+ All Categories
Home > Documents > crack.seismo.unr.educrack.seismo.unr.edu/ftp/pub/louie/class/theses/Mayhew... · Web viewThe Astor...

crack.seismo.unr.educrack.seismo.unr.edu/ftp/pub/louie/class/theses/Mayhew... · Web viewThe Astor...

Date post: 22-Mar-2018
Category:
Upload: trinhdien
View: 217 times
Download: 3 times
Share this document with a friend
29
Structural Controls of the Astor Pass Geothermal Field, northwestern Nevada Thesis proposal for the degree of Master of Science in Geology Brett Mayhew Advisor: Dr. James Faulds Mackay School of Earth Sciences and Engineering
Transcript

Structural Controls of the Astor Pass Geothermal Field, northwestern Nevada

Thesis proposal for the degree ofMaster of Science in Geology

Brett Mayhew

Advisor: Dr. James Faulds

Mackay School of Earth Sciences and EngineeringUniversity of Nevada, Reno

April 25, 2012

Introduction

The Great Basin, USA, provides an unusual combination of geologic components that

foster geothermal activity, including an anomalously high geothermal gradient, active faulting,

and regional extension. Although the northern Basin and Range has experienced active

volcanism as recently as 10 to 3 Ma, this volcanism is thought to have little effect on the current

geothermal gradient (Blackwell, 1983; Faulds et al., 2006). Only a small number of systems

(Roosevelt Hot Springs, Coso, Steamboat) on the margins of the Great Basin show evidence of

active magmatism. Thus, the majority of the ~430 known geothermal systems in the Great Basin

are considered amagmatic, relying on elevated crustal heat flow of ~75±5 mWm-2 (Blackwell,

1983; Blackwell and Richards, 2004) and conduits for hydrothermal fluids along fault zones

(Faulds et al., 2006).

Recent studies of these amagmatic systems have shown that certain structural settings act

as more effective conduits. Steeply dipping normal faults that have discrete steps, terminate,

intersect other faults, intermesh with oppositely dipping faults, or reside in transtensional pull-

aparts are the preferred settings (Faulds et al., 2006, 2010, 2011; Vice et al., 2007; Hinz et al.,

2008, 2010, 2011). These fault geometries generate favorable sub-vertical, high fault and fracture

density zones that act as preferential fluid conduits. The fluid flow in these fault zones is further

increased when critically stressed with respect to the ambient stress conditions (Ferrill and

Morris, 2003; Moeck et al., 2009).

Astor Pass is a blind geothermal system that lies close to the California-Nevada border at

the margin of the northern Walker Lane and the western Basin and Range (Figure 1). The system

was identified by Coolbaugh et al. (2006) based on the presence of tufa mound lineaments. Tufa

can be precipitated when calcium rich geothermal fluids contact atmospheric CO2 dissolved in

lake water. The Astor Pass tufa is ~ 6 km northwest and along strike from a large modern

geothermal outflow at Needle Rocks, where boiling water emanates from springs and geothermal

wells drilled in the 1960s.

Purpose of Study

The purpose of this study is to analyze the structural controls of the Astor Pass

geothermal system. Geothermal activity in Astor Pass is evident from the linear tufa mounds and

moderately high temperatures shown from shallow temperature surveys (Kratt et al., 2010) and

deep wells. On the basis of detailed geologic mapping that showed the distribution of major

faults and distinct lineaments of tufa towers, Vice et al. (2007) and Vice (2008) suggested that a

zone of dilation occupied the southwest quadrant of a fault intersection between a northwest-

striking oblique dextral fault and north-northwest striking, west-dipping normal-dextral fault,

southwest of the tufa towers.  Deeper wells (approximately 1300m) were drilled in the area of

hypothesized dilation in hopes of finding the upflow of the high temperature geothermal system.

These wells were unsuccessful in intercepting a high temperature upflow zone with peak well

temperatures of 94°C, lower than the estimated max temperature of 134°C from geothermometry

(Coolbaugh, unpublished data).  The temperature versus depth plots for the three wells show a

nearly isothermal gradient with increasing depth.

