<<

Downloaded from specialpapers.gsapubs.org on July 1, 2015

The Geological Society of America Special Paper 439 2008

Late Quaternary MIS 6–8 shoreline features of pluvial Owens , , eastern

A.S. Jayko* U.S. Geological Survey, U.C. White Mountain Research Station, Bishop, California 93514, USA

S.N. Bacon* Desert Research Institute, Division of Earth and Sciences, Reno, Nevada 89521, USA

ABSTRACT

The chronologic history of pluvial Owens Lake along the eastern in Owens Valley, California, has previously been reported for the interval of time from ca. 25 calibrated ka to the present. However, the age, distribution, and paleoclimatic context of higher-elevation shoreline features have not been formally documented. We describe the location and characteristics of wave-formed erosional and depositional features, as well as fluvial strath terraces that grade into an older shoreline of pluvial Owens Lake. These pluvial-lacustrine features are described between the Olancha area to the south and Poverty Hills area to the north, and they appear to be vertically deformed ϳ20 Ϯ 4 m across the active oblique-dextral Owens Valley fault zone. They occur at elevations from 1176 to 1182 m along the lower flanks of the and east of the fault zone to as high as ϳ1204 m west of the fault zone. This relict shoreline, referred to as the 1180 m shoreline, lies ϳ20–40 m higher than the previously documented Last Glacial Maximum shoreline at ϳ1160 m, which occu- pied the valley during marine isotope stage 2 (MIS 2). Crosscutting relations of wave-formed platforms, notches, and sandy beach de- posits, as well as strath terraces on lava flows of the Big Pine volcanic field, bracket the age of the 1180 m shoreline to the time interval between ca. 340 Ϯ 60 ka and ca. 130 Ϯ 50 ka. This interval includes marine oxygen isotope stages 8–6 (MIS 8–6), corresponding to 260–240 ka and 185–130 ka, respectively. An additional age esti- mate for this shoreline is provided by a cosmogenic 36Cl model age of ca. 160 Ϯ 32 ka on reefal tufa at ϳ1170 m elevation from the southeastern margin of the valley. This 36Cl model age corroborates the constraining ages based on dated lava flows and re- fines the lake age to the MIS 6 interval. Documentation of this larger pluvial Owens Lake offers insight to the hydrologic balance along the east side of the southern Sierra Nevada and will assist with future regional paleoclimatic models within the western Basin and Range.

Keywords: pluvial Owens Lake, system, deformed shorelines, marine isotope stage 6, MIS 6, Pleistocene lake, , Owens Valley.

*E-mail: Jayko: [email protected]; Bacon: [email protected]. Jayko, A.S., and Bacon, S.N., 2008, Late Quaternary MIS 6–8 shoreline features of pluvial Owens Lake, Owens Valley, eastern California, in Reheis, M.C., Hershler, R., and Miller, D.M., eds., Late Cenozoic Drainage History of the Southwestern Great Basin and Lower Colorado River Region: Geologic and Biotic Perspectives: Geological Society of America Special Paper 439, p. 185–206, doi: 10.1130/2008.2439(08). For permission to copy, contact [email protected]. ©2008 The Geo- logical Society of America. All rights reserved. 185 Downloaded from specialpapers.gsapubs.org on July 1, 2015

186 Jayko and Bacon

INTRODUCTION The graben floor is underlain by at least four fault-bound struc- tural blocks with late Quaternary and/or Holocene displacement. Owens Valley is part of the Owens River system, which, dur- The valley is asymmetric from west to east due in part to rela- ing wet periods of the Pleistocene, extended from Mono Basin in tive uplift of a horst block that lies west of the Owens Valley fault the north to in the southeast (Gale, 1914; Black- zone and east of the Sierra Nevada frontal faults, and in part to the welder, 1933; Smith et al., 1983; Jannik et al., 1991). Throughout deposition of large, late Quaternary alluvial fans along the west- Holocene time, the watershed has been hydrologically connected ern margin of the valley. The location of the Owens River mean- along only a third of its length—from Long Valley to Owens Lake der belt reflects the position of the distal margins of these fans, (Fig. 1). The lower-elevation basins, including Indian Wells, Sear- which have deflected the river eastward toward the base of the les, Panamint, and Death Valleys, have unique lacustrine histories Inyo Mountains (Fig. 1). that are a consequence of the magnitude of glacial-pluvial events, During much of the middle and late Quaternary, a freshwater which in part control the available water in the Owens River sur- lake occupied Owens Valley (Gale, 1914; Smith and Bischoff, face and groundwater systems. So, from an evolutionary perspec- 1997). The current spillway or sill in Owens Valley at ϳ1145 m tive, the duration and times of climate-controlled hydrologic north of Haiwee Reservoir (Gale, 1914) likely last overflowed ca. connectivity of the basins and their respective geochemistries 15.5 cal ka (Bacon et al., 2006) (Fig. 1). However, the sill dropped have affected the distribution of various aquatic species (Hubbs to that elevation during the late LGM following an event that and Miller, 1948; Miller, 1948; Firby et al., 1997; Reheis, 1999; eroded the berm, which confined the early LGM highstand at Reheis et al., 2002a, 2002b). ϳ1160 m.The location of the former 1180 m spillway is unknown. In this study, we report the first age constraints and basin- It may be eroded from the Haiwee area at the southernmost part of wide documentation of shoreline features and fluvial strath ter- Owens Valley, or it may lie further to the south at the southern end races of a pluvial Owens Lake at elevations between ϳ1180 and of Rose Valley, where erosional fluvial terraces formed in late 1200 m, collectively referred to as the 1180 m shoreline. The Quaternary (ca. 440–130 ka) basalt flows were noted by Duffield 1180 m shoreline is distinct from, and considerably older than, and Smith (1978a, 1978b). Although an effort was made to estab- the Last Glacial Maximum (LGM) ca. 25 calibrated (cal) 14Cka lish the location of this feature, it remains problematic. highstand at ϳ1160 m elevation, and it is also separate from the younger, lower-elevation shorelines (Carver, 1970; Beanland Methods and Clark, 1994; Bacon et al., 2006; Orme and Orme, this vol- ume). Beanland and Clark’s (1994) paleoseismic study of the Documentation of the 1180 m shoreline is based primarily on OwensValleyfault zone noted the presence of an ϳ1183 m strand- field observations, crosscutting relations with dated volcanic line along the northwest flank of the Coso Range. Other benches rocks, and a cosmogenic 36Cl erosion age on tufa provided by and notches southeast of Olancha at approximately the same Fred Phillips (New Mexico Tech, 2007, personal commun.). The elevation were also considered to be possible wave-formed plat- spatial distribution of constructional and erosional lacustrine land- forms or river terraces (Beanland and Clark, 1994, their strand- forms and erosional fluvial features associated with the 1180 m line A; sites 1–3 herein; Fig. 1). However, they were uncertain shoreline are described here. A description of each location is whether these sites represented offset from the ϳ1160 m LGM included in Table 1. Elevations estimated by global positioning strandline or if they were older, and so they reported them as system (GPS) were cross-checked with 1:24,000 topographic “Ͼ25 ka (?)” (Beanland and Clark, 1994, p. 6). quadrangle maps and 30 m digital elevation models (DEMs). A tufa sample was analyzed for 36Cl by standard procedures Tectonic Setting (Zreda, 1994), employing addition of a 35Cl spike and dissolu- tion of the carbonate sample in only nitric acid (F.M. Phillips, Owens Valley lies adjacent to the southern part of the Sierra 2007, personal commun.). Field and analytical results are given Nevada in eastern California (Fig. 1), and it is a 15–40-km-wide, in Table 2. The sample age was calculated using the spreadsheet 200-km-long complex graben shaped by active transtensional program CHLOE (Phillips and Plummer, 1996) using the 36Cl tectonism (e.g., Beanland and Clark, 1994; Savage and Lisowski, production parameters of Phillips et al. (2001) and the elevation- 1995; Reheis and Dixon, 1996). The structural setting is impor- latitude scaling of Lal (1991). No correction for secular variation tant because of the high potential for deformation of shorelines of Earth’s magnetic field was applied. The production parameteri- and spillway elevations by active tectonic processes. The valley zation of Phillips et al. (1990) was largely calibrated using samples is bounded by the Sierra Nevada frontal faults on the west and the from the same region, as well as using the same elevation-latitude White and Inyo Mountains fault zones on the east (e.g., Pakiser scaling; thus, the age calculation should be relatively robust. The et al., 1964; Bryant, 1984; Bacon et al., 2005; Slemmons et al., age was calculated assuming that the dissolved material was only

