<<

Quaternary faulting in Queen Valley, -Nevada: Implications for kinematics of -slip transfer in the eastern California shear zone– belt

Jeffrey Lee† Jason Garwood* Department of Geological Sciences, Central Washington University, Ellensburg, Washington 98926, USA Daniel F. Stockli Department of Geology, University of Kansas, Lawrence, Kansas 66045, USA John Gosse Department of Earth Sciences, Dalhousie University, Halifax, Nova Scotia B3H 4J1, Canada

ABSTRACT transfer, block rotation, and simple shear Keywords: normal faults, strike-slip faults, models, within the dextral fault system pro- thrust faults, terrestrial cosmogenic nuclide New geologic map, tectonic, geomorpho- posed for the eastern California shear zone– geochronology, fault-slip transfer, Walker Lane logic, and terrestrial cosmogenic nuclide Walker Lane belt are not applicable to this belt, Eastern California shear zone. (TCN) geochronologic data document the part of the Mina defl ection. Rather, dextral geometry, style, kinematics, and slip rates fault slip is transferred by both a restraining INTRODUCTION on late Quaternary faults within the Queen westward step and a releasing eastward step. Valley, California-Nevada area. These data Restraining and releasing bends have been Modern deformation between the North provide important insight into the kinematics extensively documented at a range of scales American and Pacifi c plates is distributed across of fault-slip transfer from the dextral White in strike-slip fault tectonic settings globally, a wide zone of the western margin of North Mountains fault zone northward into the and they have been simulated in analog mod- America, from the eastward Mina defl ection. Queen Valley is an ~16-km- els; thus, it is not surprising to document both into the western long, NE-trending basin bounded to the south within the ~630-km-long dextral shear zone (e.g., Bennett et al., 1999; Thatcher et al., 1999) by the White Mountains and underlain by that makes up the northern eastern Califor- (Fig. 1). Geodetic and geologic studies indicate four major Pleistocene to Holocene alluvial- nia shear zone–Walker Lane belt. Our results, that the San Andreas fault system accommodates fan surfaces. Four different fault types and combined with published slip rates for the ~75%–80% of the total relative plate motion. orientations cut and offset all but the young- dextral White Mountain fault zone and the The modern plate-boundary strain residual is est surfaces: (1) The normal-slip Queen Valley eastern sinistral Coaldale fault, suggest that transferred away from the San Andreas fault fault, which consists of a set of NE-striking, transfer of dextral slip into the Mina defl ec- system via the eastern California shear zone NW- and SE-dipping normal fault scarps that tion is partitioned into three different compo- northward into the Walker Lane belt and the cut across the SE side of the valley and offset nents: horizontal extension along the Queen Central Nevada seismic belt (Fig. 1). Geodetic all but the youngest surfaces; (2) discontinu- Valley fault, thrust faulting that merges into data indicate that right-lateral shear, at a rate of ous NE-striking, sinistral faults exposed on the the dominantly dextral slip along the Coy- ~9–13 mm/yr (Thatcher et al., 1999; Dixon et north side of the valley; (3) the NW-striking ote Springs fault, and dominantly sinistral al., 2000; Gan et al., 2000; Bennett et al., 2003; dextral Coyote Springs fault, which merges slip along the Coaldale fault. A velocity vec- Oldow, 2003; Hammond and Thatcher, 2007), into (4) a set of E-W–striking thrust faults. tor diagram of fault-slip partitioning across dominates within the eastern California shear Measured offsets across normal fault scarps Queen Valley predicts a small component of zone and Walker Lane belt, accounting for developed within 10Be TCN-dated surfaces contraction across the Coyote Springs and ~20%–25% of the total relative plate motion. yield minimum late Pleistocene horizontal western Coaldale faults. Contraction across In the northern eastern California shear zone extension rates of 0.1–0.3 mm/yr. Documented the Mina defl ection is consistent with global (north of the ), dextral displacement fault geometries and slip orientations across positioning system data. An observed reduc- is funneled through a relatively narrow zone of Queen Valley suggest that fault-slip transfer tion in late Pleistocene fault-slip rates at the shear along four major subparallel strike-slip models, such as the extensional displacement northern end of the eastern California shear fault zones, the Stateline, Death Valley–Furnace zone and across the southwestern part of the Creek–Fish Lake Valley, Hunter Mountain– †E-mail: [email protected] Mina defl ection may be explained by distribu- Panamint Valley, and White Mountains–Owens *Present address: 928 Roosevelt Street, Ridgecrest, tion of slip across a much broader zone than Valley fault zones (Figs. 1 and 2). At the north- California 93555, USA. generally thought. ern end of the eastern California shear zone,

GSA Bulletin; March/April 2009; v. 121; no. 3/4; p. 599–614; doi: 10.1130/B26352.1; 12 fi gures; 3 tables; Data Repository item 2008169.

For permission to copy, contact [email protected] 599 © 2008 Geological Society of America Lee et al.

formed as conjugate faults to the main dextral 124°W shear zone (Fig. 4B). Clockwise rotation of the 44°N 112°W blocks bounding the sinistral faults accommo- 44°N dated dextral slip. Wesnousky (2005) also noted Tertiary volcanic that paired normal-faulted basins exposed at the in ends of the sinistral faults imply that sinistral cover ver Pla Snake Ri faults have transferred slip between the basins and/or were formed as a consequence of clock- Cascades wise rotation. Different mechanisms for transferring slip

Juan de Fuca plate American North between subparallel dextral strike-slip faults plate and connecting faults have been proposed to the north and south of the Mina defl ection. In CNSB the northern eastern California shear zone, Lee WLB et al. (2001) used structural, kinematic, and

ISB geomorphic data from the Deep Springs fault (Fig. 2) to argue that domains of NE-striking Great Valley faults, bounded by NW-striking right-lateral strike-slip faults, exhibited no rotation and were dominated by dip-slip normal faults. Lee et al. (2001) argued that the mechanism of slip trans-

ECSZ Sa Colorado fer was one of right-stepping, high-angle normal n faults in which the magnitude of extension was 36°N A Plateau 124°W n proportional to the amount of strike-slip motion d 36°N re transferred (Fig. 4A). In contrast, in the north- a 112°W s ern Walker Lane belt, between Reno and Yer- f a ington (Fig. 2), Cashman and Fontaine (2000) Pacific plate ult used paleomagnetic data to argue that domains of NE-striking faults, bounded by NW-striking right-lateral strike-slip faults, were dominated by clockwise rotation and left-lateral strike-slip Figure 1. Simplifi ed tectonic map of the western part of the U.S. Cordillera showing the major faults. Fault movement and block rotation within geotectonic provinces and modern plate boundaries. Basin and Range extensional province is a zone of distributed deformation accommodated in dark gray; CNSB (Central Nevada seismic belt), ECSZ (eastern California shear zone), ISB the right-lateral strike-slip motion (Fig. 4B). (Intermountain seismic belt), and WLB (Walker Lane belt) are in light gray. Fault-slip transfer from a relatively narrow (~25 km wide) and simple geometric zone of dextral shear northward into a broader (~60 km wide) and complex geometric zone of faulting individual NW-striking strike-slip fault systems defl ection, which transfers dextral slip from styles and slip orientations is one of the dis- abruptly swing eastward into an array of NE- to fault zones that made up the northern end of the tinctive structural characteristics of the Mina E-W–striking faults in the southern portion of eastern California shear zone to the northern defl ection. In this paper, we combine new geo- the central Walker Lane belt, which is known Walker Lane belt and Central Nevada seismic logic mapping and tectonic, geomorphologic, as the Mina defl ection (Figs. 2 and 3). The dra- belt (e.g., Oldow, 1992; Oldow et al., 2001; and terrestrial cosmogenic nuclide (TCN) geo- matic change in Cenozoic fault orientation that Stockli et al., 2003; Wesnousky, 2005; Tincher chronologic data from the Queen Valley area defi nes the Mina defl ection is attributed to geo- and Stockli, 2008). The curvilinear Cenozoic to determine fault geometries, slip orientations, metric control by the latest Precambrian–earli- faults in the Mina defl ection form a z-shaped and slip magnitudes that bear on the mechanism est Paleozoic rifted continental margin, which extensional relay zone that transfers dextral slip. of fault-slip transfer from the relatively narrow was mimicked by Phanerozoic depositional pat- Fault-slip transfer through this regional releas- dextral White Mountains fault zone to the broad terns and mid-Paleozoic through Mesozoic con- ing bend results in the formation of multiple deformation zone that defi nes the southern part tractional structures (e.g., Stewart and Suczek, rhomboidal pull-apart structures. Oldow (1992) of the Mina defl ection (Figs. 2 and 3). 1977; Speed, 1978; Oldow et al., 1989). and Oldow et al. (1994) postulated that strike Fault-slip transfer by extension, contraction, slip and transtensional slip across the dextral QUEEN VALLEY and/or rotation is common within strike-slip faults were transferred across the Mina defl ec- fault systems worldwide (e.g., McKenzie and tion by right-stepping normal faults in which the Geological Setting Jackson, 1983, 1986; Cunningham and Mann, magnitude of extension was proportional to the 2007) and has been simulated in analog models amount of transtension transferred (Fig. 4A). Queen Valley, located at the southwest corner (e.g., Wilcox et al., 1973; Hempton and Neher, Wesnousky (2005) concluded that the exten- of the Mina defl ection and the northern end of the 1986; McClay and Dooley, 1995; McClay and sional fault-slip transfer model was not observ- White Mountains, is an ~16-km-long, NE-trend- Bonora, 2001). Fault-slip transfer by extension able at the surface today; instead, he proposed ing basin (Figs. 2, 3, and 5). This geographic set- or rotation has been postulated for the Mina that sinistral faults within the Mina defl ection ting marks where the relatively narrow zone of

600 Geological Society of America Bulletin, March/April 2009 Kinematics of fault-slip transfer, California-Nevada

40°N Figure 2. Simplifi ed fault map of the Walker Lane belt and northern part of the eastern Cali- HLF

MVF

NSNF 50 km fornia shear zone showing the major Quaternary faults. Solid ball is located on the hang-

W ing wall of normal faults; arrow pairs indicate relative motion across strike-slip faults. Fault SFZ abbreviations: BSF—Benton Springs fault; CF—Coaldale fault; DSF—Deep Springs fault; 39°N DVFCFLVFZ—Death Valley–Furnace Creek–Fish Lake Valley fault zone; GHF—Gumdrop Hills fault; HLF—Honey Lake fault; HMF—Hunter Mountain fault; MVF—Mohawk Val- Reno ley fault; NSNF—Northern Sierra Nevada fault; OVF— fault; PLF—Pyramid 121°W Lake fault; PSF—Petrifi ed Springs fault; PVF—Panamint Valley fault; QVF—Queen Valley

PLF fault; SLF—Stateline fault; SNFFZ—Sierra Nevada frontal fault zone; TPF—Towne Pass fault; WMFZ—White Mountains fault zone; WSFZ—Warm Springs fault zone.

