<<

RESEARCH

Rates of extension along the Fish Lake Valley fault and transtensional deformation in the Eastern shear zone–Walker Lane belt

Plamen N. Ganev1,*, James F. Dolan1, Kurt L. Frankel2, and Robert C. Finkel3,4 1DEPARTMENT OF EARTH SCIENCES, UNIVERSITY OF SOUTHERN CALIFORNIA, LOS ANGELES, CALIFORNIA 90089, USA 2SCHOOL OF EARTH AND ATMOSPHERIC SCIENCES, GEORGIA INSTITUTE OF TECHNOLOGY, ATLANTA, GEORGIA 30332, USA 3DEPARTMENT OF EARTH AND PLANETARY SCIENCE, UNIVERSITY OF CALIFORNIA–BERKELEY, BERKELEY, CALIFORNIA 94720, USA 4CENTRE EUROPÉEN DE RECHERCHE ET D’ENSELIGNEMENT DES GÉOSCIENCES DE L’ENVIRONNEMENT (CEREGE), 13545 AIX EN PROVENCE, FRANCE

ABSTRACT

The oblique-normal-dextral Fish Lake Valley fault accommodates the majority of Pacifi c–North America plate-boundary deformation east of the San Andreas fault in the northern part of the Eastern California shear zone. New rates for the extensional component of fault slip, deter- mined with light imaging and detection (LiDAR) topographic data and 10Be geochronology of four offset alluvial fans, indicate a northward increase in extension rate. The surface exposure ages of these fans range from ca. 71 ka at Perry Aiken Creek and Indian Creek to ca. 94 ka and ca. 121 ka at Furnace Creek and Wildhorse Creek, respectively. These ages, combined with the measured vertical components of slip at each site, an assumed 60° fault dip, and a N65°E extension direction, yield calculated late Pleistocene–Holocene horizontal extension rates of 0.1 ± 0.1, 0.3 ± 0.2, 0.7 +0.3/–0.1, and 0.5 +0.2/–0.1 mm/yr at Furnace Creek, Wildhorse Creek, Perry Aiken Creek, and Indian Creek, from south to north, respectively. Comparison of these rates with geodetic measurements of ~1 mm/yr of N65°E extension across the northern Eastern California shear zone indicates that the Fish Lake Valley fault accommodates approximately half of the current rate of regional extension. When summed with published rates of extension for faults at the same latitude, the Fish Lake Valley fault data indicate that long-term geo- logic deformation rates are commensurate with short-term geodetic extension rates. The northward increase in Pleistocene extension rates is opposite the northward decrease in dextral slip rate trend along the Fish Lake Valley fault, likely refl ecting a diffuse extensional transfer zone in northern Fish Lake Valley that relays slip to the northeast across the Mina Defl ection and northward into the Walker Lane belt.

LITHOSPHERE; v. 2; no. 1; p. 33–49; Data Repository 2009285. doi: 10.1130/L51.1

INTRODUCTION et al., 2007a, 2007b, 2008). North of the active Valley fault systems (Burchfi el et al., 1987; left-lateral Garlock fault, motion is accommo- Applegate, 1995; Dixon et al., 1995; Reheis and The Eastern California shear zone–Walker dated on four major fault systems: the Owens Dixon, 1996; Lee et al., 2001, 2009a; Oswald Lane belt is an evolving fault system east of the Valley, Panamint Valley–Saline Valley–Hunter and Wesnousky, 2002; Walker et al., 2005; San Andreas fault that accommodates ~20%– Mountain, –Fish Lake Valley, and Andrew and Walker, 2009). North of the Mina 25% (9–10 mm/yr) of Pacifi c–North America Stateline fault zones (Fig. 1; e.g., Beanland and Defl ection, strain is accommodated by a series plate-boundary deformation (e.g., Dokka and Clark, 1994; Lee et al., 2001, 2009a, 2009b; of right-lateral faults as part of the Walker Lane Travis, 1990b; Humphreys and Weldon, 1994; Oswald and Wesnousky, 2002; Kylander-Clark belt (e.g., Nielsen, 1965; Stewart, 1988; Oldow Hearn and Humphreys, 1998; Thatcher et al., et al., 2005; Bacon and Pezzopane, 2007; Fran- et al., 1989, 2001; Oldow, 2003; Wesnousky, 1999; Dixon et al., 2000, 2003; McClusky et al., kel et al., 2007a, 2007b, 2008a, 2008b; Guest et 2005a, 2005b). 2001; Miller et al., 2001; Bennett et al., 2003; al., 2007; Kirby et al., 2008). In the northern part Although dextral shear accommodates most Faulds et al., 2005; Wesnousky, 2005a, 2005b; of the Eastern California shear zone, between slip on the northwest-striking faults in Fish Lake Frankel et al., 2007a). The Eastern California latitude 37°N and 38°N, dextral motion between Valley, fault segments that strike more northerly shear zone extends northward for almost 500 km the stable Sierra block and North Amer- exhibit a relatively large extensional component from the San Andreas fault through the Mojave ica is distributed from Long Valley caldera in the of slip. Space-based geodetic studies demon- Desert and along the eastern side of the Sierra west to the Silver Peak–Lone Mountain exten- strate that the current strain fi eld is consistent Nevada. In the Mojave section of the shear sional complex in the east (e.g., Burchfi el et al., with transtensional deformation and indicate an zone, fault motion is primarily right-lateral, 1987; Brogan et al., 1991; Berry, 1997; Reheis extension rate of ~1 mm/yr oriented normal (i.e., with a subordinate component of north-south and Sawyer, 1997; Hearn and Humphreys, N65°E) to the predominant ~N25°W strike of shortening, and slip is localized along several 1998; Dixon et al., 2000; Oldow et al., 2001; the Eastern California shear zone at this latitude major north-northwest–striking faults (Dokka, Petronis, 2005; Kirby et al., 2006; Frankel et (e.g., Savage et al., 2001; Bennett et al., 2003; 1983; Bartley et al., 1990; Dokka and Travis, al., 2007a, 2007b; Lee et al., 2009b). Multiple Wesnousky, 2005a). This extension rate in the 1990a, 1990b; Schermer et al., 1996; Walker northeast-striking, down-to-the-northwest nor- northern Eastern California shear zone is most and Glazner, 1999; Glazner et al., 2002; Oskin mal faults transfer slip between the right-lateral likely distributed among the Sierra Nevada fron- Owens Valley, Panamint Valley–Saline Valley– tal fault, the faults of the Volcanic Tablelands, *Corresponding author e-mail: [email protected]. Hunter Mountain, and Death Valley–Fish Lake the White Mountains fault, and the Fish Lake

For permission to copy, contact [email protected] | © 2010 Geological Society of America 33

Downloaded from http://pubs.geoscienceworld.org/gsa/lithosphere/article-pdf/2/1/33/3050053/33.pdf by guest on 29 September 2021 GANEV ET AL.

Valley fault system (Fig. 1). The geologically determined component of the extension rates 119° W 118° W on the Sierra Nevada frontal fault (Berry, 1997; Le et al., 2006) and the White Mountains fault (Kirby et al., 2006) are both ~0.2–0.3 mm/yr, PSFP CA NV S and the rate across the distributed normal faults F of the Volcanic Tablelands is ~0.3 mm/yr (Shee- WFW BSFB S F F han, 2007). Thus, the three main fault systems image to the west of the Fish Lake Valley fault collec- RFRF tively accommodate approximately half of the current rate of extension within the northern part Mina DeflectionEFEF of the Eastern California shear zone. CFCF In this paper, we investigate how much MLFM F M of the remaining extension is accommodated L LMFL F by the Fish Lake Valley system. Specifi cally, SLFS F GPS

38° N V L QVFQ F 9.3+/-0.2 mm/yr herein we report new observations from our F P EPFE analysis of high-resolution digital topographic White Mtns. HCFH GPS airborne laser swath mapping (light detection C F FLVFF and ranging [LiDAR]) data, as well as ter- WMFW L VF restrial cosmogenic nuclide (TCN) dating of M ~1 mm/yr RVFR F faulted landforms along the Fish Lake Valley V

F F fault. Our results provide new, geochronologi- S DSFD cally determined, late Pleistocene– Holocene extension rates on this fault system that clarify

Inyo Mtns. how motion in the Eastern California shear zone is accommodated and, ultimately, trans- F V ferred northward to the Walker Lane belt. EVFE NDVFN 37° N OVFO D Moreover, these results allow us to assess the V F V NNV F M F degree to which short-term geodetic and lon- TMFT CA ger-term geologic rates are consistent, and the HMSVFH extent to which deformation is accommodated MS F Sierra Nevada P VF TPFT by slip on geomorphically well-defi ned faults Panamint Mtns. versus distributed, off-fault deformation.

Coso Range DEATH VALLEY–FISH LAKE VALLEY

AHFA FAULT ZONE H SNFS N F

F The Fish Lake Valley fault, which forms the northern 80 km of the Death Valley–Fish Lake PVFP 36° N V Valley fault system, is marked by steep, east- ALFA F L facing fault scarps, ponded drainages, and shutter NN F ridges indicative of recent fault activity (Sawyer, 1990; Brogan et al., 1991; Reheis, 1992; Reheis et ¯ GFGF al., 1993, 1995; Hooper et al., 2003; Frankel et al., 05025 2007a, 2007b, 2008b). The southern and central km sections of the Fish Lake Valley fault strike pre- dominantly north-northwest, whereas the north- Figure 1. Map of Quaternary faults in eastern California and western Nevada showing the major right step in the Eastern California shear zone–Walker Lane system across the Mina Defl ection. ern part of the fault is characterized by numer- Fish Lake Valley fault is shown in white. Black arrows show generalized fault-parallel and fault- ous north- and northeast-striking strands that perpendicular components of geodetic velocity fi eld (Bennett et al., 2003; Wesnousky, 2005a). The splay out into Fish Lake Valley from the main corners of Figure 2 are shown in black. Faults are from the U.S. Geological Survey (USGS) Quater- north-northwest–trending, range-bounding fault nary fault and fold database. AHF—Ash Hill fault, ALF—Airport Lake fault, BSF—Benton Springs (Fig. 2; Sawyer, 1991; Reheis and Sawyer, 1997). fault, CF—Coaldale fault, EF—Excelsior Mountains fault, EPF—Emigrant Peak fault, EVF—Eureka Right-lateral motion on the Fish Lake Val- Valley fault, DSF—Deep Springs fault, FLVF—Fish Lake Valley fault, GF—Garlock fault, HCF—Hil- ton Creek fault, HMSVF—Hunter Mountain–Saline Valley fault, LMF—Lone Mountain fault, MLF— ley fault is thought to have begun ~10 m.y. ago fault, NDVF—northern Death Valley fault, OVF—Owens Valley fault, PSF—Petrifi ed (Reheis and Sawyer, 1997), and the strike-slip Springs fault, PVF—Panamint Valley fault, QVF—Queen Valley fault, RF—Rattlesnake Flat fault, rate averaged over late Pleistocene–Holocene RVF—Round Valley fault, SLF—Silver Lake fault, SNF—Sierra Nevada frontal fault, SVF—Saline time is 2.5–3 mm/yr (Frankel et al., 2007b). Valley fault, TMF—Tin Mountain fault, TPF—Towne Pass fault, WF—Warm Springs fault, WMF— The extensional component of oblique-normal- White Mountains fault. dextral motion, responsible for the opening of Fish Lake Valley, most likely began ~5 m.y. ago,

34 www.gsapubs.org | Volume 2 | Number 1 | LITHOSPHERE

Downloaded from http://pubs.geoscienceworld.org/gsa/lithosphere/article-pdf/2/1/33/3050053/33.pdf by guest on 29 September 2021 Rates of extension along the Fish Lake Valley fault | RESEARCH

′ ′ 1995). These four alluvial fans were depos- N

N 118°30 W 118°00 W ′ ′ ited along the eastern White Mountains pied- mont at the mouths of (from south to north) 38°00 38°00 Furnace Creek, Wildhorse Creek, Perry Aiken F Creek, and Indian Creek. We refer to each of V our study sites relative to the respective creek QVFQ that formed them. At two of the sites, Furnace Creek and Indian Creek, Frankel et al. (2007b) F P measured large (180–290 m) dextral offsets of EPFE Silver Peak Range alluvial surfaces that they dated with cosmo- 4 genic 10Be surface exposure geochronology at Fish Lake Valley 94 ± 11 ka and 71 ± 8 ka, respectively. Their resulting right-lateral strike-slip rates at these 3 two locations are 3.1 ± 0.4 mm/yr at Furnace Creek and 2.5 ± 0.4 mm/yr at Indian Creek.