2008). The valley floor is split longitudinally by the active, oblique- CaCO3 (i.e., 56 wt% CaO and 44 wt% CO2), given that dissolu- dextral Owens Valley fault zone, which last ruptured along an tion was only in nitric acid, and thus 36Cl should not have been ϳ120 km segment with a M 7.5–7.7 earthquake in A.D. 1872 liberated from phases other than carbonate minerals (Phillips, (e.g., Beanland and Clark, 1994; Slemmons et al., 2008) (Fig. 1). 2007, personal commun.). Downloaded from specialpapers.gsapubs.org on July 1, 2015

Figure 1. Map of Owens Valley showing location of study areas and shoreline features of pluvial Owens Lake at elevations between ϳ1180 and 1200 m. Inset map shows the Owens River system and late Pleistocene , the headwaters of the Owens River system in Mono Basin, and terminus at Badwater in Death Valley (modified from Smith and Bischoff, 1997). Other locations include: B—Bishop; AB—Aberdeen; I—Independence; LP—Lone Pine; VT—Volcanic Tableland. Abbre- viations for late Quaternary faults: CFF—Centennial Flat fault; OVFZ—Owens Valley fault zone; S. IMF—Southern Inyo Mountain fault; SNFF—Sierra Nevada frontal fault; WMFZ—White Mountain fault zone. Downloaded from specialpapers.gsapubs.org on July 1, 2015

TABLE 1. DESCRIPTION OF 1200 to 1180 METER SHORELINE FEATURES AND ASSOCIATED DEPOSITS IN OWENS VALLEY Site Elevation gnihtroN gnitsaE noitpircseD erutaeF no. (m) 1 1180 4010297 413964 Wave-cut platform or paleoriver terraces of the Owens River Wave-cut platform; river terraces of Beanland and Clark (1994). 2 1180 4018762 416422 Wave-cut notches into older alluvial fans that exhibit well- Wave-cut notches developed soil having petrocalcic horizon (Bkm). Fans are older than ca. 25 cal ka based on ~1160 m shoreline (Lubetkin and Clark, 1988; Bacon et al., 2006). Wave-cut erosional features of Beanland and Clark (1994). 3 1179 ± 5 4024931 422364 Wave-cut notch and bluff; erosional feature formed into Wave-cut notch 1170 ± 5 alluvial fan with well-developed soil having petrocalcic horizon (Bkm); 1179 m elevation at upper break in slope of bluff; 1170 m at toe of bluff. 4 1185 ± 5 4027460 424105 Embayment; remnant exposure of well-sorted, subrounded- Embayment shoreline deposit subangular, medium to coarse sand composed almost entirely of quartz and feldspar lithologies; surface contains a lag of abundant fine-medium tufa fragments (rhizoliths?); ~0.6 cm below surface is boundary between an increase in clay (Bt horizon) that shows rubification and oxidation of parent material (Cox horizon). 5 1180 ± 6 4027291 424143 Wave-cut notch formed into Miocene-Pliocene basalt of Wave-cut notch Coso Formation. 6 1184 ± 4 4026825 423984 Abrasion platform or ravinement surface that has formed Abrasion platform or across metasedimentary rock. ravinement surface 7 1182 ± 6 4026825 423927 Embayment; laterally continuous incised planar surface Abrasion platform with gravel armored by a lag of sparse well-rounded gravel and lag subrounded-subangular basaltic cobble; surface underlain by rubified coarse sand. 8 1179 ± 4 4026802 423716 Tufa; large remnant exposure of tufa-cemented, well-sorted Tufa coarse sand with a gravelly to fine boulder lag at the surface; area of tufa fringes basaltic bedrock at edge of embayment. 9 1179 ± 4 4026986 423573 Beach rock; moderately exposed and ~1 m thick, well-sorted Beach rock coarse sand that is tufa-cemented; beach rock formed on shallow basaltic bedrock. 10 1169 ± 5 4027255 423493 Tufa mound; remnant exposure forming a topographic point Tufa mound that is primarily composed of ~1–2-m-thick tufa with silica veins. .tlasab otni demrof hcton tuc-evaW 548324 4208204 5 ± 5711 11 5711 ± 5 4208204 548324 tuc-evaW hcton demrof otni .tlasab tuc-evaW hcton 12 1181 ± 5 4027988 423855 Wave-cut notch between two embayments that is formed Wave-cut notch into basalt. 13 1176 ± 3 4028230 424681 Abrasion platform; subrounded-rounded gravel, cobble, and Abrasion platform

tufa (pedogenic CaCO3?) lag at surface. 14 1181 ± 5 4028304 424733 Embayment; laterally continuous incised planar surface Abrasion platform armored by a lag of well-rounded gravel and subrounded- subangular basaltic cobble; surface is underlain by rubified coarse sand with an increase in clay content (Bt horizon). 15 1175 ± 4 4028493 424602 Abrasion platform forming concordant surface that forms Abrasion platform topographic point; surface has subrounded-rounded gravel and cobble lag. 16 1181 4040381 421739 6.7–4.3 Ma fanglomerate of Slate Canyon (Stone et al., Truncated landform 2004) truncated at 1181 m. .kcordeb etimolod otni demrof hcton tuc-evaW 418814 4503404 5 ± 9711 71 9711 ± 5 4503404 418814 tuc-evaW hcton demrof otni etimolod .kcordeb tuc-evaW hcton .kcordeb etimolod derehtaew ro afuT 797814 7613404 3 ± 3811 81 3811 ± 3 7613404 797814 afuT ro derehtaew etimolod .kcordeb fuT a

.kcordeb etimolod otni demrof hcton tuc-evaW 317814 6603404 5 ± 0811 91 0811 ± 5 6603404 317814 tuc-evaW hcton demrof otni etimolod .kcordeb tuc-evaW hcton

20 1177 ± 3 4043260 418584 Pocket barrier; thin veneer of well-rounded bladed and Pocket barrier spherical gravel composed of different lithologies; gravel deposit is situated within saddle. 21 1180 ± 5 4043360 418726 Pocket barrier; one welldednuor- levarg .tsalc tekcoP reirrab 22 1177 ± 5 4043340 418806 Abrasion platform; surface is incised and contains trace well- Abrasion platform rounded gravel. (continued) Downloaded from specialpapers.gsapubs.org on July 1, 2015

TABLE 1. DESCRIPTION OF 1200 to 1180 METER SHORELINE FEATURES AND ASSOCIATED DEPOSITS IN OWENS VALLEY (continued) Site Elevation gnihtroN gnitsaE noitpircseD erutaeF no. (m) 23 1181 4052346 410429 Prominent wave-cut notch into bedrock (Haystack knob). Wave-cut notch

24 1180 4055266 410014 North Long John Canyon debris slide truncated at 1180 m. Truncated landform Slide material exhibits soil development; argillic horizon present (Blair, 1999). 25 1180 ± 5 4073946 401078 Wave-cut abrasion platform formed into bedrock adjacent to Wave-cut platform Mazourka Canyon Road.

26 1184 ± 5 4089709 394148 Wave-cut abrasion platform and weakly developed notch Wave-cut notch and platform into bedrock at base of range front. 27 1181 ± 5 4090783 393828 Abrasion platform into bedrock at base of range front.

28 1183 ± 5 4093367 393205 Abrasion platform and associated wave-cut notch formed Abrasion platform into bedrock at base of range front. 29 1200 ± 6 4096225 389680 Higher and older strath surfaces formed on Red Mountain Strath terraces lava flow (340 ± 60 ka) that extend to north. Surfaces are incised relatively more deeply and dip to west. 30 1195 ± 5 4096062 389700 Strath surfaces formed on Red Mountain lava flow (340 ± 60 Strath terraces ka) with local well-rounded granitic cobble clasts. 31 1196 ± 5 4095984 389754 Strath surfaces formed on Red Mountain lava flow (340 ± 60 Strath terraces ka); no gravel. 32 1197 ± 5 4095844 389564 Accumulation of spherical and well-rounded gravel and Deposit on strath terrace cobble consisting of granitic and metamorphic lithologies.

33 1195 ± 5 4095871 389428 Strath surfaces formed on Red Mountain lava flow (340 ± 60 Strath terraces ka) with sparse spherical and well-rounded to rounded gravel and cobble (3–12 cm long axis). Surface moderately incised by 2–4-m-deep channels that dip to east. 34 1186 ± 5 4091059 388094 Wave-cut notch into northern part of Armstrong Creek lava Wave-cut notch flow (<300 ka). Many ephemeral channels are graded to this elevation, which are formed on an ~6-m-high geomorphic scarp. Base of scarp is at ~1180 m. 35 1185 ± 5 4090721 388041 Wave-cut notch into northern part of Armstrong Creek lava Wave-cut notch flow (<300 ka). 36 1192 ± 5 4087019 389101 Wave-cut notch and platform into Armstrong Creek lava flow Wave-cut notch and platform (<300 ka). 37 1193 ± 5 4086966 388618 Wave-cut notch and platform into Armstrong Creek lava flow Wave-cut notch and platform (<300 ka) with sparse ~2 cm rounded granitic clasts. 38 1192 ± 5 4087075 388518 Abrasion surface into Armstrong Creek lava flow (<300 ka). Abrasion surface

39 1193 ± 5 4087046 388446 Very well-rounded volcanic stone in crag of pressure ridge of Shoreline deposit Armstrong Creek lava flow (<300 ka). 40 1193 ± 6 4087138 388351 Wave-cut notches in Armstrong Creek lava flow (<300 ka) Wave-cut notches/shoreline and associated deposit that contains very well-rounded, deposit pea-sized gravel consisting mostly of volcanic lithologies.

