern
Walker Lane, United States: Journal of Geophysical Research, v. 109, no.
B08402, 16 p.
Caine, J.S., Evans, J.P., and Forster, C.B., 1996, Fault zone architecture and permeability
structure: Geology, v. 24, no. 11, p. 10251028.
Colgan, J.P., Dumitru, T.A., McWilliams, M., and Miller, E.L., 2006, Timing of
Cenozoic volcanism and Basin and Range extension in the northwestern Nevada:
New constraints from the northern Pine Forest Range: Geological Society of
America Bulletin, v. 118, no. 1/2, p. 126-139.

44

Coolbaugh, M.F., and Shevenell, L.A., 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., Arehart, G.B., Faulds, J.E., and Garside, L.J., 2005, Geothermal
systems in the Great Basin, western United States: Modern analogues to the roles
of magmatism, structure, and regional tectonics in the formation of gold deposits,
in Rhoden, H.N., Steininger, R.C., and Vikre, P.G., eds., Geological Society of
Nevada Symposium 2005: Window to the World, p. 10631081.
Coolbaugh, M.F., Raines, G.L., and Zehner, R.E., 2007, Assessment of exploration bias
in data-driven predictive models and the estimation of undiscovered resources:
Natural Resources Research, 9 p.
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.
Drakos, P.S., 2007, Tertiary stratigraphy and structure of the southern Lake Range,
northwest Nevada: Assessment of kinematic links between strike-slip and normal
faults in the northern Walker Lane [M.S. thesis]: University of Nevada, Reno, 165
p.
Duffield, W.A., and McKee, E.H., 1986, Geochronology, structure, and basin-range
tectonism of the Warner Range, northeastern California: Geological Society of
America Bulletin, v. 97, p. 142146.
Eneva, M., Falorni, G., Teplow, W., Morgan, J., Rhodes,G., and Adams, D., 2011,
Surface deformation at the San Emidio geothermal field, Nevada, from satellite
Radar Interferometry: Geothermal Resources Council Transactions, v.35, p. 16471653.
Etchecopar, A., Vasseur, G., and Daignieres, M., 1981, An inverse problem in
microtectonics for determination of stress tensors from fault striation analyses:
Journal of Structural Geology, v. 3, p. 51-65.
Fairley, J.P., and Hinds, J.J., 2004, Rapid transport pathways for geothermal fluids in an
active Great Basin fault zone: Geology, v. 32, no. 9, p. 825-828.
Faulds, J.E., Garside, L.J., and Oppliger, G., 2003, Structural analysis of the Desert PeakBrady geothermal fields, northwest Nevada: Implications for understanding links
between northeast-trending structures and geothermal reservoirs in the Humboldt
structural zone: Geothermal Resources Council Transactions, v. 27, p. 859-864.
Faulds, J.E., Coolbaugh, M., Blewitt, G., and Henry, C.D., 2004, Why is Nevada in hot
water? Structural controls and tectonic model of geothermal systems in the

45

northwestern Great Basin: Geothermal Resources Council Transactions, p. 649654.


Faulds, J.E., Henry, C.D., and Hinz, N.H., 2005a, Kinematics of the northern Walker
Lane: An incipient transform fault along the Pacific-North American plate
boundary: Geology, v. 33, no.6, p. 505-508.
Faulds, J.E., Henry, C.D., Coolbaugh, M.F., Garside, L.J., and Castor, S.B., 2005b, Late
Cenozoic strain field and tectonic setting of the northwestern Great Basin, western
USA: Implications for geothermal activity and mineralization, in Rhoden, H.N.,
Steininger, R.C., and Vikre, P.G., Geological Society of Nevada Symposium
2005: Window to the World, p. 1091-1104.
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 Resource Council Transactions, v. 30, p. 69-75.
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 22, p. 437-470.
Faulds, J.E., and Melosh, G., 2008, A preliminary structural model for the Blue Mountain
geothermal field, Humboldt County, Nevada: Geothermal Resources Council
Transactions, v. 32, p. 273-278.
Faulds, J.E., Moeck, I., Drakos, P., and Zemach, E., 2010, Structural assessment and 3D
geological modeling of the Bradys geothermal area, Churchill county (Nevada,
USA): A preliminary report: 35th Workshop on Geothermal Reservoir
Engineering, Stanford University, Stanford, California, February 1-3, 2010, 5 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 Resource Council Transactions, v. 35, p. 777-783.
Ferrill, D., and Morris, A.P., 2003, Dilational normal faults, Journal of Structural
Geology, v. 25, p. 183-196.
Gauthier, B., and Angelier, J., 1985, Fault tectonics and deformation: a method of
quantification using field data: Earth and Planetary Science Letters, v. 74, p. 137148.
Grant, F.S., 1985, Aeromagnetics, geology and ore environments, II. Magnetite and ore
environments: Geoexploration, v. 23, p. 335-362.

46

Gudmundsson A., Fjeldskaar I., and Brenner S.L., 2002, Propagation pathways and fluid
transport in jointed and layered rocks in geothermal fields: Journal of
Volcanology and Geothermal Research, v. 116, p. 257-278.
Hammond, W.C., and Thatcher, W., 2004, Contemporary tectonic deformation of the
Basin and Range province, western United States: 10 years of observation with
the Global Positioning System: Journal of Geophysical Research, v. 109, 21 p.
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, 12 p.
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., and 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.
Henkel, H., and Guzmin, M., 1977, Magnetic feature of fracture zones: Geoexploration,
v. 15, p. 173-181.
Henry, C.D., and Perkins, M.E., 2001, Sierra NevadaBasin and Range transition near
Reno, Nevada: Two-stage development at 12 and 3 Ma: Geology, v. 29, p. 719
722.
Henry, C.D., Faulds, J.E., dePolo, C.M., and Davis, D.A., 2004, Geologic map of the
Dogskin Mountain Quadrangle, Washoe County, Nevada: Nevada Bureau of
Mines and Geology Map 148, 1:24,000, 13 p.
Hinz, N.H., Faulds, J.E., and Oppliger, G.L., 2008, Structural controls of Lee-Allen Hot
Springs, southern Churchill County, western Nevada: A small pull-apart in the
dextral shear zone of the Walker Lane: Geothermal Resources Council
Transactions, v. 32, p. 285-290.
Ichinose, G. A., Anderson, J. G., Smith, K. D., and Zeng, Y., 2003, Source parameters of
eastern California and western Nevada earthquakes from regional moment tensor
inversion: Bulletin of the Seismological Society of America, v. 93, p. 61-84.
Kreemer, C., Blewitt, G., and Hammond, W.C., 2006, Using geodesy to explore
correlations between crustal deformation characteristics and geothermal
resources: Geothermal Resources Council Transactions, v. 30, p. 441-446.
Kreemer, C., Blewitt, G., and, Hammond, W.C., 2009, Geodetic constraints on
contemporary deformation in the northern Walker Lane: 2. Velocity and strain
rate tensor analysis, in Oldow, J.S, and Cashman, P.H., eds., Late Cenozoic
structure and evolution of the Great Basin-Sierra Nevada transition: Geological
Society of America Special Volume 447, p. 17-31.