VolcanicV Tablelands Multiple normal fault scarps are also present in o l both of these alluvial fans, as we discuss in the c 2 a White Mtns. following sections. n i 1 FLVFF c L Our other two study sites, Wildhorse Creek

V N N T ′ ′ and Perry Aiken Creek, are located between the a F b Furnace Creek and Indian Creek fans (Fig. 1). l WMFW e 37°30 37°30 l Both the Wildhorse Creek and Perry Aiken a M n Creek sites have numerous normal fault scarps. d F s The tallest single normal fault scarp (minimum vertical component of displacement of 73.5 m)

F along the entire Death Valley–Fish Lake Valley S Eureka Valley fault zone is found just north of Perry Aiken DSFD Owens Valley Creek (Reheis et al., 1993; Reheis and Sawyer, 1997). Reheis et al. (1993) used tephrochronol- Sierra Nevada ogy to suggest that alluvial fans at Wildhorse Creek and Perry Aiken Creek were deposited during middle to late Pleistocene time.

TERRESTRIAL COSMOGENIC NUCLIDE Inyo Mtns. GEOCHRONOLOGY

OVFO

V We used terrestrial cosmogenic nuclide geo-

F N N ′ chronology (TCN) to date the offset alluvial fans along the Fish Lake Valley fault. TCN geo- ¯ 37°00 chronology allows us to determine the age of abandonment of an alluvial surface (e.g., Gosse 0105 km and Phillips, 2001). TCN geochronology mea- ′ sures the concentration of nuclides produced in 118°00 W a rock by the interaction between cosmic rays Figure 2. Detailed map of the study area showing the Fish Lake Valley fault in white. The loca- and minerals at Earth’s surface (e.g., Lal, 1991; tions of our study sites are labeled 1 through 4 from south to north: 1—Furnace Creek; 2—Wild- Gosse and Phillips, 2001). In order to obtain an horse Creek; 3—Perry Aiken Creek; 4—Indian Creek. DSF—Deep Springs fault, EPF—Emigrant Peak accurate age for abandonment of each sampled fault, FLVF—Fish Lake Valley fault, OVF—Owens Valley fault, QVF—Queen Valley fault; WMF—White geomorphic surface, several criteria must be sat- Mountains fault. isfi ed: (1) the sampled boulder must be in the same geometry as it was at the time of deposi- tion; (2) the sampled boulder should not have as suggested by the observation that the bound- oblique-slip rate is ~0.9 mm/yr, parallel to a net prior exposure history (inheritance); and (3) ing faults in the northern section of the valley slip vector plunging ~20° toward N10°–20°W boulders with evidence of erosion should not cut across sedimentary rocks of the Miocene (Kirby et al., 2006). be sampled since they will provide an attenu- Esmeralda basin (Reheis and Sawyer, 1997; The focus areas of our study are normal ated concentration of cosmogenic isotopes and, Petronis et al., 2002, 2009; Petronis, 2005). fault scarps formed in four alluvial fans along hence, an anomalously young age (Gosse and On the west side of the White Mountains, the the Chiatovich Creek, Dyer, and Oasis sections Phillips, 2001). right-lateral White Mountains fault originated of the fault, which have been mapped in detail The isotope of interest for this study is later, at ca. 3 Ma (Stockli et al., 2003). The late by a number of researchers (Fig. 2; Brogan et 10Be, which is produced through spallation and Pleistocene–Holocene White Mountains fault al., 1991; Reheis, 1992; Reheis et al., 1993, muon-induced reactions with Si and O in quartz.

LITHOSPHERE | Volume 2 | Number 1 | www.gsapubs.org 35

Downloaded from http://pubs.geoscienceworld.org/gsa/lithosphere/article-pdf/2/1/33/3050053/33.pdf by guest on 29 September 2021 GANEV ET AL.

Beryllium-10 is well retained in quartz, so we

collected quartz-bearing samples from granitic .2 ,††† 6.4 6.3 6.7 boulders embedded in the surface of the faulted 5.6 6.1 6.4 5.2 *** ± 8.3 ± 8.3 alluvial fans. Fifteen samples were collected ##, 3.3 ± 9.1 15.2 ± 10.2 15.2

from the top 1–5 cm of large granitic boulders Lal (1991); 118.2 ± 10.5 118.2 (Table 1). These boulders came from the sta- (2000) Stone Age

ble parts of fan surfaces mapped by Reheis et ,††† ***

al. (1993, 1995) as Qfm (alluvium of McAfee ##, Creek) at Wildhorse Creek and Qfi (alluvium of (2005) Indian Creek) at Perry Aiken Creek (Fig. 3). We Age

carefully selected well-varnished boulders that ,††† ***

lacked evidence of erosion (for example, we did ##, not sample “sombrero-shaped” boulders). If, Dunai (2001) al. Lifton et however, any erosion has occurred, which we rate Time-varying production Age

think is probably unlikely due to the arid cli- ,††† *** a.

mate in the study area, the boulders would yield 6 ##, (ka) anomalously young ages, which would in turn (2006) Age Desilets and Desilets and Desilets al. et make our calculated slip rates too fast. Zreda (2003); §§

Quartz was purifi ed by standard techniques, and Be was extracted using ion-exchange ,††† *** ##, chromatography, precipitated as BeOH, and (ka) Constant converted to BeO at the Georgia Institute of Age production rate production Be half-life of 1.36 × 10 of 1.36 Be half-life ess.ess.washington.edu/).

Technology cosmogenic nuclide geochronology 10 ) 2 laboratory (e.g., Kohl and Nishiizumi, 1992; ,†† Bierman and Caffee, 2002). The 10Be/9Be ratio Be decay constant. 10 ± 0.04 ± 0.04 ± 0.03 ± 12.7 140.0 ± 10.8 119.0 ± 15.5 129.4 ± 15.1 127.1 ± 13.3 111.6 ± 12.5 124.6 ± 13.0 109.7 126.1 ± 11.2 ± 10.8 107.9 109.0 ± 9.6 for each sample was measured at the Center for Be 10

Accelerator Mass Spectrometry at Lawrence atoms/g SiO 6 10

Livermore National Laboratory, and model Be concentration** (10

ages were calculated using the CRONUS-Earth

# 10Be-26Al exposure age calculator (version 2.1; ) Be –13 http://hess.ess.washington.edu) and constant 9 Be/ (x10

10 10 Measured Be production rates (Lal, 1991; Stone, 2000; LAKE VALLEY Balco et al., 2008) (Table 1). In Table 1, we § report the ages using a constant production rate, Be GEOCHRONOLOGY FOR THE WILDHORSE CREEK AND PERRY AIKEN CREEK ALLUVIAL FANS IN ALLUVIAL FISH CREEK AIKEN PERRY THE Be GEOCHRONOLOGY WILDHORSE CREEK FOR AND 10 Be and a 4% uncertainty in the the in uncertainty 4% a Be and (g)

as well as time-varying production rates calcu- 10 solution lated using four different models (Lal, 1991; Be carrier counting statistics. counting

Stone, 2000; Dunai, 2001; Desilets and Zreda, † 0843 0.3030 ± 0.22 7.72 ± 0.04 1.34 ± 7.3 78.6 ± 9.0 74.9 ± 8.8 73.6 ± 7.4 72.4 ± 73.0 2003; Lifton et al., 2005; Desilets et al., 2006). (g) mass In reporting ages and slip rates herein, we use Quartz the constant-production-rate model developed Be standards prepared by Nishiizumi et al. (2007) and using a using a and (2007) et al. by Nishiizumi prepared Be standards

by Lal (1991) and Stone (2000) to ensure con- 10 sistency in our comparisons with the previously (cm) Thickness COSMOGENIC NUCLIDE COSMOGENIC determined ages of Frankel et al. (2007b), who , carrier mass (1%), and mass (1%), , carrier

used the same model for the Furnace Creek and of rate production in the ertainty Indian Creek sites. Time-varying production (m) rates could result in younger ages by up to 10%, Elevation which would yield commensurately faster rates of extension.

Perry Aiken Creek Fan Age (°N/°W) RESULTS OF TERRESTRIAL RESULTS OF was used for the granitic surface samples. surface granitic the for was used 3

Nine TCN samples were analyzed from the longitude and Latitude Qfi alluvial-fan surface at Perry Aiken Creek.

The Qfi surface is characterized by subdued to unit* moderately incised channels, well-developed desert pavement, and a continuous, thick des- ert varnish on clasts. Clasts are tightly packed No geometric shielding correction for topography was necessary (horizon <20° in all directions). directions). all <20° in necessary (horizon was for topography correction No geometric shielding Propagated uncertainties include error in the blank in error include Propagated uncertainties (2000). Stone and (1991) Lal Based on unc include ages model the in Propagated error A density of 2.7 g/cm 2.7 of A density Carrier concentration = 1155 ppm. ppm. = 1155 Carrier concentration Inc., ICN Pharmaceutical, to were compared ratios Isotope § †† ## † # §§ ††† and well sorted in a continuous pavement, and ANALYTICAL TABLE 1. PA-308-12 Qfi 37.6621/118.0997 Wildhorse Creek 1684 4 10.0993 0.3003 ± 0.16 6.82 ± 0.03 1.17 ± 6.7 72.5 ± 8.3 69.5 ± 8.1 68.3 ± 6.8 67.3 ± 67.4

PA-308-5 Qfi 37.6629/118.1049 1753 5 10.2114 0.303 0.14 5.3 ± 0.03 0.9 ± ± 4.9 53.6 ± 6.1 50.8 ± 5.9 49.9 ± 5.0 48.9 ± 4.5 49.3 **A mean blank value of 66,644 ± 13,532 atoms was used to correct for background. background. for used to correct atoms was 13,532 ± 66,644 of blank value mean **A Sample name Fan

IC-508-3 IC-508-4 Qfm IC-508-6 Qfm 37.5946/118.0570 37.5947/118.0589 1752 Qfm 1730 37.5943/118.0537 2 1760 2 20.3301 2 20.0780 0.3085 0.3069 ± 0.32 24.6 20.1210 ± 0.34 26.36 0.3050 ± 0.04 2.16 ± 0.28 22.09 2.33 ± 11.5 126.4 ± 0.03 1.93 14.1 ± 118.0 ± 13.8 116.1 ± 11.5 114.0 ± 10.2 112.4 1 ± 12.6 105.8 ± 12.3 103.9 ± 10.3 102.3 10 IC-508-2 Qfm 37.5947/118.0571 1750 1 20.0923 0.3082 ± 0.36 25.16 ± 0.04 2.23 ± 11.8 129.9 14.5 ± 121.1 14.2 ± 119.1 ± 11.8 116.9 *Map units are from Reheis et al. (1993, 1995). See Figures 7–10. 7–10. Figures See 1995). (1993, al. Reheis et from are units *Map

PA-308-13 PA-308-14 PA-308-15 Qfi Qfi 37.6628/118.0991 Qfi 37.662/118.0987 1674 37.6619/118.0987 1675 1676 1 1 2 10.0745 0.3033 10.3113 10.0097 ± 0.17 7.27 0.3032 0.3028 ± 0.24 7.18 ± 0.03 1.26 0.14 6.3 ± ± 0.04 1.21 ± 0.03 1.10 ± 7.1 77.0 ± 7.0 74.1 ± 8.8 73.8 ± 6.2 67.3 ± 8.6 72.5 ± 8.7 71.0 ± 7.7 64.7 ± 7.2 71.4 ± 8.5 69.8 ± 7.6 63.7 ± 71.6 ± 7.1 68.7 ± 6.3 62.7 ± 68.9 ± 62.8 Perry Aiken Creek Perry Aiken ***Beryllium-10 model ages were calculated with the CRONUS-Earth online calculator version 2.1 (Balco et al., 2008; http://h et al., 2008; (Balco 2.1 version calculator online CRONUS-Earth the calculated with were ages model ***Beryllium-10 PA-308-3 Qfi 37.6625/118.1063 1756 1 10.0292 0.3032 0.17 7.4 ± ± 0.03 1.29 ± 6.8 74.2 ± 8.4 70.8 ± 8.3 69.6 ± 6.9 68.4 ± 6 68.9 PA-308-1 Qfi 37.6626/118.1066 1775 5 10. 1775 5 37.6626/118.1066 PA-308-1 Qfi

PA-308-4 Qfi 37.6628/118.1049 1755 5 10.1533 0.3024 ± 0.17 7.43 ± 0.03 1.27 ± 7.0 76.0 ± 8.7 72.5 ± 8.5 71.2 ± 7.1 70.1 ± 70.5 PA-308-10 Qfi 37.6620/118.0998 1682 2 10.0598 0.3021 ± 0.12 5.94 ± 0.02 1.03 ± 5.7 62.6 ± 7.2 60.4 ± 7.0 59.4 ± 5.9 58.3 ± 58.4 IC-508-5 Qfm 37.5948/118.0588 1753 3 20.1309 0.3035 23.18 ± 0.30 2.02 ± 23.18 0.3035 20.1309 1753 3 37.5948/118.0588 IC-508-5 Qfm IC-508-7 Qfm 37.5950/118.0537 1729 3 20.1563 0.3050 ± 0.31 19.21 ± 0.03 1.68 ± 9.1 100.2 ± 11.4 95.4 ± 11.1 93.7 ± 9.2 92.4 93.0 prominent solifl uction treads and risers are also present (Reheis and Sawyer, 1997). In addition,