41 1204 4051096 403487 Strath terraces on bedrock at the mouth of Lone Pine Creek. Strath terraces Alluvial fans dated between ca. 25 cal ka and 17 ka by Lubetkin and Clark (1988) and Bierman et al. (1995) are inset into, and are graded to, elevations between 1160 and 1145 m; therefore, strath surfaces older than ca. 25 cal ka. 42 1199 4033141 406390 Degraded abrasion surface into bedrock at base of Sierra Abrasion surface Nevada range front. 43 1182 4010355 410780 Erosional benches developed on old alluvial-fan deposits Abrasion surface and trimline east of Sage Flat; trimline with gentle shoreline angle. 44 1173 4009510 411128 Tightly packed, clast-supported cobble and boulder deposit Possible boulder beach berm along east side of 1182 m benches, east of Sage Flat Downloaded from specialpapers.gsapubs.org on July 1, 2015

190 Jayko and Bacon

TABLE 2. FIELD AND ANALYTICAL DATA FOR 36Cl SAMPLE OL96-1 formed abrasion surfaces, platforms, and notches. Fluvial strath FROM SITE 10, CENTENNIAL FLAT EMBAYMENT terraces are located near the shoreline features but at slightly Variable Definition Units Value 36 36 15 higher elevation north of Independence. All these features are R36 Sample Cl/Cl ratio Cl/10 Cl 46,073 36 36 15 present on Paleozoic-Mesozoic bedrock or Miocene-Pliocene ±R36 Sample Cl/Cl ratio uncertainty Cl/10 Cl 1250 –3 rb Bulk density g cm 2.65 volcanic and sedimentary rocks of the Coso Formation (e.g., l Sample thickness cm 3 s Stinson, 1977), on Miocene-Pliocene fanglomerate (e.g., Stone θ Water content cm3/cm3 0.1 avg et al., 2004) and late Pleistocene alluvium, or on late Pleistocene velE m 9611 taL N° 22983.63 lava flows of the Big Pine volcanic field (e.g., Cox et al., 1963; gnoL W° 71358.711 Turrin and Gillespie, 1986, with localities and ages reported in

Ssnow Snow shielding unitless 1.00 Connor and Conway, 2000, Figure 7 therein; Dorn et al., 1987; S Total shielding unitless 0.999 T Martel et al., 1987). Λ –2 f Effective attenuation length g cm 170 LOI Loss on ignition wt% 42.39 Dating of the 1180 m shoreline is based on crosscutting rela-

Na2O wt% 0.04 tions of lacustrine features with dated lava flows of the Big Pine MgO wt% 1.06 volcanic field within the northern part of the valley, and from a Al O wt% 0.21 36 2 3 Cl exposure age on tufa at ϳ1170 m elevation in the south (F. SiO wt% 0.93 2 Phillips, New Mexico Tech, 2007, personal commun.). Recon- P O wt% 0.02 2 5 naissance observations of weathering also distinguish the clastic K2O wt% 0.03 CaO wt% 52.92 deposits associated with the 1180 m shoreline from younger LGM TiO wt% 0.01 2 clastic lacustrine deposits at ϳ1160 m, based mainly on the MnO wt% 0.00 degree of weathering of the soil profile. Fe2O3 wt% 0.06 Cl ppm 7.86 B ppm 14.00 CENTENNIAL FLAT EMBAYMENT Sm ppm 0.00 Gd ppm 5.00 Depositional and erosional remnants of the 1180 m shoreline U ppm 2.00 Th ppm 2.00 are located in southern Owens Lake basin along the northwestern flank of the Coso Range near Centennial Flat (Figs. 1 and 2). This Msample Sample mass g 110.200 35 Mspike Mass Cl spike solution g 4.003 embayment contains erosional wave-formed platforms or abra- C Concentration spike solution g g–3 0.997 spike sion surfaces (Qab) and associated wave-formed notches and S/S Analytical stable isotope ratio [35Cl/(35Cl + 37Cl)] 20.94 36 36 15 trimlines that formed on massive, indurated late Cenozoic basalt R/S Analytical Cl/Cl ratio Cl/10 Cl 8120.00 and Mesozoic metamorphic basement rock (Fig. 3). Similar fea- tures are also preserved on friable, fine-grained late Neogene clas- tic rocks of the Coso Formation (Table 1, sites 5–7 and 11–15). 1180 METER SHORELINE AND FLUVIAL FEATURES These remnant wave-formed surfaces are distinctly flat-lying IN OWENS VALLEY and typically terminate at abrupt breaks in slope that form trim- lines or notches (Fig. 4). Rare, erratic, and very well-rounded, We document lacustrine and fluvial features at ϳ1180 m ele- cobble- and pebble-sized clasts of Mesozoic metasediment, in- vation in Owens Valley and focus on four areas, two sites to the cluding hornfelsic siltstone, mudstone, and intermediate meta- east of the active Owens Valley fault zone and two to the west volcanic rock, occur just below the 1180 m trimlines and notches. (Fig. 1). These sites are referred to as: (1) the Centennial Flat These clasts also form a cobbly to pebbly lag on the surface of embayment, on the northeastern corner of the Coso Range; (2) the sandy embayment deposits or on abrasion surfaces (sites 6 and 7; Swansea pocket barrier, in the southern Inyo Mountains range Table 1; Figs. 3A and 3B). front near Swansea; (3) the southern end of the Big Pine volcanic The most extensive depositional feature at the embayment field in north-central Owens Valley, which includes fluvial ero- consists of a 3–4-m-thick, massive and moderately sorted, coarse sional features on lava flows north of Taboose Creek and shore- arkosic sand deposit (unit Qem; Figs. 2 and 3B). The Qem deposit line erosional features on lava flows near Sawmill Creek; and is widely distributed; it extends for nearly 1 km, and it ranges in (4) the Alabama Hills strath terrace and wave-formed platforms elevation from 1165 to 1182 m. The northwestern half of the near Lone Pine. We also briefly describe other erosional shoreline embayment is truncated by LGM and younger gravelly beach features observed in Owens Valley that are associated with the ridges (Qg) below ϳ1160 m (Beanland and Clark, 1994; Bacon 1180 m shoreline. et al., 2006; Orme and Orme, this volume). Depositional features associated with the shoreline include The Qem deposits grade to the same elevation near an iso- remnant cobble and gravel beach deposits, sandy embayment lated, 3–5-m-thick outcrop of whitish-gray, bedded silt and fine deposits with a lag of gravel and tufa fragments, and calcareous sand with a gravelly lag of tufa fragments mapped as unit Qm in deposits in the form of tufa mounds with local secondary lami- the western margin of the embayment (site 8; Table 1; Figs. 2 nated silica, as well as sandy beach rock cemented with tufa. and 3C). The lithologic association of the Qm sediment is typical Erosional wave-generated landforms include trimlines, wave- of a shallow-water to depositional environment linked Downloaded from specialpapers.gsapubs.org on July 1, 2015

Late Quaternary MIS 6–8 shoreline features of pluvial Owens Lake 191

Figure 2. Map showing location of Centennial Flat embayment along the northeast corner of the Coso Range. Numbered site locations (stars) are described in text and Table 1. CFF—Centennial Flat fault, modified from Jayko (2008). LGM—Last Glacial Maximum; MIS—marine oxygen isotope stage. to groundwater discharge. Outcrops of beach rock cemented by surfaces range in thickness from ϳ0.5 m at the ϳ1179 m trim- tufa, gravelly tufa lag, and isolated well-rounded pebbles and line and notches to ϳ3–4 m downslope at 1169 m near out- gravel clasts lie along a trimline that crosses a bedrock slope at crops of tufa mounds. These massive mounds consist of frac- ϳ1179 m. These features lie along the southwestern margin of tured and bedded tufa with secondary siliceous laminations as the embayment between sites 8 and 9 (Table 1; Figs. 2 and much as 2–6 mm thick that infill dissolution voids (site 10; 3C). The in situ tufa beach rock and fragmental lag on bedrock Table 1; Figs. 2 and 3D). Downloaded from specialpapers.gsapubs.org on July 1, 2015

Figure 3. Photographs showing shoreline features and embayment deposits at the Centennial Flat embayment (see Figs. 1 and 2). (A) Abrasion platform in foreground and /embayment deposit in background; person for scale, view looking west. (B) Embayment deposits of coarse sand overlain by lag gravel, which consists of very well-rounded beach pebble and gravel clasts; person for scale, view looking west. Inset shows very well-rounded beach pebbles in the gravel lag deposit. (C) Strandline of fine-grained clastic sediments composed of marl and tufa deposited on bedrock on the upthrown block of the Centennial Flat fault; view looking northwest. (D) Thick accumulation of bedded tufa and local secondary silica lamina; backpack for scale, view looking northwest. (E) Chlorine-36 exposure age for sample OL96-1 as a function of total erosion since the tufa formed. Dashed lines represent age uncertainties calculated from the 1␴ analytical standard deviation. The arrows indicate the ages associated with the estimated limits on total tufa erosion. Downloaded from specialpapers.gsapubs.org on July 1, 2015

Late Quaternary MIS 6–8 shoreline features of pluvial Owens Lake 193

Figure 4. Photographs showing abrasion platforms and wave-formed notches at the Centennial Flat embayment (see Figs. 1 and 2). (A) Abrasion platforms and wave-formed notches looking west from site 13 to site 11. (B) Well-developed abrasion surface consisting of wave-formed notch and abrasion platform on a local bedrock headland; view looking west.

