<<

SHORT RESEARCH

Low-temperature thermochronology of the Black and Panamint mountains, , : Implications for geodynamic controls on Cenozoic intraplate strain

Tandis S. Bidgoli1, Erika Amir2, J. Douglas Walker1, Daniel F. Stockli3, Joseph E. Andrew1, and S. John Caskey2 1DEPARTMENT OF GEOLOGY, UNIVERSITY OF KANSAS, LAWRENCE, KANSAS 66045, USA 2DEPARTMENT OF EARTH AND CLIMATE SCIENCES, SAN FRANCISCO STATE UNIVERSITY, SAN FRANCISCO, CALIFORNIA 94132, USA 3JACKSON SCHOOL OF GEOSCIENCES, UNIVERSITY OF TEXAS, AUSTIN, TEXAS 78712, USA

ABSTRACT

We use apatite and zircon (U-Th)/He thermochronometry to evaluate space-time patterns and tectonic drivers of Miocene to Pliocene defor- mation within the Death Valley area, eastern California. Zircon He ages from the footwall of the Amargosa–Black Mountains detachment in the Black Mountains record continuous cooling and exhumation from 9 to 3 Ma. Thermal modeling of data from the central Black Moun- tains suggests that this cooling took place during two intervals: a period of rapid footwall exhumation from 10 to 6 Ma, followed by slower (<5 mm/yr) exhumation since 6 Ma. Cumulative exhumation is estimated to be 10–16 km. Paleodepth reconstruction of cooling ages from the footwall of the Panamint-Emigrant detachment, in the central Panamint Range, also show two periods of cooling. Zircons record late Miocene cooling, whereas apatite He ages show punctuated exhumation at ca. 4 Ma. The results suggest the Panamint Range experienced a minimum of 7.2 km of exhumation since ca. 12 Ma. The new data, when evaluated within the context of published fault timing data, suggest that the transition from Basin and Range extension to dextral transtension is spatially and temporally distinct, beginning at ca. 11–8 Ma in ranges to the east and north of the Black Mountains and migrating westward into eastern Death Valley at 6 Ma. Initiation of dextral transten- sion was coincident with a major change in plate-boundary relative motion vectors. Data from Panamint Range and several ranges to the west of Death Valley indicate transtension initiated over a large area at ca. 3–4 Ma, coeval with proposed lithospheric delamination in the central and southern Sierra Nevada Range. Our results suggest that the transition from extension to dextral transtension may reflect an evolution in tectonic drivers, from plate-boundary kinematics to intraplate lithospheric delamination.

LITHOSPHERE; v. 7; no. 4; p. 473–480; GSA Data Repository Item 2015194 | Published online 21 May 2015 doi:10.1130/L406.1

INTRODUCTION ca. 10–8 Ma, the plate boundary changed from suggesting that the transition from intraplate NW oblique to dominantly coast-parallel mo- extension to intraplate dextral transtension may Plate motions may be accommodated across tion (Atwater, 1970; Atwater and Stock, 1998). not have been simply related to plate-boundary broad zones of diffuse and variable-style defor- It is proposed that this motion change initiated kinematics, and that other geodynamic factors mation within continental lithosphere (e.g., Eng- the change from Basin and Range–style de- may have played a role in the temporal and spa- land, 1987). One of the best-known examples of formation to the transcurrent Eastern Califor- tial pattern of deformation. One possible factor a diffuse zone is the Pacific–North American nia shear zone and attendant structures (e.g., is lithospheric delamination, which is proposed plate boundary, where ~25% of the current plate Dokka and Travis, 1990). (Plate-boundary and to have occurred in the central and southern motion is accommodated far inboard of the intraplate deformation patterns are shown in parts of the Sierra Nevada at ca. 3.5 Ma (Man- San Andreas transform fault (Minster and Jor- the animations of Atwater and Stock [1998], ley et al., 2000; Jones et al., 2004; Zandt et al., dan, 1987; Dokka and Travis, 1990; Bennett et available at http://emvc.geol.ucsb.edu/down- 2004; Gilbert et al., 2007). al., 2003[). Geodetic data from across western load/nepac.php, and McQuarrie and Wernicke In this paper, we present new zircon and North America show that most of this intraplate [2005]). However, timing data from individual apatite (U-Th)/He thermochronologic data from strain is focused in the Eastern California shear structures within the Eastern California shear the Black and Panamint mountains border- zone (also the Walker Lane belt), a system of ac- zone range from middle Miocene to late Pleis- ing Death Valley in eastern California (Fig. 1). tive, mostly northwest-striking, strike-slip faults tocene (see Burchfiel et al., 1987; Hodges et al., Death Valley lies within a region that has expe- that account for ~9–12 mm/yr of displacement 1989; Holm and Dokka, 1991, 1993; Holm et rienced both Basin and Range extension (e.g., of the Sierran microplate relative to the eastern al., 1992; Hoisch and Simpson, 1993; Snow and Stewart, 1983; Snow and Wernicke, 2000) and (Fig. 1; Dokka and Travis, 1990; Lux, 1999; Snyder and Hodges, 2000; Niemi et more recent dextral transtension related to the Dixon et al., 1995; Bennett et al., 2003). al., 2001; Monastero et al., 2002; Stockli et al., Eastern California shear zone and Walker Lane Global circuit reconstructions of the Pacific 2003; Guest et al., 2007; Lee et al., 2009; Mahan belt (e.g., Dixon et al., 1995; Reheis and Saw- plate relative to North America suggest that at et al., 2009; Beyene, 2011; Walker et al., 2014), yer, 1997; Bennett et al., 2003; Frankel et al.,

LITHOSPHERE© 2015 Geological | Volume Society 7 of| AmericaNumber 4| |For www.gsapubs.org permission to copy, contact [email protected] 473

Downloaded from http://pubs.geoscienceworld.org/gsa/lithosphere/article-pdf/7/4/473/3040498/473.pdf by guest on 30 September 2021 BIDGOLI ET AL.