Yerington

38°N

Fig. 3 119°W 118°W 120°W

GHF Wassuk

PSF Range BSF 01020

CF BSF km Sierra Nevada QVF

37°N Bishop

WMFZ N 119°W DSF

OVF

DVFCFLVFZ

HMF

SNFFZ 36°N Columbus Mono Salt TPF Lake SLF Mina Deflection Marsh 118°W CF PVF 117°W 116°W 38°N Adobe Valley Montgomery Figure 3. Shaded relief map of the Mina Pass defl ection showing the major Quaternary QVF faults. Solid ball is located on the hanging wall of normal faults; arrow pairs indicate WMFZ EPFZ relative motion across strike-slip faults. Fault FLVFZ Mountains abbreviations: BSF—Benton Springs fault; White CF—Coaldale fault; EPFZ—Emigrant Peak Long fault zone; FLVFZ—Fish Lake Valley fault Valley zone; QVF—Queen Valley fault; WMFZ— White Mountains fault zone. Location of map is shown in Figure 2.

NW-striking dextral faults that defi ne the north- and Kistler, 1970; Crowder and Sheridan, 1972; ery Pass, the Queen Valley fault merges with ern eastern California shear zone abruptly swing Stockli et al., 2003; Tincher and Stockli, 2008). the Coaldale fault, an E-W–striking, left-lateral into a broad zone of NE- to E-W–striking faults The SE side of the basin is bounded by the NE- strike-slip fault (Bradley, 2005; Tincher and that defi ne the Mina defl ection (e.g., Oldow, striking, NW-dipping Queen Valley normal Stockli, 2008) (Fig. 3). 1992; Stockli et al., 2003). Paleozoic metasedi- fault, which forms the northern termination of Apatite (U-Th)/He thermochronological data mentary rocks, Mesozoic granitic intrusions, the NW-striking, dextral White Mountains fault from the footwall of the Queen Valley fault have and Tertiary basalt lava and rhyolite tuff underlie zone (Fig. 3). Fault kinematic data are dominated yielded ages of ca. 3–5 Ma that decrease with the ranges surrounding the valley, and alluvial- by nearly pure down-to-the-NW dip slip along decreasing elevation and exhibit an infl ection fan deposits underlie the Queen Valley basin the Queen Valley fault; the extension direction point at ~2200 m, suggesting the existence of an (e.g., Dalrymple and Hirooka, 1965; Evernden is ~335° (Stockli et al., 2003). Near Montgom- exhumed Pliocene partial retention zone (Stockli

Geological Society of America Bulletin, March/April 2009 601 Lee et al.

sinistral fault A normal fault dextral fault B dextral fault

normal fault rotating block dextral fault dextral fault Figure 4. Schematic block diagrams illustrating two fault-slip transfer mechanisms between subparallel strike-slip faults proposed for the east- ern California shear zone and Walker Lane belt. (A) Displacement transfer model whereby the magnitude of extension along the connecting nor- mal faults is proportional to the amount of strike-slip motion transferred (modifi ed from Oldow et al., 1994). (B) Block rotation model in which clockwise rotation of blocks, bounded by dextral faults, is accommodated by sinistral faults (model of McKenzie and Jackson, 1983, 1986).

et al., 2000). The concordant and invariant ages ± pyroxene phryic basalt lavas (Fig. 5). We sug- Qf2 surfaces appear light toned and relatively below ~2200 m directly date the onset of exhu- gest that these lavas are Pliocene in age because smooth in aerial photography (Figs. 6 and 7). mation, initiation of normal slip, and the forma- they are similar to ca. 3–4 Ma basalt lavas that Proximal to the range front, Qf2 deposits are tion of the Queen Valley basin at 3.0 ± 0.5 Ma. occur in the Benton Range and throughout the composed primarily of sparse, ≤2-m-wide gra- Because the Queen Valley fault forms the north- Queen Valley area (Dalrymple and Hirooka, nitic boulders embedded in coarse, angular, gra- ern termination of the dextral White Mountains 1965; Crowder and Sheridan, 1972; Brad- nitic-derived pebbly sand. Bouldery, ≤3-m-wide fault zone, the opening of the basin at ca. 3 Ma ley, 2005; Blackburn et al., 2007; Tincher and debris fl ows are locally present. Distal Qf2 sur- refl ects the onset of right-lateral strike-slip fault- Stockli, 2008). Locally overlying unit Tb, there faces are characterized by weakly developed, ing. In addition, along the western range front is unit Trt, a quartz + feldspar, lithic-bearing widely spaced (several meters) channels that of the White Mountains, apatite samples did rhyolite tuff. In turn, this unit is overlain by unit dissect a relatively smooth surface. Deposits not yield apatite (U-Th)/He ages younger than Tts, a tan weathering, weakly bedded, basalt, that underlie this part of the surface contain ca. 12 Ma, suggesting that Pliocene normal dis- granite, metasedimentary, and obsidian pebble- <1% angular, weathered dominantly granite placement decreased southward away from the bearing quartz + feldspar tuffaceous sandstone. boulders, with lesser basalt, andesite, and scarce Queen Valley fault (Stockli et al., 2003). Based metasedimentary boulders, embedded in an on height of the escarpment and thickness of Quaternary Units angular granite-derived pebbly sand. the Queen Valley basin fi ll (~200 m) (Black Qf3 surfaces are darker than Qf2 surfaces in and Stockli, 2006), the minimum vertical offset The Queen Valley basin is underlain by four aerial photography and possess a moderately across the Queen Valley fault is 1370 m. major Pleistocene to Holocene alluvial-fan dissected surface composed of bar and swale surfaces. These surfaces have been subdivided morphology and bouldery debris fl ows adja- ROCK UNITS AND AGES into two groups, Qf surfaces, which cover the cent (within ~2.5 km) to the range front (Figs. 6 southwestern two-thirds of the valley, and Qqf and 7). Based primarily on inset geometry, the Queen Valley is underlain by late Quater- surfaces, which cover the northeastern third of Qf3 surfaces are subdivided into three subunits nary alluvial-fan deposits, which to the south- the valley (Figs. 5 and 6). Alluvial-fan surfaces (Figs. 5, 6, and 7). The oldest surface, Qf3a, pos- east are in fault contact with Paleozoic marine in each group are further subdivided based sesses the most well-developed bar and swale metasedimentary rocks (Crowder and Sheridan, on tone and fan surface morphology, includ- morphology, whereas the younger Qf3b surface 1972) that are in turn intruded by a variety of ing terrace height, presence or absence of bar possesses moderately developed bar and swale Jurassic and Triassic granitic bodies (Anderson, and swale morphology, degree of fan dissec- morphology. Three ages of debris fl ows are 1937; Harris, 1967; Evernden and Kistler, 1970; tion, and inset geometry (Figs. 6 and 7). The found within or on Qf3b surfaces. The oldest are Crowder and Sheridan, 1972). To the east and oldest Qf surface, Qf1, defi ned by its highly remnant pre-Qf3b debris fl ows that are partially northeast of the valley, late Quaternary alluvial- eroded ridge and ravine morphology, is pre- buried by the Qf3b surface, Qf3b debris fl ows, fan deposits are in fault contact with Paleozoic served primarily in the southwestern part of the and scarce, widely spaced, recent, single(?)- and Mesozoic rocks that are unconformably fi eld area (Fig. 5). Relatively dense vegetation event younger debris-fl ow channels deposited overlain by or fault-juxtaposed against Tertiary likely precludes desert pavement development on top of Qf3b surfaces (see TCN surface age conglomerate, Oligocene to Pliocene rhyolite, (e.g., Quade, 2001), and desert varnish was not dating section). andesite, and basaltic andesite (Dalrymple and observed. The underlying alluvial-fan sedi- Qf3c surfaces are the most aerially extensive Hirooka, 1965; Tincher and Stockli, 2008). ments, composed of granite, andesite, basalt, in Queen Valley (Fig. 5). These surfaces appear and metasedimentary rock–derived angular somewhat lighter in black-and-white aerial pho- Tertiary Units pebble-rich sand and typically weathered tographs compared to older Qf3 surfaces (Fig. 6). and fractured cobbles and boulders up to 2 m These surfaces possess well-developed plumose Unit Tb is exposed at the western end of the across, are deposited on pre-Tertiary bedrock texture and widely spaced, weak to moderately valley; it is composed of plagioclase ± olivine or Tertiary basalt lava. developed bar and swale morphology. Qf3c

602 Geological Society of America Bulletin, March/April 2009 Kinematics of fault-slip transfer, California-Nevada

Explanation Quaternary Qe Eolian deposits Qf4 Active or recently abandoned stream channels Qls Landslide Queen Canyon alluvial fans Qf3c Alluvial-fan surface and deposits Qqf3d Alluvial-fan surface and deposits Qf3b Alluvial-fan surface and deposits Qqf3c Alluvial-fan surface and deposits Qf3a Alluvial-fan surface and deposits Qqf3b Alluvial-fan surface and deposits Qf2 Alluvial-fan surface and deposits Qqf3a Alluvial-fan surface and deposits Qp Playa deposit 118.35°W 37.96°N Qf1 Alluvial-fan surface and deposits Qqf2 Alluvial-fan surface and deposits Tertiary Qqf1 Alluvial-fan surface and deposits PMT Tts Tuffaceous sandstone Trt Rhyolite tuff PMT PMT Tb Basalt lava Qf3c Paleozoic-Tertiary Paleozoic metasedimentary, Mesozoic intrusive, and PMT Tertiary conglomerate and volcanic rocks, undifferentiated Qf4 Map Symbols