Relative Age of Deposits The tufa mound sampled was close to the highest shoreline The relative degree of weathering, in the form of rubification deposits, may have been reef-forming, and is therefore interpreted and accumulation of pedogenic clay, is used to differentiate the to be a nearshore tufa deposit rather than a deep-water tufa tower. higher sandy deposits of the embayment (Qem) at ϳ1180 m from Similar shoreline tufa deposits at , Panamint Valley, the lower LGM and younger gravelly beach deposits (Qg) and Lake Lahontan have thicknesses around 0.5–3.5 m (e.g., Ben- between 1145 and 1160 m. From a reconnaissance-level investi- son, 1994). The 36Cl exposure age of the tufa is estimated assum- gation of the soil profile of these deposits, the moderately devel- ing that 30–75 cm of tufa has been removed by chemical and oped soils associated with the 1180 m shoreline exhibit colors that physical erosion. The total erosion limits correspond to limiting are brownish yellow (10YR 6/6) within soils developed from erosion rates of 2.2 mm/k.y. (for 30 cm erosion) and 4.1 mm/k.y. arkosic parent material. Furthermore, an increase in clay is appar- (for 75 cm erosion), which are assumed to be constant (Phillips, ent in the upper part of the profiles, which also exhibit no dis- 2007, personal commun.). Figure 3E shows the calculated 36Cl cernible carbonate accumulation. In contrast, the LGM and exposure age of the tufa as a function of total erosion depth from younger deposits are similar in color to the parent material (i.e., Table 2. The lower and upper limits of the estimated erosion yield no rubification) and lack apparent increase in clay content, but age limits between 192 and 127 ka, respectively; therefore, we get they do have an incipient development of pedogenic carbonate an average age of ca. 160 Ϯ 32 ka. Application of the alternative that forms thin, discontinuous coatings under clasts. Locally, dis- 36Cl production rate of Stone et al. (1996) for spallation from cal- continuous outcrops of beach rock are cemented by tufa carbon- cium would give a somewhat older age of ca. 218 Ϯ 33 ka ate (as opposed to pedogenic carbonate). (Phillips, 2007, personal commun.). These results, although from only a single sample from tufa sampled at ϳ1170 m (site 10), are Cosmogenic 36Cl Age Estimate consistent with a marine oxygen isotope stage (MIS) 6 age. This In 1996, Fred M. Phillips in the company of Antony and elevation lies below the highstand elevation for this locality. The Amalie Jo Orme sampled the tufa mound at site 10 in the Centen- sampled tufa is continuously exposed up to the beach deposits and nial Flat embayment for chlorine-36 (36Cl) surface exposure dat- relicts of tufa at 1180 m. We infer that the tufa at this elevation ing. The amount of 36Cl production in this sample was influenced formed a reef that was established as the lake was rising, as we by spallation of calcium. As a result, the calculated age is sensi- have not found evidence for a lake highstand that was stable tive to surface erosion, as is typical of spallation-dominated enough to erode shoreline features in bedrock at 1170 m elevation nuclides such as 10Be. Interpretation of the cosmogenic nuclide elsewhere in the basin. age therefore requires careful evaluation of the total erosion of the tufa, which proceeds at a relatively high rate due to the high sol- Lacustrine Setting ubility of calcite (Phillips, 2007, personal commun.). The tufa The Centennial Flat embayment of pluvial Owens Lake was present as scattered lag and remnant tufa heads projecting was open to the northwest and contained at least two islands, as ϳ20 cm above the soil surface (Fig. 3D). well as beach cliffs or erosional escarpments distinguished by Downloaded from specialpapers.gsapubs.org on July 1, 2015

194 Jayko and Bacon trimlines, notches, and abrasion surfaces formed on bedrock. The with down-to-the-west displacement, is located west of the pre- widespread accumulation of sand and silty sediment characteris- served embayment (Figs. 1 and 2). Relative uplift on the east side tic of both beach and marsh or wetlands deposits in the embay- of the Centennial Flat fault apparently preserved the lacustrine ment means they likely formed in a bay head and (or) bay-side record within the embayment. The LGM and younger beach beach depositional environment similar to that described by deposits are also well preserved on the upthrown side of the Cen- Adams and Wesnousky (1998) for the Jessup embayment of Lake tennial Flat fault, whereas to the west of the fault, they are mostly Lahontan in Nevada. The deposition of clastic sediment and buried by alluvium or eroded. development of erosional features within the embayment likely was influenced by large wave energy and an unobstructed fetch Swansea Pocket Barrier directed from the north. Since the development of the Centennial Flat embayment, The best example of a beach deposit associated with the younger LGM beach ridges and barriers associated with the 1180 m shoreline is located within the foothills of the southern ϳ1160 m water level formed below cliffed headlands and trun- Inyo Mountains near the historic site of Swansea, and it is re- cated the older embayment deposits. The Centennial Flat fault, ferred to as the Swansea pocket barrier (Figs. 1 and 5). At this which is a northwest-trending and possibly oblique-dextral fault site, a remnant gravel deposit (Qg1) occurs on a gently sloping

Figure 5. Map showing 1180 m beach deposit and associated shoreline features, including wave-cut notch and trimlines, at the Swansea pocket barrier (see Fig. 1 for location). LGM—Last Glacial Maximum. Downloaded from specialpapers.gsapubs.org on July 1, 2015

Late Quaternary MIS 6–8 shoreline features of pluvial Owens Lake 195 abrasion surface at ϳ1177 m between two bedrock knobs com- nated with 2–4-mm-wide secondary silica, which partly fills and posed of Paleozoic dolomite (site 20; Table 1; Figs. 5 and 6A). cements fractured dolomite bedrock. This pinkish tufa contrasts Well-rounded, discoidal clasts lie on a thin veneer of ϳ2–4-cm- with the whitish or grayish white pedogenic carbonate observed wide fragmented and angular tufa at this site (site 20; Table 1; elsewhere in the valley. Most of the abrasion surfaces are denuded Figs. 5 and 6B). The tufa appears to have been brecciated in situ of beach gravel and associated tufa, but a few rare, well-rounded by postdepositional mechanical weathering. The Qg1 deposit beach cobbles and pebbles were found on bedrock surfaces to the also rests on bedrock benches (Qab) that terminate against wave- east (sites 21, 22; Table 1; Fig. 5). cut notches (e.g., site 17; Table 1; Figs. 5 and 6C). A trimline ad- The Qg1 deposit and associated Qab surfaces are locally jacent to a wave-formed notch locally separates a hillslope with buried by undifferentiated post–15 cal ka alluvial sediment and rough microtopography and abundant colluvium from a notice- colluvium (Qa2), and they are dissected by active alluvial washes ably much smoother and more sparsely colluviated hillslope (Qa3) along the southeastern edge of the deposit. Bedrock cliffs below (Figs. 5, 6A, and 6C). truncate the western edge of the deposit. The latest Quaternary Pinkish tufa covers several abrasion surfaces on bedrock (Ͻ15 cal ka) lake plain and beach ridges (Carver, 1970; Bacon slopes south of the Qg1 deposit between 1177 and 1179 m eleva- et al., 2006; Orme and Orme, this volume) lie ϳ40 m below the tion (sites 17–19, 23; Table 1; Figs. 5 and 6A). The tufa is lami- cliffs at elevations below ϳ1140 m.