To elucidate the relationship between faults and hydrothermal fluids at Astor Pass, I plan

to construct and analyze a detailed three-dimensional structural model of the faults and

stratigraphy of the study area. The model will incorporate 1) detailed geologic mapping of the

northern Virginia Mountains and Terraced Hills, 2) interpretations of 16 lines of seismic

reflection profiles, 3) detailed analysis of well cuttings, and 4) analysis of borehole stress and

kinematic data.  The construction of this 3D model will allow for a detailed understanding of the

geothermal system and the structural controls therein.  This will also allow for a more complete

understanding of geothermal areas with similar fault-intersection structural environments and aid

in exploration and development of future sites.  

Tectonic Setting

Astor Pass is located in the northwest Great Basin in the western Basin and Range.  Astor

Pass lies directly northwest of Pyramid Lake between the Virginia Mountains and Terraced Hills

to the west and east, respectively.  Astor Pass lies along the northeastern boundary of the

Pyramid Lake domain of the Walker Lane structural belt (Stewart, 1988; Faulds and Henry,

2008). This places Astor Pass in the “transitional Walker Lane”, a transtensional structural

region involving northwest-striking, left-stepping, right-lateral faults of the Walker Lane and

kinematically linked north to north-northeast-striking Basin and Range normal faults (Figure 2)

(Faulds et. al.,, 2005a, b; Vice 2008). Along this boundary, strain is transferred between dextral

shear in the Walker Lane and west-northwest directed extension in the Basin and Range (Oldow,

1992; Faulds et al., 2005, 2006).

The Walker Lane accounts for ~20% of the strain between the Pacific and North

American plates (Hammond and Thatcher, 2004; Faulds and Henry, 2008; Hammond et al.,

2011). Geodetic studies currently suggest that the Pyramid Lake fault in the northern Walker

Lane accommodates ~1.0 ±0.3 mm/yr of dextral shear, whereas the structural blocks of the Basin

and Range (Pyramid Lake to Big Smokey Valley blocks) just east of Pyramid Lake fault are

extending ~N55W at a rate of 4.6 mm/yr relative to stable North America (Hammond et al.,

2011). Hammond et al. (2011) modeled the Pyramid Lake fault as the border between these two

structural areas, separating the Walker Lane transtensional and Basin and Range extensional

domains. A similar dextral slip rate of 2.6 ± 0.3 mm/year for the Pyramid Lake fault has been

measured by analyzing offset of geomorphic features along the fault trace (Briggs and

Wesnousky, 2004).

Basin and Range extension in this region began ~13-10 Ma, concurrently with a major

east to northeast extensional event that also occurred along much of the Walker Lane and eastern

flank of the Sierra Nevada (Trexler et al., 2000; Henry et al., 2007;).  Dextral movement on the

major faults of the northern Walker Lane, including the Pyramid Lake and Warm Springs Valley

faults began ~ 9-3 Ma. (Henry et al., 2007; Faulds and Henry, 2008).  Dextral faults in the

northern Walker Lane accommodated approximately 20-30 km of cumulative displacement

during this time (Faulds et al., 2005; Henry et al., 2007; Faulds and Henry, 2008; Hinz et al.,

2009).  

Stratigraphy

Basement-

The basement of the northern Virginia Mountains and Astor Pass consists of Mesozoic

metasedimentary rocks and Cretaceous and/or Jurassic granodiorites. The Mesozoic

metasedimentary rocks are similar to those found in the Nightingale and Peavine sequences

(Bonham and Papke, 1969; Henry et al., 2004). These rocks do not crop out in the northern

Virginia Mountains but are identified in well cuttings at ~1200 m depth at Astor Pass and crop

out at Dogskin Mountain to the southwest and the southern Lake Range to the southeast (Henry

et al., 2004; Drakos, 2007).  

Tertiary-

The majority of Astor Pass and to a lesser extent the northern Virginia Mountains are

composed of a thick section of middle Miocene volcanic rocks informally known as the Pyramid

sequence (Bonham and Papke, 1969; Faulds and Henry, 2002; Henry et al., 2004; Vice, 2008).