118° W 116° W 5-3 Nevada AHe ABMD = Amargosa-Black Mtn. Detach. BCD = Boundary Canyon Detachment 14-12 Nevada Test Site AFT EPFS = East Panamint fault system Calif Region ZHe FCF = Furnace Creek fault GF = Garlock fault ornia ZFT HMF = Hunter Mountain fault NDVF = Northern Death Valley fault OVF = fault 37° N In 37° N yo M PED = Panamint -Emigrant Detach. NDVF 11-10 SDVF = Southern Death Valley fault ountains SLF = State Line fault 15-11 SNFF = Sierra Nevada Frontal fault CW 9-7 3 BCD 7-6 CA NV UT OVF 11

HM 5.1 AZ HMF +/-0.3 4.8 +/-0.6 4.2 Death 4 +/-2.8 4.2 +/-0.2 4.8 Spring Mountains +/-0.5

PED V FCF Argus alley

EPFS 5.5 5-3 ABMD +/-0.3 SL Sierra Nevada 34-6 F 11-7 4.3 4.6 Range +/-0.5 C C’A A’ +/-1.0 BM 36° N 9-4 36° N Panamint 10-7 3.3 NR SNFF 19-3 +/-0.3 Range SDVF 45-7 B B’ 4 9-5 14-12 5.0 +/-0.6 Kilometers SR GF 7-4 CM 8-5 0212.5 550 118° W 116° W Figure 1. Color-shaded relief map of the greater Death Valley area with major detachments (ticks), and normal (ball and stick) and strike-slip (arrows with relative motion) faults shown in black. Faults are after Wernicke et al. (1988). Published fission-track and (U-Th)/He ages (Ma) are shown in ovals. Mean zircon (black text) and apatite (U-Th)/He ages (red text) and age ranges are shown for samples from this study. Inset map shows area of figure (dashed box) with respect to western North America. AHe—apatite helium; AFT—apatite fission track; BM—Black Mountains; CM—Clark Mountains; CW—Cottonwood Mountains; HM—Hunter Mountain; NR—Nopah Range; ZHe—zircon helium; ZFT—zircon fission track. Data sources: McCullough Range—Mahan et al. (2009); Avawatz Mountains—Reinert et al. (2003); Black Mountains—Holm et al. (1992), Holm and Dokka (1993); Bare Mountain and Bullfrog Hills— Ferrill et al. (2012); —Holm and Dokka (1991), Hoisch and Simpson (1993); Panamint Range—B. Wernicke (2011, personal commun.); —Walker et al. (2014); Inyo Mountains—Lee et al. (2009); White Mountains—Stockli et al. (2003).

2007), and it is therefore an ideal place in which 1.7 Ga gneiss and late Precambrian metasedi- sition beginning at 6 Ma, suggesting a major to study the inboard part of the evolving Pa- mentary rocks intruded by the 11.6 Ma Wil- reorganization of the basin at that time (Wright cific–North American plate boundary. The new low Springs pluton (Asmerom et al., 1994) and et al., 1999, and references therein for details on timing data are integrated with published ther- 10.4 Ma Smith Mountain Granite (Miller et al., the Furnace Creek basin). mochronologic and sedimentologic data of a re- 2004). Published 40Ar/39Ar and zircon and apa- The Panamint Range, on the western flank of gional scale to show that the initiation of dextral tite fission-track ages from the Black Mountains Death Valley, is the relatively intact hanging wall shear varied from east to west across the region, show that exhumation and cooling along the to the Amargosa–Black Mountains detachment. possibly reflecting an interplay between plate- Amargosa–Black Mountains detachment began The range hosts two major Cenozoic structures: boundary effects and intraplate drivers such as at ca. 12 Ma in the southern Black Mountains; the Eastern Panamint fault system and the Pana- lithospheric delamination. however, major unroofing of the northern and mint-Emigrant detachment (Fig. 1). These struc- central parts of the range took place between tures expose a core of 1.7 Ga gneiss, middle and GEOLOGIC BACKGROUND ca. 8 and 6 Ma (Holm et al., 1992; Holm and late Precambrian metasedimentary rocks, and Dokka, 1993). The sedimentological record of Cretaceous and Cenozoic plutons. The Eastern The Black Mountains are located on the the lower part of the adjacent Furnace Creek ba- Panamint fault system is a low-angle normal eastern margin of Death Valley (Fig. 1). The sin matches the timing of exhumation recorded fault system, exposed in the eastern part of the range is host to the Amargosa–Black Mountains in the thermochronology data; however, a sig- range. The timing of motion on this fault sys- detachment, a normal fault system that exposes nificant portion of the basin shows rapid depo- tem is poorly constrained, but it likely postdates

474 www.gsapubs.org | Volume 7 | Number 4 | LITHOSPHERE

Downloaded from http://pubs.geoscienceworld.org/gsa/lithosphere/article-pdf/7/4/473/3040498/473.pdf by guest on 30 September 2021 Low-temperature thermochronology of Death Valley, California | SHORT RESEARCH