36 www.gsapubs.org | Volume 2 | Number 1 | LITHOSPHERE

Downloaded from http://pubs.geoscienceworld.org/gsa/lithosphere/article-pdf/2/1/33/3050053/33.pdf by guest on 29 September 2021 Rates of extension along the Fish Lake Valley fault | RESEARCH

A B

Figure 3. Representative examples of sampled boulders from (A) the Qfm deposit at Wildhorse Creek, and (B) the Qfi deposit at Perry Aiken Creek.

the Qfi deposit is distinguished by a well-devel- subcluster of six 10Be ages centered at ca. 75 ka from the 71 ± 8 ka age range for all nine samples oped soil with a 5- to 10-cm-thick silty vesicular suggests that these ages may most accurately would yield values (both normal and strike-slip) A horizon, a weak argillic B horizon with thin refl ect the age of abandonment of the Qfi sur- that are ~5% too fast. clay fi lms, and stage II to III carbonate develop- face at Perry Aiken Creek. These six ages yield ment (e.g., Reheis and Sawyer, 1997). an age for the surface of 75.4 ± 2.2 ka. In this Wildhorse Creek Fan Age Our Qfi surface exposure ages exhibit a interpretation, the three younger ages outside of pronounced, single peak distribution with a the very tight 75 ka cluster would refl ect either We analyzed six TCN samples collected mean age and standard deviation of 71 ± 8 ka slight, unrecognized erosion of the sampled from the Qfm alluvial-fan deposits at Wild- (Fig. 4A). The ages range from 54 ± 5 ka to 79 tops of these boulders and/or minor exhuma- horse Creek. The Qfm surface, referred to as ± 7 ka (Table 1). We take the tight clustering tion of these three clasts. Although we suspect “alluvium of McAfee Creek” by Reheis and of ages as an indication that the Qfi surface has that the slightly older 75.4 ± 2.2 ka age based Sawyer (1997), is characterized by subdued to remained relatively stable and that there is neg- on the tight subcluster of six samples may more moderately incised channels, well-developed ligible inheritance. The age we obtained for Qfi accurately refl ect actual age of the surface, in desert pavement, and a continuous, thick desert at Perry Aiken Creek is indistinguishable from the following discussion, we use the 71 ± 8 ka varnish on clasts. The Qfm deposit exhibits a the previously determined 10Be age of 71 ± 8 ka age based on the entire suite of nine samples for thick, silty vesicular A horizon, strong argillic B for the Qfi deposit at Indian Creek by Frankel et consistency with our other measurements and horizon, and stage IV laminar carbonate devel- al. (2007b), and falls within the 50–130 ka age those of Frankel et al. (2007b). We note, how- opment (e.g., Reheis and Sawyer, 1997). The range estimated by Reheis and Sawyer (1997) ever, that if the tight ca. 75 ka cluster of six ages pavement is locally moderately well packed and on the basis of soil development and surface does represent the true age of abandonment for well sorted, and abundant carbonate rubble is morphology. The presence of an extremely tight the Qfi surface, then the slip rate we calculate present throughout (Reheis and Sawyer, 1997).

Perry Aiken Creek A Wildhorse Creek B Figure 4. Probability density n = 9 n = 6 functions of the 10Be ages from surface boulders. (A) Probabil- 70.6 ± 8.1 ka 121.2 ± 13.7 ka ity density function of the nine samples used to determine the age of Qfi at Perry Aiken Creek. (B) Probability density function of the six samples used to deter- mine the age of Qfm at Wild-

Relative probability Relative probability horse Creek. The uncertainties are reported as the mean and standard deviation.

0 20 40 60 80 100 120 140 0 40 80 120 160 200 10Be model age (ka) 10Be model age (ka)

LITHOSPHERE | Volume 2 | Number 1 | www.gsapubs.org 37

Downloaded from http://pubs.geoscienceworld.org/gsa/lithosphere/article-pdf/2/1/33/3050053/33.pdf by guest on 29 September 2021 GANEV ET AL.

The six 10Be samples we dated from this sur- 25 profi les perpendicular to the strike of each this analysis at the Furnace Creek and Indian face range in age from 100 ± 9 ka to 140 ± 13 ka set of scarps at the four study sites. The vertical Creek sites. As discussed later, we provide a (Table 1), forming a relatively tight cluster at displacement components for each set of faults new strike-slip restoration and strike-slip rate at 121 ± 14 ka (Fig. 4B). As at Perry Aiken Creek, were measured by multiple topographic profi les, the Perry Aiken site. In marked contrast to the we take the clustered distribution of these ages and the means of these measurements were cal- other three study sites, we could not determine as evidence that the Qfm surface has remained culated (Fig. 5; Fig. DR1 and Table DR11). The a unique strike-slip restoration at the Wildhorse relatively stable since the time of deposition total vertical component of displacement for Creek site, so we used the strike-slip rate of the and that inheritance is minimal. The age we each site was determined by summation of the fault determined by Frankel et al. (2007b) and determined for the Qfm surface is signifi cantly mean values. The horizontal component of each the age of the offset alluvial surface to provide younger than the 600?–760 ka age range pro- displacement was subsequently calculated using an approximate restoration. Given the overall posed by Reheis and Sawyer (1997) on the basis simple trigonometric relationships by assuming similarity of the three well-constrained strike- of soil development and surface morphology. a 60° dip of the fault plane for each scarp (e.g., slip rates that are now available along the fault Kirby et al., 2006; Le et al., 2006, 2009a). Uncer- (Furnace Creek and Indian Creek from Fran- Furnace Creek and Indian Creek Fan Ages tainties associated with measurements of the ver- kel et al. [2007b]; Perry Aiken Creek as docu- tical components of each displacement include mented herein), we think that this restoration is The ages of the alluvial fans at Furnace surface roughness (~20 cm) and the vertical accu- likely to be approximately correct. Creek and Indian Creek were recently deter- racy (~10 cm) of the LiDAR data. Combined, mined by Frankel et al. (2007b) using cosmo- these two uncertainties are <5% of the total mean Furnace Creek genic 10Be geochronology, and we utilized their vertical component of displacement at each site, results to determine the extension rates at these and thus we report a conservative displacement Two fault sets are prominent at the Furnace two locations. At Furnace Creek, Frankel et al. error of 5%. This 5% error is also carried over to Creek site: north-northwest–trending faults of (2007b) reported an age for surface Qfi o (modi- the extensional component of displacement. We the main, predominantly right-lateral strand fi ed from Qfi of Reheis and Sawyer, 1997) of report the horizontal (extensional) component of of the Fish Lake Valley fault, and a northeast- 94 ± 11 ka, whereas at Indian Creek, they deter- displacement across the fault zone at each of the trending set of distributed normal faults to mined the age of surface Qfi y (modifi ed from four sites, rather than the total vertical separation, the east. In order to account for the difference Qfi of Reheis and Sawyer, 1997) to be 71 ± 8 ka. due to the presence of multiple scarps associated in strike between the two sets of faults at this Both of these ages are in agreement with soil with antithetic faults. Antithetic faults would lead site, we resolved the extension measured by the and fan morphology data reported previously to a reduction in the net vertical displacement profi les from the two fault sets as a vector sum from those sites by Reheis and Sawyer (1997). across the fault zone. Therefore, we feel that the oriented toward N83°E (Fig. 7). Transects PP′- total extension represents a more robust measure P1P1′ and QQ′-Q1Q1′, and R-R′ through W-W′ GEOMORPHIC ANALYSIS OF NORMAL of the net fault slip. At each site, we determined are superimposed in order to capture the vertical FAULT SCARPS an extension direction that is normal to the aver- component of displacement across two scarps age strike of the faults. Due to the presence of that juxtapose the right-laterally offset alluvial The analysis of LiDAR digital topographic multiple sets of faults with different orientations fan (Figs. 5A and 7). Inasmuch as right-lateral data is an integral component of our study. The at Furnace Creek, Wildhorse Creek, and Indian offset of the alluvial fan at this location will LiDAR data were collected in the fall of 2005 by Creek, we resolved the cumulative extension at result in apparently smaller scarp heights due the National Center for Airborne Laser Mapping these three sites as the sum of the extension vec- to juxtaposition of laterally variable topogra- (NCALM) using an Optech Inc. model ALTM tors for each fault set (Fig. 6). phy, these profi les provide us with the minimum 2033 laser mapping system. The laser was One concern in this type of analysis is the vertical displacement. In order to determine a installed on a Cessna 337 twin-engine airplane, oblique nature of slip on the Fish Lake Valley more accurate measurement of potential dip- which fl ew over the fault trace at an average ele- fault (Figs. 7–10). In the absence of direct fi eld slip motion on the mountain-front strand, we vation of 600 m above ground level and an aver- evidence for the rake of oblique slip (e.g., slick- used the strike-slip restoration of Frankel et al. age speed of 60 m/s. The pulse rate frequency ensides), we analyzed the difference between (2007b). These authors showed that restoration of the Optech ALTM 2033 was set at 33 kHz, horizontal extension and right-lateral slip along of 290 ± 20 m of right-lateral slip realigns the and it recorded the fi rst and last returns of each the main right-lateral fault strand at each of our cone-shaped geometry of the Qfi fan as well as pulse, plus the relative intensity of each return. four study sites. Specifi cally, at each site, we a major incised drainage (their Fig. 2). Our mea- The average shot density for the LiDAR data used restorations of total strike-slip motion to surement of residual scarp heights on the Fran- was ~3 points/m2. The aircraft was equipped defi ne the pre-offset geometry of the dated allu- kel et al. (2007b) restoration indicates a vertical with a dual-frequency global positioning system vial-fan surfaces. Any scarps that exhibit a verti- component of displacement of 8 ± 1 m. This (GPS) receiver and a real-time display of the cal component of slip on the restored surfaces small amount of oblique-normal displacement fl ight path and area coverage. High-resolution must be caused by true dip-slip motion, and we indicates that the main mountain-front fault digital elevation models (DEMs) with 5–10 cm used measurements of those scarp heights in strand at Furnace Creek is a near-pure right-lat- vertical accuracy and 1 m horizontal resolution our analysis of extension. We used the strike- eral strike-slip fault (strike-slip/normal: 290/8), were produced using a kriging algorithm in slip restorations of Frankel et al. (2007b) for with a slip direction oriented N48°W, plunging SURFER software (version 8.04; Carter et al., at 3° toward the northwest. 2007; Sartori, 2005). In marked contrast to the predominantly 1 ArcGIS (version 9.2) was used to produce GSA Data Repository Item 2009285, Figures DR1 right-lateral oblique-slip mountain-front strand, and DR2 and Table DR1, is available at www hill-shaded relief and topographic maps to aid .geosociety.org/pubs/ft2009.htm, or on request from the scarps we mapped to the east do not exhibit in the identifi cation and mapping of all normal [email protected], Documents Secretary, GSA, any apparent strike slip, suggesting that they fault scarps at each site. We analyzed a total of P.O. Box 9140, Boulder, CO 80301-9140, USA. are nearly pure dip-slip faults. We used profi les

38 www.gsapubs.org | Volume 2 | Number 1 | LITHOSPHERE

Downloaded from http://pubs.geoscienceworld.org/gsa/lithosphere/article-pdf/2/1/33/3050053/33.pdf by guest on 29 September 2021 Rates of extension along the Fish Lake Valley fault | RESEARCH

A Furnace Creek horizontal extension toward N83°E: 13.0 ± 0.7 m (assuming a 60° fault dip of each fault strand) Qfi 1630 Qfi 10.9 ± 0.6 m 3.6 ± 0.2 m 1645 Profile T-T΄ 1620 Profile P-P΄ 1635

1610 1625 Elevation (m) 0 40 80 120 160 0 20 40 60 80 100

Horizontal distance (m)

B 1680 65.4 ± 3.3 m Wildhorse Creek Profile O-O΄ 1660 horizontal extension toward N53°E: 42.1 ± 2.1 m (assuming a 60° fault dip of each fault strand)

1640

1620 Elevation (m)

1600 Qfm

1580

1560 0 50 100 150 200 250 300 350 400 450 500 Horizontal distance (m)

12.2 ± 0.6 m C 1680 Perry Aiken Creek Qfi Profile L-L΄ horizontal extension toward N68°E: 49.1 ± 2.5 m 1640 (assuming a 60° fault dip of each fault strand )

1600 73.5 ± 3.7 m Elevation (m)

1560

1520 0 100 200 300 400 500 600 700 800 900 1000 Horizontal distance (m) 1700 Indian Creek 1880 4.1 ± 0.2 m 4.5 ± 0.2 m Profile A-A΄ D 1870 horizontal extension toward S88°E: 43.7 ± 2.1 m Qfi (assuming a 60° fault dip of each fault strand ) 1860 2.2 ± 0.1 m 1660 Qfi 3.8 ± 0.2 m 1850 0 100 200 300 3.9± 0.2 m Qfi 1620 8.3 ± 0.4 m Profile G-G΄ 1810 Profile B-B΄ 12.4 ± 0.6 m 21.0 ± 1.0 m 3.3 ± 0.2 m 1790 1580 3.5 ± 0.2 m 11.6 ± 0.6 m Elevation (m) 9.1 ± 0.5 m 1770 1.5 ± 0.1 m

1540 1750 0 500 1000 1500 2000 2500 3000 0 200 400 600 800 Horizontal distance (m)

Figure 5. Selected topographic profi les across alluvial-fan surfaces and calculated vertical components of displacement from the four study sites: (A) Furnace Creek; (B) Wildhorse Creek; (C) Perry Aiken Creek; and (D) Indian Creek. See Fig- ures 7–10 for locations of the profi les. The gray and black lines on profi le P-P′ in Figure 5A represent the vertical compo- nent of displacement between two laterally juxtaposed features.