Contact Qqf3d Normal fault; solid circle on the hanging wall Qf4 Fig. 9b Qf4 PMT Qqf3a 10.5 (v) Normal fault scarp; hachure on downthrown side 0.9 (v) Qf3c Qf4 PMT Strike-slip fault; arrows indicate relative motion across Qqf2 5.2 (v) fault; hachures point down dip Tts Qqf2 A' 1.4 (v) Qqf2 14.0 (v) Thrust fault; teeth on hanging wall Qf2 Qqf3b A Qqf2 13.6 1.4 (v) Qf2 TCN boulder location and age in ka Qf3c Qqf3d 1.2 (v) Qqf2 Qqf3a Measured fault offset, in meters; l = left Qqf2 Qqf2 50.4 (l) Qf2 3.6 (v) 1.1 (v) PMT lateral; v = vertical 72.3 52.5 Qf4 50.4 (l) Qqf3c Qqf3b 54.8 Queen 0.1 (v) Qqf2 Canyon Qf3c 6 16.3 (v) Qqf1 y 98.2 wa PMT h ig H 13.6 11.6 (v) 14.2 22.9 Qf2 118.50°W Orchard Spring Qf4 6.1 (v) 37.92°N Qf4 Qf4 Qf2 Qf2 Qf1 72.6 77.7 75.8 37.89°N PMT Qf3c Qf3c Qf2 PMT 80.8 78.5 Tb Qf2 Qf3c Qf4 118.375°W Trt Qls Qf4 Qf4 Tb Tts Tb Qf3a 70.6 9.1 1.6 Qf3ca Qe Qf1 Qf3b Morris Creek Qf3c Tb Qf3b Qf4 Qf3a

Qf2 Qf3b Qf3c Qp Qf2 Qf3c N Qf2 Qf2 Qf1 Qf2 PMT Qf2 Qe Qf2 Tb Qf1 Qf1Tb 37.86°N Qf3c Tb 02.55.0 118.45°W kilometers

Figure 5. Simplifi ed geologic map of the Queen Valley area showing Tertiary volcanic rocks, Quaternary sedimentary deposits and alluvial-fan surfaces, and Quaternary faults. Locations and ages of cosmogenic radionuclide samples are shown; locations and mag- nitudes of measured offset across fault scarps are shown. TCN—terrestrial cosmogenic nuclide.

Geological Society of America Bulletin, March/April 2009 603 Lee et al.

118.37°W Qqf3d 118.44°W 37.93°N 37.93°N Qqf3a

Qqf2 Qf2

Qqf2

Qf3c Qf2 QuQQueeneen CCaCanyonnyyon Orchard Spring scarp

Qls

0 1 Qf3b N kilometers (scale is approximate) Morris Creek MorrisMoM rrrriss CanyonCanyoy n 37.88°N 37.88°N 11118.37°W18.8 37°W°W 118.44°W

Figure 6. Aerial photograph of the central part of Queen Valley highlighting different alluvial-fan surfaces and prominent fault scarps (arrows). See Figure 5 for unit abbreviations.

surfaces are underlain by angular, dominantly nished granitic and metamorphic boulders and ing these surfaces are composed of angular granite-derived pebbly sand containing <1% cobbles, fresh boulder levee bars, and sand. granite, andesite, basalt, and metasedimentary generally large (≤2 m across), angular granite The Queen Canyon sequence of alluvial- rock–derived pebble-rich sandy matrix with cobbles and boulders embedded within. fan surfaces is similar to surfaces elsewhere typically weathered and fractured cobbles and The youngest surfaces, Qf4, are generally light in Queen Valley, but the deposits underlying boulders up to 2 m across. As observed on the toned in aerial photography and defi ne active or these surfaces can be traced to Queen Canyon Qf1 surface, desert pavement did not develop, recently abandoned stream channels cut into (Fig. 6). The oldest alluvial-fan surface, Qqf1, probably because the density of vegetation is older surfaces. Active channels are commonly is exposed as remnants deposited on pre-Ter- too high (e.g., Quade, 2001). relatively densely vegetated, whereas abandoned tiary bedrock at the apex of ridge and ravine The next youngest surface, Qqf2, is somewhat ones are not. Channel deposits include unvar- morphology (Fig. 5). The deposits underly- darker than Qf2 but lighter than Qqf3a, and it

604 Geological Society of America Bulletin, March/April 2009 Kinematics of fault-slip transfer, California-Nevada

118.50°W 37337.91°N.91°N 118.44°W TbTb 37.91°N

Coyote Springs fault N

Qf3b 012 kilometers (scale is approximate)

Qf2 Qf3c Qf3a Qf2 Qp Qf2 Tb Qf3b

37.65°N

118.50°W Qf1/2 37.65°N Qf3a/b 118.44°W

Figure 7. Aerial photograph of the southern part of Queen Valley highlighting different alluvial-fan surfaces and the Coyote Springs fault (arrows). See Figure 5 for unit abbreviations.

possesses a smooth surface morphology with with no channels. These surfaces contain plumose texture and weak to moderate bar and no channels (Fig. 6). These surfaces contain dominantly granite-derived, with lesser basalt swale morphology, and it is underlain by domi- dominantly granite-derived, with lesser basalt and andesite, angular pebbly sand with <1% nantly granite-derived, angular pebbly sand. and andesite, angular pebbly sand with scarce to 15–60 cm embedded angular cobbles and boul- In addition to the alluvial-fan deposits, land- ~5% strongly weathered and fractured primarily ders. The surface characteristics of the next two slide, playa, and eolian deposits are exposed granite boulders up to 2 m across. younger surfaces, Qqf3b and Qqf3c, are similar within the valley (Figs. 5, 6, and 7). An ~2.7 km2 Based on inset geometry, Qqf3 surfaces are to Qf3a. These surfaces were mapped as inde- landslide, Qls, exposed near the mouth of Mor- subdivided into four different subunits (Fig. 5). pendent units based on crosscutting and inset ris Canyon, is characterized by hummocky The oldest, Qqf3a, is darker compared to Qf2 geometries. The youngest of the Qqf3 surfaces, topography underlain by dominantly granite- surfaces. The surface morphology is smooth Qqf3d, possess a moderately to well-developed derived coarse angular sand and <1%–50%,

Geological Society of America Bulletin, March/April 2009 605 Lee et al.

TABLE 1. SUMMARY OF 10Be MODEL AGES Boulder Sample Quartz Be carrier 10Be/9Be 10Be/9Be error Uncertainty Sample Surface Lat. Long. Elev. height thickness Mass Mass Prod. rate (x10–13 atom (x10–13 atom Conc. Age 2σ number unit (°N) (°W) (m) (m) (cm) (g) (g) (atom g–1 yr–1) atom–1) atom–1) (x105 atom g–1) (ka) (ka) QV04 Qqf1 37°54.60 118°22.00 2136 0.9 2 51.1483 0.4993 23.8 34.94 1.016 22.84 98.2 5.6

QV01 Qqf2 37°54.70 118°22.10 2133 0.9 2 46.0416 0.5096 23.6 16.54 3.323 12.26 52.5 2.1 QV02* Qqf2 37°54.71 118°22.17 2124 0.7 3 60.2504 0.6431 23.2 23.09 4.456 16.51 72.3 2.7 QV03 Qqf2 37°54.60 118°22.00 2110 1 6 60.0088 0.4001 22.6 27.4 7.490 12.23 54.8 2.9 Mean (± s.e.#) 53.6 ± 1.2

QV07 Qf2 37°54.06 118°25.28 1888 2 4 55.0888 0.3605 19.6 31.94 1.563 14 72.6 7 QV08 Qf2 37°53.85 118°24.93 1952 1.4 3 49.9321 0.5108 20.6 23.13 4.448 15.85 78.5 3 QV09 Qf2 37°53.87 118°24.94 1946 1.4 2 59.3973 0.6512 20.7 22.33 4.308 16.4 80.8 3.1 QV10 Qf2 37°53.90 118°24.98 1940 1.4 2 48.4406 0.6738 20.5 16.41 4.030 15.28 75.8 3.6 QV11 Qf2 37°53.95 118°25.02 1930 1.2 4 50.0076 0.5114 20.2 22.51 3.913 15.41 77.7 2.6 Mean (± s.e.) 77.1 ± 0.8

QV13 Qf3b 37°53.27 118°25.15 1981 2.2 3 15.4193 0.4150 21 8.088 1.932 14.58 70.6 3.3 QV14 Qf3b 37°53.27 118°25.14 1983 0.9 3 59.3861 0.6458 21 2.622 6.456 1.909 9.1 0.4 QV15 Qf3b 37°53.18 118°24.88 2008 2.4 4 59.6890 0.6500 21.3 0.4732 2.693 0.3451 1.6 0.2

QV05 Qf3c 37°53.60 118°23.00 2093 0.9 2 23.7671 0.4093 23.1 4.557 1.131 5.255 22.9 1.1 QV06 Qf3c 37°54.09 118°23.00 2112 0.7 3 50.0017 0.5122 23 4.525 1.111 3.104 13.6 0.7 QV12 Qf3c 37°54.07 118°23.05 2116 0.9 2 58.3805 0.6433 23.2 4.434 1.054 3.271 14.2 0.7 Mean (± s.e.) 16.9 ± 0.2 Note: Production rate at sea level, high latitude (= 5.1 atom g–1 yr–1) was scaled to site according to Lal (1991) as modifi ed by Stone (2000), ignoring temporal variations ρ –3 Λ –2 in geomagnetic fi eld, which do exist but for which there is no consensus for adjustment. Rates were averaged for sample thickness (2–6 cm, = 2.7 g cm , fn = 160 g cm ), and less than 1% topographic shielding effect, but not for erosion, possible snow or ash cover, or inheritance. The 10Be t½ used (1.5 Ma) was recently shown to be 15% too high, but because the Accelerator Mass Spectometry standards (LLNL3000) were also calibrated to the 1.5 Ma half-life, there is no adjustment required (Nishiizumi et al., 2007). Concentration of Be carrier was 1015 μg/mL. 10Be/9Be for the four process blanks for these samples averaged 1.4 x 10–14, about 3× higher than the long-term average ratio for the shielded beryl crystal used as a carrier and measured at Center for Accelerator Mass Spectrometry, Lawrence Livermore National Laboratory. Sample QV15 was measured with a low precision (12%, 2σ) compared to the others (~4%, 2σ), and because the blank correction was ~25%, we cautiously report its age and recommend that it not be used in this or other studies. Considering all sources of error, the total systematic and random uncertainty in the ages is likely >20% 2σ. *Sample not used in calculating mean age. #s.e.—standard error.