47

Larsen, P.H., 1988, Relay structures in a Lower Permian basement-involved extensional


system, East Greenland: Journal of Structural Geology, v. 10, p. 3-8.
Locke, C.A., Johnson, S.A., Cassidy, J., and Mauk, J.L.,1999, Geophysical exploration of
the Puhipuhi epithermal area, Northland, New Zealand: Journal of Geochemical
Exploration, v. 65, p. 91109.
Lowell, R.P., and Rona, P.A., 2005, Hydrothermal activity: Encyclopedia of Geology, v.
5, p. 362-372.
Marrett, R., and Allmendinger, R.W., 1990, Kinematic analysis of fault-slip data: Journal
of Structural Geology, v. 12, p. 973-986.
Meschede, M., 1994. Methoden der Strukturgeologie; ein Leitfaden zur Aufnahme und
Auswertung strukturgeologischer Daten im Gelaende und im Labor. Enke,
Stuttgart, Federal Republic of Germany, 175 p.
Moeck, I., Kwiatek, G., and Zimmermann, G., 2009, Slip tendency, fault reactivation
potential and induced seismicity in a deep geothermal reservoir: Journal of
Structural Geology, v. 31, p. 1174-1182, doi:10.1016/j.jsg.2009.06.012.
Moeck, I., Brandt, W., Bruhn, D., Munoz, G., Ritter, O., Bauer, K., Weber, M., Backers,
T., Kwiatek, G., and Huenges, E., 2010, From prospecting to drilling: New
exploration strategies for enhanced geothermal systems: Proceedings of the World
Geothermal Congress, Bali, Indonesia, 25-29 April, 5 p.
Moore, J.N., 1979, Geology map of the San Emidio geothermal area: U.S. DOE:
Department of Geothermal Energy, 78-1701.b.1.2.2, 8 p.
Morris, A., Ferrill, D.A., and Henderson, D.B., 1996, Slip tendency analysis and fault
reactivation: Geology, v. 24, no. 3, p. 275-278.
Optim, Inc., 2011, Finding large aperture fractures in geothermal resource areas using a
three-component long-offset surface seismic survey, San Emidio geothermal
resource, Washoe County, Nevada: Seismic Data Processing Report for USDOEFOA-0000019 Grant, 90 p.
Peacock, D.C.P., and Sanderson, D.J., 1994, Geometry and development of relay ramps
in normal fault systems: American Association of Petroleum Geologists Bulletin,
v. 78, p. 147-165.
Petit, J.P., 1987, Criteria for the sense of movement on fault surfaces in brittle rocks:
Journal of Structural Geology, v. 9, no. 5/6, p. 597-608.
Reiter, F., and Acs, P., 1999, TectonicsFP 1.5: A computer program for structural
geology: Universitt Innsbruck.

48

Ressel, M.W., 1996, A transitional basaltic center in west-central Nevada: Petrochemistry


and constraints on regional middle Miocene magmatism and tectonism [M.S.
Thesis]: University of Nevada, Reno, 197 p.
Sheridan, J., and Hickman, S., 2004, In situ stress, fracture, and fluid flow analysis in
well 38C-9: An enhanced geothermal system in the Coso geothermal field:
Proceedings, 29th workshop on Geothermal Reservoir Engineering, Stanford
University, January 26-28 2004, SGP-TR-175.
Sperner, B., Ott, R., and Ratschbacher, L, 1993, Fault-striae analysis: a Turbo Pascal
program package for graphical presentation and reduced stress tensor calculation:
Computer & Geosciences, v. 19, p. 1361-1388.
Trexler, J.H., Jr., Cashman, P.H., Henry, C.D., Muntean, T.W., Schwartz, K., TenBrink,
A., Faulds, J.E., Perlins, M., and Kelly, T.S., 2000, Neogene basins in western
Nevada document 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: Boulder, Colorado, Geological
Society of America Field Guide 2, p. 97-116.
Vice, G.S., Faulds, J.E., Ehni, W.J., and Coolbaugh, M.F., 2007, Structural controls of a
blind geothermal system in the northern Pyramid Lake area, northwestern
Nevada: Geothermal Resources Council Transactions, v. 31, p. 133-137.
Vice, G., 2008, Structural controls of the Astor Pass-Terraced Hills geothermal system in
a region of strain transfer in the western Great Basin, northwestern Nevada [M.S.
thesis]: University of Nevada, Reno, 114 p.
Wood, G., and Thomas, M.D., 2002, Modeling of high-resolution gravity data near the
McArthur River uranium deposit, Athabasca Basin; in Summary of Investigations
2002, Volume 2: Saskatchewan Geological Survey, Saskatchewan Industry
Resources, Miscellaneous Report 2002-4.2, CD-ROM, Paper D-16, 9 p.
Wood, J.D., 1990, Geology of the Wind Mountain gold deposit Washoe County, Nevada,
in Raines, G.L., Lisle, R.E., Schafer, R.W., and Wilkinson, W.H., eds., Geology
and ore deposits of the Great Basin: Symposium proceedings: Geological Society
of Nevada, p. 1051-1061.
Yamaji, A., 2000, The multiple inverse method; a new technique to separate stresses
from heterogeneous fault-slip data: Journal of Structural Geology, v. 22, no. 4, p.
441452.
Yamaji, A., Sato, K., and Otsubo, M., 2005, Multiple inverse method main processor,
Version 5.31: Division of Earth and Planetary Sciences, Kyoto University, Kyoto.
Zoback, M.L., Anderson, R.E., and Thompson, G.A., 1981, Cenozoic evolution of the
state of stress and style of tectonism of the Basin and Range province of the

49

western United States: Philosophical Transactions of the Royal Society of London


A, v. 300, p. 407-434.

50

APPENDIX A: DESCRIPTION OF MAP UNITS

Anthropogenic Features and Deposits


Qx Disturbed and modified areas Highly disturbed, modified, and graded areas
including the Wind Mountain open-pit mine (T30N R23E Section 34) and the U.S.
Geothermal power plant (T29N R23E Section 16).

Playa and Related Deposits


Qp Playa deposits (latest Holocene to late Pleistocene) White to yellow-white, finegrained silt, sand, and clay form very flat, whitish, sparsely vegetated surfaces and cover
much of the floor of the San Emidio Desert; also fill small closed depressions.
Desiccation cracks are common as well as eolian deposits clustered around sparse
vegetation. Generally little to no soil development, but commonly moderate salt
(gypsum) accumulation. Thicknesses poorly known due to lack of exposure, but
generally thin (<1 m).
Qpf Playa fringe deposits (Holocene to late Pleistocene) Whitish to light grey silts,
sands, and pebble gravel; typically mantled by eolian deposits derived from adjacent
playa surfaces. Form very gently sloping, smooth surfaces; typically gradational with
playa deposits and located at the transition from fan alluvium to playa deposits.
Thicknesses poorly known, but generally thin (<1 m).

Hillslope Deposits
Qc Colluvium (Holocene to Pleistocene) Colluvial and talus deposits generally
occurring at the base of steep slopes. Deposits typically consist of poorly sorted, angular
to subangular, clast-supported cobbles and boulders. Deposits are generally less than 5 m
thick.

Alluvial Deposits
Qa Young alluvium, undivided (Holocene to late Pleistocene) Alluvium in recently to
annually active washes generally consisting of poorly to well-sorted sands to cobblepebble gravels; locally contain boulders.
Surfaces have rough bar-and-swale
morphology. Clasts are generally subangular to subrounded. Little to no soil
development. Thicknesses poorly known, but generally a few meters or less.