LITHOSPHERE | Volume 2 | Number 1 | www.gsapubs.org 39

Downloaded from http://pubs.geoscienceworld.org/gsa/lithosphere/article-pdf/2/1/33/3050053/33.pdf by guest on 29 September 2021 GANEV ET AL.

10 45 10 A B1 B2 5 40 5 19.9 ± 1.0 m at N77°E 13.0 ± 0.7 m at N83°E 0 35 0 western eastern fault set 42.1 ± 2.1 m at N53°E -5 fault set 30 -5 Extension north (m) Extension north (m) -10 25 -10 -5 0 5 10 15 -5 0 5 10 15 20 25 Extension east (m) Extension east (m) 20

15 eastern fault Extension north (m) 10

5

0 western fault set -5 -50 5 10 15 20 25 30 35 Extension east (m)

C 25 20

15 5 49.1 ± 2.5 m at N68°E D 43.7 ± 2.1 m at S88°E 10 0

5 -5 eastern western fault set fault set

Extension north (m) 0 -10

-5 Extension north (m) -15 -50 5 10 15 20 25 30 35 40 45 50 -5 0 5 10 15 20 25 30 35 40 45 Extension east (m) Extension east (m)

Figure 6. Vector resolutions for horizontal extension at (A) Furnace Creek; (B1) Wildhorse Creek—profi les N-N′ and O-O′; (B2) Wildhorse Creek—profi les NN-NN′ and OO-OO′; (C) Perry Aiken Creek; and (D) Indian Creek. Error circles represent 5% of total amount of displacement. Note that for (A) Furnace Creek and (D) Indian Creek, we resolve the vector sums for two different sets of faults in order to calculate total horizontal extension.

X-X′ and Y-Y′ to capture the horizontal compo- Qfm surface to estimate total strike-slip offset The two profi les (N-N′ and O-O′) we ana- nent of extensional displacement across these of the Qfm surface of ~350 m. This restoration lyzed north of Wildhorse Creek were oriented northeast-oriented scarps. By combining the matches a shutter ridge to the east with the Qfm to most effectively capture all of the identifi - extension we measured across these faults with deposits from the main alluvial fan formed out able fault scarps (Figs. 5B and 8). Profi le N-N′ that on the main, mountain-front fault strand, of Wildhorse Creek (Fig. 11). This admittedly measured four smaller scarps striking approxi- we get a cumulative mean horizontal exten- crude strike-slip restoration indicates that there is mately N5°W, whereas profi le O-O′ measured sion across all faults at Furnace Creek of 13.0 some oblique displacement along the main fault the scarp along the main mountain-front fault ± 0.7 m toward N83°E (Fig. 6A). strand, which is best observed on the restored sur- strand, which strikes N40°W. Although some face north of the creek. We estimate the height of the scarps are antithetic, which would lead Wildhorse Creek of the scarp on the restored fan surface north of to a reduction in the net vertical displacement Wildhorse Creek to be less than 15 m. We note across the fault zone, it does not affect our cal- Unlike the other three study sites, we could that this observation is relatively insensitive to culations in terms of horizontal displacement. not determine a well-defi ned strike-slip offset the details of our approximate strike-slip restora- The cumulative mean horizontal component of along the mountain-front strand (Fig. 8) at Wild- tions, as changing the restoration by 40 m does extensional displacement at Wildhorse Creek is horse Creek. We therefore used the well-con- not result in a signifi cant change in scarp height. 52.3 ± 2.6 m toward N53°E (Fig. 5B). strained 3.1 ± 0.4 mm/yr of Frankel et al. (2007b) This suggests that motion along the main moun- The two profi les (NN-NN′ and OO-OO′) from the nearby Furnace Creek site (5 km south) tain-front fault strand at Wildhorse Creek is pre- that we measured across the normal fault scarps and the 121 ± 14 ka age of the Wildhorse Creek dominantly strike slip. south of Wildhorse Creek provide additional

40 www.gsapubs.org | Volume 2 | Number 1 | LITHOSPHERE

Downloaded from http://pubs.geoscienceworld.org/gsa/lithosphere/article-pdf/2/1/33/3050053/33.pdf by guest on 29 September 2021 Rates of extension along the Fish Lake Valley fault | RESEARCH

118°00'28”W 117°59'32”W

X

37°34'42”N Y

Figure 7. Hill-shaded, LiDAR-derived digital X΄X΄ elevation model (DEM) of the Furnace Creek Qfi site showing the topographic profi les ana- 3 4 lyzed. The number by each fault scarp cor- Y΄Y΄ responds to the fault number in Table 1. The mapped Quaternary deposits were modifi ed 1 from Reheis et al. (1993). See Table DR1 for the detailed measurements of scarp heights. 2 P΄P΄ Solid lines are fault scarps (used in analy- Q΄Q΄ sis), dashed lines are fault scarps visible on R P S Q P1΄P1΄ shaded relief maps but not detectible on T Q1΄Q1΄ topographic profi les (not used in analysis), UV arrows show sense of motion on strike-slip W P1P1 Q1Q1 faults, tick marks are on the hanging wall of normal faults. 5 6 ¯N 0 m 500 118°00'28”W 37°33'38”N

N ′ ″ ′ ″ ″ 118°03 35 W 118°02 20 W 05 ′

O΄O΄ 37°36

5 3 4 Figure 8. Hill-shaded, LiDAR-derived digital O elevation model (DEM) of the Wildhorse N N΄N΄ Creek site showing the topographic profi les analyzed. The white dots indicate the loca- 1 tions of the dated samples. The number by 2 each fault scarp represents the scarps that we included in our analysis, and it corre- Qfm sponds to the fault number in Table 1. The NN΄NN΄ mapped Quaternary deposits were modifi ed OO΄OO΄ from Reheis et al. (1993). See Table DR1 for the detailed measurements of scarp heights. Solid lines are fault scarps (used in analy- Qft ? sis), dashed lines are fault scarps visible on shaded relief maps but not detectible on ? topographic profi les (not used in analysis), 9 8 arrows show sense of motion on strike-slip faults, tick marks are on the hanging wall of normal faults, and dashed lines with ques- 7 tion marks denote inferred fault scarps and NNNN N 6 sense of motion. OOOO ¯ N ″ 00 ′ 0 m 500 118°03′35″W 37°35

LITHOSPHERE | Volume 2 | Number 1 | www.gsapubs.org 41

Downloaded from http://pubs.geoscienceworld.org/gsa/lithosphere/article-pdf/2/1/33/3050053/33.pdf by guest on 29 September 2021 GANEV ET AL.

118°06'27”W 118°05'10”W 37°40'40”N

Figure 9. Hill-shaded, LiDAR-derived digital elevation model (DEM) of the Perry Aiken Creek site showing the topographic profi les J΄J΄ analyzed. The white dots indicate the loca- tions of the dated samples. The number by each fault scarp represents the scarps that K΄K΄ we included in our analysis, and it corre- sponds to the fault number in Table 1. The L΄L΄ mapped Quaternary deposits were modifi ed from Reheis et al. (1993). See Table DR1 for 4 Qfm M΄M΄ the detailed measurements of scarp heights. J K Solid lines are fault scarps (used in analy- L sis), dashed lines are fault scarps visible on M 2 1 3 shaded relief maps but not detectible on topographic profi les (not used in analysis), arrows show sense of motion on strike-slip faults, tick marks are on the hanging wall of normal faults. N Qfi ¯ 0 m 500 118°06'27”W 37°39'00”N

N ′ ″ ′ ″ ″ 118°11 00 W 118°08 00 W 00 ′ 37°48 6 11

8 9 A΄A΄ C A Figure 10. Hill-shaded, LiDAR-derived digital D 10 12 1 E 13 elevation model (DEM) of the Indian Creek 2 B F C΄C΄ B΄B΄ Qfi G D΄D΄ site showing the topographic profi les ana- E΄E΄ lyzed. The number by each fault scarp rep- H resents the scarps that we included in our I 11 F΄F΄ 3 analysis, and it corresponds to the fault 4 G΄G΄ number in Table 1. The mapped Quaternary 7 15 deposits were modifi ed from Reheis et al. H΄H΄ (1995). See Table DR1 for the detailed mea- I΄I΄ surements of scarp heights. Solid lines are 5 13 fault scarps (used in analysis), dashed lines 6 9 are fault scarps visible on shaded relief maps 8 10 12 but not detectible on topographic profi les 14 (not used in analysis), arrows show sense of motion on strike-slip faults, tick marks are on the hanging wall of normal faults. N N ″ 00 ′ 0 m 500 118°11′00″W 37°45

42 www.gsapubs.org | Volume 2 | Number 1 | LITHOSPHERE

Downloaded from http://pubs.geoscienceworld.org/gsa/lithosphere/article-pdf/2/1/33/3050053/33.pdf by guest on 29 September 2021 Rates of extension along the Fish Lake Valley fault | RESEARCH

N ¯N

0 m 500 00 m 500m 500

A B

N N

0 m 500 0 m 500

C D

Figure 11. Hill-shaded geologic (A and C), and topographic (B and D) maps of Wildhorse Creek (top) and Perry Aiken Creek (bottom) alluvial fans derived from light detection and ranging (LiDAR) data. Although no unequivocal right-lateral restoration is apparent, we retrodeformed the Wildhorse Creek fan by ~360 m by combining the strike-slip rate from the nearby Furnace Creek site (Frankel et al., 2007b) and the age of the Wildhorse Creek fan. This yielded an acceptable restoration of the main channel of Wildhorse Creek, as well as for an incised drainage ~400 m to the north. At Perry Aiken Creek, restoration of 250 m of right-lateral strike slip restored the channel margins of the incised Perry Aiken Creek, yielding a strike-slip rate of ~3.0–3.5 mm/ yr, similar to the 3.1 ± 0.4 mm/yr rate determined by Frankel et al. (2007b) at the Furnace Creek site.

LITHOSPHERE | Volume 2 | Number 1 | www.gsapubs.org 43

Downloaded from http://pubs.geoscienceworld.org/gsa/lithosphere/article-pdf/2/1/33/3050053/33.pdf by guest on 29 September 2021 GANEV ET AL.