Figure 6. Photographs of the Swansea pocket barrier area. (A) Location of extensive beach cobble lag in association with weathered charophytic tufa; view to north-northeast. Scale: Area outlined is around site 20, Figure 5, unit Qg1. Outlined area is smaller than full mapped extent to wave-cut notch in Figure 5 and is about 50 m ϫ 50 m. (B) Very thin, very well-rounded lag of beach cobbles surrounded by angular tufa that is brecciated by mechani- cal weathering. (C) Well-developed wave-cut notch separating smoother from rougher microtopography upslope from the notch; view to west. Downloaded from specialpapers.gsapubs.org on July 1, 2015

196 Jayko and Bacon

Lacustrine Setting between these creeks, as well as near Aberdeen, ranges in eleva- The morphologic setting of the remnant gravel deposit and tion from at least 1170 m south of the Tinemaha Narrows to at associated wave-formed abrasion features at this locality appear least 1160 m near Independence west of the Owens Valley fault to be similar to relatively protected pocket barriers formed within zone within the active meander belt of the Owens River (Qfl) pluvial lake systems (e.g., Russell, 1885; Adams and Wesnousky, (Fig. 1). Erosional shoreline features lie above the valley floor at 1998). The Swansea pocket barrier was open to the southwest ϳ1180 m east of the Owens Valley fault zone along the Inyo through a saddle of bedrock (Fig. 5) and appears to have formed Mountains, whereas erosional shoreline and fluvial features are the western margin of a relatively steep-sided and deep embay- found within the southern Big Pine volcanic field at ϳ1185– ment that is now filled with latest Quaternary alluvial fans and 1200 m west of the fault zone (sites 27–41; Table 1; Fig. 7). colluvial slopes. Volcanic rocks within the southern and central parts of the Big Pine volcanic field provide age control for the associated Sites near the Big Pine Volcanic Field, lacustrine and fluvial geomorphic features (Fig. 1). Basalt lava Northern Owens Valley flows in the volcanic field have yielded K/Ar and 40Ar-39Ar dates that range from 1.2 Ma to 280 ka (for Qv1) and an age of 130 Ϯ Lacustrine and fluvial features observed in the northern part 50 ka for Qv3, whereas the intermediate unit Qv2 is undated (Cox of the study area include an extensive upland area dominated by et al., 1963; Gillespie, 1982; Gillespie et al., 1983, 1984; Turrin fluvial strath surfaces that are overlain by lag cobbles and sandy and Gillespie, 1986, see compiled ages in Connor and Conway, deposits near Taboose Creek. This upland area is succeeded to the 2000; Dorn et al., 1987; Martel et al., 1987) (Fig. 7). The relative south by a remnant wave-formed shoreline platform and over- degree of weathering and surface morphology on the Qv1 and lying sandy beach deposits near Sawmill Creek. The valley floor Qv2 lava flows distinguish them from the younger Qv3 lava flow.

Figure 7. Map showing location of shoreline features at the north end of the study area near Aberdeen, the Tinemaha Nar- rows, and the Big Pine volcanic field (see Fig. 1 for location). LGM—Last Glacial Maximum. Downloaded from specialpapers.gsapubs.org on July 1, 2015

Late Quaternary MIS 6–8 shoreline features of pluvial Owens Lake 197

The Qv2 lava flows exhibit relatively smooth surfaces that are the east, and is dissected by well-developed, 3- to 4-m-deep chan- moderately dissected (Figs. 8A and 9A), whereas the younger nels that trend east toward the Owens River (sites 31–34; Figs. 9A 180–80 ka (Qv3) flow (Dorn et al., 1987) tends to have a rough and 9B). North of the lower surface, there is a ϳ3-m-high south- and blocky surface with irregular meter-scale microtopography, a facing slope resembling a terrace riser that separates the lower and much fresher, darker appearance, and a thinner accumulation of upper surfaces. The upper surface lies at ϳ1200 m and exhibits overlying eolian sand and silt. We use the dated basalt flows to westward-trending channels as much as 6–8 m deep (sites 30–34; help constrain the relative age of inset strath terraces and wave- Table 1; Figs. 7, 9A, and 9B). The erosional surfaces are ϳ10–15 m formed platforms in the following sections. higher than Quaternary alluvium (Qa) to the south on active fan surfaces emanating from the Sierra Nevada. The relatively sub- Taboose Creek Strath Terraces dued ridges or knobs appear to be the higher portions of arcuate Two distinct erosional surfaces are developed north of pressure ridges of the flow surface that have been beveled to Taboose Creek along the southern margin of the Red Mountain nearly flat benches. These pressure ridges are evident on aerial lava flow west of the Owens Valley fault zone (Qv1; ca. 340 Ϯ photographs and satellite imagery, and they become progres- 60 ka; Turrin and Gillespie, 1986, see map Figure 7 in Connor and sively less preserved and more eroded to the south (Fig. 8A). Conway, 2000; Dorn et al., 1987) (Fig. 7). The lower erosional Isolated ridges on the 1195–1197 m surface exhibit a well- surface ranges in elevation from 1195 to 1197 m, slopes gently to developed pavement composed of weathered basalt breccia. The

Figure 8. Site localities and erosional surfaces developed on ca. 340 ka basalt Qv1 east of the Tinemaha Narrows. (A) Portion of digital orthophoto quad showing location of sites described in text and Table 1. (B) Photo showing locations of fluvial strath terraces developed on late Pleistocene basalt flows near Tinemaha Narrows. (C) Photo showing isolated subrounded and rounded clasts of a fluvial lag deposit, as well as angular basalt brecciated clasts on a strath terrace. The brecciation and fragmental character of the basalt are secondary and from weathering of more massive basalt flows. Downloaded from specialpapers.gsapubs.org on July 1, 2015

198 Jayko and Bacon

Figure 9. Site localities and abrasion surfaces developed on basalt younger than 300 ka (Qv2) north of Sawmill Creek. (A) Portion of digital orthophoto quad showing location of sites described in text and Table 1. (B) Photo showing locations of wave-formed surface developed on late Pleistocene basalt flows north of Sawmill Creek; photo taken standing at site 41 and looking toward sites 37 and 38. (C) Photo showing very well- rounded, 3–5 mm pebbles of basalt in beach lag deposit on a wave-formed surface at site 41. The brecciation and fragmental character of the other generally coarser, angular basalt clasts are secondary and from weathering of larger basalt clasts.

brecciated basalt clasts form a patchy mosaic of roughly equant- absent from these surfaces, suggesting the gravel may have been shaped 20–40-cm-wide fragments lying between larger basaltic deposited by local streams, such as Taboose Creek, flowing east outcrops. The breccia clasts were formed in situ from weathering from the Sierra Nevada rather than by the Owens River. and disaggregation of the basalt flows and have not been trans- Relative Age Constraints. The strath terraces were formed ported significant distances. There are well-rounded cobbles and during times when sedimentation rates were relatively low com- boulders, usually observed as isolated clasts rather than clusters, pared to vertical uplift rates. These straths, cut on the ca. 340 Ϯ which are coated by an iron oxide stain, above and slightly 60 ka flows of Red Mountain, have subsequently been incised by downslope of the erosion surfaces (site 34; Figs. 8A and 8C). channels. The degree of weathering on basaltic flow surfaces Preservation of the iron oxide coating implies that the clasts are described earlier and the development of weathering rinds on reworked out of an older, local deposit, and have not been trans- gravel and cobble clasts support an age older than ca. 25 cal ka. ported a great distance. The lithologies of observed clasts are gran- Furthermore, LGM outwash and younger alluvial fans of the steep odioritic and metasedimentary, which are characteristic of Sierra piedmont that lie east and south (?) of the site are graded lower Nevada plutonic rocks and roof pendants. Clasts composed of than the terraces, providing additional evidence for an older age. Bishop Tuff lithology are ubiquitous in middle and late Quater- Depositional Setting. The two planar and gently east- and nary Owens River and Owens Lake deposits. This lithology is west-sloping erosional surfaces are developed across the southern Downloaded from specialpapers.gsapubs.org on July 1, 2015