intrusion of the 10.6 Ma Little Chief Stock in He AGE RESULTS 2009). The software searches for viable thermal the center of the range, and most of the Trail histories for multisample transects by modeling Canyon volcanic sequence (ca. 10.5–9.4 Ma), Black Mountains He production, stopping, and diffusion for ran- exposed on east side of the range (McKenna and domly generated temperature-time paths using Hodges, 1990; Hodges et al., 1990). The Trail Cross-section A–A′ (Fig. 2) is oriented east- algorithms and goodness-of-fit tests described Canyon volcanic sequence is tilted 20°–40°, west and shows the geometry of the Amargosa– in Ketcham (2005). Details of the required indicating significant eastward tilting of the Black Mountains detachment and the projected model inputs and constraints are given in Ap- Eastern Panamint fault system and range since positions of our thermochronology samples pendix DR1 (see footnote 1). ca. 9 Ma (McKenna and Hodges, 1990). from Sheep Canyon in the central Black Moun- The modeling results for the central Black The Panamint-Emigrant detachment, ex- tains. Zircon (U-Th)/He mean ages range from Mountains are shown in Figure 4. The models posed on the west side of the Panamint Range, 5.4 ± 0.6 Ma to 9.3 ± 3.9Ma and systematically show continuous exhumation since ca. 10 Ma includes the Emigrant and Towne Pass faults, increase with increasing elevation (Figs. 2 and and a distinct change in the slope of model fits and the Panamint detachment. Timing con- 3; Table DR1 [see footnote 1]). Samples also at ca. 6 Ma. This change is also reflected in ex- straints for the Panamint-Emigrant detach- systematically increase in age eastward across humation rates, with rates as high as 10–5 mm/ ment come from the sedimentary succession the range (Fig. 2). A similar pattern of cooling yr for segments of the thermal history prior to of the Nova Formation preserved in its hang- ages is observed from the footwall of the Ama- 6 Ma. By contrast, exhumation rates are con- ing wall, and they suggest fault initiation and rgosa–Black Mountains detachment in Confi- sistently lower (<5 mm/yr) after 6 Ma. The cu- basin development after ca. 12 Ma (Hodges et dence Wash in the southern Black Mountains mulative exhumation from our best-fit models al., 1989; Snyder and Hodges, 2000). However, (cross-section B–B′ in Fig. 2). Zircon (U-Th)/ (30 °C/km geothermal gradient) ranges from 9 the bulk of the Nova Formation was deposited He mean ages there range from 5.4 ± 0.2 Ma to 16 km, which is similar to earlier geobaro- rapidly between ca. 4.4 and 3.0 Ma, indicating to 8.5 ± 1.4 Ma. Mean ages from this transect metric estimates for the Willow Springs pluton that much of the activity related to this fault also systematically increase from west to east; (Holm et al., 1992). The modeling results also system is Pliocene in age (Snyder and Hodges, however, the age-elevation relationship is much show that most of the model fits correspond to 2000). These sedimentologic observations are more subdued when compared with data from reasonable geothermal gradients between 20 consistent with a palinspastic restoration of ba- the central Black Mountains (Figs. 2 and 3). and 50 °C/km, although model fits with geother- salt flows across northern that In addition to these transects, we analyzed mal gradients as high as ~140 °C/km were not also suggests significant post–4 Ma fault mo- several samples from along the range front of excluded by the analysis. tion (Burchfiel et al., 1987) the Black Mountains (Fig. 1). Zircon (U-Th)/ In contrast, thermal models for the Pana- He mean ages span from 3.3 ± 0.3 Ma to 5.5 mint Range are best satisfied by model fits at (U-Th)/He THERMOCHRONOMETRY ± 0.3 Ma, with most samples clustering at very high geothermal gradients (100+ °C/km; around 4 Ma (Fig. 1). These samples also show Fig. 4). Both intrusion of the Little Chief Stock We analyzed 48 samples from the footwalls a modest correlation between mean age and el- at 10.6 Ma (Hodges et al., 1990) prior to fault of the Amargosa–Black Mountains detachment evations (Fig. 3). initiation and isotherm advection during rapid and Panamint-Emigrant detachment using zir- exhumation could elevate geothermal gradients con and apatite (U-Th)/He thermochronometry Panamint Mountains in the area, but these transient effects would not (Fig. 1). Two additional samples were also ana- be representative of time-averaged geothermal lyzed: one from the western base of the Nopah Cross-section C–C′ (Fig. 2) shows the ge- gradients. Another possible explanation for the Range and another from the northern base of ometry of the Miocene to Pliocene faults in the high geothermal gradients in our models is tilt- the Avawatz Mountains (Fig. 1). Samples rep- Panamint Range and projected positions of our ing or horizontal-axis rotation of our samples, resent a range of units, including Proterozoic thermochronology samples from the footwall of which would narrow the vertical distance gneiss and metasedimentary rocks, Cretaceous the Panamint-Emigrant detachment in Surprise between samples and, consequently, paleo- gneiss and leucogranite, and Cenozoic diorite Canyon and Pleasant Canyon in the central part isotherms. Such a scenario seems likely given and granite. Laboratory and analytical work was of the range. Zircon (U-Th)/He mean ages range that 20°–40° of tilting is documented for the performed at Isotope Geochemistry Laborato- from 6.4 ± 1.0 Ma to 45.3 ± 4.9 Ma and sys- range (McKenna and Hodges, 1990). To further ries at the University of Kansas, the (U-Th)/He tematically increase with increasing elevation evaluate the data and account for this rotation, Geo- and Thermochronometry Laboratory at the (Table DR1 [see footnote 1[). Samples also in- we restored samples to their middle Miocene University of Texas at Austin, and the (U-Th)/ crease in age eastward across the range. In con- paleodepths using the unconformity below the He Thermochronology Laboratory at the Uni- trast, mean apatite (U-Th)/He ages from these east-tilted Trail Canyon volcanic sequence as a versity of California, Santa Cruz. Data tables, transects appear to cluster at ca, 3–4 Ma. The pre-extensional approximately paleohorizontal descriptions of analytical procedures, and error exception is our structurally highest sample, datum. The restoration assumes that the range is reporting are given in Tables DR1 and DR2 and which had a mean age of 19.4 ± 0.3 Ma. a simple, tilted fault block; therefore, the hinge Appendix DR1 in the GSA Data Repository.1 position for tilting is not a factor (e.g., Arm- THERMAL MODELING AND PALEODEPTH strong et al., 2003). Our reconstruction adds 1 GSA Data Repository Item 2015194, Table DR1 RECONSTRUCTION ~500 m of overburden above the unconformity (summary of [U-Th]/He data), Table DR2 ([U-Th]/He based on the thickness of the sequence in the data), and Appendix DR1 (analytical procedures, er- In order to evaluate possible cooling histo- north-central part of the range (McKenna and ror reporting, and thermal modeling inputs and con- ries and apparent exhumation rates, we modeled Hodges, 1990). A simple check on the accuracy straints), is available at www.geosociety.org/pubs​ footwall transects from the central Black Moun- of the reconstruction is that it places exposures /ft2015.htm, or on request from editing@geosociety​ .org, Documents Secretary, GSA, P.O. Box 9140, Boul- tains and Panamint Range using the numerical of the Little Chief Stock at 1.8–2.2 km below der, CO 80301-9140, USA. modeling software HeMP (Hager and Stockli, the middle Miocene surface, consistent with

LITHOSPHERE | Volume 7 | Number 4 | www.gsapubs.org 475

Downloaded from http://pubs.geoscienceworld.org/gsa/lithosphere/article-pdf/7/4/473/3040498/473.pdf by guest on 30 September 2021 BIDGOLI ET AL.

14 A A’ a) 12 10 Northern Black Mountains Ts Tertiary alluvial and uvial units 8 Tv Tertiary volcanic units 6 Tv Tg Tertiary granite Th)/He age (M 4 2 Pz Td Tertiary diorite (U- 2 Ts m) Zu Pz Paleozoic undierentiated 0 ABMD 1 . (k 12BM08 Ts Ts Tg Tv YZu Precambrian undierentiated Td Xgn Td Elev 0 Xgn Precambrian gneiss SC7 SC5a,b 0 SC3 SC2 1600 SC1 GV2 1200 Zircon (U-Th)/He age and errors A SC6 SC4 SC8 A’

200 Sheep Canyon1000 1400 Apatite (U-Th)/He age and errors SC1011 1400 N 400 SC9 600 800 1200 GV1 1400 01510 5 12 10 B B’ 8 Southern Black Mountains 6 Ts 2 m)

h)/He age (Ma) 4

1 . (k (U -T 2 ABMD Ts Ts Ts Xgn Xgn Xgn YZu Elev 0 0 600 400 800 12CW01 200 Con dence Wash 12CW03 1000 12CW02 B 12CW05 12CW04 B’ 12CW07 12CW06