information about the distribution of slip on suggesting that the defl ection may be purely of juxtaposition of laterally variable topography. this section of the Fish Lake Valley fault. Spe- fl uvial origin, perhaps at a bend in the stream As at Furnace Creek, we used the strike-slip res- cifi cally, although we could not effectively mea- localized at one of the minor normal fault scarps toration of Frankel et al. (2007b) to determine sure the amount of extension across the main, that extend through this area. a more accurate measurement of potential dip- mountain-front strand south of Wildhorse Creek The presence of a signifi cant component of slip motion on the mountain-front strand. Fran- because of the presence of a prominent pressure strike slip at this site raises the possibility that the kel et al. (2007b) showed that restoration of 178 ridge along this part of the fault system, we were very tall scarp just north of Perry Aiken Creek is ± 20 m of right-lateral slip realigns numerous able to measure the amount of extension accom- due partially to strike-slip offset of laterally vary- incised drainages, as well as the overall limits modated to the west of the mountain-front ing topography. However, measurements of four of the Qfi fan (their Fig. 3). Our measurement strand along several north-northwest–trending topographic profi les across all strands of the fault of residual scarp heights on the Frankel et al. normal faults that add slip to the mountain-front system, including across the tallest part of the (2007b) restoration indicates a vertical com- system from the southwest. Specifi cally, profi les scarp north of Perry Aiken Creek, yield strikingly ponent of displacement of 23 ± 1 m. By com- NN-NN′ and OO-OO′ indicate that these faults similar amounts of total vertical separation on the bining the horizontal and vertical components have accommodated a total of 19.9 ± 1.0 m Qfi surface of 85.1 ± 4.3 m (Figs. 5C and 9). The of slip, we fi nd a total oblique displacement of extension oriented toward N77°E since similarity of these measurements along different of the Qfi surface of 179 ± 20 m, with a slip abandonment of the Qfm surface (Fig. 6B). profi les suggests that this is a robust observation direction plunging 7° toward N30°W. Profi les As discussed later, northward addition of this of the total vertical component of oblique slip C-C′ through I-I′ were used to measure the extra component of extension from distributed at the Perry Aiken Creek site. Moreover, along displacement across a set of distributed, north- extension within the White Mountains partially profi les L-L′ and M-M′, the Qfi surface is buried northeast–trending normal faults to the east. accounts for the northward increase in extension beneath the gently north-sloping, northern shoul- In order to account for the difference in strike rates that we observe along the Fish Lake Valley der of a younger fan east of the fault zone, indi- between the main strand and the eastern fault fault system. cating that the vertical measurements on those set, we resolved the extension measured by the profi les are minima. The vertical component of profi les from the two fault sets as a vector sum. Perry Aiken Creek total displacement is in agreement with previous The cumulative mean horizontal extensional measurements by Reheis and Sawyer (1997), displacement at Indian Creek is 43.7 ± 2.1 m The fault zone at Perry Aiken Creek exhibits who suggested 85.5 m of vertical separation on toward S88°E (Fig. 6D). The cumulative verti- a complex pattern of subparallel, anastomosing the Qfi surface. Based on our 85.1 ± 4.3 m mea- cal displacement across all fault strands at this fault strands (Fig. 9). This section of the fault surement of total vertical separation, the cumu- site (75.4 ± 3.8 m) is almost twice as large as includes one of the tallest single fault scarps lative mean horizontal extensional component the previously reported preferred vertical com- in Fish Lake Valley, located just north of the of displacement at Perry Aiken Creek is 49.1 ponent of displacement of 40 m by Reheis and mouth of Perry Aiken Creek. This 73.5-m-tall ± 2.5 m toward N68°E (Fig. 6C). A comparison Sawyer (1997). fault scarp provides clear evidence for signifi - of the extensional and strike-slip components cant normal displacement. However, the fault of slip indicates that the Fish Lake Valley fault SUMMARY OF RATE DATA ALONG THE strand also exhibits evidence for large right-lat- at Perry Aiken Creek is primarily a right-lateral FISH LAKE VALLEY FAULT eral strike-slip displacement. Specifi cally, back- strike-slip fault, with a strike-slip to dip-slip ratio slipping the main, mountain-front fault strand of ~5:1 (i.e., 250 m:49 m). We calculated extension rates at each of our by 250 m restores the right-laterally offset Qfi four study sites by combining vertical compo- terrace riser along the margins of Perry Aiken Indian Creek nents of displacement, fan surface ages, and an Creek (Fig. 11). Combining the 250 m offset assumed 60° fault dip. Specifi cally, our extension with the 71 ± 8 ka age for the nine 10Be samples At the Indian Creek site, we analyzed nine rates were computed by combining probability from the Qfi surface yields a strike-slip rate of topographic profi les across two different sets density functions of the measured displacements ~3.2–4.0 mm/yr, similar to, but slightly faster of fault scarps (Figs. 5D and 10). Profi le A-A′ and TCN ages. These extension rates were fi rst than, the 3.1 ± 0.4 mm/yr rate calculated by extends across the two normal fault scarps, one calculated in their original extension direction, Frankel et al. (2007b) from the Furnace Creek facing east and one facing west, to the west of and then resolved to an extension direction of site to the south. Alternatively, if, as we suspect, the range-front dextral fault strand. We used N65°E to facilitate comparison with geodetic the 75.4 ± 2.2 ka age for the very tight cluster of profi le B-B′ to capture the horizontal exten- measurements of extension rates (Table 2). We six ages from the Perry Aiken fan surface more sional displacement across the main right-lateral used a Gaussian uncertainty model (e.g., Bird, accurately refl ects the age of abandonment of strand of the fault. This profi le provides us with 2007; Kozaci et al., 2009; McGill et al., 2009; the Qfi surface, then the strike-slip rate would a minimum vertical displacement, as right-lat- Zechar and Frankel, 2009), and the uncertainties be slightly slower at ~3.2–3.4 mm/yr. We also eral offset of the alluvial fan at this location will in the extension rates are reported at the 2σ con- note that on the restored image (Figs. 11C– result in apparently smaller scarp heights due to fi dence interval. The resulting extension rates 11D; Fig. DR2 [see footnote 1]), there remains a pronounced right defl ection of the Perry Aiken Creek canyon ~300 m west of (i.e., upstream TABLE 2. EXTENSION RATES FOR THE FOUR STUDY SITES RESOLVED TOWARD N65°E Site α a ECSZ extension Extension rate from) the main strike-slip fault strand. It is pos- extension= (a)xcos(α) (°) (m ± 5%) resolved at N65°E (mm/yr) sible that this represents an additional strike-slip (m ± 5%) strand, which would increase the strike-slip rate Furnace Creek 18 13.0 ± 0.7 12.4 ± 0.6 0.1 ± 0.1 Wildhorse Creek 12 42.1 ± 2.1 41.2 ± 2.1 0.3 ± 0.2 ECSZ extension resolved at N65°E at this site. However, the westernmost main fault Perry Aiken Creek 3 49.1 ± 2.5 49.0 ± 2.5 0.7 +0.3/–0.1 α strand that extends southward to near this right Indian Creek 27 41.5 ± 2.1 37.0 ± 1.9 0.5 +0.2/–0.1 a defl ection does not appear to cross the canyon, Note: ECSZ—Eastern California shear zone.

44 www.gsapubs.org | Volume 2 | Number 1 | LITHOSPHERE

Downloaded from http://pubs.geoscienceworld.org/gsa/lithosphere/article-pdf/2/1/33/3050053/33.pdf by guest on 29 September 2021 Rates of extension along the Fish Lake Valley fault | RESEARCH

from south to north are: Furnace Creek = 0.2 Aiken Creek and Indian Creek sites, respec- cannot rule out a temporal change in extension ± 0.1 mm/yr toward N83°E, Wildhorse Creek tively. These extension rates are similar to those rates, we think that a more likely explanation lies = 0.4 ± 0.2 mm/yr toward N53°E, Perry Aiken estimated by Reheis and Sawyer (1997), who in the overall geometry of the fault system and Creek = 0.7 +0.3/–0.1 mm/yr toward N68°E, reported a preferred late Pleistocene vertical the manner in which slip is transferred north- and Indian Creek = 0.6 +0.2/–0.1 mm/yr toward component of oblique slip at Furnace Creek of ward from the Eastern California shear zone into S88°E. 0.3 mm/yr and 0.8 mm/yr at Indian Creek, on the the Walker Lane belt across the Mina Defl ection. The 10Be dates from all four of our sites should basis of tephrochronology. If we assume a 60° be considered maximum ages for calculating the dip for the fault plane, their preferred extension Geodetic versus Geologic Rates of extension rates because the normal fault scarps rates at Furnace Creek and Indian Creek would Extension must have developed after the deposition and be 0.2 mm/yr and 0.5 mm/yr, respectively. No abandonment of the Qfi and Qfm alluvial-fan preferred vertical component of total slip rate In order to compare short-term geodetic deposits. However, inasmuch as extension has was reported by Reheis and Sawyer (1997) for and longer-term geologic extension rates at the been active at similar rates along the Fish Lake Wildhorse Creek and Perry Aiken Creek. latitude of Fish Lake Valley, we resolved the Valley fault system for at least 760 k.y. (Reheis At our Wildhorse Creek site, we were able geologic extension-rate vectors from our four and Sawyer, 1997), any time lag between fan to quantify the amount of extension that has study sites toward the extension direction of abandonment and initial extensional deforma- been added to the Fish Lake Valley fault system N65°E defi ned by the geodetically constrained tion of the fan surface is likely to have been brief between that site and the Furnace Creek site to models of Bennett et al. (2003) and Wesnousky compared with the late Pleistocene ages of the the south. Specifi cally, we used two profi les (2005a). The resulting N65°E extension rates at measured fans. Thus, this source of potential (NN-NN′ and OO-OO′) just south of Wildhorse our four sites are 0.1 ± 0.1 mm/yr at Furnace error will have a negligible effect on our calcu- Creek that capture normal fault scarps that pro- Creek, 0.3 ± 0.2 mm/yr at Wildhorse Creek, 0.7 lated extension rates. However, although we are vide additional extension to the fault system. +0.3/–0.1 mm/yr at Perry Aiken Creek, and 0.5 confi dent that we have captured all of the main When resolved to N65°E, the extension rate +0.2/–0.1 at Indian Creek (Table 2). fault strands that exhibit a normal component of on the fault set west of the main strand is 0.2 As noted already, at the latitude of Fish Lake slip, some distributed deformation that does not ± 0.1 mm/yr. Profi les N-N′ and O-O′, located Valley, over geologic time scales, approximately manifest itself as recognizable fault scarps could to the north of Wildhorse Creek, demonstrate half of the current 1.0 mm/yr (Wesnousky, be present. Furthermore, in our analysis, we do that the cumulative extension rate oriented 2005b) extension rate defi ned by short-term not include minor fault strands that are visible on toward N65°E at the northern part of the Wild- geodetic data is accommodated by faults to the LiDAR data, but which exhibit scarp heights horse Creek site is 0.3 ± 0.2 mm/yr. Thus, the the west, including the White Mountains fault of <0.5 m, which we consider to be close to the addition of the extension accommodated by the (Kirby et al., 2006), distributed normal fault- resolution limit of this technique. In addition, western fault set increases the overall extension ing in the Volcanic Tablelands (Kirby et al., assuming a steeper fault dip angle, for example, rate at Wildhorse Creek by 0.1 ± 0.1 mm/yr. 2006; Greene et al., 2007; Sheehan, 2007; data 75°, would lower each of the extension rates by This is consistent with our observation that the of Greene and Kirby in Frankel et al., 2008a), ~50%. The oblique, predominantly strike-slip extension rate oriented toward the geodetically and the Sierra Nevada frontal fault system (Le motion along the Fish Lake Valley fault might at defi ned current extension direction of N65°E et al., 2006), including the Round Valley and fi rst suggest that the dip of the faults may indeed increases northward from 0.1 ± 0.1 mm/yr at Hilton Creek faults north of Owens Valley be somewhat steeper than 60°. As noted already, Furnace Creek to 0.3 ± 0.2 mm/yr at Wildhorse (Fig. 12; Berry, 1997). The oblique-normal- however, the fault system is largely strain par- Creek, and it suggests that signifi cant extension dextral White Mountains fault exhibits a late titioned into strike-slip and normal faults, lend- may be accommodated by distributed normal Pleistocene extension rate at the latitude of our ing confi dence to our assumption of a 60° dip faulting within the White Mountains to the Furnace Creek site of ~0.2 mm/yr (Kirby et angle for the extensional faults. If steeper faults southwest of the Wildhorse Creek site. al., 2006), whereas at approximately the same do exist, they would most likely be the moun- It is possible that the northward increase latitude, there is clear evidence for distributed tain-front strands at Wildhorse Creek and Perry in extension rates that we measured along the normal faulting across the Volcanic Tablelands Aiken Creek, where they exhibit oblique motion. Fish Lake Valley fault is a manifestation of a (e.g., Sheehan, 2007; Pinter and Keller, 1995). Thus, the extension rates we calculate for these temporally variable slip rate, rather than along- Sheehan (2007) reported an extension rate sites may be maxima. We emphasize, however, strike variations in slip, inasmuch as we com- across the Volcanic Tablelands of ~0.3 mm/yr. that we do not have any direct measurements of pared offset alluvial fans that span a wide range Further west at the same latitude, Berry (1997) the dip of these faults. of ages between 71 ± 8 ka and 121 ± 14 ka. If reported a 0.5–0.6 mm/yr late Pleistocene verti- this change is due to temporal variation in slip cal component of slip on the Round Valley fault. DISCUSSION rate, then the extension rate must have acceler- This is equivalent to an extension rate on a 60° ated since ca. 94 ka (the age of Furnace Creek dipping fault of ~0.3 mm/yr. The new rate data described here allow us alluvial fan) by a factor of 2–3 times to account Combining all of these extension rates with to place quantitative constraints on the style and for the more rapid extension observed in the our rates from the Fish Lake Valley fault, the location of northward strain transfer through younger Perry Aiken Creek and Indian Creek N65°E component of extension for all of the the Mina Defl ection from faults of the Eastern fans. Moreover, the overall consistency between major fault systems in the northernmost Eastern California shear zone to structures in the Walker our late Pleistocene extension rates (averaged California shear zone, from the Sierra Nevada Lane belt. The extension rates we obtained on over 71 k.y. to 121 k.y., depending on the site) fault to the west and the Fish Lake Valley fault to the Fish Lake Valley fault increase northward and the longer-term rates of Reheis and Sawyer the east, is approximately equal to the geodeti- from 0.2 ± 0.1 mm/yr at Furnace Creek, to 0.4 (1997; averaged over 760 k.y.) suggests that the cally determined extension rate at the latitude ± 0.2 and Wildhorse Creek, and to 0.7 +0.3/– extension rate may be relatively constant over of central/northern Fish Lake Valley. This com- 0.1 mm/yr and 0.6 +0.2/–0.1 mm/yr at the Perry a wide range of time scales. Thus, although we parison suggests that the rate of extension at the

LITHOSPHERE | Volume 2 | Number 1 | www.gsapubs.org 45

Downloaded from http://pubs.geoscienceworld.org/gsa/lithosphere/article-pdf/2/1/33/3050053/33.pdf by guest on 29 September 2021 GANEV ET AL.