Late Quaternary MIS 6–8 shoreline features of pluvial Owens Lake 199

Red Mountain lava flow. The occurrence of erratic cobbles and (sites 35 and 36; Table 1; Figs. 1 and 7). Compared to the surface gravel to small boulders in conjunction with these surfaces indi- of the 340 Ϯ 60 ka (Qv1) Red Mountain lava flow (Dorn et al., cates that they formed by fluvial rather than lacustrine processes. 1987), the Qv2 flow appears to be younger, has pressure ridges that Therefore, we interpret these erosion surfaces as fluvial terraces. are moderately to well-preserved, and is not as dissected by alluvial The fluvial terraces are on the upthrown block of the Owens Val- channels. Compared to the surface of the ca. 130 Ϯ 50 ka (Qv3) ley fault zone and are ϳ1.5 km west of the mapped trace (Fig. 7). lava flow to the north (Dorn et al., 1987), the Qv2 flow is more Near these terraces, the Owens Valley fault zone steps to the east weathered, has more muted, meter-scale microtopography, and is and changes to a more northerly strike, with an implied increase more dissected. Therefore, based on relative degrees of surface in the vertical component of slip. It appears that the terraces were weathering characteristics, Qv2 is intermediate in age (Fig. 8A). initially cut during two different phases and were subsequently Depositional Setting. The numerous breaks in slope at the incised by a later cutting phase that formed the channels, which intersection with flat platforms along the southern margin of the likely were influenced by relative uplift associated with the lava flow, combined with a deposit we interpret as beach pebbles Owens Valley fault zone. As a result, these planar surfaces eroded and sand, suggest that these features are wave-formed notches or on lava flows are interpreted as strath terraces (e.g., Bull, 1991). angles. Development of these features likely was influenced by These strath terraces are graded to the south, to the Sawmill Creek large wave energy and unobstructed fetch directed from the shoreline, and are used as a datum to constrain the elevation of a south-southeast. Although the evidence is not robust for widely lake level below ϳ1195–1197 m. distributed beach deposits associated with the ϳ1192–1193 m platforms in the area, we interpret these features as a shoreline that Sawmill Creek Shoreline appears to be older than adjacent post-LGM alluvium. A planar and flat erosional surface on the southern margin of the Qv2 flow is locally exposed under a thin and discontinuous Alabama Hills Strath Terraces veneer of alluvium between the elevations of 1192 and 1193 m (sites 37–41; Table 1; Figs. 7 and 9A). In this area, erosional features In the vicinity of Lone Pine, there are several planar erosional are characterized as isolated benches in the form of flat platforms, features formed on indurated Mesozoic metasedimentary rocks as well as more continuous platforms that intersect low-relief between the elevations of 1200 and 1204 m along the eastern knobs with significant slope breaks that appear as notches (sites Alabama Hills range front (Fig. 1). These erosional features are 37 and 38; Fig. 9B). At one site, a relatively small lag (ϳ40 m2) discontinuous and occur in the form of bedrock benches and of very well-rounded, roughly equant, 3–5-mm-sized basalt peb- notches. In addition, a relatively flat and well-exposed fluvial cut- bles is preserved against a well-developed wave-formed notch and-fill terrace on bedrock was observed at the mouth of Lone (site 41; Table 1; Figs. 9B and 9C). The well-rounded, size-sorted, Pine Creek (site 42; Table 1; Fig. 1). equant nature of the basalt pebbles, the close proximity to their The terrace cut on bedrock is overlain by a fill of clast-sup- source rock, and the absence of detrital clasts of the Bishop Tuff, ported, well-rounded, boulder and cobble conglomerate at its indicative of an Owens River fluvial deposit, suggest that the northern edge along the Whitney Portal road west of Lone Pine basaltic pebbles could have been deposited in a beach (shoreline) (Fig. 10A). The relatively fresh and loose conglomerate was ini- environment. tially mapped by Blair (2001) and later by Benn et al. (2006) as A Ͼ1-m-thick deposit of well-rounded and moderately possible evidence for catastrophic outburst floods from failure of sorted coarse pebbly sand underlies the well-rounded basaltic lag moraine-damned lakes within the steep eastern range front of the pebbles. This sand exhibits considerable rubification (7.5YR 5/6) Sierra Nevada. The same deposit is mapped and dated nearby as but has a weakly developed soil, indicated by a lack of soil struc- outwash/alluvium by Bierman et al. (1995), who reported cosmo- ture, that appears to be currently influenced by active bioturbation genic 10Be surface ages from boulders that were later recalculated from rodents. The well-rounded and weathered clasts of this with up-to-date production rates by Bacon and Pezzopane (2007) deposit contrast with very light-colored, arkosic, fresh sandy and to between ca. 26 and 11 ka. These 10Be surface ages are cor- pebbly alluvium on the alluvial-fan surface only 2–3 m lower. roborated by additional age constraints of the same fan between Based on the small fraction of arkosic sediment within the beach ca. 25 and 12 cal ka by Lubetkin and Clark (1988) and Beanland deposit, it appears that there has been little recent eolian or allu- and Clark (1994), based on relative weathering characteristics vial contribution at the site. and crosscutting relations of 14C-dated lithoid tufa deposits north Age Constraints. The beach deposits overlie a bench or abra- of Lone Pine Creek. In contrast, further upstream, the cut terrace sion platform that lies along the southern margin of an undated lava appears to be overlain by a sequence of interbedded gravelly flow (Qv2) west of the Owens Valley fault zone (Fig. 7). This plat- channel bed deposits and massive sandy and gravelly to bouldery form is well developed along the southern margin of the Qv2 flow deposits that are reddened or deeply weathered (Blair, 2001). The but is not discernible toward the axis of the flow, except at the north- weathered deposits are in sharp contact with and overlain by the ern flow margin, where a portion of the adjacent Qv1 lava flow is loose and fresh alluvium dated as old as ca. 26 ka. exposed (Fig. 7). Notches interpreted as either fluvial or shoreline This terrace is interpreted as a fluvial strath terrace be- features at ϳ1185–1186 also occur on the Qv1 flow to the north cause it is cut into indurated bedrock and it is located on the Downloaded from specialpapers.gsapubs.org on July 1, 2015

Figure 10. Site localities in southern Owens Valley. (A) Photo showing strath terrace at ϳ1200 m elevation developed on Mesozoic bedrock. Strath is overlain by grayish weathered Last Glacial Maximum (LGM) glacial outwash at this locality and by both the LGM and older reddish weathered outwash deposit upstream. (B) Photo showing wave-formed notches at 1160, 1180, and ϳ1300 m elevation on granitic bedrock headlands of a promontory called Haystack. (C) View looking east toward Coso Range from ϳ1182 m elevation bench west of Sage Flat. Smooth bench at 1182 m is succeeded by a very gen- tle shoreline angle and trimline that separates the eroded surface from the bouldery slope. (D) View to the north with Sierra Nevada on the left showing benches at 1182 m elevation. (E) Photo showing tightly packed, clast-supported cobble-boulder deposit (interpreted as a possible berm) adjacent to the benches at ϳ1177 m elevation. Field notebook (ϳ10 cm ϫ 17 cm) for scale near the center of the photo. Downloaded from specialpapers.gsapubs.org on July 1, 2015