0 N 800

12CW08 0

12CW09 800

1000 12 0 1000 600 0 5 10 15 20 25 60 C C’

a) 50 40 Central Panamint Range 30 20 Th)/He age (M

(U- 10 0 YZu 2 YZu Zu Zu YZu EPFS km) Tg (

PED . Xgn Xgn Tv 1 Ts Xgn YZu Ts ev YZu YZu El YZu Xgn Tg 0 SC01 SC08 SC05 SC04 Surprise Canyon SC09 SC10 PC04 PC02 SC12 PC10 PC09 PC06 PC05 PC03 PC01 Projected positions of 2200 SC13 2400 Pleasant Canyon samples N 1800 20 0 0 2800 0102030 Distance (km)

Figure 2. Geologic cross sections across the central Black Mountains (A–A′), southern Black Mountains (B–B′), and central Panamint Range (C–C′). Thermochronology samples, shown as dots, are projected into the plane of each cross section. In this projected view, samples may lie above or below the topography of the section line. Mean zircon (black diamonds) and apatite (white boxes) (U-Th)/He ages and errors (1s standard devia- tion) are shown above each sample. Strip maps below each cross section show sample locations with respect to the line of the section. Cross section locations are shown in Figure 1. Geology is from Drewes (1963), Hunt and Mabey (1966), and the compilation of Workman et al. (2002). ABMD—Amargosa–Black Mountain detachment; PED—Panamint-Emigrant detachment; EPFS—East Panamint fault system.

476 www.gsapubs.org | Volume 7 | Number 4 | LITHOSPHERE

Downloaded from http://pubs.geoscienceworld.org/gsa/lithosphere/article-pdf/7/4/473/3040498/473.pdf by guest on 30 September 2021 Low-temperature thermochronology of Death Valley, California | SHORT RESEARCH

1700 1800 Sheep Canyon Zr A B 1500 Con dence Wash Zr 1600 1300 BM rangefront Ap 1400 ) ) BM rangefront Ap 1100 1200 (m (m 900 1000

700 800 Elevation Elevation 500 600 Surprise Canyon Zr 300 400 Surprise Canyon Ap Pleasant Canyon Zr 100 200 Pleasant Canyon Ap -100 0 0246810 12 14 0102030405060 (U-Th)/He age (Ma) (U-Th)/He age (Ma)

Figure 3. (A) Plot of (U-Th)/He age versus elevation for the Black Mountains showing a strong age-elevation relationship for the Sheep Canyon data, while data from Confidence Wash show a more modest correlation. (B) (U-Th)/He age versus elevation plot for the Panamint Range. The zircon (Zr) data show a strong age versus elevation relationship, while apatite (Ap) data are mostly eleva- tion invariant.

0 0 Figure 4. (A) Modeled temperature- 50 AC50 time (T-t) paths (acceptable fits) for ZHe PRZ the Sheep Canyon transect, central 100 100 Black Mountains. Most of the mod- 150 150 eled paths have an inflection point ZHe PRZ or change in slope and cooling rate 200 200 at ca. 6 Ma. Black boxes show the 250 250 model constraints based on biotite and hornblende 40Ar/39Ar ages from T (deg C) 300 T (deg C) 300 Holm et al. (1992) and an assumed 350 350 mean annual surface temperature. 400 400 (B) Exhumation rates from modeled (t-T) histories that satisfy a 30 °C/ 450 450 km geothermal gradient (1000 fits). 500 500 Results suggest that rates have 10 9 8 7 6 5 4 3 2 1 0 100 80 60 40 20 0 decreased since 6 Ma. Inset chart Time (Ma) Time (Ma) shows the number of acceptable 35 30 fits for geothermal gradients rang- B 1200 D 1200 ing from 10 °C to 200 °C. (C) Mod- 30 800 25 800 eled (T-t) paths (acceptable fits) for the Panamint Range. Black boxes

Model ts 400 25 Model ts 400 show the model constraints based 20 on muscovite and hornblende K-Ar 0 0 te (mm/yr) 50 100 150 200 te (mm/yr) 50 100 150 200 ages from Lanphere (1962) and an 20 Geothermal gradient Ra Geothermal gradient Ra 15 assumed mean annual surface tem- (deg C/km) (deg C/km) 15 perature. (D) Exhumation rates from modeled (t-T ) histories colored 10 10 according to geothermal gradient. Inset chart shows the number of Exhumation 5 Exhumation 5 acceptable fits for geothermal gra- dients ranging from 10 °C to 200 °C. 0 0 ZHe—zircon helium; PRZ—partial 10 9 8 7 6 5 4 3 2 1 0 100 80 60 40 20 0 retention zone. Time (Ma) Time (Ma)

LITHOSPHERE | Volume 7 | Number 4 | www.gsapubs.org 477

Downloaded from http://pubs.geoscienceworld.org/gsa/lithosphere/article-pdf/7/4/473/3040498/473.pdf by guest on 30 September 2021 BIDGOLI ET AL.