This thick sequence nonconformably overlies the Mesozoic basement. One source of this

sequence appears to be a Miocene shield volcano in the northern Virginia Mountains, possibly

associated with ancestral Cascade arc volcanism (Faulds and Henry, 2002). Locally, the Pyramid

sequence is broken into an upper and lower sequence, separated by the middle Miocene tuff of

Mullen Pass (Vice, 2008). This tuff crops out stratigraphically up-section of Miocene rocks

exposed at Astor Pass to the east in the Terraced Hills and does not appear in the drill cuttings.

Thus, the volcanic and volcaniclastic units found in Astor Pass appear to correlate with the lower

Pyramid sequence. This lower part of the sequence is primarily composed of interbedded basalt

and basaltic andesite flows, volcaniclastic rocks, as well as sparse rhyolitic flows (Faulds et al.,

2007; Henry et al., 2004). The northern Virginia Mountains also contain a large rhyolitic dome,

presumably coeval with the Pyramid sequence. Cuttings from wells drilled at Astor Pass show

that the Pyramid sequence volcanic rocks are approximately 1100 m thick.  The mafic lavas

include a variety of textures, including aphyric to finely porphyritic, porphyritic with phenocrysts

up to 9 mm long, and aphanitic flows, some of which are intercalated with volcaniclastic rocks.

The clastic deposits found in the Astor Pass cuttings are matrix supported, poorly to moderately

sorted, light gray conglomerates, breccias, and sandstones.  The clasts found in the sedimentary

rocks were derived primarily from the Pyramid sequence volcanic rocks, with the sandstone and

matrix in the conglomerates also consisting of feldspars and lithics sourced from the Pyramid

sequence.  These findings are in accord with previous descriptions of the Pyramid sequence

volcanic and sedimentary units in the surrounding area (Faulds et al., 2003, 2007; Henry et al.,

2004; Drakos, 2007; Vice, 2008).

Quaternary-

The majority of Astor Pass and some parts of the northern Virginia Mountains are

overlain by Quaternary deposits, primarily lacustrine deposits, tufa deposits, and young alluvial

fans (Vice, 2008). The tufa in Astor Pass is of particular importance. The tufa mounds that crop

out on the floor of Astor Pass occur in linear trends, suggesting a fault-controlled origin and

recent geothermal activity. These tufa mounds can also grow to over 90 m in height as seen at

the Needle Rocks ~6 km to the southeast of Astor Pass (Coolbaugh et al., 2006).

Structural Framework

Astor Pass occupies a narrow ~0.5 km wide pass between the Terraced Hills to the east

and north and the Virginia Mountains to the southwest.  The Terraced Hills are composed of a

series of moderately east-tilted fault blocks composed primarily of the middle Miocene Pyramid

sequence (Vice, 2008).  Based on kinematic and seismic reflection data, along with offset

stratigraphy, the north-northwest-striking faults appear to have accommodated both normal and

right-lateral motion, whereas the WNW-striking faults probably accommodated primarily dextral

displacement (Vice, 2008).  

Fault blocks in the northern-most Virginia Mountains also have a moderate easterly tilt.

To the south, the Virginia Mountains contain a north-trending extensional anticline with NNW-

striking normal faults (Faulds and Henry, 2002; Faulds et al., 2003; Henry et al., 2004). The

limbs of the anticline dip ~30º, similar to the east-tilted fault blocks in the Terraced Hills.

Preliminary analysis of Astor Pass reveals a similar structural framework as the

surrounding area; seismic reflection data shows multiple east- and west-dipping normal faults

linked with apparent sub-vertical right-lateral or oblique-slip faults. The sub-vertical northwest

striking fault shows oblique motion, whereas the north-striking faults appear to be primarily

normal faults based on fault dip (Vice, 2008). Paleomagnetic data also show that at least part of

the Terraced Hills have been rotated slightly clockwise (~12º) (Vice, 2008).  

Objectives

The objective of this project is to develop a conceptual structural model of the Astor Pass

geothermal system by: 1) constructing a detailed 3D structural model, 2) identifying faults and

fault intersections conducive to fluid flow, and 3) continuing to develop the stratigraphic,

kinematic, and stress models of the area.  