51

Qfy Young active fan alluvium and recently abandoned active alluvial surfaces
(Holocene) Poorly to well-sorted pebble to cobble gravel and sands with angular to
subangular clasts; locally contain boulders, especially at fan heads. Surfaces slightly to
fully smoothed. Minimal to cambic soil development. Deposits are up to 1 m thick.
Qfy1 Young active fan alluvium (late Holocene) Poorly to well-sorted pebble to cobble
gravel; locally contain boulders, especially at fan heads. Surfaces generally slightly
smoothed. Little to no soil development and thin (5-10 cm) silt caps. Poorly to
moderately stratified, matrix-supported deposits with angular to subangular clasts.
Surfaces commonly have bar-and-swale morphology. Deposits are as much as 2 m thick.
Qfy2 Young fan alluvium, undivided (Holocene to late Pleistocene) Poorly to wellsorted pebble to cobble gravel with isolated boulders; poorly to moderately stratified,
matrix-supported deposits with angular to subangular clasts. Surfaces are generally fully
smoothed with poorly to moderately developed pavements and sparse cobble or boulder
gravel bars. Typically have cambic soils and 10-20 cm silt caps. Deposits are as much as
3 m thick.
Qf Fan alluvium, undivided (Holocene to late Pleistocene) Small- to large-size, deeply
incised fans. Poorly to well-sorted pebble to cobble gravel; locally contain boulders,
especially at fan heads; typically matrix supported and dominated by subangular to
angular clasts. Variable surface morphology and soil development. Surfaces range from
smooth to rough. Well-developed desert pavement. Deposit thickness may locally
exceed 15 m.
Qfi Intermediate fan alluvium, undivided (late to middle Pleistocene) Cobbly,
pebbly, silty, coarse sands with isolated boulders. Generally weakly indurated, poorly to
moderately stratified, moderately sorted, matrix-supported deposits with angular to
subangular clasts. Surfaces are generally fully smoothed, erosionally rounded near
surface edges, and well dissected. Typically have argillic soils and Stage III-IV
carbonate horizons. Thicknesses poorly known, but up to 6 m.
Qfo Old fan alluvium, undivided (middle to early Pleistocene) Poorly to well-sorted
pebble to cobble gravel; locally contain boulders, especially at fan heads; poorly to
moderately stratified, matrix-supported deposits with angular to subangular clasts.
Surfaces broadly rounded and moderately to well dissected. Stage IV carbonate horizons
up to a meter or more thick with upper soil horizons typically stripped. Deposits are up
to 5 m thick.
Qss Silicified alluvium, undivided (Holocene to Pleistocene) Well-indurated, poorly
sorted, matrix-supported silicified alluvial fan deposits. These deposits include siliceous
sinter and opal, as well as silicified reeds and gastropods. These are only exposed along a
north-south trending zone of alteration, known as the San Emidio fault, along the eastern
edge of the San Emidio Desert. T29N R23E Sections 4, 9, 16, 21.

52

Qas Acid-sulfate altered alluvium, undivided (Holocene to Pleistocene) Nonindurated, poorly sorted, matrix-supported acid-sulfate altered alluvial fan deposits.
These deposits include clay alteration (kaolinite), as well as native sulfur, cinnabar, and
gypsum. These are only exposed along a north-south trending zone of alteration, known
as the San Emidio fault, along the eastern edge of the San Emidio Desert. T29N R23E
Sections 4, 9, 16, 21.
Qasfg Acid-sulfate altered fanglomerate (Holocene to Pleistocene) Well indurated,
weakly acid-sulfate altered, light-gray pebbly to bouldery alluvial fan deposits dominated
by clasts of Mesozoic phyllite (TrJn). This unit is only sparsely exposed along a northsouth trending zone of alteration, known as the San Emidio fault, along the eastern edge
of the San Emidio Desert and consists of poorly stratified, matrix-supported, poorly
sorted, fanglomerate with subangular to subrounded clasts of phyllite and quartzite with
lesser aphanitic basaltic andesite (Tpb). These deposits include clay alteration
(kaolinite), as well as native sulfur, cinnabar, and gypsum. T29N R23E Sections 4, 9, 16,
21.
Qtg Terrace gravels, undivided (late Holocene to Pleistocene) Poorly sorted gravels
that form a thin lens, ~5 m thick, above the Tertiary rhyolitic tuff (Tr). These gravels
consist of subrounded to rounded aphanitic basaltic andesite (Tpb), phyllite and quartzite
(TrJn), and weathered, subrounded granite cobbles from the nearby Selenite Range.
These gravels are only found above the identified Late Pleistocene Lake Lahontan
highstand.
QTa Basin fill alluvium, undivided (late Holocene to late Miocene/Pliocene)
Undivided sedimentary strata, including Quaternary alluvium and silicified and
unsilicified late Miocene to Pliocene sandstone, siltstone, and clay-rich mudstone (Tss
and Ts). Sandstone is light-tan to brown, thinly bedded, poorly- to well-indurated,
poorly- to moderately-sorted, clast-supported litharenite with angular to subrounded
grains up to ~3 cm long of fine-grained volcanic rocks and metasedimentary rocks (TrJn).

Quaternary Lacustrine Deposits


Qtu Subaqueous spring-related tufa deposits (Holocene to late Pleistocene) Lightbrown dendritic tufa mounds crop out along a north-south trending zone of alteration,
known as the San Emidio fault, along the eastern edge of the San Emidio Desert. These
deposits are spatially associated with a north-trending Holocene fault scarp. T29N R23E
Sections 4, 9, 16, 21.

53

Qtu/s Subaqueous spring-related tufa and siliceous sinter deposits (Holocene to late
Pleistocene) Light-brown dendritic tufa mounds along with lenses of siliceous sinter crop
out along a north-south trending zone of alteration, known as the San Emidio fault, along
the eastern edge of the San Emidio Desert. These deposits are spatially associated with a
north-trending Holocene fault scarp. T29N R23E Sections 4, 9, 16, 21.
Ql Shallow lake sediments (Holocene to late Pleistocene) Generally fine-grained, white
to light-tan, well sorted, very weakly indurated silty sand, silt, and clay associated with
pluvial Lake Lahontan. Thinly bedded <1 mm. Generally little to no soil development
due to eolian deflation. Deposits are up to 3 m thick.
Qb Beach deposits (Holocene to late Pleistocene) Shoreline beach berms associated
with Ql lake levels consisting of well sorted cobble pebble gravel to coarse sand. Clasts
consist of well-rounded gray phyllite, white to pink quartzite, and aphanitic dark gray
basalt. Typically have cambic soils with Stage II carbonate. Deposits are up to 6 m
thick.

Tertiary Rocks
Tgc Fault related gypsum and calcite deposits Gypsum and calcite forming an
unstratified crust within alluvium near a mapped fault. This unit only crops out in the
northwestern part of the map area. This crust is 1-4 cm thick. T30N R23E Section 23.
Tcv Fault related sheeted calcite veins In the northwest part of the map area, mapped
faults are commonly associated with topographically elevated sheeted calcite veins. The
fabric of these veins matches the dip of the related fault. Calcite in these veins is
generally in rhombohedral form up to 15 cm long. The veins range in thickness from 1
cm to ~20 m.
Tsf Sediments and fan alluvium This unit crops out in only the northwest part of the
map area and consists of unconsolidated Tertiary sediments (Ts) and alluvium.
Thickness ~5 m.
Tss Silicified sediments with siliceous sinter Pervasively silicified clastic Tertiary
sedimentary rocks (Ts) with intercalated sinter horizons of probable fluvial and lacustrine
origin. This unit makes up the majority of the Wind Mountain deposit. These rocks are
white to light tan and gray with some red oxidation. The intense silicification of these
rocks has erased sedimentary textures other than faint bedding. Individual beds range in
thickness from <1 cm to more than 10 m. The entire unit is roughly 200 m thick.