crystallization pressures estimated for the stock of the 190 °C isotherm in our paleodepth recon- tion, and volume of suites of igneous rocks in the (McDowell, 1974). struction. Therefore, the total vertical exhuma- region, with copious intermediate to felsic mag- The resulting paleodepth reconstruction tion since the late Miocene is probably closer matism during the middle to late Miocene giv- (Fig. 5) shows two distinct periods of cooling. to 8.4 km. ing way to small-volume mafic eruptions during The structurally lowest zircon ages overlap the Pliocene to Quaternary (e.g., Asmerom et al., within errors and show a period of late Miocene DISCUSSION AND CONCLUSIONS 1994). (Space-time patterns are well expressed cooling. Samples above 7.5 km then steadily in the palinspastic reconstruction of the North increase in age with decreasing paleodepth, de- The new data from the Black Mountains, American Volcanic Database by McQuarrie and fining a zircon He partial retention zone. The combined with earlier thermochronologic con- Oskin [2010].) shallowest zircon samples cluster at ca. 45 Ma straints from the range (Holm et al., 1992; Holm The new thermochronology data, when and are not reset, defining the top of the partial and Dokka, 1993), show a continuous exhuma- placed within the context of published fault retention zone. In contrast, apatite ages from tion history that begins at ca. 12 Ma. A marked timing data, demonstrate that the very eastern these same samples are invariant at ca. 4 Ma, change in the thermal history and inferred ex- part of the Eastern California shear zone expe- consistent with rapid exhumation at that time. humation rate occurred at ca. 6 Ma, which is rienced a transition to dextral transtension that Assuming a mean surface temperature of coincident with a major change in the character was coincident with and likely triggered by the 10 °C, our paleodepth reconstruction indicates of sediment and depositional rates within the plate-boundary kinematic change at 10–8 Ma, that the Panamint Range has experienced a min- Furnace Creek basin (Wright et al., 1999, and while areas to the west lagged this change by imum of 180 °C of cooling since the late Mio- references therein). The Panamint Range simi- several million years. A westward progression cene. The estimated pre-extensional geothermal larly shows two periods of exhumation: one in in fault initiation is predicted by and may sup- gradient from our reconstruction is ~25 °C/km, the late Miocene and the other in the Pliocene. port a rolling hinge model for the structural de- which translates to ~7.2 km of vertical exhuma- The Pliocene cooling coincides with a period of velopment of the greater Death Valley area (e.g., tion. However, our structurally deepest sample rapid sedimentation within the adjacent Nova Stewart, 1983; Wernicke et al., 1988; Snow and resides 1.2 km below the approximate position basin (Snyder and Hodges, 2000), suggesting a Wernicke, 2000; Niemi et al., 2001; Ferrill et major structural reorganization at that time. al., 2012), but given the two-phase pattern of Timing constraints for the major faults within deformation documented by this and other stud- 0 Tv the Eastern California shear zone are variable. ies (e.g., Stockli et al., 2003; Lee et al., 2009; Faults located east and north of the Black Moun- Walker et al., 2014), such a model may be too 1 tains (e.g., Stateline fault, Boundary Canyon de- simplistic. As an alternative, the westward pro- tachment, Furnace Creek fault, Northern Death gression of fault initiation could reflect changes Valley–Fish Lake Valley fault) initiated between in crustal strength that closely follow the locus 2 11–8 Ma, around the time of the plate-boundary of earlier extension and crustal thinning (Frid- change (Fig. 1; Holm and Dokka, 1991, 1993; rich and Thompson, 2011). In this case, the

m) 3 Hoisch and Simpson, 1993; Reheis and Sawyer, earlier pattern of strain, which may have been 1997; Snow and Lux, 1999; Niemi et al., 2001; controlled by the migration of a rolling hinge Mahan et al., 2009; Beyene, 2011; Ferrill et al., or some other mechanism, and its impact on 4 2012). It is unclear if these structures initiated at crustal rheology are the dominant controls on the onset of regional extension or dextral trans- the initiation of dextral transtension. 5 tension, a problem that is compounded by the Although both of these models adequately fact that some of the published data are maxi- describe the westward progression in fault ini- 130°C mum ages rather than well-defined fault initiation tiation, neither model explains the near-simulta- iocene paleodepth (k 6 ages (e.g., Reheis and Sawyer, 1997; Snow and neous inception of the Eastern California shear M ZHe PRZ Lux, 1999; Niemi et al., 2001). However, tim- zone west of Death Valley at 3–4 Ma, suggesting 7 ing data across much of the Eastern California that other dynamic drivers may need to be con- 190°C shear zone suggest that extension and transten- sidered. One factor that may be imparting a con- sion occurred in discrete phases (see summary in trol on the temporal and spatial pattern of intra- 8 Norton, 2011). The Northern Death Valley–Fish plate strain is lithospheric delamination, which Lake Valley fault system, for example, initiated at is interpreted to have occurred in the central and southern Sierra Nevada at ca. 3.5 Ma (Manley et 9 ca. 10 Ma, but reorganized at ca. 6 Ma, similar to the Black Mountains (Reheis and Sawyer, 1997; al., 2000; Jones et al., 2004; Zandt et al., 2004; 0102030405060 Stockli et al., 2003). Like the Panamint Range, Gilbert et al., 2007). Mantle tomographic studies (U-Th)/He age (Ma) the Slate Range, and Inyo and White mountains have identified a large, dense lithospheric body Figure 5. Mean (U-Th)/He ages vs. middle Mio- show two phases of rapid exhumation: one in (the Isabella anomaly) spreading out from the cene paleodepth for samples from the central the middle Miocene and another at ca. 3–4 Ma base of the crust into the asthenosphere beneath Panamint Range. Zircon ages are shown as (Fig. 1; Walker et al., 2014; Lee et al., 2009; the western Sierras and Great Valley, while the black diamonds; apatite ages are shown as Stockli et al., 2003). The earlier of these events topographically high eastern Sierras are made white boxes. Error bars are 1s standard devia- is linked to Basin and Range extension, while the up of relatively thin crust (~35 km) and low- tions. Shaded area corresponds to the approxi- latter is linked to the initiation of dextral trans- velocity mantle (Gilbert et al., 2012, and refer- mate position of the zircon He partial retention zone (ZHe PRZ). Two periods of rapid exhuma- tension (Stockli et al., 2003; Lee et al., 2009; ences therein). Aside from the geophysical evi- tion are evident in the data: one in the late Mio- Walker et al., 2014). A two-phase history is also dence for delamination, garnet-rich xenoliths, cene and another in the Pliocene. supported by differences in the timing, composi- present in 12–8 Ma basalts, are conspicuously

478 www.gsapubs.org | Volume 7 | Number 4 | LITHOSPHERE

Downloaded from http://pubs.geoscienceworld.org/gsa/lithosphere/article-pdf/7/4/473/3040498/473.pdf by guest on 30 September 2021 Low-temperature thermochronology of Death Valley, California | SHORT RESEARCH