This study will involve: 1) detailed 1:24,000 geologic mapping in the northern Virginia

mountains, 2) construction of detailed cross sections using the detailed mapping and

interpretations of seismic reflection data, 3) incorporation of these cross sections into a new fault

trace map for Astor Pass, 4) a detailed analysis of well cuttings and borehole stress data, 5)

integration of new and previous mapping, cross sections, petrographic analysis, and borehole

breakout data into a 3D structural model using Earthvision software, and 6) GIS compilation of

an updated geologic map of the Astor Pass geothermal area.

Methods

The methods involved in this study include geologic mapping, interpretation of seismic

reflection data, petrographic analysis, stress data analysis, 3D structural modeling, and GIS

compilations.

Geologic Mapping - Detailed geologic mapping of ~25 km2 of the northern Virginia Mountains

will be conducted at a scale of 1:24,000. This map area will lie primarily within the Astor Pass

7.5 minute topographic quadrangle. Contacts and faults will be mapped in the field on

stereographic air photos and transferred into digital form using Vr software.  The Vr software

allows the stereo air photos to be georeferenced and ortho-rectified yielding an accurate 3D

terrain model overlain with geologic data. These data will then be exported to ArcGIS 10.1 to

produce a final cartographic publication. The GIS production will then be added to the ongoing

compilation of a geologic map for the Pyramid Lake area.   The geologic map and available

subsurface data (e.g., well cuttings and seismic reflection data) will be used to construct six

detailed cross-sections for both the Astor Pass area and across the entire study area.  

Logging of drill holes - Three potential geothermal production wells and several shallow

temperature gradient holes will be logged for lithology and faults using a research-grade

binocular microscope and hand lens. Cuttings of particular importance are those showing

evidence of faults, including gouge and zones of lost circulation during drilling. Lithology from

the logs will then be used to constrain the subsurface geology in the 3D geologic model.   

Structural Analysis - Analysis of the geometry and kinematics of faults at Astor Pass and the

Virginia Mountains will be conducted using surface and borehole indicators. Borehole breakout

data will be the primary method of analyzing stress and dilation at Astor Pass. Borehole imaging

was collected during drilling of the APS-2 and APS-3 wells. By using wellbore images,

directions of the maximum and least principle horizontal stresses can be determined using

breakouts (Zoback et al., 2003). Breakouts occur at the azimuth of the minimum horizontal

stress, appearing in an unwrapped image as a dark band of low reflectance on opposing sides of

the well. The width of the breakout (opening angle) and the orientation are then easy to

determine. Tensile fractures may also exist 90° from the breakout showing failure in

compression as well as tension (Zoback et al., 2003). Particular attention will be paid to the

orientation of the least principle stress and if possible, its measured magnitude. The orientation

and magnitudes of the principal stress allows for determination of the slip and dilation tendency

of individual faults or fault segments (e.g., Moeck et al., 2009). The analysis of the images and

calculations of stresses from the breakout data will most likely be conducted by colleagues at

GNS Science in New Zealand, but I will then synthesize this work with the other data sets to

develop a comprehensive stress model for Astor Pass.

Structural features, including bedding attitudes, flow foliations, and compaction

foliations, will all be recorded while mapping where available. Preliminary mapping indicates

that well-exposed fault surfaces are scarce throughout Astor Pass and the Virginia Mountains,

but fault traces and offsets can be mapped accurately and will constrain fault attitudes and

kinematics. Fault kinematic indicators, such as striations, accretionary mineral steps, Riedel

shears, and rough facets will also be measured if available (e.g., Angelier et al., 1985; 1994).

Accretionary mineral steps are a simple way of determining the slip direction on fault surfaces,

with surface crystal growth during fault movement forming steps in the same direction as

motion. Riedel shears also act as an excellent kinematic indicator, forming at narrow acute

angles to the fault plane (15-25o). The acute angle of the Riedel points in the direction of

movement of the host block (Figure 3). If enough fault kinematic data can be obtained, principle

stress directions will be calculated and compared to those from the borehole breakout data using

the methodology described by Marrett and Allmendinger (1990).  