54

Tr Rhyolitic tuff Isolated exposure of partially welded and slightly silicified, glassy,
light-pink rhyolitic tuff. Dominantly glass with pumice and lithic fragments (up to ~1 cm
long), of aphyric dacite and basaltic andesite. Phenocrysts (<5%) consist of fractured,
slightly resorbed plagioclase (up to ~1 mm long) and lesser biotite and sanidine. This
flow, ~5 m thick, has limited exposure at the top of the late Miocene-Pliocene
sedimentary section (Ts) just east of the Wind Mountain epithermal mineral deposit and
outside of the pervasive silicification. Plagioclase has yielded an 40Ar/39Ar age of
4.80.9 Ma (sample R09-49; Appendix B).
Ts Sedimentary rocks (late Miocene to Pliocene) Light-tan to brown, fine- to coarsegrained clastic sedimentary strata, including sandstone, clay-rich siltstone and mudstone,
and conglomerate. Sandstone is thinly bedded, poorly- to well-indurated, poorly- to
moderately-sorted, clast-supported litharenite with angular to subrounded grains up to ~3
cm long of fine-grained volcanic rocks and metasedimentary rocks (TrJn). Conglomerate
is moderately-indurated, poorly-sorted, and matrix-supported with subangular to
subrounded clasts of Tpb and TrJn up to ~5 cm long. These rocks crop out east of Wind
Mountain and are locally weakly silicified. Thickness 450 m.
Tcla Clay-rich sedimentary rocks Poorly-indurated, poorly stratified, white to light-tan,
clay-rich alluvium of probable lacustrine origin exposed along the northeastern side of
Wind Mountain; includes montmorillonite and gypsum that may be related to argillic
alteration of late Miocene-Pliocene clastic sedimentary rocks (Ts). Individual beds are
difficult to identify but are generally marked by cross-fiber gypsum deposits. Thickness
30 m.
Tapb Hydrothermally altered lower Pyramid sequence volcanic rocks Light gray
aphanitic basaltic andesite, which has been hydrothermally altered to clay and iron
oxides. Thickness 10 m.
Thbr Hydrothermal breccia White to light-blue silicified breccia exhibiting
hydrothermal flow textures and rounded clasts of chalcedony and Mesozoic
metasedimentary rocks in a chalcedony matrix (TrJn). This unit is only exposed near
mapped fault zones in the northwestern part of the map area. Pyrite up to 2 mm long can
be found in the matrix and lithic fragments. Thickness 1-10 m.

Pyramid Sequence
Tpd Dacite Aphyric to sparsely porphyritic, medium gray dacite containing <10%
phenocrysts of plagioclase up to 1.5 mm long. Contains xenoliths of gray aphanitic
andesite up to 5 cm long. These lavas crop out in only the eastern part of the map area

55

and form Falcon Hill near State Highway 447. Plagioclase has yielded an 40Ar/39Ar age
of 14.41.6 Ma (sample R10-60, Appendix B). Thickness 50 m.
Tpts Tuffaceous and volcaniclastic sedimentary rocks Heterogeneous package of
well stratified, matrix and clast-supported, poorly to well sorted, light-brown to yellow
and gray tuffaceous sandstone with lesser conglomerate. Sandstone is thinly to thickly
bedded, moderately-indurated, non-calcareous to slightly calcareous, coarse-grained
litharenite with tuffaceous matrix and subrounded to angular clasts of basaltic andesite,
dacite, and pumice up to ~1 cm long. Beds range in thickness from less than ~2 cm to
more than ~1 m. Sandstone exhibits some cross-bedding and scour marks with local softsediment deformation. Conglomerate is clast-supported and moderately sorted with
angular clasts of aphanitic basalt (Tpb) up to ~10 cm long. These beds occur as lenses
within the Pyramid sequence lavas (Tpb). The cross-bedding and scour marks in the
sandstone and conglomerate suggest a fluvial origin. Thickness isolated lenses up to 3
m.
Tpb Sparsely porphyritic basaltic andesite This unit consists of numerous flows and
flow breccias of aphyric and aphanitic to sparsely porphyritic basaltic andesite with lesser
olivine basalt. Flows are dark gray to red-brown and typically foliated. This foliation is
more apparent in glassy aphyric lavas. Individual flows are ~2 to 25 m thick and
generally laterally continuous. Phenocrysts (generally <5%) consist of plagioclase
(sparsely up to ~4 mm long), augite, and lesser olivine (typically altered to iddingsite)
and orthopyroxene. Groundmass consists of fine- to coarse-grained plagioclase laths and
olivine with fine-grained interstitial pyroxene. Glassier flows exhibit slightly flattened
vesicles. Flow breccias contain lithics of mainly angular to subangular aphanitic basalt
and scoria. Flow foliation indicates ~30 of tilting to the east. Plagioclase has yielded an
40
Ar/39Ar age of 15.10.2 Ma (sample R09-167, Appendix B). Thickness 700 m.
Tpts Tuffaceous and volcaniclastic sedimentary rocks Heterogeneous package of
massive to well stratified, matrix- and clast-supported, poorly to well sorted, light-brown
to gray tuffaceous sandstone and non-welded ash-flow tuff with lesser conglomerate and
ash-fall tephra. Sandstone is thinly to thickly bedded, moderately indurated, noncalcareous to slightly calcareous, coarse-grained litharenite with tuffaceous matrix
containing subrounded to angular clasts of basaltic andesite, dacite, and pumice up to ~1
cm long. Beds range in thickness from less than ~2 cm to more than ~1 m. Sandstone
exhibits some cross-bedding and scour marks with local soft-sediment deformation. A
non-welded pumiceous ash-flow tuff up to 5 m thick lacks bedding and contains basaltic
and dacitic lithics up to ~10 cm long, angular pumice up to 5 cm long with lesser fiamme
and vitrophyric clasts, and <5% phenocrysts of biotite, hornblende, and plagioclase.
Conglomerate is clast-supported and moderately sorted with angular clasts of fine-

56

grained oxidized basalt (Tpb) cobbles up to ~50 cm long. Ash-fall tephras are slightly
reworked and up to ~15 cm thick with some rounding and flattening of vitrophyric
shards. The sandstone and conglomerate indicate a fluvial origin for much of the Tpts
unit. These rocks locally pinch out against paleo-topographic highs in the TriassicJurassic Nightingale sequence and occur in a north-trending topographic saddle between
Mesozoic metasedimentary rocks (TrJn) to the west and Pyramid sequence lavas (Tpb)
to the east. Thickness 300 m.
Tptsu Tuffaceous and volcaniclastic sedimentary rocks, undivided Same
heterogeneous package as Tpts but includes three semi-continuous basalt to basaltic
andesite lenses (Tpbl, Tpp, Tpb) that crop out in a north-trending topographic saddle in
the northern Lake Range.
Tpbl Sparsely porphyritic basaltic andesite Dark gray to black, glassy, aphanitic
basaltic andesite flow with typically sparse phenocrysts (<5%) of partially resorbed
elongate plagioclase phenocrysts up to ~3 mm long and lesser augite and olivine. Olivine
(<2 mm long) is typically altered to iddingsite. This flow, up to ~5 m thick, is a semicontinuous lens in the Pyramid sequence tuffaceous sedimentary unit (Tpts). Weak
alteration of this flow has produced calcite in voids and disseminated cubic opaque
minerals (most likely pyrite) throughout the groundmass.
Tpp Porphyritic basaltic andesite Dark gray to black, glassy, slightly vesicular, slightly
porphyritic granular basalt with ~15% phenocrysts of slightly resorbed plagioclase laths
up to ~2 mm long with lesser olivine altered to iddingsite. Groundmass is comprised
primarily of fine-grained plagioclase and glass. This flow, up to ~10 m thick, is a semicontinuous lens in the Pyramid sequence tuffaceous sedimentary unit (Tpts). Plagioclase
has yielded an 40Ar/39Ar age of 16.10.4 Ma (sample R09-39, Appendix B).
Tpb Basalt Black, aphanitic, aphyric basalt flows with <1% altered plagioclase
phenocrysts up to ~2 mm long. This flow, up to ~10 m thick, is highly oxidized with a
yellow and red staining and is a semi-continuous lens intercalated in the Pyramid
sequence tuffaceous sedimentary unit (Tpts).

Lower Miocene Volcanic Rocks


Td Dacite Poorly exposed, light-gray porphyritic dacite with ~40% phenocrysts of
hornblende (~40% of total phenocrysts; up to ~1 cm long), plagioclase (~35%, up to ~5
mm long), and biotite (~15%, up to ~3 mm long). This flow crops out in only the southcentral part of the map area (T29N R23E Section 34). Faulting prevents an accurate

57

estimate of the thickness. Hornblende yielded an 40Ar/39Ar age of 24.10.4 Ma, (sample
R10-48, Appendix B).