missing from eruptions younger than ca. 4 Ma, An update: International Geological Review, v. 40, Hodges, K.V., McKenna, L.W., Stock, J.M., Knapp, J.H., and suggesting that a portion of the lithosphere was p. 375–402. Walker, J.D., 1989, Structural relationships between Bennett, R.A., Wernicke, B.P., Niemi, N.A., Friedrich, A.M., Neogene extensional basins, Panamint Valley area, removed during the intervening interval (Du- and Davis, J.L., 2003, Contemporary strain rates in the southeast California: New perspectives of Basin and cea and Saleeby, 1996, 1998). The timing of northern Basin and Range Province from GPS data: Tec- Range topography: Tectonics, v. 8, p. 453–467, doi:​ tonics, v. 22, p. 1008, doi:10.1029/2001TC001355. 10.1029​/TC008i003p00453. delamination is inferred from a brief pulse of Beyene, M., 2011, Mesozoic Burial, and Mesozoic and Ceno- Hodges, K.V., McKenna, L.W., and Harding, M.B., 1990, Struc- alkalic volcanism at ca. 3.5 Ma, interpreted to zoic Exhumation of the Funeral Mountains Core Com- tural unroofing of the central Panamint Mountains, have been caused by asthenospheric upwelling plex, Death Valley, Southeastern California [Ph.D. the- Death Valley region, southeastern California, in Wer- sis]: Las Vegas, Nevada, University of Nevada, 362 p. nicke, B.P., ed., Basin and Range Extensional Tectonics in the wake of lithospheric removal (Manley et Burchfiel, B.C., Hodges, K.V., and Royden, L.H., 1987, Geology near the Latitude of Las Vegas, Nevada: Geological al., 2000; Farmer et al., 2002; Jones et al., 2004). of Panamint Valley– pull-apart system, Cali- Society of America Memoir 176, p. 377–390, doi:​10.1130​ The similar timing of lithospheric delamina- fornia: Palinspastic evidence for low-angle geometry /MEM176-p377. of a Neogene range-bounding fault: Journal of Geo- Hoisch, T.D., and Simpson, C., 1993, Rise and tilt of metamor- tion and the development of the Eastern Cali- physical Research, v. 92, p. 10,422–10,426, doi:10.1029/ phic rocks in the lower plate of a detachment fault in the fornia shear zone west of Death Valley suggests JB092iB10p10422. Funeral Mountains, Death Valley, California: Journal of Dixon, T.H., Robaudo, S., Lee, J., and Reheis, M.C., 1995, Con- Geophysical Research, v. 98, p. 6805–6827, doi:​10.1029​ these events are linked. Such a connection has straints on present-day Basin and Range deformation /92JB02411. been postulated and modeled by a number of from space geodesy: Tectonics, v. 14, p. 755–772, doi:​ Holm, D.K., and Dokka, R.K., 1991, Major late Miocene cool- authors (e.g., Jones et al., 2004; Oldow et al., 10.1029/95TC00931. ing of the middle crust associated with extensional oro- Dokka, R.K., and Travis, C.J., 1990, Role of the Eastern Califor- genesis in the Funeral Mountains, California: Geophysi- 2008; Le Pourhiet et al., 2006; Saleeby et al., nia shear zone in accommodating Pacific–North Ameri- cal Research Letters, v. 18, p. 1775–1778, doi:10.1029​ ​ 2012), but major gaps in fault timing constraints can plate motion: Geophysical Research Letters, v. 17, /91GL02079. have rendered the relationship between the two p. 1323–1326, doi:10.1029/GL017i009p01323. Holm, D.K., and Dokka, R.K., 1993, Interpretation and tec- Drewes, H., 1963, Geology of the Funeral Peak Quadrangle, tonic implications of cooling histories: An example ambiguous. Data from this study, combined California, on the Eastern Flank of Death Valley: U.S. from the Black Mountains, Death Valley extended ter- with published fault timing constraints (Holm Geological Survey Professional Paper 413, 78 p. rane, California: Earth and Planetary Science Letters, Ducea, M.N., and Saleeby, J.B., 1996, Buoyancy sources for v. 116, p. 63–80, doi:10.1016/0012-821X(93)90045-B. and Dokka, 1991, 1993; Hoisch and Simpson, a large, unrooted mountain range, the Sierra Nevada, Holm, D.K., Snow, J.K., and Lux, D.R., 1992, Thermal and 1993; Reheis and Sawyer, 1997; Snow and Lux, California: Evidence from xenoliths thermobarometry: barometric constraints on the intrusive and unroofing 1999; Niemi et al., 2001; Stockli et al., 2003; Journal of Geophysical Research, v. 101, p. 8229–8244, history of the Black Mountains—Implications for tim- doi:10.1029/95JB03452. ing, initial dip, and kinematics of detachment faulting Lee et al., 2009; Mahan et al., 2009; Beyene, Ducea, M.N., and Saleeby, J.B., 1998, A case for delamina- in the Death Valley region, California: Tectonics, v. 11, 2011; Ferrill et al., 2012; Walker et al., 2014), tion of the deep batholithic crust beneath the Sierra p. 507–522, doi:10.1029/92TC00211. indicate that the initiation of dextral transtension Nevada, California: International Geology Review, v. 40, Hunt, C.B., and Mabey, D.R., 1966, Stratigraphy and Struc- p. 78–93, doi:10.1080/00206819809465199. ture, Death Valley, California: U.S. Geological Survey occurred as a westward-migrating wave, with England, P.C., 1987, Diffuse continental deformation: Length Professional Paper 494A, 162 p. transtensional structures initiating in the east scales, rates and metamorphic evolution: Philosophi- Jones, C.H., Farmer, G.L., and Unruh, J., 2004, Tectonics of cal Transactions of the Royal Society of London, v. 321, Pliocene removal of lithosphere of the Sierra Nevada, at ca. 11–8 Ma, in Death Valley and Fish Lake p. 3–22, doi:10.1098/rsta.1987.0001. California: Geological Society of America Bulletin, Valley at 6 Ma, and in areas west of Death Val- Farmer, G.L., Glazner, A.F., and Manley, C.R., 2002, Did litho- v. 116, p. 1408–1422, doi:10.1130/B25397.1. ley at 4–3 Ma. This pattern may be explained by spheric delamination trigger late Cenozoic potassic Ketcham, R.A., 2005, Forward and inverse modeling of low volcanism in the southern Sierra Nevada, California?: temperature thermochronometry data: Reviews in an evolution in geodynamic drivers, from plate- Geological Society of America Bulletin, v. 114, p. 754– Mineralogy and Geochemistry, v. 58, p. 275–314, doi:​ boundary kinematics to intraplate factors like 768, doi:10.1130/0016-7606(2002)114<0754:DLDTLC>2 10.2138/rmg​ .2005.58.11.