0.5–2.0-m-sized angular cobbles and boulders. alluvial-fan surfaces (Fig. 5) (Table 1).1 Unfor- not be quantifi ed. A 1 mm/k.y. boulder erosion In contrast to all other Quaternary surfaces, peb- tunately, in many instances, three or fewer boul- rate would decrease the age of 15 ka boulders ble-sized clasts are a minor (<1%) component. ders were considered appropriate due to weath- by ~1% and 85 ka boulders by ~7%. By avoid- Playa deposits, Qp, are poorly exposed at the ering or position. In all instances, the exposure ing boulders with evidence of signifi cant ero- western end of Queen Valley and are composed ages were interpreted without correction for the sion (for example, “sombrero shaped boulders,” of fi ne- to medium-grained, subrounded to sub- possibility of inheritance (discussed later) or gnammas, and deep rillen), we minimized the angular, granite-derived quartzo-feldspathic sand boulder erosion, snow cover, exhumation from effect of erosion on the exposure ages, and interlayered with fi ne sand and mud with detri- the fan surface, or neutron loss due to small therefore the contribution of erosion to age tal mica. Locally, sand horizons are tuffaceous, boulder geometry (possible 10% effect), and error is <7%. Taking into account these errors, possessing pumice and obsidian grains. Hori- only in two instances were adjustments neces- we report a conservative 20% error (2σ) in the zontal bedding varies from 70-cm-thick sand sary for topographic shielding (<1% effect for accuracy of surface ages when calculating fault- beds at the base to a 40-cm-thick mix of mud both). Although some of the boulders were slip rates (see Fault Scarps, Magnitude of Off- and sand beds, which is capped by >2.5-m-thick shorter than desired (H = 0.7–2.4 m) to mini- set, and Slip Rates section). massive playa muds. Subrounded granite pebble mize the possible infl uence of burial or exhuma- and cobble fl oat suggests the presence of peb- tion, there was no apparent positive age cova- Morris Creek Catchment bly lens deposits, although these deposits are riance with boulder height and no geomorphic not exposed. Finally, medium- to fi ne-grained indication of burial or exhumation, indicating Five boulders from the Qf2 fan surface (Mor- pumice-bearing eolian sand, unit Qe, is exposed that partial burial may not have been important. ris Creek catchment) were sampled to evalu- along the western edge of the map area. Erosion has occurred on all boulders, yet none ate the possibility of inheritance. We interpret exhibited differential erosion greater than 5 cm. the low (5%) coeffi cient of variation about the TERRESTRIAL COSMOGENIC NUCLIDE Nevertheless, the total amount of erosion could mean age of those samples (Table 1) to indicate (TCN) SURFACE EXPOSURE AGES that this population of boulders does not con- tain inheritance—it would be very improbable Fifteen quartz-rich granite samples were 1GSA Data Repository Item 2008169, terrestrial for all fi ve boulders to contain the same amount cosmogenic nuclide (TCN) surface exposure ages and collected from the top of the largest and best- magnitude of offset across fault scarps, is available at of inheritance. We cannot preclude inheritance 10 preserved in situ boulders for Be TCN sur- www.geosociety.org/pubs/ft2008.htm. Requests may from the other fan boulder samples; however, face exposure age determination of fi ve of the also be sent to [email protected]. inheritance is unlikely on the basis of the low

606 Geological Society of America Bulletin, March/April 2009 Kinematics of fault-slip transfer, California-Nevada coeffi cient of variation about the mean age of an older debris fl ow that was buried by the Fault geometry varies from a single, W-facing of the Qf2 samples. The observed disparities 53.6 ka surface. Therefore, we take the mean of fault scarp along the southwestern portion that among boulder ages on the same fan surfaces the youngest two boulders to represent the time juxtaposes alluvial-fan deposits upon bedrock can be explained on the basis of fi eld observa- of last deposition before faults cut this surface. to a set of NW-facing, with subordinate SE- tions (see following). facing, en echelon fault scarps that cut alluvial- Boulder samples collected from the Mor- FAULT SCARPS, MAGNITUDE OF fan surfaces at the mouth of Queen Canyon. To ris Creek Qf2 surface yield a mean age of 77.1 OFFSET, AND SLIP RATES the north of Queen Canyon, the fault geometry ± 0.8 ka (n = 5, standard error) (Fig. 5) and a returns to a W- to SW-facing fault that juxta- range in ages between 72.6 ± 7.0 ka and 80.8 Four active fault types and fault orientations poses alluvial-fan deposits upon bedrock. Here, ± 3.1 ka. We interpret the mean age of these cut undifferentiated pre-Tertiary bedrock, Ter- older, smaller-offset, approximately E-W–strik- dates to indicate the timing of deposition of the tiary basalt lava, rhyolite tuff, and tuffaceous ing normal faults are exposed in the footwall faulted Qf2 fan surface. The mean age is con- sandstone, and the late Pleistocene alluvial-fan of the Queen Valley fault (Tincher and Stockli, sistent with a depositional event during oxygen surfaces: (1) a series of NE-striking, NW- and 2008). A few, discontinuous normal fault scarps isotope stage 4 glacial. SE-dipping normal faults and fault scarps cut that cut alluvial-fan surfaces are exposed within Two approximately contemporaneous Qf3 and offset bedrock and Qf1, Qf2, Qf3, Qqf1, the southern two-thirds of the valley (Fig. 5). surfaces on the fan from the Morris Creek Qqf2, and Qqf3 surfaces along the SE side of Height of the footwall escarpment ranges from catchment yield exposure ages ranging from the valley; (2) a set of discontinuous NE-strik- a maximum of ~1810 m at Morris Creek to a 70.6 ± 3.3 ka to 1.6 ± 0.2 ka. Field relations ing, sinistral faults cut Qf2 surfaces on the north minimum of ~620 m at the north end of the val- of boulders collected from the Qf3b surface side of the valley; (3) a NW-striking dextral ley (Fig. 5). Recent slip along the Queen Valley clearly indicate that the 70.6 ka boulder is from fault cuts and offsets Tertiary basalt lava, and fault is indicated by (1) well-developed geomor- a remnant debris fl ow and the 1.6 ka boulder is Qf1, Qf2, and Qf3 surfaces along the SW side phic features, such as triangular facets (Fig. 8) from the youngest debris fl ow in the fi eld area, of the valley; and this fault merges with (4) two, and wineglass-shaped drainages, and (2) fault possibly revealing how some fans are character- en echelon WNW- to E-W–striking, N-dipping scarps developed across late Pleistocene allu- ized by punctuated aggradation over a long time thrust faults that cut Tertiary basalt lava and Qf1 vial-fan surfaces. interval. Note, however, that the 1.6 ka boul- and Qf2 surfaces (Figs. 5, 6, and 7). Normal fault scarps cut Qf1, Qf2, Qf3a, der sample was measured with low precision, Qf3b, and Qfc, Qqf1, Qqf2, and Qqf3a, Qqf3b, and the age may be geologically meaningless Normal Faults and Qqf3c surfaces but not younger surfaces. (Table 1). Therefore, the only age that may be Where fault scarps offset small debris-fl ow useful for constraining the age of the Qf3b sur- Exposed along the SE side of the valley, the channels or erosional escarpments cut into face is sample QV14 at 9.1 ka. Boulder samples NE-striking Queen Valley normal fault extends alluvial-fan surfaces, geomorphic evidence collected from a Qf3c surface in the hanging for ~14 km from southwest of Morris Creek indicates normal dip slip but no dextral slip. wall of the Orchard Spring normal fault scarp to the NE end of the valley (Fig. 5). Between Most scarps are mildly to moderately eroded yield a mean age of 16.9 ± 0.2 ka (n = 3, stan- the southwestern end of the valley and Queen and degraded and moderately vegetated dard error) (Fig. 5), although it is possible that Canyon, the fault strikes NE, but it swings to a (Figs. 8 and 9; see footnote 1). In profi le, fault- the older of the three ages (sample QV05) may nearly N-S strike at the NE end of the valley. scarp morphology varies from relatively short contain an inherited concentration, in which case, the mean age of Qf3c would be 13.9 ka. Because we cannot distinguish the presence of inheritance, we will take the mean of all three NE SW ages to represent the age of the Qf3c fan.

Queen Canyon Catchment Orchard Spring fault scarp At the mouth of Queen Canyon, a single boul- der on Qqf1 has an exposure age of 98 ± 6 ka (±2σ). This was mapped as the oldest fan surface in the fi eld area. Three boulders were dated from the next-youngest alluvial fan surface, Qqf2, from the Queen Canyon catchment. Two boul- ders from this surface yield a mean age of 53.6 ± 1.2 ka (standard error) (Fig. 5). One boulder sample from a debris fl ow on the Qqf2 surface yields an age of 72.3 ± 2.7 ka, 20 ka older than the other two samples. On the basis of the experi- ment on the Qf2 samples from Morris Creek catchment, it is unlikely that chemistry or accel- erator mass spectrometry AMS error would con- tribute this error. We cannot preclude inheritance Figure 8. Photograph of the Orchard Springs fault scarp, which exhibits in the 72 ka boulder, but fi eld relations suggest offset of 6.3 m; hanging wall alluvial fan surface Qf3c in the foreground. that this boulder is from an exposed remnant Note triangular facet to the southwest of the fault scarp.

Geological Society of America Bulletin, March/April 2009 607 Lee et al.

A 15 Profile 1 Qf2

10 vertical offset 6.1 ± 0.3 m 5 Qf3c Vertical distance (m) Vertical 0 0 10 20 30 40 50 60 70 Horizontal distance (m)

Qqf2 30 Profile 2

20 vertical offset 14.0 ± 0.7 m

10 Qqf3d Vertical distance (m) Vertical 0 0 20 40 60 80 100 120 140 Figure 9. (A) Representative Horizontal distance (m) topographic profi les across 16 normal fault scarps and cal- Profile 3 Qqf2 12 vertical offset culated surface offset or scarp 8 10.5 ± 0.5 m offset. See B for location of profi les and Table 2 for a com- 4 Qqf2 plete list of measured profi les.