Late Quaternary MIS 6–8 shoreline features of pluvial Owens Lake 201 upthrown block of the Owens Valley fault zone with a mapped lent exposures of tightly packed, clast-supported cobble and trace ϳ100 m to the east (Fig. 1). The surface of the strath ter- boulder deposits. These clast-supported deposits may represent race projects out to the east above the valley floor and ϳ40– boulder berms that were formed by reworking of the older fan 44 m above the LGM shoreline at ϳ1160 m. The Lone Pine material along a shoreline (site 44, Table 1; Fig. 10E). If so, these Creek appears to have subsequently cut a channel in bedrock gravelly, clast-supported deposits are the best example of clastic through the strath terrace down to ϳ1160 m, which may reflect sediment associated with the 1180 m shoreline on the west side base-level changes of pluvial Owens Lake as it dropped from of the valley. ϳ1160 m to ϳ1145 m between ca. 25 and 15.5 cal ka. Because the strath is locally overlain by very weathered deposits, the Abrasion Platforms—Inyo Range Front strath terrace must have been cut well before deposition of the Remnant erosional landforms that appear to be abrasion younger ca. 25–12 ka gravels. We interpret the old, pre-LGM platforms formed by wave erosion are cut on bedrock along the strath terrace and other benches along the range front of the lower range front of the Inyo Mountains between Independence Alabama Hills as evidence of a nearby shoreline and adjacent and just south of Tinemaha Narrows (sites 27–29; Table 1; Figs. 1 streams entering a lake near Lone Pine that had a maximum ele- and 7). These relict abrasion surfaces gently slope basinward and vation of ϳ1200 m. This elevation is similar to the elevation of occur on relatively narrow bedrock noses and isolated knobs that features documented at the southern Big Pine volcanic field on protrude from the range front at elevations between ϳ1180 and the west side of the Owens Valley fault zone. 1184 m (sites 26–29; Table 1; Figs. 1 and 11). Most of these sur- faces are devoid of surficial sediment and occur on Paleozoic Erosional Wave-Formed Features in Owens Valley metasedimentary or Mesozoic granitic bedrock. These landforms terminate toward the range front at distinct breaks in slope as Several erosional shoreline features are recorded at localities relatively well-preserved, wave-formed notches (sites 27 and 29; along the range front of the Inyo Mountains at ϳ1180 m in eleva- Fig. 11), or at saddles between the steeper slopes of the range front tion and along the Sierra Nevada at ϳ1200 m (sites 16, 25, and (sites 26 and 28; Fig. 11). The morphologic setting of the shore- 43; Table 1; Fig. 1). These sites preserve wave-formed features line platforms indicates that they were developed on headlands consisting of abrasion platforms and notches in resistant bedrock, and islands similar to shoreline features in Mono Lake basin (e.g., and truncated landforms composed of unconsolidated Quaternary Russell, 1889; Reheis et al., 2002a) and Lake Lahontan basin and Tertiary alluvial deposits. (e.g., Russell, 1885; Adams and Wesnousky, 1998). Development of these features likely was influenced by a large fetch and wave Bedrock Notches energy directed from the south. Notches cut on quartz monzonite are located at Haystack (Kern Knob) along the Inyo Mountains directly east of Lone Truncated Landforms Pine (site 24; Table 1; Figs. 1 and 10B). Three distinct notches South of the Swansea pocket barrier locality, along the range occur at this site at elevations of 1160 and 1180 m, as well as a front of the southern Inyo Mountains, there is a relict and heavily higher one at ϳ1300 m (Fig. 10B). During lake levels above dissected alluvial-fan complex composed of late Tertiary (6.7– ϳ1120 m, the area of Haystack formed a headland that pro- 4.3 Ma) alluvium and interbedded volcanic tuff (Stone et al., 2004) truded out from the range front. South of Lone Pine along the (site 16; Table 1; Figs. 1 and 12A). The base of the fan is abruptly Sierra Nevada range front at Cottonwood Creek, there is a rem- truncated, forming a nearly linear slope break with distinct facets nant bench on bedrock resembling a wave-formed platform at at ϳ1180 m elevation (Fig. 12A). We infer that the slope break rep- ϳ1199 m elevation (site 43; Table 1; Fig. 1). There is a large resents a relict shoreline angle formed by the 1180 m shoreline. gully (as much as 20 m deep) below this bench that exposes a An additional landform that exhibits truncation at ϳ1180 m sequence of strongly indurated fanglomerate overlain by a suc- is the North Long John Canyon slide complex, which is located cession of weathered alluvium. along the Inyo Mountains range front east of Lone Pine (Blair, 1999) (site 26; Table 1; Fig. 1). We infer that the abrupt truncation Sage Flat Benches of this landform represents a shoreline angle, as it is similar in Erosional features in the form of benches occur at ϳ1182 m elevation and morphology to other features described along the elevation. They are inset within old alluvial-fan deposits east of eastern and southern margins of the valley. Blair (1999) inter- Sage Flat in southern Owens Valley (site 43, Table 1; Figs. 1 and preted the ϳ25 million m3 slide complex as early Holocene (?) 10C). The benches are relatively smooth and terminate against a age. However, based on a moderately developed soil that in- trimline that forms what appears to be a gentle to moderate shore- cludes a Bt horizon (Blair, 1999), rubification, as well as a muted line angle (Fig. 10D). The alluvial-fan deposits that underlie the surface morphology, we interpret this landslide as older than benches are locally exposed in railroad cuts and are composed LGM and possibly equivalent in age to the ϳMIS 6–aged glacial mostly of debris-flow deposits consisting of matrix-supported outwash/ alluvial fans mapped in the Fish Springs area by Zehfuss sands and silts with minor gravel. In addition, there are less preva- et al. (2001). Downloaded from specialpapers.gsapubs.org on July 1, 2015

202 Jayko and Bacon

Figure 11. Photos of localities where noticeable platforms are developed on bedded, inclined Paleozoic metasedimentary rock at 1180 m elevation east of the Owens Valley fault zone along the base of the Inyo Range. We interpret the platforms to be abrasion or wave-eroded surfaces. No relict or lag clasts of beach material were observed at the sites along the Inyo Range. (A) Surface on Paleozoic bedrock hill in Owens Valley near Mazourka Canyon road. (B) Surface at ϳ1184 m on bedrock at site 27. (C) Surface at ϳ1181 m at site 28. (D) Surface at ϳ1183 m on bedrock at site 29.

DISCUSSION sure of relative age. Soils and weathered clasts found within sedi- ment deposited by the 1180 m shoreline at the Centennial Flat Age of the 1180 m Shoreline embayment and sediment deposited prior to, or coeval with, the shoreline at Taboose and Sawmill Creeks, at the Alabama Hills The age of the pluvial Owens Lake that formed the ϳ1180 m strath terraces, and on the North Long John Canyon slide complex shoreline is constrained to be younger than the 340 Ϯ 60 ka Red appear to have similar weathering characteristics. These deposits Mountain flow and older than the 130 Ϯ 50 ka basaltic lava flows also have soil development similar to that described on alluvial of the Big Pine volcanic field, based on crosscutting relations of fans south of Big Pine at Fish Springs, where cosmogenic 10Be the Sawmill Creek shoreline, Taboose Creek strath surfaces, and and 26Al dates on boulders yielded ages that range from 171 Ϯ the lack of features on the younger lava flow. Thus, the shoreline 21 ka to 136 Ϯ 17 ka (Zehfuss et al., 2001). developed sometime between MIS 8 and 6, which occurred dur- An additional age estimate for the 1180 m shoreline may be ing the intervals of 280–260 ka and 185–130 ka, respectively determined by dividing the amount of measured displacement (Winograd et al., 2006). The 36Cl model age of ca. 160 Ϯ 32 ka between shoreline features across the Owens Valley fault zone by (CHLOE production parameters; Phillips et al., 2001) or 218 Ϯ the average vertical slip rate of ϳ0.12 m/k.y. on the Owens Val- 33 ka (calcium spallation rate of Stone et al., 1996) also supports ley fault zone at Lone Pine (from Bacon and Pezzopane, 2007) the MIS 8–6 age as determined from crosscutting relations to solve for time. From this calculation, the time necessary to between dated lava flows. accommodate ϳ20 Ϯ 4 m of apparent vertical deformation across The degree of soil development in the form of rubification, the Owens Valley fault zone ranges from ϳ200 to 130 k.y. This presence of pedogenic clay, and weathering rinds provide a mea- independent age estimate of the 1180 m shoreline is within the Downloaded from specialpapers.gsapubs.org on July 1, 2015

Late Quaternary MIS 6–8 shoreline features of pluvial Owens Lake 203

area shown on Figure 7, Gillespie (1982) and Gillespie et al. (1984) constrained the older Tahoe moraines to between ca. 460 and 130 ka using K/Ar dates on interbedded basalt flows. Reli- able age determinations for moraines older than MIS 6 have yet to be reported from the southern Sierra Nevada (Kaufman et al., 2004). Based on available age control, it is likely the 1180 m shoreline was coeval with the older Tahoe glaciation in the south- ern Sierra Nevada, which from the best available age estimates is equivalent to MIS 6.

Owens Lake Core OL-92 during MIS 6–8

Depths to MIS 6–8 horizons in the Owens Lake core OL-92 have been identified by paleoclimate proxies including lithologic, chemical, mineralogic, geophysical, and paleontological data (Smith and Bischoff, 1997). The cycling between closed- and open-lake systems defined both by paleoecology and mineralogy of lacustrine sediment refines the duration and magnitude of the cycles between interglaciations and glaciations. The Owens Lake core reached a depth of 323 m, penetrating the Bishop tuff (ca. 760 ka) at ϳ304–309 m depth, and it terminated near the Brunhes- Matayama magnetic reversal (ca. 783 ka) at ϳ310–320 m depth (Smith and Bischoff, 1997). Sediment correlated with MIS 6 ranges from ϳ80–92 m depth, whereas sediment corresponding to MIS 8 ranges from ϳ111–131 m depth. The absence of fish remains in the core between ϳ96 and 112 m depth combined with the dominance of saline-water diatom assemblages in that inter- val denote MIS 7. During the interval between MIS 6 and MIS 8, ostracode and diatom proxies indicate that there was a stronger Figure 12. (A) Aerial photograph showing abrupt truncation of Tertiary and longer duration of cold and freshwater during MIS 6 than dur- alluvial fans along the southern Inyo Range at ϳ1180 m elevation (site 16; ing MIS 8 (Bradbury, 1997; Carter, 1997). Fig. 1). (B) Aerial photograph showing truncation of the Long John Can- In addition to biologic proxy indicators, the total inorganic ϳ yon fan slide of Blair (1999) along the southern Inyo Range at 1180 m content (TIC) and cation exchange capacity (CEC) of the clay elevation (site 25; Fig. 1). fraction in core OL-92 shows abrupt increases during closed- basin lake conditions, which represent an interglacial cycle (Bischoff et al., 1997a, 1997b). Based on this proxy data set in range of constraining ages from dated lava flows, and it is simi- combination with depth-age estimates in the core, full glacial con- lar to the 36Cl model age on tufa. ditions in Owens Valley began at ca. 190 ka (MIS 6/7) and ended at ca. 120 ka (MIS 5/6) (Bischoff et al., 1997a). Other proxy indi- Correlation to Sierran Glacial Chronologies cators in core OL-92, such as the presence of rock flour (a glacial erosional by-product) and salinity, support the TIC and CEC vari- The penultimate glaciation in the Sierra Nevada, referred to ations. These additional proxies also show that between ca. 260 as the Tahoe glaciation, had two major advances, which are re- and 220 ka, there were three short-duration cold intervals, with ferred to as the younger Tahoe (Tahoe II) and older Tahoe (Tahoe I) the coldest around 220 ka (Bischoff et al., 1997c; Menking 1997), (e.g., Kaufman et al., 2004). The younger Tahoe advance is esti- which is also strongly expressed in the Devils Hole and sea-surface mated at ca. 50–42 ka from 36Cl cosmogenic exposure ages on temperature paleoclimate records (Winograd et al., 2006). boulders from moraine crests at Bloody Canyon in Mono Basin (Phillips et al., 2001; Kaufman et al., 2004; corrected from Importance for Aquatic Species ca. 66–55 ka reported in Phillips et al., 1990). Boulders on older Tahoe moraines range between 218 and 133 ka in age (Phillips The most recent overflow of the Owens River system via et al., 1990; Kaufman et al., 2004) (Fig. 1). Along the range to Searles and Panamint Valleys to its terminus in Death Valley the south, the younger Tahoe is constrained between 119 ka and probably last occurred during MIS 6 or older time (Jayko et al., 53 ka and is therefore most likely MIS 4 in age (Gillespie, 1982; this volume). U-series tufa dates from the highstand Blackwelder Gillespie et al., 1984). Near Sawmill Canyon, located west of the shoreline in Death Valley range in age from 216 to 120 ka (Hooke Downloaded from specialpapers.gsapubs.org on July 1, 2015