Cretaceous Intrusions
KTr Flow-banded rhyolite dikes Poorly exposed, light-tan, slightly porphyritic rhyolite
(<15% phenocrysts) with quartz phenocrysts up to ~2 mm. These dikes, up to 10 m thick
and 30 m long, cross-cut the foliation in the Mesozoic metasedimentary rocks (TrJn) with
generally an east to west strike with a weak flow-banding that generally parallels dike
margins. The dikes do not cross-cut the overlying Pyramid sequence.

Triassic and Jurassic Nightingale Formation


TrJn Metasedimentary rocks, including phyllite, quartzite, and marble
Metamorphosed and folded low-grade, argillaceous, dark gray phyllite, with some black
slate and schist and intercalated light gray carbonate and light-tan to pink quartzite
horizons. Phyllite and slate are highly micaceous with an average foliation dipping
roughly 50 west. Quartzite horizons are up to 10 m thick, typically cross-cut by quartz
veins, and have sparse micaceous lenses. Individual quartz grains have recrystallized and
have a glassy appearance. Quartzite weathers to a reddish-brown due to iron-oxide
content. Carbonate lenses, up to 1 m thick, are medium-grained with stringers of finergrained more resistant sandy carbonate. White bull quartz is common as veins
throughout TrJn.

58

APPENDIX B: 40Ar/39Ar GEOCHRONOLOGY


Five samples were selected from the northern Lake Range for 40Ar/39Ar analysis.
Samples were analyzed both macroscopically and microscopically to determine the
degree of alteration and glass content. Only unaltered samples with little or no glass were
selected for dating. Samples were crushed and sieved in order to obtain a desirable size
fraction. A Frantz magnetic separator was used to separate samples according to general
degree of magnetization. Plagioclase grains were separated for 4 samples and hornblende
for 1. All samples were washed with solutions of HCl and/or HF.
Separates were analyzed by Dr. Michael Cosca at the US Geological Survey lab
in Denver, Colorado using 40Ar/39Ar laser fusion and incremental heating of individual
grains and multi-grain aliquots. The plagioclase separates contain low potassium values
as reflected in the relative imprecision in both the single grain 40Ar/39Ar fusions and
incremental heating experiments. A summary of individual sample results is described
below (all uncertainties reported as 2 sigma).

R09-49

42 MSWD=0.47 Prob.=9.80
916 MSWD=17.81 Prob.=0 .00

This sample reflects significant age dispersion in 40Ar/39Ar fusion analyses of single
grains and multi-grain aliquots. This dispersion is attributed to multiple inputs of
plagioclase of varying age. The youngest age of about 4 Ma recorded in both the single
grain fusions and incremental heating of multi-grain aliquots probably records the last
volcanic episode contributing plagioclase to this sample. Petrographic examination of
this sample reflects variable source input.

59

R10-60

14.41.6 MSWD=1.94 Prob.=0.07


15.41.7 MSWD=7.32 Prob.=0.00

This sample yields 40Ar/39Ar plateau ages of 15.5 +/- 0.2 and 15.9 +/- 0.4 Ma, a mean
value of 15.4 +/- 1.7 Ma from three fusions of multi-grain aliquots, and a mean value of
14.4 +/- 1.6 Ma from single grain analyses. These different analyses overlap at the 2
sigma level but may indicate multiple inputs of plagioclase of varying age. 40Ar/39Ar
fusion of individual grains displays some dispersion, but the large analytical uncertainties
associated with these analyses (because of low K content) overlap a value of ca. 14.415.4 Ma.

R09-167

154 MSWD=3.18 Prob.=0.01


15.10.2 MSWD=1.75 Prob.=0.17

Incremental step heating and fusion of individual plagioclase grains and multi-grain
aliquots yield overlapping ages of 15.1 Ma. This age reflects volcanic plagioclase of one
source.

60

R09-39

16.10.4 MSWD=0.35 Prob.=0.7


14.61.0 MSWD=2.07 Prob.=0.07

Two multi-grain aliquots were incrementally heated by CO2 laser heating. Similar
degassing behavior and total gas ages are reflected in both aliquots with one aliquot
yielding an 40Ar/39Ar age plateau of 16.0 +/- 0.7 Ma. 40Ar/39Ar fusions of individual
grains yield a range of 13 to 22 Ma, with significant imprecision on individual analyses.
Age uncertainties between individual grains exceed analytical precision in some cases
which may suggest multiple inputs. The large uncertainties associated with these
analyses render them of limited value. 40Ar/39Ar fusions of multi-grain aliquots yield
overlapping ages of 16.1 +/0.4 Ma, consistent with the 40Ar/39Ar plateau ages of the
multi-grain incremental heating experiments, and therefore are interpreted as the most
reliable age for this volcanic feldspar.

61

R10-48

A multi-grain incremental heating of amphibole from this sample provides an 40Ar/39Ar


plateau age of 24.1 +/- 0.4 Ma over approximately 95% of the argon released. This age
represents cooling of the amphibole at this time.

62

APPENDIX C: SEISMIC REFLECTION PROFILES AND INTERPRETATION

As part of this study, ~32 km of active source seismic data were acquired in the
eastern San Emidio Desert (Optim Inc., 2011). Ten profiles, each approximately 3048 m
(10,000 feet) in length, had 49 shot points per line with a geophone spacing of 17 m (55
feet) and a shot point spacing of 67 m (220 feet). Shots were generated using a vibroseis
truck. First arrival times picked from raw shot gathers were used to derive P-wave and Swave velocities using SeisOpt 2D processing software. The P-wave velocity model
was then put through a Kirchhoff pre-stack depth migration to derive images of faults and
fractures.
For interpretive purposes, these depth-migrated seismic reflection profiles along
with velocity models were incorporated with available well data into OpendTect, open
source seismic interpretation software from dGB Earth Sciences. Profiles were
interpreted by analyzing prominent reflectors, identifying offset reflectors, and applying
detailed geologic map, well data, seismic velocity, and gravity survey information.
Faults were interpreted on the basis of offset reflection packages, changes in velocity,
and/or some steeply dipping reflectors within the profiles. Interpreted profiles were
directly imported into illustration software to facilitate construction of detailed cross
sections. The following profiles (seismic lines 1-10) show both the seismic reflection
data with velocity overlay and the interpretations of faults and stratigraphy.