Seismic Reflection Interpretations - During the summer of 2010, seismic reflection data were

collected and/or reprocessed in the Astor Pass area. Previous seismic data collected by

ZAPATA-Blackhawk in 2006, 5 lines totaling ~8.4 km, was reprocessed by Optim in 2010. In

addition, 11 new profiles totaling ~40 km in length were collected. All 16 seismic lines were

processed by Optim, including first-arrival velocity optimization and pre-stack depth migration

analysis (Louie et al., 2011).

In an attempt to more clearly image the steeply dipping normal faults in the area,

acquisition of the 11 new seismic lines used long offsets (i.e., larger distances from sources to

receivers). These long offsets focused the survey toward imaging structure and stratigraphy at

predicted reservoir depths of 600-1500 m instead of higher resolution data at shallower depths as

would be the case with shorter offsets.  All data collected showed reflections and terminations to

1500 m, as well as velocity control to 900 m. Processing was conducted using Optim’s

proprietary Advanced Seismic Imaging, as well as SeisOpt®. Depth-migrated images were then

interpreted for faults and stratigraphy using OpendTect software. Fault picks were based

primarily on obvious trends in terminations and offsets of sub-horizontal to dipping reflections,

as well as fault-plane reflections (Louie et al., 2011).

Petrographic analysis - Along with the cuttings analysis, hand samples and thin sections will be

described in detail. Hand samples will be analyzed to correlate units across the region. Thin

sections will also be used for correlating units, as well as for selecting samples for

geochronologic dating.  

3D modeling – A three-dimensional model will be constructed using a combination of geologic

mapping, seismic reflection interpretations, well cuttings analysis, and the structural analysis of

the borehole breakout data. The Astor Pass model will be focused on an approximately 10 km2

area centered on the tufa mounds. The fault model is built primarily using fault traces from the

seismic interpretation in combination with mapped faults in the model area. The fault model is

also correlated with data from the well cuttings, specifically lost circulation zones and cuttings

showing fault gouge or slickenlines. The stratigraphic horizon model is then built into the

completed fault model, drawing data from well cuttings, geologic mapping, and a cross section

produced by Vice (2008). This completed 3D model may then be used to visualize and detect

structural zones important to a more detailed understanding of the geothermal system. Specific

fault segments and fault intersections and their locations can be studied in detail.

Structural analysis of the fault segments and fault intersections can also be performed by

exporting the modeling data into 3D Stress. The stress measurements calculated from the

borehole breakout data are then applied to the fault segments and intersections, showing

localized areas of slip and dilation tendency.

GIS compilation – Line work from the detailed field mapping will be digitized using Vr

software to display the work in three-dimensional space. The resulting shape files will then be

exported into ArcGIS for further modification of attributes and formatting. Field mapping

datasets will be combined with temperature, magnetic, and other geophysical data from Astor

Pass to complement the existing GIS geologic database for the Pyramid Lake area.

PRELIMINARY SCHEDULE

Semester GoalSpring 2011 Choose project

Begin literature reviewSummer 2011 Field reconnaissance and detailed geologic mapping

Begin logging well cuttingsBegin interpreting seismic reflection profiles

Fall 2011 Complete seismic reflection profilesComplete logging well cuttingsComplete cross sectionsUpdate current geologic map

Spring 2012 Submit program of studyComplete preliminary 3D model

Summer 2012 Complete field workFinish digitizing mapFinish 3D model

Fall 2012 Finalize mapSubmit first thesis draftPresent paper at Geothermal Resources Council MeetingAGU Poster with Drew Siler

Spring 2013 Complete thesisDefend ThesisPresent work at Exxon Global Geoscience meetingPresent work at Nevada Petroleum and Geothermal Society meeting

EXPECTED PRODUCTS

A GIS database that includes a geologic map and well data of the Astor Pass area.

A geologic map of the northernmost Virginia Mountains. This map may be published as

a Nevada Bureau of Mines and Geology open-file report.

A completed 3D structural model of Astor Pass and immediate surrounding area.

Publication of a paper in the Geothermal Resources Council Transactions.

Publication of an article in a peer-reviewed journal.