​ lithospheric delamination. .0.CO;2. Lanphere, M.A., 1962, I. Geology of the Wildrose area, Pana- Ferrill, D.A., Morris, A.P., Stamatakos, J.A., Waiting, D.J., mint Range, California; II. Geochronologic studies in Donelick, R.A., and Blythe, A.E., 2012, Constraints on Death Valley- region, California [Ph.D. ACKNOWLEDGMENTS exhumation and extensional faulting in southwestern thesis]: Pasadena, California Institute of Technology. This research was supported by a University of Kanas Pat- Nevada and eastern California, USA, from zircon and Lee, J., Stockli, D.F., Owen, L.A., Finkel, R.C., and Kislitsyn, terson Fellowship and student research grants from the apatite thermochronology: Lithosphere, v. 4, p. 63–76, R., 2009, Exhumation of the Inyo Mountains, Califor- Geological Society of America and the American Associa- doi:10.1130​ /L171.1.​ nia: Implications for the timing of extension along the tion of Petroleum Geologists. We would like to thank Michael Frankel, K.L., Brantley, K.S., Dolan, J.F., Finkel, R.C., Klinger, western boundary of the Basin and Range Province and Giallorenzo for providing apatite separates from the Nopah R.E., Knott, J.R., Machette, M.N., Owen, L.A., Phillips, distribution of dextral fault slip rates across the Eastern Range, and Roman Kislitsyn and Jeremey Hourigan for as- F.M., Slate, J.L., and Wernicke, B.P., 2007, Cosmogenic California shear zone: Tectonics, v. 28, p. TC1001, doi:​ 10 36 sistance with sample preparation and analysis. We are grate- Be and Cl geochronology of offset alluvial fans along 10.1029/2008TC002295.​ ful to Chad La Fever, Edward Morehouse, Gabriel Creason, the northern Death Valley fault zone: Implications for Le Pourhiet, L., Gurnis, M., and Saleeby, J., 2006, Mantle and Rachelle Warren for assistance with sample collection, transient strain in the Eastern California shear zone: instability beneath the Sierra Nevada mountains in and to Jessica Savage, Matthew Myers, and Ty Tenpenny for Journal of Geophysical Research, v. 112, p. 2156–2202, California and Death Valley extension: Earth and Plan- help with sample processing. This manuscript was greatly doi:10.1029​ /2006JB004350.​ etary Science Letters, v. 251, p. 104–119, doi:10.1016/j​ improved by thoughtful reviews by Barbara Carrapa, Nathan Fridrich, C.J., and Thompson, R.A., 2011, Cenozoic Tectonic .epsl.2006.08.028.​ Niemi, and Arlo Weil. Reorganizations of the Death Valley Region, Southeast Mahan, K.H., Guest, B., Wernicke, B., and Niemi, N.A., 2009, California and Southwest Nevada: U.S. Geological Sur- Low-temperature thermochronologic constraints on vey Professional Paper 1783, 36 p. the kinematic history and spatial extent of the Eastern REFERENCES CITED Gilbert, H., Jones, C., Owens, T.J., and Zandt, G., 2007, Imag- California shear zone: Geosphere, v. 5, p. 483–495, doi:​ Armstrong, P.A., Ehlers, T.A., Chapman, D.S., Farley, K.A., and ing Sierra Nevada lithospheric sinking: Eos (Transac- 10.1130/GES00226.1.​ Kamp, P.J., 2003, Exhumation of the central Wasatch tions, American Geophysical Union), v. 88, p. 225, Manley, C.R., Glazner, A.F., and Farmer, G.L., 2000, Timing Mountains, Utah: 1. Patterns and timing of exhuma- doi:10.1029/2007EO210001. of volcanism in the Sierra Nevada of California: Evi- tion deduced from low-temperature thermochronology Gilbert, H., Yang, Y., Forsyth, D.W., Jones, C.H., Owens, T.J., dence for the Pliocene delamination of the batholithic data: Journal of Geophysical Research, v. 108, p. 2172, Zandt, G., and Stachnik, J.C., 2012, Imaging lithospheric root?: Geology, v. 28, p. 811–814, doi:10.1130/0091- doi:10.1029​ /2001JB001708.​ foundering in the structure of the Sierra Nevada: Geo- 7613(2000)28<811:TOVITS>2.0.CO;2. Asmerom, Y., Jacobsen, S.B., and Wernicke, B.P., 1994, Varia- sphere, v. 8, p. 1310–1330, doi:10.1130/GES00790.1. McDowell, D., 1974, Emplacement of the Little Chief Stock, tions in magma source regions during large-scale con- Guest, B., Niemi, N., and Wernicke, B., 2007, Stateline fault Panamint Range, California: Geological Society of tinental extension, Death Valley region, western United system: A new component of the Miocene–Quaternary America Bulletin, v. 85, p. 1535–1546, doi:10.1130/0016​ States: Earth and Planetary Science Letters, v. 125, Eastern California shear zone: Geological Society of -7606​(1974)​85​<​1535​:EOTLCS>2.0.CO;2. p. 235–254, doi:10.1016/0012-821X(94)90218-6. America Bulletin, v. 119, p. 1337–1347, doi:10.1130/0016- McKenna, L.W., and Hodges, K.V., 1990, Constraints on the Atwater, T., 1970, Implications of plate tectonics for the Ceno- 7606(2007)119[1337:SFSANC]2.0.CO;2. kinematics and timing of late Miocene–Recent exten- zoic tectonic evolution of western North America: Geo- Hager, C., and Stockli, D.F., 2009, A new MATLAB©-based sion between the Panamint and Black Mountains, logical Society of America Bulletin, v. 81, p. 3513–3536, helium modeling package (“HeMP”) for thermal history southeastern California, in Wernicke, B.P., ed., Basin doi:10.1130/0016-7606(1970)81[3513:IOPTFT]2.0.CO;2. recovery from single and multi-thermochronometer and Range Extensional Tectonics near the Latitude Atwater, T., and Stock, J., 1998, Pacific–North America plate (U-Th)/He data and data arrays: Geological Society of of Las Vegas, Nevada: Geological Society of America tectonics of the Neogene southwestern United States: America Abstracts with Programs, v. 41, no. 7, p. 487. Memoir 176, p. 363–376, doi:10.1130/MEM176-p363.