63

QTa

QTa

QTa

QTa
QTa
QTa

QTa

QTa

Tpb
Tpb

Tptsu

Tpb
Tpb

Tptsu

Tptsu

Tpb

Tpb

Tptsu
5000

10000

Tptsu

Tptsu

Tptsu

Tptsu
TrJn

Tpb

Tpb

TrJn

15000

TrJn

20000 Velocity (ft/s)

TrJn

TrJn

64

QTa

QTa

QTa

QTa

Tptsu

5000

Tptsu

Tptsu
TrJn

Tpb

Tpb

Tpb

Tpb

TrJn

QTa

QTa

Tpb
Tptsu
TrJn

TrJn

10000

15000

20000 Velocity (ft/s)

65

QTa

QTa
QTa

Tpb
Tpb

Tpb

Tpb

Tptsu

Tpb

TrJn

Tptsu

Tptsu

Tptsu

TrJn

Tpb

Tpb

QTa

Tptsu

QTa

QTa

QTa

TrJn

TrJn

TrJn

Tptsu

5000

10000

15000

20000 Velocity (ft/s)

66

QTa

QTa
QTa
QTa
QTa

Tpb

Tpb

Tpb

QTa

Tpb
Tpb

QTa

Tptsu

Tptsu

Tptsu

Tptsu

Tpb

TrJn

Tptsu
Tpb

TrJn

TrJn
Tptsu

TrJn

TrJn

5000

10000

15000

20000 Velocity (ft/s)

67

Well Kosmos 1-9

Well Kosmos 1-8

Well SE-2

QTa
TrJn
QTa

QTa

QTa

QTa QTa

QTa

QTa

Tpb
Tpb

Tpb

Tpb

Tpb

Tpb

TrJn

Tptsu

Tpb
Tptsu
Tptsu

Tptsu

TrJn

TrJn

5000

TrJn

Tptsu

Tptsu

TrJn

10000

15000

20000 Velocity (ft/s)

68

QTa

QTa

QTa

QTa

Tpb

Tpb Tpb

Tptsu

TrJn

Tpb

Tpb

TrJn

Tptsu
Tptsu

Tptsu

Tptsu Tptsu

TrJn TrJn

Tpb

Tpb

Tpb

Tptsu

QTa QTa

QTa

QTa

TrJn

Tptsu

TrJn

TrJn

TrJn

5000

10000

15000

20000 Velocity (ft/s)

69

1SPEVDUJPO8FMMTWell 75B-16
Well 76-16 Well 75-16

QTa

QTa
QTa

QTa

Tpb

Tpb

Tptsu

Tptsu
TrJn

TrJn

Tptsu
TrJn

Tptsu
TrJn

Tptsu

Tptsu

Tptsu
TrJn

TrJn

TrJn

Tpb

Tpb

Tpb

Tpb

Tpb

Tpb
Tptsu

QTa

QTa

TrJn

Tpb

Tpb

Tptsu

QTa

QTa

QTa

QTa

QTa QTa

TrJn

TrJn

TrJn

5000

10000

15000

20000 Velocity (ft/s)

70

QTa

QTa
QTa

QTa

Tpb

TrJn

Tpb
TrJn

TrJn

Tptsu

Tptsu

Tptsu
Tptsu

Tpb
Tptsu

Tpb

Tpb

Tptsu

Tpb

QTa

QTa

TrJn

TrJn

TrJn

TrJn

5000

10000

15000

20000 Velocity (ft/s)

71

QTa

QTa

QTa
Tpb

QTa
Tpb

Tpb

QTa

QTa
Tpb

Tpb

QTa

QTa

QTa

Tpb

Tpb

Tpb
Tptsu

TrJn

Tptsu
Tptsu

Tptsu

Tptsu

Tptsu
TrJn

Tptsu

TrJn
TrJn

TrJn
TrJn

5000

10000

15000

20000 Velocity (ft/s)

72

10

10

5000

10000

15000

20000 Velocity (ft/s)

NEVADA BUREAU OF MINES AND GEOLOGY

OPEN-FILE REPORT 11-X


PRELIMINARY GEOLOGIC MAP OF THE NORTHERN LAKE RANGE, SAN EMIDIO GEOTHERMAL AREA, WASHOE COUNTY, NEVADA

Prepared with support from the U.S. Department of Energy

96

97

11922'30"W 99

98

300

01

02

%%%%%%%

Anthropogenic Features and Deposits

Qc

: %%

:
@

:
@

S% S% %

Fault related gypsum and calcite deposits

Tcv

Fault related sheeted calcite veins

Tsf

Non-indurated sedimentary rocks and fan alluvium

Tss

Silicified sedimentary rocks

Tr

Rhyolitic tuff

Ts

Sedimentary rocks (late Miocene to Pliocene)

Tcla

Clay-rich sedimentary rocks

Tapb

Hydrothermally altered lower Pyramid sequence volcanic rocks

Thbr

Hydrothermal breccia

Tpd

Dacite

Tpts'

Tuffaceous sedimentary and volcaniclastic rocks

Tpts

Tuffaceous sedimentary and volcaniclastic rocks

Tptsu

Tuffaceous sedimentary and volcaniclastic rocks, undivided

Tpp

Porphyritic basaltic andesite

Tpb

Basalt

:
@

47

Inclined

Strike and dip of foliation in igneous rock

Oblique strike-slip fault Solid where certain and location accurate,


long-dashed where approximate; queried if identity or existence uncertain.
Arrows show relative motion.

33

Inclined

Strike and dip of foliation in metamorphic rock


25

Former shoreline Long-dashed where approximate.


!

82
!

Inclined

!
40

39

Ar/ Ar sample locality

Dacite

Cretaceous Intrusions

A'

Well locations

Line of cross section

Flow-banded rhyolite dikes

<150m

1'

Line of seismic reflection profile

150-1000m

>1000m

<150C

Metasediments including phyllite, quartzite, and marble

>/=150C

Gravity contour

S S%

56

33

Tpb
Tpts
Tpb

67

44

52

Tpb

73

56

78

Tpb

55

58
38

Tpts

Qfy1

Qfy2

B'

"

15.10.2 Ma

Qb Qb
Qb Qfy1
Qfy2
Qfy1
Qb

Qf

72

Tpts

S%

@
@

Qf

58

@
@

"

Tpb'

Tpb'
Tpb'

^Jn

Tptsu

^Jn
^Jn

Tptsu

^Jn

^Jn
Qb

^Jn

^Jn

-2000

^Jn

69

2000

^Jn

Tpb'

Tptsu
Tpb'

^Jn

Tptsu

Tptsu
Tptsu
Tptsu

Tpp

Qb

^Jn

Tpb'

A T

20

Qfy1

4000

Tptsu
4470

Tpb'
Tpts
TptsTpbl'
Tpp
Tpts

Qb

Qb

^Jn

Qfy2

19

Qb

Tptsu

Tptsu

Tpp

16.10.4 Ma

Qfy2

Tpb'

Tpb'

Tpts

Qfy2

Tpb'

$
$
$

Qfy2

Qfy1

Qa
40 ^Jn
Tpts
Qf

Qf

Tpts'

^Jn

^Jn

Tpts

^Jn ^Jn

Tpts'

30

Tpb

Qc

Tpb'

Tpb'

Tpb'

73

Tptsu

Tpbl'
Tpp Tpts
Tpts

Tptsu

Tpb'

Tpb'

37

Qfy2

Tpts

57 ^Jn

Qb

QTa

QTa

Qfy

Tpb

Qf

Qfy1

42

21

Qf

Qfy2
Qfy1

QTa

QTa

58

Qf

QTa

Tpb'

Qa
QTa

Tptsu

Ts

QTa

QTa

QTa

QTa

QTa

QTa

86

Qfy2

1000

Qa Qf

71

Qfy

Tss

Tr

Qb

29

QbQfy2

Qfy2

Qfy

Qb

Qb Qx

QTa

Qfy1

Qpf

Qp

Qp

Qp

Qp

$
$

43

Qb

Qx

Qf

62

Qfy2
Qb

Qc

34

65

TptsQf Tpbl'
TppTpts

@
@

56

$
$

Qfy1

Tpb'

Qf

60

30

Ql

Qfy2

Tpb'

Tpb'

60

Qb

Qfy2

Tpts'

50

Tpb

6000

4022'30"N

@
@

Qc

Qfy2

Qb

52

Qc

34

47

Qfy1

63

Tpts

Tpts'35

Tpb'
40

50

Qb

2000

25

Qf ^Jn
Qfy1 Qc
^Jn 78 52
Qfy2
Qfy2
Qfy1
Qf
Qfy^Jn
2
Qfy1
Qfy2 57

Qfy2

feet

meters

Qfy2

A'

Qfy2

^Jn

Qfy2

Qb

Tpb'

@@

Qf
Tpts

Tpts'

Tpb Tpts

Qb

-1000

Tpb

Tpb'

oo

Tpts'

34

Tpts'

Tpp

:@

55

16

Qf

Qf

32

Qf
Qf

30

"

Td

Tpts

40

Td

%%%

Tpts

Tpts'
25

B'

meters

feet

50

Tpb
Tpb
^Jn

68

Qf

Qf

Tpb'

Tpts'

Td

%%%

Qf

Qb

Tpts'

Tpts

Tpb'

Tpb'

o
o

Tpd

^Jn Qf

2000

Qfy2

Strike and dip of jointing

Tpts'
Tpts'

Tpb

Tpts

Qfy2

24.10.4 Ma

Qfo

Inclined

34

74

Qb

62

Qf

39

%%%

%% % %
% %%
%
%
%%%

Qfy1

41

Qf

Qfy2

Qb

Qb

:%

Sparsely porphyritic basaltic andesite

Sparsely porphyritic basaltic andesite

?
Normal fault Long-dashed where approximate, dotted where concealed;
queried if identity or existence uncertain. Ball on downthrown side.