Vertical distance (m) Vertical 0 020406080100(B) Simplifi ed geologic map of Horizontal distance (m) a portion of the Queen Valley area highlighting location and 118.425°W 118.375°W magnitude of measured offset across fault scarps. Abbrevia- B tions: 3.6 (v7) indicates magni- PMT 0.9 (v25) Qqf3d 1 kilometer Qqf3a tude of vertical offset and pro- fi le number (see Table 2). See 10.5 (v3) Qf2 Qqf2 37.93°N Figure 5 for location of map, N Qqf2 map units, and symbols. A Qqf3b 1.4 (v8) Qf4 1.2 (v9) Qqf2 A' 5.2 (v4) 50.4 (l) 3.6 (v7) Qqf2 1.4 (v6) Qqf2 Qqf3a Qqf3c 14.0 (v2) Qqf3b Qf3c 1.1 (v5) Qqf3d Qqf2 Qqf2 Qf2 0.1 (v23)

6.1 (v1) 16.3 (v21) Qqf1 11.6 (v22) 37.90°N Qf2 Qf2 PMT Qf1 Orchard Spring Qf3c Qf3b Qls

118.425°W 118.375°W

608 Geological Society of America Bulletin, March/April 2009 Kinematics of fault-slip transfer, California-Nevada

scarps with a single sharp knickpoint, suggest- TABLE 2. MEASURED NORMAL FAULT OFFSET ing that these scarps formed by a single earth- Footwall Hanging wall quake event, to relatively tall scarps with either Farfield Farfield Fault scarp Vertical offset† one or two knickpoints, suggesting that one or Profile Northing* Easting* Surface slope (°) Surface slope (°) slope (°) (m) 1 0378397 4195639 Qf2 7 Qf3c 5 25 6.1 ± 0.3 two events resulted in their forma- 2 0379323 4197315 Qqf2 7 QCf3d 6 28 14.0 ± 0.7 tion (Wallace, 1977) (Fig. 9). Measured surface 3 0379395 4190137 Qqf2 1 QCf2 3 22 10.5 ± 0.5 offset or scarp offset across scarps ranges from 4a 0374517 4197510 Qqf2 5 QCf3d 0 16 6.6 ± 0.3 a few tens of centimeters to 14.0 m (Table 2). 4b 0374517 4197510 Qqf2 3 QCf3d 0 3 1.4 ± 0.1 ′ 5 0378219 4197576 Qqf2 1 QCf3a 1 2 1.1 ± 0.1 Transect A-A (Fig. 9B), perpendicular to the 6 0378241 4197772 Qqf2 1 QCf3a 1 3 1.4 ± 0.1 strike of fault scarps cutting Qqf2 surfaces at 7 0377894 4197900 Qqf2 4 QCf3a 5 19 3.6 ± 0.2 the mouth of Queen Canyon, yields a minimum 8 0378310 4197928 Qqf3a 3 QCf3a 2 2 1.4 ± 0.1 9 0376870 4198360 Qf3c 2 Qf3c 0 3 1.2 ± 0.1 9.4 ± 4.4 m of horizontal extension (assuming a 21 0380034 4196474 Qf1 7 Qf1 4 13 16.3 + 6.4/–5.9 fault dip of 60° ± 10°) toward 282°. Combining 22 0379810 4196695 Qf1 7 Qf2 6 18 11.6 + 4.2/–3.7 this calculated horizontal extension measure- 23a 0379449 4196758 Qf1 7 Qf2 2 3 ~0.1 ment with the 10Be model Qqf2 surface age of 23b 0379449 4196758 Qf1 7 Qf2 2 6 3.5 +1.8/–1.7 25 0376874 4198793 Qqf3a 3 Qqf3a 2 7 0.9 ± 0.4 53.6 ± 10.7 ka, we get a minimum horizontal *UTM coordinates (WGS84; Zone 11) are from the intersection of the topographic profile with the center of extension rate of 0.2 ± 0.1 mm/yr. At Orchard the fault scarp or the center of the graben (profiles 4a, 4b, and 23a, 23b). Spring, Qf3c has been deposited across the †Vertical offset measurements are minimum for those scarps with younger surfaces in the hanging wall. Queen Valley fault, which offsets a Qf2 sur- face a minimum of 6.1 ± 0.3 m (Figs. 8 and 9). This offset measurement combined with 10Be separated by ~2.5 km. Because alluvial-fan of hills may defi ne a pop-up or positive fl ower model surface ages of 77.1 ± 15.4 ka for Qf2 in surfaces may be time transgressive, we suggest structure in which oblique-slip thrust faults are the footwall and 16.9 ± 3.4 ka for Qf3c in the that the distal part of the fan surface is consid- characterized by a convex-upward shape and hanging wall yields an estimated vertical slip erably older than the dated proximal part, and, link down-dip (e.g., Sylvester, 1988; McClay rate bracketed between ~0.1 and 0.4 mm/yr and therefore, the calculated slip rate of 0.7 mm/yr and Bonora, 2001). Playa deposits (unit Qp) are a horizontal extension rate bracketed between is likely an overestimate. exposed on the NE, upslope side of these hills, ~0.1 and 0.3 mm/yr (assuming a fault dip of Exposed along the western end of Queen suggesting that these deposits ponded behind 60° ± 10°) toward 300°. These late Pleistocene Valley, there is a NW-striking, dominantly the developing hanging wall. horizontal extension rates are the same, within SW-facing with lesser NE-facing, ≤7.5-m-high error, to somewhat slower than the minimum fault scarp that we informally refer to as the KINEMATICS OF FAULT-SLIP 0.3 ± 0.1 mm/yr Pliocene horizontal extension Coyote Springs fault (Figs. 5 and 7). This fault TRANSFER ACROSS THE MINA rate across the Queen Valley fault based on apa- cuts Tertiary volcanic rocks, Quaternary playa DEFLECTION tite (U-Th)/He age of 3.0 ± 0.5 Ma for onset of deposits, and Qf1, Qf2, Qf3a, and Qf3b allu- normal slip (Stockli et al., 2003) and minimum vial-fan surfaces, although measurable offset Queen Valley forms the southwestern part of vertical offset of 1370 m across the fault. markers are not present. An elevated, weakly the Mina defl ection and the northern termination convex-up Qf2 surface is located at a left step of the dextral White Mountains fault zone, and, Strike-Slip Faults along the fault trace, which we interpret as a therefore, the faults exposed in the Queen Val- compressional ridge, thus indicating the fault ley area transfer slip from the White Mountains Exposed on the north-central side of the is characterized by dextral slip. fault zone into the Mina defl ection. Geomor- valley, there is a NE-striking, NW- and SE- phically, the Queen Valley normal fault is the facing, left-lateral strike-slip fault that cuts a Thrust Faults best-developed fault in Queen Valley. Tincher distal Qf2 surface. Evidence for strike-slip and Stockli (2008) argued that since ca. 3 Ma, faulting includes left-laterally offset chan- To the southeast, the Coyote Springs fault the magnitude of extension has decreased from nel edges, shutter ridges, and both NW- and merges into a set of ~WNW- to E-W–trending southwest to northeast along the Queen Val- SE-facing, ≤1-m-high fault scarps developed hills underlain by Tertiary basalt lava and Qf1 ley fault, implying that other structures within along the same fault trace. Two channel edges and Qf2 surfaces (Figs. 5 and 7). A sharp break Queen Valley are accommodating transfer of are left-laterally offset 48.9 ± 5.6 m and 51.7 in slope is exposed along the base of the S-fac- slip to the Coaldale fault. Our tectonic geomor- ± 5.7 m across the same fault trace, yielding a ing side of the hills, which we interpret as the phic observations indicate that the southern half weighted mean offset of 50.3 ± 4.0 m (Figs. 5 trace of N-dipping thrust faults. A break in slope of the Queen Valley fault is active, whereas the and 10; Table 3). This offset measurement is also exposed on the N-facing side of the hills, northern half is not, consistent with Tincher and combined with an age for the Qf2 surface of although this one is not as well developed as on 77.1 ± 15.4 ka yields a sinistral slip rate of 0.7 the S-facing side. This break in slope may rep- ± 0.2 mm/yr. This estimated slip rate is unusu- resent the trace of a S-dipping thrust fault. The TABLE 3. MEASURED ally high because it is nearly twice that esti- morphology of the hills suggests that they defi ne LEFT-LATERAL FAULT OFFSET mated for the sinistral Coaldale fault (cf. Brad- a hanging-wall anticline; however, poor exposure Fault no. Northing* Easting* Left-lateral ley, 2005; Tincher and Stockli, 2008), a fault to and absence of bedding in the basalt lava fl ows offset (m) which the ~50 m of offset is likely transferred and alluvial-fan deposits make it impossible to S1 374792 4198456 48.9 ± 5.6 S2 374644 4198240 51.7 ± 5.7 (see following). Although this distal Qf2 sur- determine if they are folded. If our interpretation *UTM coordinates (WGS84; Zone 11) for the face morphologically appears continuous with of an uplifted antiformal structure bounded by southwest end of the fault. the proximal Qf2 surface we dated, they are thrust faults with opposite dips is correct, the set

Geological Society of America Bulletin, March/April 2009 609 Lee et al.

Stockli’s interpretation. The sinistral slip faults we observed on the northern side of Queen Val- ley may be accommodating the missing strain. The Queen Valley fault, and other major 4,198,700 NE-striking, NW-dipping normal faults such as the Deep Springs and Towne Pass faults to the south, defi ne right steps or releasing bends (e.g., 05025 Sylvester, 1988; McClay and Dooley, 1995) and N transfer slip between the dextral White Moun- 4,198,650 meters tains–Owens Valley, Hunter Mountain–Panamint Valley, Death Valley–Furnace Creek–Fish Lake Valley, and Benton Springs–Petrifi ed Springs fault zones (Figs. 2 and 3). Right steps in the 4,198,600 Death Valley and Hunter Mountain–Panamint Valley fault zones are associated with the Death Valley and Panamint Valley pull-apart basins, 48.9 ± 5.6 m respectively (Burchfi el et al., 1987; Burchfi el and Stewart, 1966; Oswald and Wesnousky, 4,198,550 channel 1 edge 2002), and Queen Valley is, in part, analogous to these larger basins. However, in contrast to the Death Valley and Panamint Valley pull-apart basins, Queen Valley does not defi ne a simple 4,198,500 right-stepping extensional step between two dextral fault systems. Rather, sets of faults with different geometries and slip directions, includ- ing the NE-striking Queen Valley normal fault, ENE-striking sinistral faults, and a NW-striking 4,198,450 dextral fault that merges into WNW- to E-W– striking thrust faults, transfer slip from the nar- row White Mountains fault zone to the broad 51.7 ± 5.7 m zone of faults that defi ne the Mina defl ection. 4,198,400 Three kinematic fault-slip transfer models channel 2 edge have been proposed for the dextral slip–dom- inated eastern California shear zone–Walker Lane belt. In the extensional displacement fault transfer model (Fig. 4A), proposed for fault- 4,198,350 slip transfer across the Mina defl ection (Oldow, 1992) and across the Deep Springs and Towne Pass faults (Lee et al., 2001; Reheis and Dixon, 1996) (Figs. 2 and 3), the magnitude of exten- sion along right-stepping normal faults is pro- 4,198,300 portional to the amount of strike-slip motion transferred between subparallel dextral strike- slip faults. In the block rotation model (Fig. 4B), proposed for the central Walker Lane belt (Cash- 4,198,250 man and Fontaine, 2000) (Fig. 2), dextral strike- slip motion is accommodated by dominantly sinistral slip along faults that bound blocks that rotate clockwise within a zone of distributed deformation (McKenzie and Jackson, 1983, 4,198,200 374,600 374,650 374,700 374,750 374,800 374,850 374,900 1986). The simple shear couple model (e.g., Wilcox et al., 1973; Sylvester, 1988), in which Figure 10. Global positioning system (GPS) survey of a sinistral fault and offset channel edges the antithetic sinistral faults between subparallel developed in a distal Qf2 surface. Survey points are shown as open circles; survey point errors dextral faults undergo clockwise rotation during are shown as a light-gray band; best-defi ned fault and channel edges are shown by black lines. slip, was proposed for the Mina defl ection by Coordinates are UTM zone 11 (WGS84). Wesnousky (2005). The range of fault styles, orientations, and slip directions exposed in the Queen Valley area pre- cludes the extensional displacement transfer and block rotation models. On the other hand, the