LITHOSPHERE | Volume 7 | Number 4 | www.gsapubs.org 479

Downloaded from http://pubs.geoscienceworld.org/gsa/lithosphere/article-pdf/7/4/473/3040498/473.pdf by guest on 30 September 2021 BIDGOLI ET AL.

McQuarrie, N., and Oskin, M., 2010, Palinspastic restoration of Reheis, M.C., and Sawyer, T.L., 1997, Late Cenozoic history and of America Bulletin, v. 115, p. 788–816, doi:10.1130/0016​ NAVDat and implications for the origin of magmatism slip rates of the Fish Lake Valley, Emigrant Peak, and -7606(2003)115<0788:CTEOTW>2.0.CO;2.​ in southwestern North America: Journal of Geophysical Deep Springs fault zones, Nevada and California: Geo- Walker, J.D., Bidgoli, T.S., Didericksen, B.D., Stockli, D.F., and Research, v. 115, p. B10401, doi:10.1029​ /2009JB006435.​ logical Society of America Bulletin, v. 109, p. 280–299, Andrew, J.E., 2014, Middle Miocene to Recent exhuma- McQuarrie, N., and Wernicke, B.P., 2005, An animated tectonic doi:10.1130/0016-7606(1997)109<0280:LCHASR>2.3.CO;2. tion of the Slate Range, eastern California, and impli- reconstruction of southwestern North America since 36 Reinert, E., Cowan, D.S., and Reiners, P.W., 2003, Transpres- cation for the timing of extension and the transition to Ma: Geosphere, v. 1, p. 147–172, doi:10.1130/GES00016.1.​ sion within the Eastern California shear zone and low- transtension: Geosphere, v. 10, p. 276–291, doi:10.1130​ Miller, M., Friedman, R.M., and Dee, S., 2004, Pre–10 Ma temperature thermochronometry of the Avawatz Moun- /GES00947.1. TIMS U-Pb zircon age for the Smith Mountain Granite, tains, , U.S.A.: Geological Society Wernicke, B.P., Axen, G.J., and Snow, J.K., 1988, Basin and Death Valley, California: Implications for timing of major of America Abstracts with Programs, v. 35, no. 4, p. 64. Range extensional tectonics at the latitude of Las extension: Geological Society of America Abstracts Saleeby, J., Le Pourhiet, L., Saleeby, Z., and Gurnis, M., 2012, Vegas, Nevada: Geological Society of America Bulle- with Programs, v. 36, no. 5, p. 501. Epeirogenic transients related to mantle lithosphere tin, v. 100, p. 1738–1757, doi:10.1130/0016-7606(1988)​ Minster, J.B., and Jordan, T.H., 1987, Vector constraints on west- removal in the southern Sierra Nevada region, Califor- 100​<​1738​:BARETA​>2.3.CO;2. ern US deformation from space geodesy, neotectonics, nia: Part I. Implications of thermomechanical modeling: Workman, J.B., Menges, C.M., Page, W.R., Taylor, E.M., Ekren, and plate motions: Journal of Geophysical Research, Geosphere, v. 8, p. 1286–1309, doi:10.1130/GES00746.1. E.B., Rowley, P.D., Dixon, G.L., Thompson, R.A., and v. 92, p. 4798–4804, doi:10.1029/JB092iB06p04798.​ Snow, J.K., and Lux, D.R., 1999, Tectono-sequence stratig- Wright, L.A., 2002, Geologic Map of the Death Valley Monastero, F.C., Walker, J.D., Katzenstein, A.M., and Sabin, raphy of Cenozoic rocks in the Cottonwood Mountains Ground-water Model Area, Nevada and California: U.S. A.E., 2002, Neogene evolution of the Indian Wells Valley, and northern Death Valley area, California and Nevada, Geological Survey Miscellaneous Field Studies Map east-central California, in Glazner, A.F., Walker, J.D., and in Wright, L.A., and Troxel, B.W., eds., Cenozoic Basins of MF-2381, http://pubs.usgs.gov/mf/2002/mf-2381/ (last Bartley, J.M., eds., Geologic Evolution of the Mojave the Death Valley Region: Geological Society of America accessed December 2013). Desert and Southwestern Basin and Range: Geological Special Paper 333, p. 17–64, doi:10.1130/0​-8137​-2333-7.17. Wright, L.A., Greene, R.C., Cemen, I., Johnson, F.C., Prave, A.R., Society of America Memoir 195, p. 199–228, doi:​10.1130​ Snow, J.K., and Wernicke, B.W., 2000, Cenozoic tectonism and Drake, R.E., 1999, Tectonostratigraphic development /0​-8137​-1195-9.199. in the central Basin and Range: Magnitude, rate, and of the Miocene–Pliocene Furnace Creek Basin and related Niemi, N.A., Wernicke, B.P., Brady, R.J., Saleeby, J.B., and distribution of upper crustal strain: American Journal features, Death Valley region, California, in Wright, L.A., Dunne, G.C., 2001, Distribution and provenance of the of Science, v. 300, p. 659–719, doi:10.2475/ajs.300.9.659. and Troxel, B.W., eds., Cenozoic Basins of the Death Val- middle Miocene Eagle Mountain Formation, and impli- Snyder, N.P., and Hodges, K.V., 2000, Depositional and tec- ley Region: Geological Society of America Special Paper cations for regional kinematic analysis of the Basin and tonic evolution of a supradetachment basin: 40Ar/39Ar 333, p. 87–114, doi:10.1130/0-8137-2333-7.87. Range Province: Geological Society of America Bulletin, geochronology of the Nova Formation, Panamint Zandt, G., Gilbert, H., Owens, T.J., Ducea, M., Saleeby, J., and v. 113, p. 419–442, doi:10.1130/0016-7606​ (2001)​ 113<​ 0419​ ​ Range, California: Basin Research, v. 12, p. 19–30, doi:​ Jones, C.H., 2004, Active foundering of a continental arc :DAPOTM>2.0.CO;2. 10.1046/j​ .1365-2117.2000.00108.x.​ root beneath the southern Sierra Nevada in California: Norton, I., 2011, Two-stage formation of Death Valley: Geo- Stewart, J.H., 1983, Extensional tectonics in the Death Valley Nature, v. 431, p. 41–46, doi:10.1038/nature02847. sphere, v. 7, p. 171–182, doi:10.1130/GES00588.1. area, California: Transportation of the Panamint Range Oldow, J.S., Geissman, J.W., and Stockli, D.F., 2008, Evolu- structural block 80 km northwestward: Geology, v. 11, tion and strain reorganization within late Neogene p. 153–157, doi:10.1130/0091-7613(1983)11<153:ETITDV​ MANUSCRIPT RECEIVED 9 JULY 2014 structural stepovers linking the central Walker Lane and >2​ .0​ .CO;2.​ REVISED MANUSCRIPT RECEIVED 27 APRIL 2015 northern Eastern California shear zone, western Great Stockli, D.F., Dumitru, T.A., McWilliams, M.O., and Farley, MANUSCRIPT ACCEPTED 13 MAY 2015 Basin: International Geology Review, v. 50, p. 270–290, K.A., 2003, Cenozoic tectonic evolution of the White doi:10.2747​ /0020​ -6814.50.3.270.​ Mountains, California and Nevada: Geological Society Printed in the USA

480 www.gsapubs.org | Volume 7 | Number 4 | LITHOSPHERE

Downloaded from http://pubs.geoscienceworld.org/gsa/lithosphere/article-pdf/7/4/473/3040498/473.pdf by guest on 30 September 2021