Tpb'

Tpbl'

Strike and dip of bedding

Contact Solid where certain and location accurate, long-dashed


where approximate.

Qf

Ql
Ql

Ql Ql

Tpts

68

Symbology (per FGDC-STD-013-2006)

Triassic and Jurassic Nightingale Formation

35

Tpb

Qfy1

Qfy2

Qfy1 Qfy
2
Qfy2

Qb
Tpb'

Qfy2
Qf

Qfy1

Holocene

Tgc

KTr

Tpts
Tpb
Tpb Tpp
Tpts

46

^Jn

Qfy2

Qfy2

^Jn

S% S% S
S%S

S S% S
SS% S
SS

@%

:
S

%:

%
@

@ @

oo
@

@
@

@
@
@

Beach deposits (middle Holocene to late Pleistocene)

Tpts'

Qfy2

Pliocene

o
o

@@

::
@
@

:
:@

"

S%

:
%

:@
@
@

@ :

Qb

KTr

Triassic /
Jurassic

"

S%

:%

@
@

:
@

:
@

Tpp

@
@

Tpts'

o o
o o

::

% SS

:
@

:
@

@
@

%%%

@
@

Shallow lake sediments (middle Holocene to late Pleistocene)

^Jn

Ql

Cretaceous

Calcite vein Solid where certain.

Subaqueous spring-related tufa and siliceous sinter deposits (middle Holocene to late Pleistocene)

Lower Miocene Volcanic Rocks

Ql

@
Qfy2

Td

Tpp

Qf

Qfy1

Qb
Qb

@
Qb

Tpb

Subaqueous spring-related tufa deposits (middle Holocene to late Pleistocene)

Qtu

Tpp
Tpp

Tpts

Tptsu

Tpp

Qf

Qfy1

Tpts

63

Qf

Tpb

Qb

@
Ql

Tpbl'

Qfy1

Qfy1

Tpbl'
Tpbl'

S%:

@
@

:
@

Basin fill alluvium, undivided (late Holocene to late Tertiary)

Td

54 54

QTa

Tpts Tpb

Terrace gravels, undivided (late Holocene to Pleistocene)

76

75

@@

@
@

Qtg

71

Acid-sulfate altered fanglomerate (Holocene to Pleistocene)

66

Ql

Qfy2

Tpd
Tpts'

91

32

51 82

Tpts

Qb
44
Qfy1

Tpp

KTr

Qfy1

Qfy2

69

Tertiary

%%

%%%

:
@

@
@

Qfy2

87

Acid-sulfate altered alluvium, undivided (Holocene to Pleistocene)

38

Tpts

Tpb'

S%

Qas

Silicified alluvium, undivided (Holocene to Pleistocene)

69

Tpp

Qss

Pyramid Sequence

31

62
59
50 43

Qf

Qb

Qtu
Qtu

Qfy2

75B-16

Qfy2

Qb

57

Tpbl'

66

Ql

Ql

Tpb

68

Qf

7'

Ql Qfy1

Qtu

Old fan alluvium, (middle to early Pleistocene)

36

32

6'

Qb

Phillips

Ql

28

A'

Qfy1

77

Tpbl'

54

49
75

51

Tpts

Qf

Ql

Qfy1

Qfy1

Ql

Ql
Ql Ql

Qfo

Qf

Qfy1

Ql

41

Qfy1 Qfy1

Qtu

4470

Qfy1

Ql

Qpf

34

Qfy2

71

% %% %

75

5'

Qtu
Qtu

Qx

Tcla

Tpp

Qb
Qfy2
Qb
Qfy2
Qfy2
Qfy1
Qfy1
Qfy2

Qfy1

Geothermal
power plant

Qfy2

H
H Qtu

68

66
87

76

Qfy1
Qfy2Qf
Qfy
2
Qfy2
Qfy1
Qf Qf 46 46
Qfy1
Qb

Qb

75-16

$:

Qfy1

Qx

%
%%%

% % % % % % % % %% %
%
%
%
% %% %
%%%%% %% % %%
% % %%% %
%%%
% % %%
%% %% % % % % % %
%%%%%
%% % % % % % % % % % % % %
%%%
% % %%% % % % % %

72

Qb

49
40 43 40

56

Tpbl'

Tpp

Qf

4022'30"N

56

Tpp

60

Qfy1

Tss
Tcv

30

Qf

Qfy1
Qfy2

Tsf
Tr

24

44

Tpp

33

61

49 57

Qfy1

Qfy1

Qtu
Qfy2
Qas Qtu
Qtu
Qss
@ Qtu
Qpf Qb
Qb
Qas
Qpf
Qtu
Qas Qtu
Qfy1
Qfy1
Qfy1
Qas
Qas Qfy1
Qb
Qtu
Qfy2

Qpf

54

Tpbl'

Tpp

30

64

Qf

Qfo

Qfy2

Qfy2

Qb

65-16

Qfy2

Qfy2

70

88

Qfy2

76-16

Qb Qb
Qfy2

53

@
%% % % %
% %%
%
@
%
%
%
% % % %%
%
%
%
%
%
%%%

Qb

Qas Ql
Qb Qb
Qfy1
Qfy2
Qas
Qfy1
Qtu
Qp Qtu/s
Ql
Ql
Qfy1
Qfy1
Qb Qfy1

61

S!

Qtu
Qss
Qas
Qfy1 Qfy1

Qtu
Qfy2
Qfy2
Ql
Qtu
Qtu
Ql
Qfy
Qpf
2
Qtu
Qtu Ql
Qtu Qtu Ql
Qtu Qtu

Qfy2 87
Qfy2
! 54
Qfy2 7054
Qf Qf
Qfy2
87
Qfy2 Qfy2 Thbr
Qfy2 Qf
Qb
81

Qfy2

79

80

Qtu/s

Tpbl'

25

@ 45
Tsf
Qa Tpts

Tpb'

!
Thbr
75

Ts

Ts
Qfy1 Qfy1

Kosmos
1-9

73

80

Ql

34

Tpb'

Tpts

66

55

4'

Qb Qss
Qss

@
Kosmos
1-8

Qss
Qtu

Tpp

Qfy1

@ 40

87

12

73

Qfy1

Qa
Tpts
Tr

Tr

Qfy2

Qfy2

74

78

Tpb'

46

Tsf

Qb

Qtu

Qf

12

Qx

Qb

Ql

Tertiary Rocks

Tpts

Qfy2

Tr
Tr

81

MW-3

Qfy2

60

Tpts
Tsf
Qa

Qfy1

Qfy2

Qfy1

75

81
71

@
Qb
Qx Qx

Qfy2

Qp

Tpb'

Tss
Tss
Tss

Intermediate fan alluvium, undivided (late to middle Pleistocene)

Qtu/s

Qf

Tpb'

89

Ts

Qf Qx

Qfi

65

84

Qx

Qtu

79

Tpb'

Tpb'

13

3'

Qpf

76

81

Tpb'