610 Geological Society of America Bulletin, March/April 2009 Kinematics of fault-slip transfer, California-Nevada

Coaldale sinistral fault Queen Valley Figure 11. Block model illus- normal fault trating the transfer of dextral Queen Valley slip from the White Mountains pull-apart basin fault zone into the Queen Valley area. Dextral slip is partitioned Coyote Springs into two primary components: a dextral fault west-stepping restraining bend,

N leading to the development of a fl ower or pop-up structure, and an east-stepping releasing bend, leading to the development of a pop-up structure bounded by thrust faults normal fault and hanging-wall White Mountains dextral fault pull-apart basin.

simple shear couple model, with the shear cou- mechanisms proposed to explain the growth of Given these constraints, a horizontal two- ple parallel to the White Mountains fault zone, restraining and releasing bends along strike- dimensional (2-D) velocity vector diagram for predicts well the orientation, geometry, and slip faults (e.g., Mann, 2007, and references Pliocene fault-slip-rate transfer from the White style of faults documented in the Queen Valley therein). We postulate that small (relative to the Mountains fault zone into the Queen Valley area area. Simple shear is associated with rotation, Mina defl ection) preexisting crustal heteroge- makes the following predictions (Fig. 12B). (1) but we did not observe structural or geomorphic neities, likely related to mid-Paleozoic through The Benton block moved at a rate of 0.50 mm/yr evidence for rotation along the Queen Valley Mesozoic contractional structures, controlled toward 342° relative to the Queen Valley block. fault, nor did Bradley (2005) or Tincher and the development of the restraining and releasing Motion of the Benton block was 20° clockwise Stockli (2008) observe structural or geomorphic bends in Queen Valley. with respect to the NW strike of the dextral evidence for rotation along the Coaldale fault. We can estimate the magnitude of fault-slip Coyote Springs fault, suggesting that slip along The absence of such observations suggests that partitioning across the Queen Valley region this fault was dominantly dextral, with a lesser either rotation did not occur within the Mina using a combination of calculated Pliocene component of contraction. (2) The Benton block defl ection or rotation occurred to the north of to late Pleistocene slip rates along the normal moved at a rate of 0.80 mm/yr due north relative these faults within the Mina defl ection (Petro- Queen Valley fault (0.1–0.3 mm/yr) (Stockli to the Truman Meadows block. Motion of the nis, 2005; Petronis et al., 2004) but not along et al., 2003; this work), the sinistral Coaldale Benton block was 38° clockwise with respect faults that form the southern boundary of the fault (≥0.4 mm/yr) (Bradley, 2005; Tincher and to the strike of the NW-striking, dextral Coyote Mina defl ection. Thus, neither of these models Stockli, 2008), and the dextral White Mountains Springs fault, suggesting that slip along this por- can be applied to the range of fault types and fault zone (0.3–0.8 mm/yr) (Kirby et al., 2006). tion of the fault was also transpressional. (3) The geometries observed in the Queen Valley area. In particular, we can predict the style of fault- Queen Valley block moved at a rate of 0.35 mm/ We suggest, therefore, that transfer of dextral ing and slip rates along those faults, the Coyote yr toward 24° relative to Truman Meadows slip from the NW-striking White Mountains Springs fault and the western end of the Coaldale block. This vector was 46° counterclockwise fault zone into the Queen Valley area is parti- fault, for which measurable offset geomorphic with respect to the ENE-striking, sinistral tioned into two primary components (Fig. 11): markers were not observed (Fig. 12). Based on Coaldale fault; therefore, slip is predicted to (1) a restraining bend to the west, resulting in fi eld observations and structural data, we assign have been transpressional across the western development of contractional structures (pop- horizontal extensional slip rates (assuming a 60° end of this fault. up structure) that merge into the dextral Coyote dipping fault) along the Queen Valley fault of Fault-slip transfer of a calculated late Pleisto- Springs fault, and (2) a releasing bend to the 0.3 mm/yr (Pliocene) to 0.1 mm/yr (late Pleis- cene transtensional slip rate of 0.35 mm/yr along east, resulting in development of extensional tocene) toward 310° (this work) and a Pliocene the White Mountains fault zone—this slip rate structures (pull-apart basin) that merge into the pure strike-slip rate along the eastern part of the is approximately half of the assumed Pliocene sinistral Coaldale fault. Restraining and releas- Coaldale fault of 0.4 mm/yr toward 70° (Brad- slip rate—yields predicted slip rates and orienta- ing bends have been extensively documented at ley, 2005; Tincher and Stockli, 2008). For the tions that are different from those predicted for a range of scales in strike-slip fault tectonic set- northern part of the White Mountains fault zone, the Pliocene (Fig. 12C). (1) The Benton block tings (e.g., Mann, 2007, and references therein), we extrapolated calculated slip rates from the moved at a rate of 0.25 mm/yr, half the predicted and they have been simulated in analog models southern part where Kirby et al. (2006) reported Pliocene slip rate, toward 324° relative to the (e.g., McClay and Bonora, 2001; McClay and a middle Pleistocene dextral slip rate of 0.7– Queen Valley block. This vector was parallel to Dooley, 1995). Thus, it is not surprising to docu- 0.8 mm/yr toward 330°, parallel to the strike of the NW-striking, dextral Coyote Springs fault, ment both restraining and releasing bends within the northern part of the White Mountains fault suggesting that slip along this fault was dextral. the ~630-km-long dextral shear zone that makes zone, and a late Pleistocene transtensional slip (2) The Benton block moved at a rate of 0.4 mm/ up the northern eastern California shear zone– rate of 0.3–0.4 mm/yr toward 320°–340°. We yr toward 21° relative to the Truman Meadows Walker Lane belt. The development of the broad assume that the Pliocene strike-slip rate along block, also signifi cantly slower than the Plio- deformation zone that defi nes the Mina defl ec- the eastern part of the Coaldale fault and that cene rate. This slip vector was oriented 59° tion has been attributed to preexisting crustal the middle Pleistocene slip rate along the White clockwise with respect to the strike of the Coy- heterogeneities (e.g., Stewart and Suczek, 1977; Mountains fault zone have remained constant ote Springs fault, suggesting that slip along this Speed, 1978; Oldow et al., 1989), one of many since the Pliocene. fault was also transpressional but dominantly

Geological Society of America Bulletin, March/April 2009 611 Lee et al.

Figure 12. Kinematic fault model for dis- placement transfer from the dextral White eastern Coaldale fault Mountains fault zone to faults within the A 0.4 mm/yr Queen Valley region. (A) Simplifi ed geom- Truman Meadows etry of major Quaternary faults in the block Queen Valley region (compare to geologic map in Fig. 5). Solid lines indicate Pliocene slip rates (mm/yr) based on measured offset 0.35 mm/yr markers and age of marker, and dashed lines 0.8 mm/yr western Coaldale fault are predicted slip rates based on the velocity Coyote Springs fault Queen Valley vector diagram in B. Large arrow indicates block motion of the Sierra Nevada microplate (SN) with respect to a fi xed North American 0.5 mm/yr 0.3 mm/yr plate (NA) (Dixon et al., 2000). Solid ball is Queen Valley fault on the hanging wall of normal faults; teeth are on the hanging wall of thrust faults; arrow pairs indicate sense of relative strike- White Mountains slip motion. (B) Pliocene horizontal veloc- N block (fixed) ity vector diagram for fault blocks Benton Benton (B), Queen Valley (QV), Truman Meadows block White Mountains fault (TM), and White Mountains (WM) shown

SN-NA in A. Slip rates between fault blocks are 0.75 mm/yr in mm/yr. Solid lines indicate Pliocene slip rates (mm/yr) based on measured offset markers and age of marker, and dashed lines are predicted slip rates based on this velocity diagram. (C) Late Pleistocene hori- zontal velocity vector diagram for the fault blocks shown in A. Abbreviations and sym- bols same as in B. See text for discussion.

0.9 0.9 B C

B

0.5 0.5 0.75 (B/WM)

0.8 (B/TM) B 0.5 (B/QV)

QV 0.3 (QV/WM) 0.25 (B/QV) 0.35 (B/WM) N-velocity (mm/yr) N-velocity 0.35 (TM/QV) (mm/yr) N-velocity QV 0.1 (QV/WM) 0.4 (B/TM) 0 0 WM WM 0.35 (TM/QV) 0.4 (TM/WM) 0.4 (TM/WM) TM TM -0.2 -0.2 -0.5 0 -0.5 0 W-velocity (mm/yr) W-velocity (mm/yr)