119°00′W 118°00′W

Mina Deflection N N ′ ′

Volcanic 38°00 38°00 Hills F V QVFQ F P EPFE 0.5 mm/yr Silver Peak Range HCFH 0.7 mm/yr GPSGPS WMFW White Mtns. C VVoVolcanic TablelandsTTa r F ool /y aba M m l m b cca 0.3 mm/yr ~1~1 mm/yr lle aan F e n lla 0.6 mm/yr aand ici c nnd 0.1 mm/yr

RVFR dds

V s

F 0.3 mm/yr FLVFF LV Sierra Nevada 0.3 mm/yr 0.2 mm/yr F

F S Eureka Valley DSFD

N Inyo Mtns. N N ′ ′ ¯ OVFO 02512.5 V km F 37°00 37°00

119°00′W 118°00′W Figure 12. Rates of horizontal extension on faults in the Eastern California shear zone between 37°N and 38°N latitude. The rates on the Round Valley fault (RVF) and Hilton Creek fault (HCF) are from Berry (1997), the rate on the Volcanic Tablelands is from Sheehan (2007), the rate on the White Mountains fault (WMF) is from Kirby et al. (2006), and the rates on the Fish Lake Valley fault are from this study. In some of these publications, the authors report only the vertical component of slip; therefore, we resolved the extension rates by assuming a 60° fault dip on the faults. DSF—Deep Springs fault, EPF—Emigrant Peak fault, FLVF—Fish Lake Valley fault, HCF—Hill Creek fault, OVF—Owens Valley fault, QVF—Queen Valley fault, RVF—Round Valley fault.

latitude of northern Fish Lake Valley may have be thought of as a major (~80-km-wide), east- site at Indian Creek. The northward decrease in remained relatively constant over the past 104– trending right step in a dominantly right-lateral, right-lateral slip rate is even more pronounced 105 yr, although it is also possible that temporal north-northeast–trending fault system (Stewart, in the northernmost part of Fish Lake Valley, variations in slip rate have occurred over shorter 1988; Dixon et al., 1995; Reheis and Dixon, where the surface expression of the fault zone time scales on the various faults that accommo- 1996; Oldow et al., 1989, 1994, 2001, 2009; Lee ends abruptly ~10 km north of Indian Creek. date extension across the region. Additional slip et al., 2001, 2006, 2009a, 2009b; Petronis et al., The observation that extension rates increase rate calculations on all of these faults at a wider 2002, 2009; Oldow, 2003; Stockli et al., 2003; northward along the Fish Lake Valley fault, span of time scales are necessary to test whether Wesnousky, 2005a, 2005b; Kirby et al., 2006; whereas dextral rates decrease, has important such temporal variations in rate have occurred. Frankel et al., 2007b; Sheehan, 2007). Within implications for the distribution of strain along the Mina Defl ection, deformation is accom- this section of the Pacifi c–North America plate Strain Transfer at the Eastern California modated by east-trending left-lateral faults and boundary, and more generally for mechanisms Shear Zone–Walker Lane Transition clockwise block rotations (Stewart, 1985; Cash- of slip transfer along evolving, structurally com- man and Fontaine, 2000; Faulds et al., 2005; plex fault systems (Fig. 12). The Fish Lake Valley fault terminates just Wesnousky, 2005a). In general, both the northward increase in north of Indian Creek, and slip from this fault, as As documented by Frankel et al. (2007b) extension rate that we document and the north- well as the White Mountains–Queen Valley and and the Perry Aiken strike-slip rate we pres- ward decrease in dextral slip rate documented by Tablelands fault systems to the west, is trans- ent herein, the strike-slip rate decreases north- Frankel et al. (2007b) refl ect transfer of slip off ferred northeastward across the Mina Defl ection ward along the Fish Lake Valley fault, from 3 the predominantly right-lateral Fish Lake Val- onto oblique-normal right-lateral faults of the to 3.5 mm/yr at Furnace Creek and Perry Aiken ley fault and onto north- and northeast-trending Walker Lane belt. Thus, the Mina Defl ection can Creek, to 2.5 mm/yr at our northernmost study normal faults as part of a distributed zone of slip

46 www.gsapubs.org | Volume 2 | Number 1 | LITHOSPHERE

Downloaded from http://pubs.geoscienceworld.org/gsa/lithosphere/article-pdf/2/1/33/3050053/33.pdf by guest on 29 September 2021 Rates of extension along the Fish Lake Valley fault | RESEARCH

transfer located in the ~40-km-long by 30-km- into the Mina Defl ection along a diffuse set of wide, triangular area east of the Fish Lake Val- highly distributed normal faults beneath the Vol- ley fault between the Emigrant Peak fault and canic Hills, then this would imply that clock- the east-trending left-lateral faults of the Mina wise rotations and/or left-lateral slip rates on the Defl ection (Fig. 2). For example, the north- Mina Defl ection faults would increase eastward. northeast–trending normal faults that cut the fan One possibility that we consider is that the to the east of the main range-front fault strands northeast-trending normal faults that character- at the Furnace Creek site appear to “pull” slip ize the northern part of Fish Lake Valley rep- off the Fish Lake Valley fault system and trans- resent an early stage in the evolution of faults

fer it northeastward onto the Emigrant Peak fault similar to the east-trending left-lateral faults of BSF system (Fig. 6). Similarly, the north-northeast– the Mina Defl ection. In such a scenario, these trending normal faults at Indian Creek serve to faults would develop as northeast-trending nor- transfer slip off the Fish Lake Valley fault and mal faults that act to transfer strain across the into the zone of distributed normal faulting in major right step of the Mina Defl ection. As RF this corner of Fish Lake Valley, leaving only noted in Wesnousky’s (2005a) earlier model 2.5 mm/yr of right-lateral strike-slip motion on for the structural evolution of the Mina Defl ec- EF the Fish Lake Valley fault at this site (Fig. 9; tion, in response to ongoing right-lateral shear, Frankel et al., 2007b). This diffuse normal fault- these northeast-trending normal faults would ing along with normal displacements on the gradually rotate clockwise into a more east-west CF prominent Emigrant Peak fault system account orientation, switching to left-lateral strike-slip for the development of the deep basin that motion as a result of this reorientation. However, defi nes the northeast-trending part of northern both the well-established nature of the northeast- Fish Lake Valley (including the dry “Fish Lake” trending basin along the north side of the Silver SP proper; Fig. 12). The most pronounced decrease Peak Range and long-term activity of the north- FLVF in right-lateral strike-slip rate along the Fish to northeast-trending Emigrant Peak fault sys- Lake Valley fault occurs just north of the Indian tem (Petronis et al., 2002, 2009; Petronis, 2005) Creek site, where the geomorphic expression of argue that these are well-established, long-lived the fault system dies out completely within a features that do not appear to be actively rotat- zone of extensive recent lava fl ows in the Vol- ing. In either case, slip transfer across the north- canic Hills (Fig. 12). Between Indian Creek and ern end of Fish Lake Valley into and across the the northwestern limit of faulting observable on Mina Defl ection appears to involve a large com- the LiDAR data within the southern part of the ponent of distributed normal faulting, as well Volcanic Hills, the Fish Lake Valley fault system as left-lateral strike-slip faulting, perhaps quite splays northward into an increasingly distributed distributed at the northwest corner of the valley, set of numerous small-scale normal faults. Thus, and clockwise rotations (Fig. 13). between the north end of the geomorphically well-defi ned Fish Lake Valley fault at Indian CONCLUSIONS Creek, and the left-lateral faults of the Mina Figure 13. Fault model showing the development Defl ection to the north, it appears that distributed New LiDAR topographic data and cosmo- of northeast-striking normal faults transferring normal faulting may accommodate as much as genic 10Be geochronology of offset alluvial-fan strain in a right stepover between two north- 2.5 mm/yr of dextral motion. The coincidence of deposits on the dextral-oblique Fish Lake Valley west-striking zones of right-lateral shear. Clock- this zone of apparent distributed normal faulting fault yield well-determined late Pleistocene– wise rotation of the normal faults is necessary to achieve the highest effi ciency in slip transfer. and the extensive volcanism in the Volcanic Hills Holocene extension rates on this major oblique- Along the Fish Lake Valley fault, the right-lateral suggests that the volcanism may be localized by normal-dextral fault system. The surface expo- slip rate decreases northward near the Mina this slip transfer zone. The young volcanism in sure ages of four sites, Furnace Creek, Wildhorse Defl ection, where there is an increase in the this area may also serve to obscure geomorphic Creek, Perry Aiken Creek, and Indian Creek extension rate of the fault zone. These observa- evidence for recent normal faulting. (from south to north), range from 71 ± 8 ka to tions indicate the presence of a zone of distrib- Ultimately, much of this motion must be 121 ± 14 ka, and the mean horizontal exten- uted strain transfer between the northern Eastern California shear zone and Walker Lane belt in the accommodated along the east-west–trend- sional components of displacement at these sites northwestern part of Fish Lake Valley. BSF—Ben- ing left-lateral faults of the Mina Defl ection range from 12.4 ± 0.6 m to 49.0 ± 2.5 m toward ton Springs fault; CF—Coaldale fault; EF—Excel- (Fig. 1; e.g., the Coaldale, Excelsior Mountains, N65°E. By combining probability density func- sior Mountains fault; FLVF—Fish Lake Valley fault; and Rattlesnake Flat faults) (Stewart, 1985; tions of these displacements and ages, we fi nd RF—Rattlesnake Flat fault; SP—Silver Peak. Wesnousky, 2005a; Lee et al., 2009a). How- that extension rates averaged over late Pleisto- ever, the manner in which this slip transfers cene–Holocene time vary from 0.1 ± 0.1 mm/yr northward onto the left-lateral faults remains at Furnace Creek and 0.3 ± 0.2 mm/yr at Wild- unclear because there are no geomorphically horse Creek in the south, to 0.7 +0.3/–0.1 and regionwide current rate of extension measured well-expressed faults in the 10-km-wide zone 0.5 +0.2/–0.1 mm/yr at Perry Aiken Creek and geodetically. When summed with extension between the northern end of the Fish Lake Val- Indian Creek, respectively, to the north. rates on faults along the western White Moun- ley fault and the Coaldale fault (Fig. 12). If, as These rates suggest that the Fish Lake Valley tains piedmont, the Sierra Nevada frontal fault, we suspect, this slip is transferred northward fault accommodates approximately half of the and distributed deformation across the Volcanic

LITHOSPHERE | Volume 2 | Number 1 | www.gsapubs.org 47

Downloaded from http://pubs.geoscienceworld.org/gsa/lithosphere/article-pdf/2/1/33/3050053/33.pdf by guest on 29 September 2021 GANEV ET AL.