Qfy1

Qpf

Qf

Tpts Tpb'

Tpp

83

Qx

80

Fan alluvium, undivided (Holocene to late Pleistocene)

Qasfg

89 63

Qfy
Qtu

Qfy2

Qx

Qf

Tpp

74

74

MW-1

83

2'

MW-2

Qb

Qb

Qtu
Qtu

Qb
Qtu

3
77

71
68

Young fan alluvium, undivided (Holocene to late Pleistocene)

Qb

QTa

Quaternary Lacustrine Deposits

62

Tpb'

58

Qb

10

Qb
Qtu
Qfy Qb
Qtu Qtu

Tpb'

Tpb'

Tpp

17

Qf Qf

86

1'

Qfy2

Qss

Tpb'

Qf

Tpp

Young active fan alluvium (late Holocene)

Qasfg

Tpp

Tpp

KTr

KTr

Qfy

Qfy

64

Qf
KTr

o
o

Tcla
Tss

78

Tr

4.80.9 Ma
78

Qf

Qtu

Qfy1

Qas

Ts

4480

Tpp

33

Qtg

83

79

41
41

Qtu

Qpf

KTr
KTr

Ts

Qx

Qpf

KTr
61

80 71
67

17

Qpf

Young active fan alluvium and recently abandoned active alluvial surfaces (Holocene)

Qf

Tpts

76

Qx

Tpts

61

^Jn

Qf

45

Qfi

Thbr

Tpb'

48

78

Tss

Qfy

Qpf

Qc

70

52

Qf

Tcla

85

Qf

11

: % : %

25

Young alluvium, undivided (Holocene to late Pleistocene)

Qf

Tpb'

66

!
!
!
! !
!
!

79

Qa

Qp

49

Tapb

Tss

Qf

69

70

Tpb'

35

Qf

Qf

10

71

Tss

78

Tpb'

Qf

Qf

Tpb'

%% %%

%
% % S%

51

Tpb'
Tpb'
Qtu

Tpb'

74
83

70

78

Ts

Tapb
Tapb
Tapb

Tpts

52

Qpf

67

^Jn

Tpb'

34
57

Tpb'

33

Tpb'

Tpb'

Qf

Ts

46
67 47

4480

33

Ts

83

47

40

Tpts

% :%

Tpb'

Qpf

Tpb' 42

40

Qb

Qb

Alluvial Deposits

Tss

Qfy1

::

32

81
Qa

72

%
Qfy

Qfy2

Qfy

% % %%
%% %%%
%
%
% % % % % % % % %%%
%
%% %
%
%
%
%
%%%%%%
%
%
%
%
%
%
%%%%

72

:%

Qfy1

Qfy1

Qf

Qf

Colluvium (Holocene to Pleistocene)

Qc

%%%

QbQfy2 Qfy1
Qfy2

Playa fringe deposits (Holocene to late Pleistocene)

Hillslope Deposits

%%%%
%

Qfy2

81

Qpf
Qf

%%%
% %%%

% %% % % %% %%

Playa deposits (latest Holocene to late Pleistocene)

%%

Qp

Cenozoic

Qfy1
Qfy2

%
%% %% % % % %

% %

Qtg

Qx

Qa

Pleistocene

82

Qfy

Miocene

%%

Playa and Related Deposits


Quaternary

%%

%% % % %

% % %% %

Disturbed and modified areas

Mesozoic

% %% % % % % %

82

Qx

95

94

93

6000

^Jn

%%

Qfi

%
@

%%

%%%

Qfi Tpts

Qfo

%%% %

Tpb'

Tpb'

"

14.41.6 Ma

Tpd

%%%

QTa

QTa

Tpb'
Tpts'
Tpb'

Tpb'

Tpb'

Tptsu

Tpb'

Tpb'

Tpb'

Tpb'

Tptsu
Tptsu

Tptsu

Tptsu

%%%%
% %%

%%

Tptsu

Qfy1

0
Qfy1

Tpts'

Qfy2

Tptsu

Tptsu

^Jn

Tptsu
^Jn

%%
@

^Jn

-2000

%%

^Jn

Qfo

% %% % % %

Qfy2

^Jn

$
$

%%%
%

%%

^Jn
^Jn

Tpb'

^Jn

@
65

%%%
%%

Qfy1

Qf

^Jn

^Jn

Tptsu

2000

Tpb'

QTa

%%%%

QTa

Tpb'

Tpts

Qfo

Qfy1

Qfi

4000

Tpb'

Tpts

%%
%
%
%
%
% % %%

@
@

65

$
$
$

QTa

66

Qfo

Qfy1

QTa

QTaQTa QTa QTa

QTa
Qf

Qfi

%%%

Qfy1

43

Qfi

Qfy2
QTa

1000

Qfo

Qfy1

QbQTa
QTa

Qfy1

Qb

QTaQx

Qp

Qp Qp Qp

Qp

Qf

%%

Qtu

Qfi

Qf

Qfy2

Qfi

% % % %%% %
% % %%

Qp

66

Tpb'

31

Qfy1

%%%%

Qtu

67

41

Qf

Qfo

Qc

Tpts'

Qfo

Qb

Qfy

Qfy2

%%%

67

Qfo
Qfi

%
% %

Qfy

@@

Qfo

Qfo

Qfy2

-1000

Qfo

Qfy1

64

64
Qtu

Qfy1

63

62000m. N

63

44

62

Adjoining 7.5'
quadrangle names
293 000m. E

94

95

96

97

98 1922'30"W
98
11922'30"W

99

300

01

02

10

11

12

13

14

15

16

Scale 1:24,000
0

PRELIMINARY GEOLOGIC MAP OF THE NORTHERN LAKE RANGE,


SAN EMIDIO GEOTHERMAL AREA, WASHOE COUNTY, NEVADA
Gregory T. Rhodes, James E. Faulds, Alan R. Ramelli
2011

^
GN

1 35'

MN

16 00'

UTM GRID AND


1990 MAGNETIC NORTH

Smith Canyon
West of Empire
Empire
Tenmile
Pah-Rum Peak
San Emidio Desert North
Kumiva Peak
Betty Creek
Fox Canyon
San Emidio Desert South
Purgatory Peak
Jayhawk Well
Pyramid NE
Tohakum Peak NW
Tohakum Peak NE
Tunnel Spring

1 kilometer
Field work done in May-December 2009
Supported by the U.S. Department of Energy
(Agreement No.DE-FG36-02ID14311).

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16

0.5

0.5

1000

2000

3000

4000

5000

mile

feet

CONTOUR INTERVAL 20 FEET


Projection: Universal Transverse Mercator, Zone 11,
North American Datum 1927 (m)
Base map: U.S. Geological Survey San Emidio Desert
North 7.5' quadrangle (1990)
U.S. Geological Survey San Emidio Desert
South 7.5' quadrangle (1990)
U.S. Geological Survey Kumiva Peak
7.5' quadrangle (1990)
U.S. Geological Survey Purgetory Peak
7.5' quadrangle (1990)

DRAFT
Preliminary geologic map
Has not undergone office or field review
Will be revised before publication
Edited by XXXXXXXXXXXXXXXXX
Compilation by Gregory T. Rhodes
Cartography and map production in ESRI ArcGIS v9.3 (ArcGeology v1.3)
by Irene M. Seelye
First Edition, November 2011
Printed by Nevada Bureau of Mines and Geology
This map was printed on an electronic plotter directly from digital files. Dimensional calibration may
vary between electronic plotters and X and Y directions on the same plotter, and paper may change
size; therefore, scale and proportions may not be exact on copies of this map.

For sale by:


Nevada Bureau of Mines and Geology
2175 Raggio Pkwy.
Reno, Nevada 89512
ph. (775) 682-8766
www.nbmg.unr.edu; nbmg@unr.edu
This product is available free on our website

2199

You might also like