612 Geological Society of America Bulletin, March/April 2009 Kinematics of fault-slip transfer, California-Nevada thrust slip, unlike the longer-term slip vector. The NE-SW contraction perpendicular to surfaces includes normal, sinistral, dextral, and (3) The Queen Valley block moved at a rate of the general NW trend of the eastern California thrust slip components. The morphologically 0.35 mm/yr relative to the Truman Meadows shear zone–Walker Lane belt was observed in dominant fault in the area is the NE-striking block, the same as the predicted Pliocene slip present-day global positioning system (GPS) normal-slip Queen Valley fault. East- and west- rate, toward 56°, 32° clockwise with respect to data (Bennett et al., 1999) and fault kinematic facing normal-slip fault scarps that cut alluvial- the predicted Pliocene slip vector. This slip vec- models across the eastern California shear zone fan surfaces exhibit vertical offset of ~14 m to tor was 14° counterclockwise with respect to the (Hearn and Humphreys, 1998). The competition a few tens of centimeters. If we combine these ENE-striking, sinistral Coaldale fault; thus slip between westward motion of the central Great data with the assumption of 60° ± 10° fault dip, is predicted to have been transpressional across Basin, driven by buoyancy (Sonder and Jones, we fi nd late Pleistocene horizontal extension the western end of this fault. 1999), and NW shear along the Pacifi c–North rates of 0.1–0.3 mm/yr across the Queen Valley During the late Pleistocene, dextral slip rates American transform boundary may explain this fault, which are the same, within error, to the appear to have decreased by ~50% across the episode of contraction (Hearn and Humphreys, estimated Pliocene extension rate. Tectonic geo- northern eastern California shear zone. Along 1998; Bennett et al., 1999). morphology indicates that slip along the NW- the White Mountains fault zone, dextral slip Late Pleistocene dextral slip rates have striking Coyote Springs fault is dextral and that rates decrease from 0.7–0.8 mm/yr during the decreased by ~50% across the two major strike- this fault merges into approximately E-W–strik- middle Pleistocene to 0.3–0.4 mm/yr during the slip faults, the White Mountains and Fish Lake ing thrust faults that defi ne a fl ower structure. late Pleistocene (Kirby et al., 2006) and along Valley fault zones, in the northern eastern Cali- Discontinuous ENE-striking sinistral faults are the Death Valley–Furnace Creek–Fish Lake fornia shear zone. Horizontal extension rates exposed on the north side of the valley. Valley fault zone, late Pleistocene slip rates may have decreased along the Queen Valley fault Queen Valley forms the southwestern part decrease northward from 4.2 +1.9/–1.1 to 4.7 as well. This reduction in dextral fault slip rate of the Mina defl ection, a relatively broad, right +0.9/–0.6 mm/yr along the northern Death Val- has been attributed to either a strain transient, by step in the dextral slip–dominated eastern Cali- ley fault (Frankel et al., 2007a) to ~2.5–3.0 mm/ accommodation of deformation along structures fornia shear zone–Walker Lane belt. Queen Val- yr along the Fish Lake Valley fault zone (Fran- to the east of the Fish Lake Valley fault zone (i.e., ley also forms the northern termination of the kel et al., 2007b). Based on estimated hori- the Silver Peak–Lone Mountain extensional com- dextral White Mountains fault zone, the dextral zontal extension rates along the Queen Valley plex), by distributed strain across Owens Valley, slip of which is partitioned across Queen Val- normal fault, a minimum of 0.3 ± 0.1 mm/yr and/or deformation through Long Valley caldera ley into two primary components: (1) a restrain- since the Pliocene and 0.1–0.3 mm/yr since the (Frankel et al., 2007b; Kirby et al., 2006). The ing bend to the west, resulting in development late Pleistocene, it is permissible that slip rates western ends of many of the EEN-WWS–strik- of contractional structures that merge into the have decreased along this fault as well. Further- ing sinistral faults within the Mina defl ection are dextral Coyote Springs fault, and (2) a releas- more, the 2-D velocity vector diagrams show exposed to the west of the Coyote Springs fault ing bend to the east, resulting in development that the predicted dextral slip rates along the and White Mountains fault zone, as far west as of extensional structures that merge into the Coyote Springs fault decreased during the late the western margin of Adobe Valley and eastern sinistral Coaldale fault. Restraining and releas- Pleistocene as a consequence of the reduction margin of Mono Lake, but not as far west as ing bends are typical in strike-slip fault zones in slip along the White Mountains fault zone Long Valley caldera (Fig. 3) (e.g., Gilbert et al., worldwide and have been simulated in analog and Queen Valley normal fault (Figs. 12B and 1968; Wesnousky, 2005; Lee, 2003, personal models of strike-slip fault systems. This geom- 12C). The lack of fault scarps within Quater- commun.). These sinistral faults are subparallel etry of slip transfer in the Queen Valley region nary deposits along the eastern Coaldale fault to the Coaldale fault and have been interpreted precludes alternative fault-slip transfer models, (Bradley, 2005; Tincher and Stockli, 2008) as accommodating clockwise rotation between a such as the extensional displacement transfer, may indicate a reduction in slip rate during the dextral shear couple (Wesnousky, 2005). If true, block rotation, and simple shear couple model, late Pleistocene. Assuming the sinistral slip this kinematic confi guration implies that the proposed for elsewhere in the eastern Califor- rate along the Coaldale fault also decreased observed reduction in late Pleistocene dextral nia shear zone–Walker Lane belt. Based on the by ~50% to 2.0 mm/yr during the late Pleisto- slip rates along the White Mountains and Fish fault-slip vectors for the White Mountains fault cene, then predicted slip rates along the western Lake Valley fault zones was accommodated, at zone, the Queen Valley fault, and the eastern Coaldale and Coyote Springs faults are some- least in part, by dextral shear along the western Coaldale fault, a small component of contrac- what smaller with a slightly different orienta- boundary of the eastern California shear zone tion is predicted across the Coyote Springs and tion to those just discussed. Nevertheless, as in between the Coyote Springs–White Mountains western Coaldale faults. Present-day GPS data the previous reconstructions, a small compo- fault zone and Long Valley caldera. An outcome also predict a small component of contraction nent of contraction is predicted across the west- of this interpretation is that dextral fault slip is across the Mina defl ection. ern Coaldale and Coyote Springs faults. distributed across a zone much broader than In general, the predicted contractional com- generally thought at the northern end of the east- ACKNOWLEDGMENTS ponent of slip across the Coyote Springs fault ern California shear zone and the southwestern and western end of the Coaldale fault is smaller part of the Mina defl ection. Thanks are due to Aaron Wood, who assisted with the geologic mapping, and to Kim Le and Lisa or nearly equal to the strike-slip component. Stockli, who processed the terrestrial cosmogenic This may explain why well-developed contrac- CONCLUSIONS nuclide (TCN) samples at the University of Kansas tion-related structures were not observed along Cosmogenic Nuclide Laboratory. Discussions with either of these faults (this work; Lee, 2003, New geologic mapping and 10Be alluvial-fan Tim Dixon contributed to this work. Kurt Frankel, personal commun.). Alternatively, the pre- surface abandonment ages in the Queen Valley Jeff Knott, and Steve Wesnousky provided thought- ful reviews. Funding for this project was provided by dicted contraction may be partitioned across area yield numerical ages on alluvial-fan sur- National Science Foundation grants EAR-0125782 the Mina defl ection and accommodated along faces that range from 98.2 to 9.1 ka and geomor- awarded to J. Lee and EAR-0125879 awarded to D. faults elsewhere in the region. phic evidence that slip along faults that cut these Stockli, and Central Washington University.

Geological Society of America Bulletin, March/April 2009 613 Lee et al.