Tablelands, the long-term geologic rates of Bennett, R.A., Wernicke, B.P., Niemi, N.A., Friedrich, A.M., along the northern Death Valley fault zone: Implica- and Davis, J.L., 2003, Contemporary strain rates in the tions for transient strain in the Eastern California shear extension are commensurate with the short-term northern Basin and Range province from GPS data: Tec- zone: Journal of Geophysical Research–Solid Earth, rates determined from GPS data. tonics, v. 22, doi: 10.1029/2001TC001355. v. 112, B06407, doi: 10.1029/2006JB004350. The increase in the east-northeast–west- Berry, M.E., 1997, Geomorphic analysis of late Quaternary Frankel, K.L., Dolan, J.F., Finkel, R.F., Owen, L.A., and Hoeft, faulting on Hilton Creek, Round Valley and Coyote Warp J.S., 2007b, Spatial variations in slip rates along the southwest extensional component of slip toward faults, east-central Sierra Nevada, California, USA: Death Valley–Fish Lake Valley fault system determined the northern end of the Eastern California shear Geomorphology, v. 20, p. 177–195, doi: 10.1016/S0169 from LiDAR topographic data and cosmogenic 10Be zone refl ects a gradual northeastward transfer -555X(97)00033-0. geochronology: Geophysical Research Letters, v. 34, Bierman, P.R., and Caffee, M., 2002, Cosmogenic exposure L18303, doi: 10.1029/2007GL030549. of slip off the predominantly right-lateral Fish and erosion history of ancient Australian bedrock Frankel, K.L., Glazner, A.F., Kirby, E., Monastero, F.C., Strane, Lake Valley fault and across the Mina Defl ec- forms: Geological Society of America Bulletin, v. 114, M.D., Oskin, M.E., Unruh, J.R., Walker, J.D., Ananda- p. 787–803, doi: 10.1130/0016-7606(2002)114<0787: krishnan, S., Bartley, J.M., Coleman, D.S., Dolan, J.F., tion as part of a distributed zone of northeast- CEAEHO>2.0.CO;2. Finkel, R.C., Greene, D., Kylander-Clark, A., Marrero, trending normal faulting. Further north, in the Bird, P., 2007, Uncertainties in long-term geologic offset rates S., Owen, L.A., and Phillips, F., 2008a, Active tectonics Mina Defl ection proper, deformation is accom- of faults; General principles illustrated with data from of the Eastern California shear zone, in Duebendor- California and other western states: Geosphere, v. 3, fer, E.M., and Smith, E.I., eds., Field Guide to Plutons, modated predominantly by east-west left-lateral p. 577–595, doi: 10.1130/GES00127.1. Volcanoes, Reefs, Dinosaurs, and Possible Glaciation faults (e.g., Wesnousky, 2005a). Collectively, Brogan, G.E., Kellog, K.S., Slemmons, D.B., and Terhune, C.L., in Selected Areas of Arizona, California, and Nevada: the distributed normal faulting in northern Fish 1991, Late Quaternary Faulting along the Death Valley– Boulder, Colorado, Geological Society of America Furnace Creek Fault System, California and Nevada: Field Guide 11, p. 43–81, doi: 10.1130/2008.fl d011(03). Lake Valley, together with clockwise rotations U.S. Geological Survey Bulletin 1991, 23 p. Frankel, K.L., Dolan, J.F., Ganev, P.N., and Finkel, R.C., 2008b, and motion on the east-west left-lateral faults of Burchfi el, B.C., Hodges, K.V., and Royden, L.H., 1987, Geology Spatial and temporal constancy of seismic strain of Panamint Valley–Saline Valley pull apart system, Cali- release along the Death Valley–Fish Lake Valley fault the Mina Defl ection, serves to transfer deforma- fornia: Palinspastic evidence for low-angle geometry and Pacifi c–North America plate boundary strain dis- tion through this major right step in the Eastern of a Neogene range-bounding fault: Journal of Geo- tribution: Eos (Transactions, American Geophysical California shear zone–Walker Lane belt. physical Research, v. 92, p. 10,422–10,426, doi: 10.1029/ Union), v. 89, Fall meeting supplement, abs. T41A-1943. JB092iB10p10422. Glazner, A.F., Walker, J.D., Bartley, J.M., and Fletcher, J.M., Carter, W.E., Shrestha, R.L., and Slatton, K.L., 2007, Geo- 2002, Cenozoic evolution of the Mojave block of ACKNOWLEDGMENTS detic laser scanning: Physics Today, v. 60, p. 41–47, doi: southern California, in Glazner, A.F., Walker, J.D., and 10.1063/1.2825070. Bartley, J.M., eds., Geologic Evolution of the Mojave Cashman, P.H., and Fontaine, S.A., 2000, Strain partitioning Desert and Southwestern Basin and Range: Geological We thank Dylan Rood and Alicia Nobles for in the northern Walker Lane, western Nevada and north- Society of America Memoir 195, p. 19–41. assistance with sample preparation and analy- eastern California: Tectonophysics, v. 326, p. 111–130, Gosse, J.C., and Phillips, F.M., 2001, Terrestrial in situ cosmo- doi: 10.1016/S0040-1951(00)00149-9. genic nuclides: Theory and application: Quaternary Sci- sis, and Trevor Thomas for his assistance in the Desilets, D., and Zreda, M., 2003, Spatial and temporal distri- ence Reviews, v. 20, p. 1475–1560, doi: 10.1016/S0277 fi eld. Jeff Lee, Lewis Owen, and Michael Oskin bution of secondary cosmic-ray nucleon intensities and -3791(00)00171-2. provided thoughtful reviews that signifi cantly applications to in-situ cosmogenic dating: Earth and Greene, D., Kirby, E., Dawers, N., Phillips, F., McGee, S., Planetary Science Letters, v. 206, p. 21–42, doi: 10.1016/ and Burbank, D., 2007, Quantifying accommodation of improved the manuscript. The LiDAR data were S0012-821X(02)01088-9. active strain along distributed fault arrays in Owens Val- collected by the National Center for Airborne Desilets, D., Zreda, M., and Prabu, T., 2006, Extended scaling ley, California: Geological Society of America Abstracts Laser Mapping, and we are indebted to Michael factors for in situ cosmogenic nuclides: New measure- with Programs, v. 39, no. 6, abs. 163-8, p. 441. ments at low latitude: Earth and Planetary Science Let- Guest, B., Niemi, N., and Wernicke, B., 2007, Stateline Sartori and Ionut Iordache for help with data ters, v. 246, p. 265–276, doi: 10.1016/j.epsl.2006.03.051. fault system: A new component of the Miocene- processing. This research was made possible Dixon, T.H., Robaudo, S., Lee, J., and Reheis, M.C., 1995, Con- Quaternary Eastern California shear zone: Geological straints on present-day Basin and Range deformation Society of America Bulletin, v. 119, p. 1337–1346, doi: by the support of National Science Foundation from space geodesy: Tectonics, v. 14, p. 755–772, doi: 10.1130/0016-7606(2007)119[1337:SFSANC]2.0.CO;2. (NSF) grants EAR-0537901 and EAR-0538009, 10.1029/95TC00931. Hearn, E.H., and Humphreys, E.D., 1998, Kinematics of the a National Aeronautics and Space Administra- Dixon, T.H., Miller, M., Farina, F., Wang, H., and Johnson, D., southern Walker Lane belt and motion of the Sierra 2000, Present-day motion of the Sierra Nevada block Nevada block, California: Journal of Geophysical tion (NASA) Earth System Science Fellowship, and some tectonic implications for the Basin and Range Research, v. 103, p. 27,033–27,049, doi: 10.1029/98JB01390. and the Georgia Tech Research Foundation. Province, North American Cordillera: Tectonics, v. 19, Hooper, D.M., Bursik, M.I., and Webb, F.H., 2003, Application p. 1–24, doi: 10.1029/1998TC001088. of high-resolution, interferometric DEMs to geomor- Dixon, T.H., Norabuena, E., and Hotaling, L., 2003, Paleoseis- phic studies of fault scarps, Fish Lake Valley, Nevada- REFERENCES CITED mology and global positioning system: Earthquake- California, USA: Remote Sensing of Environment, cycle effects and geodetic versus geologic fault slip v. 84, p. 255–267, doi: 10.1016/S0034-4257(02)00110-4. Andrew, J.E., and Walker, J.D., 2009, Reconstructing late rates in the Eastern California shear zone: Geology, v. 31, Humphreys, E.D., and Weldon, R.J., 1994, Deformation Cenozoic deformation in central Panamint Valley, Cali- p. 55–58, doi: 10.1130/0091-7613(2003)031<0055:PAGPSE across the western United States: A local estimate of fornia: Evolution of slip partitioning in the Walker Lane: >2.0.CO;2. Pacifi c–North America transform deformation: Journal Geosphere, v. 5, p. 172–198, doi: 10.1130/GES00178.1. Dokka, R.K., 1983, Displacements on late Cenozoic strike-slip of Geophysical Research, v. 99, p. 19,975–20,010, doi: Applegate, D., 1995, Transform-normal extension on the faults of the central , California: Geol- 10.1029/94JB00899. Northern Death Valley fault system, California-Nevada: ogy, v. 11, p. 305–308, doi: 10.1130/0091-7613(1983)11< Kirby, E., Burbank, D.W., Reheis, M., and Phillips, F., 2006, Basin Research, v. 7, p. 269–280, doi: 10.1111/j.1365-2117 305:DOLCSF>2.0.CO;2. Temporal variations in slip rate of the White Moun- .1995.tb00110.x. Dokka, R.K., and Travis, C.J., 1990a, Late Cenozoic strike-slip tain fault zone, eastern California: Earth and Plan- Bacon, S.N., and Pezzopane, S.K., 2007, A 25,000 year record faulting in the Mojave Desert, California: Tectonics, v. 9, etary Science Letters, v. 248, p. 153–170, doi: 10.1016/ of earthquakes on the Owens Valley fault near Lone Pine, p. 311–340, doi: 10.1029/TC009i002p00311. j.epsl.2006.05.026. California: Implications for recurrence intervals, slip Dokka, R.K., and Travis, C.J., 1990b, Role of the Eastern Kirby, E., Anandakrishnan, S., Phillips, F., and Marrero, S., rates, and segmentation models: Geological Society of California shear zone in accommodating Pacifi c–North 2008, Millennial scale slip rate along the Owens Valley America Bulletin, v. 119, p. 823–847, doi: 10.1130/B25879.1. American plate motion: Geophysical Research Letters, fault, eastern California: Geophysical Research Letters, Balco, G., Stone, J.O., Lifton, N.A., and Dunai, T.J., 2008, A v. 17, p. 1323–1326, doi: 10.1029/GL017i009p01323. v. 35, L01304, doi: 10.1029/2007GL031970. complete and easily accessible means of calculating Dunai, T., 2001, Infl uence of secular variation of the magnetic Kohl, C.P., and Nishiizumi, K., 1992, Chemical isolation of surface exposure ages or erosion rates from 10Be and fi eld on production rates of in situ produced cosmo- quartz for measurement of in-situ–produced cosmo- 26Al measurements: Quaternary Geochronology, v. 3, genic nuclides: Earth and Planetary Science Letters, genic nuclides: Geochimica et Cosmochimica Acta, p. 174–195, doi: 10.1016/j.quageo.2007.12.001. v. 193, p. 197–212, doi: 10.1016/S0012-821X(01)00503-9. v. 56, p. 3583–3587, doi: 10.1016/0016-7037(92)90401-4. Bartley, J.M., Glazner, A.F., and Schermer, E.R., 1990, North- Faulds, J.E., Henry, C.D., and Heinz, N.H., 2005, Kinemat- Kozaci, O., Dolan, J.F., and Finkel, R.C., 2009, Late Holocene south contraction of the Mojave block and strike- ics of the northern Walker Lane: An incipient transform slip rate for the central North Anatolian fault, from Tahta- slip tectonics in southern California: Science, v. 248, fault along the Pacifi c–North America plate boundary: korpru, Turkey, from cosmogenic 10Be geochronology: p. 1398–1401, doi: 10.1126/science.248.4961.1398. Geology, v. 33, p. 505–508, doi: 10.1130/G21274.1. Implications for the constancy of fault loading and Beanland, S., and Clark, M.M., 1994, The Owens Valley Fault Frankel, K.L., Brantley, K.S., Dolan, J.F., Finkel, R.C., Klinger, strain release rates: Journal of Geophysical Research, Zone, Eastern California, and Surface Rupture Associ- R.E., Knott, J.R., Machette, M.N., Owen, L.A., Phillips, v. 114, B01405, doi: 10.1029/2008JB005760. ated with the 1872 Earthquake: U.S. Geological Survey F.M., Slate, J.L., and Wernicke, B.P., 2007a, Cosmogenic Kylander-Clark, A.R.C., Coleman, D.S., Glazner, A.F., and Bulletin 1982, 29 p. Be-10 and Cl-36 geochronology of offset alluvial fans Bartley, J.M., 2005, Evidence for 65 km of dextral slip