References Cited

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.

Angelier, J., 1994, Fault slip analysis and palaeostress reconstruction. In: Hancock, P.L., (Ed.), Continental Deformation, Pergamon, pp. 53–100.

Blackwell, D. D., 1983, Heat flow in the Northern Basin and Range Province. Geothermal Resources Council Special Report, 13, p. 81 – 93.

Blackwell, D. D., Wisian, K., Benoit, D., and B., G., 1999, Structure of the Dixie Valley Geothermal System, a "Typical" Basin and Range Geothermal System, From Thermal and Gravity Data: Geothermal Resource Council Transactions, v. 23, p. 525-531.

Blackwell, D.D., and Richards, M., 2004, Geothermal Map of North America: AmericanAssociation of Petroleum Geologists, 1 sheet, scale 1:6,500,000.

Bonham, HF., 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.

Coolbaugh, M. F., Faulds, J. E., Kratt, C., Oppliger, G. L., Shevenell, L., Calvin, W., Ehni, W. J., et al. 2006b, Geothermal Potential of the Pyramid Lake Paiute Reservation, Nevada, USA: Evidence of Previously Unrecognized Moderate-Temperature (150-70°C) Geothermal Systems. Geothermal Resources Council, v. 30, p. 59-67.

Faulds, J.E., and Henry, C.D., 2002, Tertiary Stratigraphy and Structure of the Virginia Mountains, Western Nevada: Implication for Development of the Northern Walker Lane: Geological Society of America: Abstracts with Programs, v. 5, p. 84.

Faulds, J.E., Henry, C.D., and dePolo, C.M., 2003, Preliminary geologic map of the Tule Peak Quadrangle, Washoe County, Nevada: Nevada Bureau of Mines and Geology Open-File Report 03-10.

Faulds, J.E., dePolo, C.M., and Henry, C.D., 2007, Preliminary geologic map of the Sutcliffe Quadrangle, Washoe County, Nevada: Nevada Bureau of Mines and Geology Open-File Report 03-17, 1:24,000.

Faulds, J. E., Coolbaugh, M. F., Blewitt, G., and Henry, C. D., 2004, Why is Nevada in hot water? Structural controls and tectonic model of geothermal systems in the northwestern Great Basin: Geothermal Resource Council Transactions, v. 28, p. 649-654.

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., Hinz, N.H., Drakos, P.S., and Delwiche, B., 2005b, Transect Across the Northern Walker Lane, Northwest Nevada and Northeast California: An Incipient Transform Fault Along the Pacific – North American Plate Boundary, in Pederson, J.,

and Dehler, C.M., eds., Interior western United States: Geological Society of America Field Guide 6, p. 129-150, doi:10.1130/2005.fld006(06).

Faulds, J. E., Coolbaugh, M. F., Vice, G. S., & 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. 69-76.

Faulds, J.E., and Henry, C.D., 2008, Tectonic influences on the spatial and temporal evolution of the Walker Lane : An incipient transform fault along the evolving Pacific – North American plate boundary: Arizona Geological Society, v. 22, p. 437-470.

Faulds, J. E., Coolbaugh, M. F., Benoit, D., Oppliger, G., Perkins, M., Moeck, I., and Drakos, P., 2010, Structural Controls of Geothermal Activity in the Northern Hot Springs Mountains, Western Nevada: The Tale of Three Geothermal Systems (Brady's, Desert Peak and Desert Queen): Geothermal Resource Council Transactions, v. 34, p. 675-683.

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. A., and Morris, A. P., 2003, Dilational normal faults: Journal of Structural Geology, v. 25, p. 183-196.

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, no. B8, p. 1-21.

Hammond, W.C., Blewitt, G., and Kreemer, C., 2011, Block modeling of crustal deformation of the northern Walker Lane and Basin and Range from GPS velocities: Journal of Geophysical Research, v. 116, no. B4, p. 1-28.

Henry, C.D., Faulds, J.E., dePolo, C.M., and Davis, D.A., 2004, Geology of the Dogskin Mountain Quadrangle, northern Walker Lane, Nevada: Nevada Bureau of Mines and Geology Map 148, scale 1:24,000, 13 p text.