REFERENCES CITED oir in Honor of Howel Williams: Geological Society of Oswald, J.A., and Wesnousky, S.G., 2002, Neotectonics and America Memoir 116, p. 275–329. Quaternary geology of the Hunter Mountain fault zone Anderson, G.H., 1937, Granitization, albitization, and related Hammond, W.C., and Thatcher, W., 2007, Crustal deforma- and Saline Valley region, southeastern California: Geo- phenomena in the northern Inyo Range of California- tion across the Sierra Nevada, northern Walker Lane, morphology, v. 42, p. 255–278, doi: 10.1016/S0169- Nevada: Geological Society of America Bulletin, v. 48, Basin and Range transition, western United States mea- 555X(01)00089-7. p. 1–74. sured with GPS, 2000–2004: Journal of Geophysical Petronis, M.S., 2005, Cenozoic Evolution of the Central Bennett, R.A., Davis, J.L., and Wernicke, B.P., 1999, Present- Research, v. 112, doi: 10.1029/2006JB004625. Walker Lane Belt, West-Central Nevada [Ph.D. the- day pattern of Cordilleran deformation in the Western Harris, D.L., 1967, Petrology of the Boundary Peak Adam- sis]: Albuquerque, New Mexico, University of New United States: Geology, v. 27, p. 371–374, doi: 10.1130/ ellite Pluton, in the Benton Quadrangle, Mono and Mexico, 286 p. 0091-7613(1999)027<0371:PDPOCD>2.3.CO;2. Esmeralda Counties, California and Nevada [M.S. Petronis, M.S., Geissman, J.W., and Oldow, J.S., 2004, Bennett, R.A., Wernicke, B.P., Niemi, N.A., Friedrich, A.M., thesis]: Berkeley, California, University of California, Paleomagnetic evaluation of crustal-scale block rota- and Davis, J.L., 2003, Contemporary strain rates in the Berkeley, 57 p. tions in the Mina defl ection of the central Walker Lane: northern Basin and Range Province from GPS data: Tec- Hearn, E.H., and Humphreys, E.D., 1998, Kinematics of the Eos (Transactions, American Geophysical Union), tonics, v. 22, doi: 10.1029/2001TC001355. southern Walker Lane belt and motion of the Sierra Nevada v. 85, suppl., Abstract GP43A-0840. Black, R.A., and Stockli, D.F., 2006, Potential fi eld interpre- block, California: Journal of Geophysical Research, Quade, J., 2001, Desert pavements and associated rock var- tation along the southern edge of the Mina defl ection, v. 103, p. 27,033–27,049, doi: 10.1029/98JB01390. nish in the Mojave Desert: How old can they be?: Geol- Nevada: Geological Society of America Abstracts with Hempton, M.R., and Neher, K., 1986, Experimental fracture, ogy, v. 29, p. 855–858, doi: 10.1130/0091-7613(2001) Programs, v. 38, no. 7, p. 351. strain and subsidence patterns over en echelon strike- 029<0855:DPAARV>2.0.CO;2. Blackburn, T.J., Stockli, D.F., and Walker, J.D., 2007, Mag- slip faults: Implications for the structural evolution of Reheis, M.C., and Dixon, T.H., 1996, Kinematics of the east- netite (U-Th)/He dating and its application to the geo- pull-apart basins: Journal of Structural Geology, v. 8, ern California shear zone: Evidence for slip transfer chronology of intermediate to mafi c volcanic rocks: p. 597–605, doi: 10.1016/0191-8141(86)90066-0. from the Owens and Saline Valley fault zones to Fish Earth and Planetary Science Letters, v. 259, p. 360–371, Kirby, E., Burbank, D.W., Reheis, M., and Phillips, F., 2006, Lake Valley fault zone: Geology, v. 24, p. 339–342, doi: 10.1016/j.epsl.2007.04.044. Temporal variations in slip rate of the White Mountain doi: 10.1130/0091-7613(1996)024<0339:KOTECS> Bradley, D., 2005, The Kinematic History of the Coaldale fault zone, eastern California: Earth and Planetary 2.3.CO;2. Fault, Walker Lane Belt, Nevada [M.S. thesis]: Law- Science Letters, v. 248, p. 153–170, doi: 10.1016/ Sonder, L.J., and Jones, C.H., 1999, Western United States rence, Kansas, University of Kansas, 96 p. j.epsl.2006.05.026. extension; how the west was widened: Annual Review Burchfi el, B.C., and Stewart, J.H., 1966, “Pull-apart” origin of Lee, J., Rubin, C.M., and Calvert, A., 2001, Quaternary fault- of Earth and Planetary Sciences, v. 27, p. 417–462, doi: the central segment of Death Valley, California: Geolog- ing history along the Deep Springs fault, California: 10.1146/annurev.earth.27.1.417. ical Society of America Bulletin, v. 77, p. 439–442, doi: Geological Society of America Bulletin, v. 113, p. 855– Speed, R.C., 1978, Paleogeographic and plate tectonic evolu- 10.1130/0016-7606(1966)77[439:POOTCS]2.0.CO;2. 869, doi: 10.1130/0016-7606(2001)113<0855:QFHAT tion of the early Mesozoic marine province of the west- Burchfi el, B.C., Hodges, K.V., and Royden, L.H., 1987, D>2.0.CO;2. ern Great Basin, in Howell, D.G., and McDougall, K., Geology of Panamint Valley–Saline Valley pull-apart Mann, P., 2007, Global catalogue, classifi cation and tectonic eds., Mesozoic Paleogeography of the Western United system, California: Palinspastic evidence for low-angle origins of restraining- and releasing-bends on active and States: Los Angeles, California, Pacifi c Section, Soci- geometry of a Neogene range-bounding fault: Journal ancient strike-slip fault systems, in Cunningham, W.D., ety of Economic Paleontologists and Mineralogists, of Geophysical Research, v. 92, p. 10,422–10,426, doi: and Mann, P., eds., Tectonics of Strike-Slip Restraining p. 253–270. 10.1029/JB092iB10p10422. and Releasing Bends: The Geological Society of Lon- Stewart, J.H., and Suczek, C.A., 1977, Cambrian and latest Cashman, P.H., and Fontaine, S.A., 2000, Strain partitioning don Special Publication 290, p. 13–142. Precambrian paleogeography and tectonics in the west- in the northern Walker Lane, western Nevada and north- McClay, K., and Bonora, M., 2001, Analog models of restrain- ern United States, in Stewart, J.H., Stevens, C.H., and eastern California: Tectonophysics, v. 326, p. 111–130, ing stepovers in strike-slip fault systems: The American Fritsche, A.E., eds., Paleozoic Paleogeography of the doi: 10.1016/S0040-1951(00)00149-9. Association of Petroleum Geologists Bulletin, v. 85, Western United States; Pacifi c Coast Paleogeography Crowder, D.F., and Sheridan, M.F., 1972, Geologic Map of p. 233–260. Symposium 1: Los Angeles, California, Pacifi c Sec- the White Mountain Peak Quadrangle, Mono County, McClay, K., and Dooley, T., 1995, Analogue models of pull- tion, Society of Economic Paleontologists and Miner- California: U.S. Geological Survey Geologic Quadran- apart basins: Geology, v. 23, p. 711–714, doi: 10.1130/0 alogists, p. 1–17. gle Map GQ-1012, scale 1:62,500. 091-7613(1995)023<0711:AMOPAB>2.3.CO;2. Stockli, D.F., Farley, K.A., and Dumitru, T.A., 2000, Calibra- Cunningham, W.D., and Mann, P., eds., 2007, Tectonics of McKenzie, D., and Jackson, J., 1983, The relationship between tion of the apatite (U-Th)/He thermochronometer on Strike-Slip Restraining and Releasing Bends: Geologi- strain rates, crustal thickening, paleomagmatism, fi nite an exhumed fault block, White Mountains, California: cal Society of London Special Publication 290, 482 p. strain and fault movements within a deforming zone: Geology, v. 28, p. 983–986, doi: 10.1130/0091-7613 Dalrymple, G.B., and Hirooka, K., 1965, Variation of potas- Earth and Planetary Science Letters, v. 65, p. 182–202, (2000)28<983:COTAHT>2.0.CO;2. sium, argon, and calculated age in a late Cenozoic basalt: doi: 10.1016/0012-821X(83)90198-X. Stockli, D.F., Dumitru, T.A., McWilliams, M.O., and Farley, Journal of Geophysical Research, v. 70, p. 5291–5296, McKenzie, D., and Jackson, J., 1986, A block model of dis- K.A., 2003, Cenozoic tectonic evolution of the White doi: 10.1029/JZ070i020p05291. tributed deformation by faulting: Journal of the Geologi- Mountains, California and Nevada: Geological Society Dixon, T.H., Miller, M., Farina, F., Wang, H., and Johnson, D., cal Society of London, v. 143, p. 349–353, doi: 10.1144/ of America Bulletin, v. 115, p. 788–816, doi: 10.1130/0 2000, Present-day motion of the Sierra Nevada block, gsjgs.143.2.0349. 016-7606(2003)115<0788:CTEOTW>2.0.CO;2. Nishiizumi, K., Imamura, M., Caffee, M., Southon, J., Finkel, Sylvester, A.G., 1988, Strike-slip faults: Geological Society of and some tectonic implications for the Basin and Range 10 Province: North American Cordillera: Tectonics, v. 19, R., and McAninch, J., 2007, Absolute calibration of Be America Bulletin, v. 100, p. 1666–1703, doi: 10.1130/00 p. 1–24, doi: 10.1029/1998TC001088. AMS standards: Nuclear Instruments and Methods, Ser. 16-7606(1988)100<1666:SSF>2.3.CO;2. Evernden, J.F., and Kistler, R.W., 1970, Chronology of B, v. 258, p. 403–413, doi: 10.1016/j.nimb.2007.01.297. Thatcher, W., Foulger, G.R., Julian, B.R., Svarc, J., Quilty, Emplacement of Mesozoic Batholithic Complexes in Oldow, J.S., 1992, Late Cenozoic displacement partitioning in E., and Bawden, G.W., 1999, Present-day deformation California and Western Nevada: U.S. Geological Survey the northwestern Great Basin, in Lane, C., and Steven, across the Basin and Range Province, western United Professional Paper P-0623, 42 p. D., eds., Geological Society of Nevada Walker Lane States: Science, v. 283, p. 1714–1718, doi: 10.1126/ Frankel, K.L., Brantley, K.S., Dolan, J.F., Finkel, R.C., Symposium: Structure, Tectonics and Mineralization of science.283.5408.1714. Klinger, R.E., Knott, J.R., Machette, M.N., Owen, L.A., the Walker Lane: Reno, Geological Society of Nevada, Tincher, C.R., and Stockli, D.F., 2008, Kinematics of multi- Phillips, F.M., Slate, J.L., and Wernicke, B.P., 2007a, p. 17–52. stage Tertiary faulting in the Queen Valley area, central Cosmogenic Be-10 and Cl-36 geochronology of offset Oldow, J.S., 2003, Active transtensional boundary zone Walker Lane belt, Nevada, in Oldow, J.S. and Cashman, alluvial fans along the northern : between the western Great Basin and Sierra Nevada P.H., eds., Late Cenozoic Structure and Evolution of the Implications for transient strain in the eastern Califor- block, western U.S. Cordillera: Geology, v. 31, p. 1033– Great Basin-Sierra Nevada Transition, Geological Soci- nia shear zone: Journal of Geophysical Research–Solid 1036, doi: 10.1130/G19838.1. ety of America Special Paper (in press). Earth, v. 112, doi: 10.1029/2006JB004350. Oldow, J.S., Bally, A.W., Lallemant, H.G.A., and Leeman, Wallace, R.E., 1977, Profi les and ages of young fault scarps, Frankel, K.L., Dolan, J.F., Finkel, R.F., Owen, L.A., and W.P., 1989, Phanerozoic evolution of the North Ameri- north-central Nevada: Geological Society of America Hoeft, J.S., 2007b, Spatial variations in slip rates along can Cordillera; United States and Canada, in Bally, Bulletin, v. 88, p. 1267–1281, doi: 10.1130/0016-7606 the Death Valley–Fish Lake Valley fault system deter- A.W., and Palmer, A.R., eds., An Overview: Boulder, (1977)88<1267:PAAOYF>2.0.CO;2. mined from LiDAR topographic data and cosmogenic Colorado, Geological Society of America, Geology of Wesnousky, S.G., 2005, Active faulting in the Walker Lane: 10Be geochronology: Geophysical Research Letters, North America, v. A, p. 139–232. Tectonics, v. 24, doi: 10.1029/2004TC001645. v. 34, doi: 10.1029/2007GL030549. Oldow, J.S., Kohler, G., and Donelick, R.A., 1994, Late Ceno- Wilcox, R.E., Harding, T.P., and Seely, D.R., 1973, Basic Gan, W., Svarc, J.L., Savage, J.C., and Prescott, W.H., 2000, zoic extensional transfer in the Walker Lane strike-slip wrench tectonics: The American Association of Petro- Strain accumulation across the eastern California shear belt, Nevada: Geology, v. 22, p. 637–640, doi: 10.1130/0 leum Geologists Bulletin, v. 57, p. 74–96. zone at latitude 36°30′N: Journal of Geophysical Research, 091-7613(1994)022<0637:LCETIT>2.3.CO;2. Oldow, J.S., Aiken, C.L.V., Hare, J.L., Ferguson, J.F., and v. 105, p. 16,229–16,236, doi: 10.1029/2000JB900105. MANUSCRIPT RECEIVED 24 OCTOBER 2007 Gilbert, C.M., Christensen, M.N., Al-Rawi, Y., and Lajoie, Hardyman, R.F., 2001, Active displacement transfer and differential block motion within the central Walker Lane, REVISED MANUSCRIPT RECEIVED 11 APRIL 2008 K.R., 1968, Structural and volcanic history of Mono MANUSCRIPT ACCEPTED 14 APRIL 2008 Basin, California-Nevada, in Coats, R.C., Hay, R.L., and western Great Basin: Geology, v. 29, p. 19–22, doi: 10.1 Anderson, C.A., eds., Studies in Volcanology: A Mem- 130/0091-7613(2001)029<0019:ADTADB>2.0.CO;2. Printed in the USA

614 Geological Society of America Bulletin, March/April 2009