48 www.gsapubs.org | Volume 2 | Number 1 | LITHOSPHERE

Downloaded from http://pubs.geoscienceworld.org/gsa/lithosphere/article-pdf/2/1/33/3050053/33.pdf by guest on 29 September 2021 Rates of extension along the Fish Lake Valley fault | RESEARCH

across Owens Valley, California, since 83 Ma: Geologi- slip belt, Nevada: Geology, v. 22, p. 637–640, doi: Quadrangles, Esmeralda County, Nevada, and Mono cal Society of America Bulletin, v. 117, p. 962–968, doi: 10.1130/0091-7613(1994)022<0637:LCETIT>2.3.CO;2. County, California: U.S. Geological Survey Map I-2464, 10.1130/B25624.1. Oldow, J.S., Aiken, C.L.V., Hare, J.L., Ferguson, J.F., and Har- scale 1:24,000. Lal, D., 1991, Cosmic ray labeling of erosion surfaces: In situ dyman, R.F., 2001, Active displacement transfer and dif- Sartori, M., 2005, ALSM Acquisition in Death Valley National nuclide production rates and erosion models: Earth and ferential block motion within the central Walker Lane, Park; Report on Data Processing: Gainesville, Florida, Planetary Science Letters, v. 104, p. 424–439, doi: 10.1016/ western : Geology, v. 29, p. 19–22, doi: National Center for Airborne Laser Mapping, Univer- 0012-821X(91)90220-C. 10.1130/0091-7613(2001)029<0019:ADTADB>2.0.CO;2. sity of Florida, 13 p. Le, K., Lee, J., Owen, L.A., and Finkel, R., 2006, Late Qua- Oldow, J.S., Elias, E.A., Ferranti, L., McClelland, W.C., and Savage, J.C., Gan, W., and Svarc, J.L., 2001, Strain accu- ternary slip rates along the Sierra Nevada frontal fault McIntosh, W.C., 2009, Late Miocene to Pliocene synex- mulation and rotation in the Eastern California shear zone, California: Slip partitioning across the western tensional deposition in fault-bounded basins within the zone: Journal of Geophysical Research, v. 106, no. B10, margin of the Eastern California shear zone–Basin and upper plate of the western Silver Peak–Lone Mountain p. 21,995–22,007, doi: 10.1029/2000JB000127. Range province: Geological Society of America Bulletin, extensional complex, west-central Nevada, in Oldow, Sawyer, T.L., 1990, Quaternary Geology and Neotectonic v. 119, p. 240–256, doi: 10.1130/B25960.1. J.S., and Cashman, P.H., eds., Late Cenozoic Structure Activity along the Fish Lake Valley Fault Zone, Nevada Lee, J., Spencer, J., and Owen, L., 2001, Holocene slip and Evolution of the Great Basin–Sierra Nevada Transi- and California [M.S. thesis]: Reno, University of rates along the Owens Valley fault, California: Impli- tion: Geological Society of America Special Paper 447, Nevada, 379 p. cations for the recent evolution of the Eastern Cali- p. 275–312, doi: 10.1130/2009.2447(14). Sawyer, T.L., 1991, Late Pleistocene and Holocene paleoseis- fornia shear zone: Geology, v. 29, p. 819–822, doi: Oskin, M.E., Le, K., and Strane, M.D., 2007a, Quantifying micity and slip rates of the northern Fish Lake Valley 10.1130/0091-7613(2001)029<0819:HSRATO>2.0.CO;2. fault-zone activity in arid environments with high-reso- fault zone, Nevada and California, in Reheis, M.C., ed., Lee, J., Stockli, D., Schroeder, J., Tincher, C., Bradley, D., lution topography: Geophysical Research Letters, v. 34, Guidebook for Field Trip to Fish Lake Valley, California- Owen, L., Gosse, J., Finkel, R., and Garwood, J., 2006, L23S05, doi: 10.1029/2007GL031295. Nevada: Golden, Colorado, Pacifi c Cell, Friends of the Fault slip transfer in the Eastern California shear Oskin, M., Perg, L., Blumentritt, D., Mukhopadhyay, S., and Pleistocene, p. 114–138. zone–Walker Lane belt: Boulder, Colorado, Geological Iriondo, A., 2007b, Slip rate of the Calico fault: Implica- Schermer, E.R., Luyendyk, B.P., and Cisowski, S., 1996, Society of America Penrose Field Guide, p. 1–26, doi: tions for geologic versus geodetic rate discrepancy in Late Cenozoic structure and tectonics of the north- 10.1130/2006.FSTITE.PFG. the Eastern California shear zone: Journal of Geophysi- ern Mojave Desert: Tectonics, v. 15, p. 905–932, doi: Lee, J., Garwood, J., Stockli, D.E., and Gosse, J., 2009a, Qua- cal Research, v. 112, B03402, doi: 10.1029/2006jb004451. 10.1029/96TC00131. ternary faulting in Queen Valley, California-Nevada: Oskin, M., Perg, L., Shelef, E., Strane, M., Gurney, E., Singer, Sheehan, T.P., 2007, Evolution of Neogene Fault Populations Implications for kinematics of fault-slip transfer in the B., and Zhang, X., 2008, Elevated shear zone loading in Northern Owens Valley, California and Implications Eastern California shear zone–Walker Lane belt: Geo- rate during an earthquake cluster in eastern California: for the Eastern California Shear Zone [Ph.D. thesis]: logical Society of America Bulletin, v. 121, p. 599–614, Geology, v. 36, p. 507–510, doi: 10.1130/G24814A.1. New Orleans, Tulane University, 203 p. doi: 10.1130/B26352.1. Oswald, J.A., and Wesnousky, S.A., 2002, Neotectonics Stewart, J.H., 1985, East-trending dextral faults in the west- Lee, J., Stockli, D.F., Owen, L.A., Finkel, R.C., and Kislitsyn, and Quaternary geology of the Hunter Mountain fault ern Great Basin; an explanation for anomalous trends R., 2009b, Exhumation of the Inyo Mountains, Califor- zone and Saline Valley region, southeastern California: of pre-Cenozoic strata and Cenozoic faults: Tectonics, nia: Implications for the timing of extension along the Geomorphology, v. 42, p. 255–278, doi: 10.1016/S0169- v. 4, p. 547–564, doi: 10.1029/TC004i006p00547. western boundary of the Basin and Range Province and 555X(01)00089-7. Stewart, J.H., 1988, Tectonics of the Walker Lane belt, west- distribution of dextral fault slip rates across the East- Petronis, M.S., 2005, Cenozoic Evolution of the Central ern Great Basin: Mesozoic and Cenozoic deformation ern California shear zone: Tectonics, v. 28, TC1001, doi: Walker Lane Belt, West-Central Nevada [Ph.D. thesis]: in a zone of shear, in Ernst, W.G., ed., Metamorphism 10.1029/2008TC002295. Albuquerque, University of New Mexico, 286 p. and Crustal Evolution of the Western United States, Lifton, N., Bieber, J., Clem, J., Duldig, M., Evenson, P., Hum- Petronis, M.S., Geissman, J.W., Oldow, J.S., and McIntosh, Rubey Volume 7: Englewood Cliffs, New Jersey, Pren- ble, J., and Pyle, R., 2005, Addressing solar modulation W.C., 2002, Paleomagnetic and 40Ar/39Ar geochrono- tice Hall, p. 683–713. and long-term uncertainties in scaling secondary cos- logic data bearing on the structural evolution of the Stockli, D.F., Dumitru, T.A., McWilliams, M.O., and Far- mic rays for in situ cosmogenic nuclide applications: Silver Peak extensional complex, west-central Nevada: ley, K.A., 2003, Cenozoic tectonic evolution of the Earth and Planetary Science Letters, v. 239, p. 140–161, Geological Society of America Bulletin, v. 114, p. 1108– White Mountains, California and Nevada: Geological doi: 10.1016/j.epsl.2005.07.001. 1130, doi: 10.1130/0016-7606(2002)114<1108:PAAAGD> Society of America Bulletin, v. 115, p. 788–816, doi: McClusky, S.C., Bjornstad, S.C., Hager, B.H., King, R.W., 2.0.CO;2. 10.1130/0016-7606(2003)115<0788:cteotw>2.0.co;2. Meade, B.J., Miller, M.M., Monastero, F.C., and Souter, Petronis, M.S., Geissman, J.W., Oldow, J.S., and McIn- Stone, J.O., 2000, Air pressure and cosmogenic isotope B.J., 2001, Present-day kinematics of the Eastern Cali- tosh, W.C., 2009, Late Miocene to Pliocene vertical- production: Journal of Geophysical Research, v. 105, fornia shear zone from a geodetically constrained block axis rotation attending development of the Silver p. 23,753–23,759, doi: 10.1029/2000JB900181. model: Geophysical Research Letters, v. 28, p. 3369– Peak–Lone Mountain displacement transfer zone, Thatcher, W., Foulger, G.R., Julian, B.R., Svarc, J., Quilty, 3372, doi: 10.1029/2001GL013091. west-central Nevada, in Oldow, J.S., and Cashman, E., and Bawden, G.W., 1999, Present-day deformation McGill, S.F., Wells, S.G., Fortner, S.K., Kuzma, H.A., and P.H., eds., Late Cenozoic Structure and Evolution of across the Basin and Range Province, western United McGill, J.D., 2009, Slip rate of the western Garlock fault, the Great Basin–Sierra Nevada Transition: Geological States: Science, v. 283, p. 1714–1718, doi: 10.1126/ at Clark Wash, near Lone Tree Canyon, Mojave Des- Society of America Special Paper 447, p. 215–253, doi: science.283.5408.1714. ert, California: Geological Society of America Bulletin, 10.1130/2009.2447(12). Walker, J.D., and Glazner, A.F., 1999, Tectonic development v. 121, p. 536–554, doi: 10.1130/B26123.1. Pinter, N., and Keller, E.A., 1995, Geomorphological analysis of the southern California deserts, in Moores, E.M., Miller, M.M., Johnson, D.J., Dixon, T.H., and Dokka, R.K., of neotectonic deformation, northern Owens Valley, Sloan, D., and Stout, D.L., eds., Classic Cordilleran 2001, Refi ned kinematics of the Eastern California California: International Journal of Earth Sciences, Concepts: A View from California: Geological Society shear zone from GPS observations, 1993–1994: Jour- v. 84, p. 200–212. of America Special Paper 338, p. 375–380. nal of Geophysical Research, v. 106, p. 2245–2263, doi: Reheis, M.C., 1992, Geologic Map of Late Cenozoic Deposits Walker, J.D., Kirby, E., and Andrew, J.E., 2005, Strain trans- 10.1029/2000JB900328. and Faults in Parts of the Soldier Pass and Magruder fer and partitioning between the Panamint Valley, Nielsen, R.L., 1965, Right-lateral strike-slip faulting in the Mountain 15′ Quadrangles, Inyo and Mono Counties, Searles Valley, and Ash Hill fault zones, California: Walker Lane, west-central Nevada: Geological Society California and Esmeralda County, Nevada: U.S. Geo- Geosphere, v. 1, p. 111–118, doi: 10.1130/GES00014.1. of America Bulletin, v. 76, p. 1301–1308, doi: 10.1130/ logical Survey Map I-2268, scale 1:24,000. Wesnousky, S.G., 2005a, Active faulting in the Walker Lane: 0016-7606(1965)76[1301:RSFITW]2.0.CO;2. Reheis, M.C., and Dixon, T.H., 1996, Kinematics of the East- Tectonics, v. 24, doi: 10.1029/2004TC001645. Nishiizumi, K., Imamura, M., Caffee, M.W., Southon, J.R., Fin- ern California shear zone: Evidence for slip transfer Wesnousky, S.G., 2005b, The San Andreas and Walker Lane kel, R.C., and McAninch, J., 2007, Absolute calibration from Owens and Saline Valley fault zones to Fish fault systems, western North America: Transpression, of 10Be AMS standards: Nuclear Instruments & Methods Lake Valley fault zone: Geology, v. 24, p. 339–342, doi: transtension, cumulative slip and the structural evolu- in Physics Research–Section B, Beam Interactions with 10.1130/0091-7613(1996)024<0339:KOTECS>2.3.CO;2. tion of a major transform plate boundary: Journal of Materials and Atoms, v. 258, p. 403–413, doi: 10.1016/j. Reheis, M.C., and Sawyer, T.L., 1997, Late Cenozoic history Structural Geology, v. 27, p. 1505–1512, doi: 10.1016/ nimb.2007.01.297. and slip rates of the Fish Lake Valley, Emigrant Peak, j.jsg.2005.01.015. Oldow, J.S., 2003, Active transtensional boundary zone and Deep Springs fault zones, Nevada and California: Zechar, J.D., and Frankel, K.L., 2009, Incorporating and between the western Great Basin and Sierra Nevada Geological Society of America Bulletin, v. 109, p. 280– reporting uncertainties in fault slip rates: Journal of block, western U.S. Cordillera: Geology, v. 31, p. 1033– 299, doi: 10.1130/0016-7606(1997)109<0280:LCHASR> Geophysical Research–Solid Earth, doi: 10.1029/2009 1036, doi: 10.1130/G19838.1. 2.3.CO;2. JB006325 (in press). Oldow, J.S., Bally, A.W., Ave Lallemant, H.G., and Leeman, Reheis, M.C., Sawyer, T.L., Slate, J.L., and Gillispie, A.R., W.P., 1989, Phanerozoic evolution of the North American 1993, Geologic Map of Late Cenozoic Deposits and Cordillera (United States and Canada), in Bally, A.W., Faults in the Southern Part of the Davis Mountain 15′ and Palmer, A.R., eds., The Geology of North America: Quadrangle, Esmeralda County, Nevada: U.S. Geologi- MANUSCRIPT RECEIVED 28 MARCH 2009 An Overview: Boulder, Colorado, Geological Society of cal Survey Map I-2342, scale 1:24,000. REVISED MANUSCRIPT RECEIVED 25 AUGUST 2009 America, Geology of North America, v. A, p. 139–232. Reheis, M.C., Slate, J.L., and Sawyer, T.L., 1995, Geologic MANUSCRIPT ACCEPTED 28 AUGUST 2009 Oldow, J.S., Kohler, G., and Donelick, R.A., 1994, Late Ceno- Map of Late Cenozoic Deposits and Faults in Parts zoic extensional transfer in the Walker Lane strike- of the Mt. Barcroft, Piper Peak, and Soldier Pass 15′ Printed in the USA

LITHOSPHERE | Volume 2 | Number 1 | www.gsapubs.org 49

Downloaded from http://pubs.geoscienceworld.org/gsa/lithosphere/article-pdf/2/1/33/3050053/33.pdf by guest on 29 September 2021