Henry, C.D., Faulds, J.E., and Craig, M., 2007, Geometry and timing of strike-slip and normal faults in the northern Walker Lane, northwestern Nevada and northeastern California: Strain partitioning or sequential extensional and strike-slip deformation?: Geological Society of America Special Paper, v. 434, no. 04, p. 59-79.

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 Resource Council Transactions, v. 32, p. 285-290.

Hinz, N.H., Faulds, J.E., and Henry, C.D., 2009, Tertiary volcanic stratigraphy and paleotopography of the Diamond and Fort Sage Mountains – Constraining slip along the Honey Lake fault zone in the northern Walker Lane, northeastern California and western Nevada, in Oldow, J.S., and Cashman, P., eds., Late Cenozoic structure and evolution of the Great Basin-Sierra Nevada transition: Geological Society of America Special Paper 447, p. 101-132, doi:10.1130/2009.2447 (07).

Hinz, N. H., Faulds, J. E., Moeck, I., Bell, J. W., and Oldow, J. S., 2010, Structural Controls of Three Blind Geothermal Resources at the Hawthorne Ammunition Depot, West-Central Nevada: Geothermal Resource Council Transactions, v. 34, p. 785-790.

Hinz, N. H., Faulds, J. E., and Stroup, C., 2011, Stratigraphic and Structural Framework of the Reese River Geothermal Area, Lander County, Nevada: A New Conceptual Structural Model: Geothermal Resource Council Transactions, v. 35, p. 827-832.

Kratt, C., Sladek, C., Coolbaugh, M.F., Valley, H., Valley, M.C., Pass, E., Vxuyh, J.P., Wkuhh, V.L.Q., Edvlqv, R., and Ri, V., 2010, Boom and Bust with the Latest 2m Temperature Surveys : Dead Horse Wells , Hawthorne Army Depot , Terraced Hills , and Other Areas in Nevada: Geothermal Resources Council Transactions, v. 34., p. 567-574

Louie, J., Pullammanappallil, S., Faulds, J., Eisses, A., Kell, A., Frary, R., & Kent, G., 2011. Astor Pass Seismic Surveys Preliminary Report, PLPT ARRA Phase II Report, p. 1-25.

Marrett, R., and Allmendinger, R.W., 1990, Kinematic analysis of fault-slip data:  Journal of Structural Geology, v. 12, p. 973-986.

Moeck, I., Kwiatek, G., and Zimmermann, G., 2009, Slip tendency analysis, fault reactivation potential and induced seismicity in a deep geothermal reservoir: Journal of Structural Geology, v. 31, no. 10, p. 1174-1182.

Oldow, J.S., 1992, Late Cenozoic displacement partitioning in the northwest Great Basin, in Stewart, J., ed., Structure, Tectonics and Mineralization of the Walker Lane: Walker Lane Symposium Proceedings Volume, Geological Society of Nevada, Reno, Nevada, p. 17-52.

Stewart, J.H., 1988, Tectonics of the Walker Lane belt, western Great Basin: Mesozoic and Cenozoic deformation in a zone of shear, in Ernst, W.G., ed., The Geotectonic development of California: Prentice Hall, Englewood Cliffs, New Jersey, p 683-713

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.

Zoback, M.D., Barton, C.A., Brudy, M., Castillo, D.A., and Finkbeiner, T., 2003, Determination of stress orientation and magnitude in deep wells: Rock Mechanics, v. 40, p. 1049-1076.

Figure 1 - Distribution of geothermal systems in the Great Basin. Astor Pass is shown on the boundary of Walker Lane in the Black Rock Desert group of systems (from Faulds et al., 2011).

Figure 2 - Regional map of the greater Pyramid Lake area, showing major northwest-striking dextral faults associated with the Walker Lane and north-striking normal faults of the Basin and Range. Box outlines the current 3D model area. Larger red box shows the entire study area.

Figure 3 - Cross section of a fault showing direction of motion of the missing block relative to preserved surface. Riedel shears (small arrows) point in the direction of movement of the block they are found in (from Angelier et al., 1985).


Recommended