204 Jayko and Bacon and Dorn, 1992; Ku et al., 1998, Hooke, 1999). Sediment core tors in core OL-92 and the timing of the older Tahoe (Tahoe I) data from near Badwater in Death Valley indicate a large lake glaciation in the southern Sierra Nevada. This pluvial Owens stand at MIS 6 and a smaller lake at MIS 2 but mainly ground- Lake was much larger and deeper than preceding younger lakes water-supported wetlands during MIS 4 (Lowenstein et al., 1999; in Owens Valley. It had a minimum area of at least ϳ880 km2, Forester et al., 2005). Results from Pleistocene pluvial studies at based on the map distribution of described features in Owens Val- Mono Basin indicate that Lake Russell may have spilled south- ley, but the exact location of the spillway has yet to be resolved, ward through Adobe Valley to Owens River during MIS 6 (Put- so the full extent of the lake is unknown. Such a large, fresh lake nam, 1947, personal commun. to Hubbs, in Hubbs and Miller, during MIS 6 is consistent with environmental interpretations of 1948) and (or) MIS 4 (Reheis et al., 2002a). Thus, during MIS 6 faunal and mineralogic indicators identified in core OL-92 (Brad- (older Tahoe), there may have been a connection from Mono Lake bury, 1997; Carter, 1997; Bischoff et al., 1997c; Menking, 1997). Basin to Death Valley, as initially speculated by Blackwelder Landscape evolution and rates of geomorphic processes can (1933). Following MIS 6 time, water from the Owens River sys- be quantified where there is sufficient age control on important tem made it no farther than Searles Valley during MIS 4 (Jayko landscape-modifying events, such as lake-level fluctuations et al., this volume) and as far downstream as Panamint Valley dur- within large lake systems during glacial-pluvial cycles. Recogni- ing MIS 2 (Smith et al., 1983; Jayko et al., this volume). Fish in tion and documentation of climate-induced landscape change the Owens River system have clearly been isolated from the other provide a baseline for studies that constrain other processes such southern and eastern populations in the Mojave and Amargosa as erosion, fault-slip rates, and water balance within a region. This drainage basins for at least ϳ120 k.y., since the waning of MIS 6 study has documented the distribution of shoreline and fluvial (Jayko et al., this volume). features that outline the highstand of a large ancestral pluvial lake Previous paleontologic and geomorphic work has shown that of probable MIS 6 age that had not previously been documented. the Owens River system was linked to the Lahontan watershed in As with many large ancestral pluvial lakes in the western United the north (Russell, 1889; Miller, 1946; Hubbs and Miller, 1948; States, which are typically named after pioneers of their field, we Smith et al., 2002; Reheis et al., 2002a). Fish remains including propose to name the MIS 6 pluvial Owens Lake as Lake GI, in Catostomus (sucker) and Siphateles (tui chub) from OL-92 and recognition of George I. Smith’s investigations in Owens Valley, older sediments in the region indicate that the Lahontan fluvial as well as other pluvial lakes within the Owens River system, over connection is older than the Bishop Tuff, dated at ca. 760 ka the past 50 yr. (Firby et al., 1997; Miller, 1946; Hubbs and Miller, 1948). Also, Prosopium (whitefish) and Oncorrhynchus (trout) at the base of ACKNOWLEDGMENTS OL-92 core between ca. 730 and 695 ka suggest colder conditions at this time (Firby et al., 1997). These cold-water Salmonidae We owe a special thanks to Fred Phillips for graciously pro- (trout) species clearly persisted following the eruption of the viding the 36Cl date from site 10 at 1170 m elevation, pro bono, Long Valley caldera (Firby et al., 1997), which deposited the when he learned about the results of this study. We also thank Alan Bishop Tuff, but they apparently could not survive subsequent Gillespie for directing us to published sources of some of the data environmental and (or) tectonic changes, which may have so often cited elsewhere in the Turrin and Gillespie (1986) included eruptive activity in the Big Pine volcanic field or disrup- abstract. We thank Wally Woolfenden, U.S. Forest Service; Alan tion of the southward drainage from the Mono Basin watershed Gillespie, University of Washington; and Scott Stine, California (cf. Russell, 1889). State University–Hayward for constructive and insightful reviews that helped strengthen the manuscript. We additionally thank Marith SUMMARY AND CONCLUSIONS Reheis for thorough editing and helpful comments. We acknowl- edge Silvio Pezzopane for engaging discussions on the 1180 m Deformed shoreline features and strath terraces graded to the shoreline in Owens Valley. S. Bacon acknowledges Eric McDonald, shorelines at elevations between ϳ1180 and 1200 m are younger Desert Research Institute, for partial support (Army Research Office than 340 Ϯ 60 ka and older than 130 Ϯ 50 ka, based on crosscut- Contract No. DAAD19-03-1-0159). We also gratefully acknowl- ting relations with dated lava flows of the Big Pine volcanic field. edge lively discussions with many Quaternary geologists who pio- Therefore, this phase of pluvial Owens Lake existed within the neered the modern lacustrine studies in the region during Pacific time interval between MIS 8 and MIS 6. Additional factors refine Cell–Friends of the Pleistocene field trips between 1996 and 2002. the age of the pluvial lake to approximately MIS 6 time. These include: (1) a 36Cl model age of 160 Ϯ 32 ka on tufa from south- REFERENCES CITED ern Owens Valley (F. Phillips, 2007, personal commun.; Table 2; Fig. 3E); and (2) an age based on a vertical slip rate on the Owens Adams, K.D., and Wesnousky, S.G., 1998, Shoreline processes and the age of the Valley fault zone at Lone Pine in combination with the apparent Lake Lahontan highstand in the Jessup embayment, Nevada: Geological Society of America Bulletin, v. 110, p. 1318–1332, doi: 10.1130/0016- vertical displacement of deformed shorelines across the Owens 7606(1998)110Ͻ1318:SPATAOϾ2.3.CO;2. Valley fault zone. Such a correlation is consistent with the rela- Bacon, S.N., and Pezzopane, S.K., 2007, A 25,000-year record of earthquakes on tive duration and intensity of MIS 6 as indicated by proxy indica- the Owens Valley fault near Lone Pine, California: Implications for recur- Downloaded from specialpapers.gsapubs.org on July 1, 2015

Late Quaternary MIS 6–8 shoreline features of pluvial Owens Lake 205

rence intervals, slip rates, and segmentation models: Geological Society of Carver, G.A., 1970, Quaternary Tectonism and Surface Faulting in Owens Lake America Bulletin, v. 119, p. 823–847, doi: 10.1130/B25879.1, doi: 10.1130/ Basin, California: Reno, University of Nevada, MacKay School of Mines B25879.1. Technical Report AT2, 103 p. Bacon, S.N., Jayko, A.S., and McGeehin, J.P., 2005, Holocene and latest Pleis- Connor, C.B., and Conway, F.M., 2000, Basaltic volcanic fields, in Sigurdsson, tocene oblique dextral faulting on the southern Inyo Mountains fault, Owens H., et al., eds., Encyclopedia of Volcanoes: San Diego, California, Acade- Lake basin, California: Bulletin of the Seismological Society of America, mic Press, p. 331–344. v. 95, p. 2472–2485, doi: 10.1785/0120040228. Cox, A., Doell, R.R., and Dalrymple, G.B., 1963, Geomagnetic polarity epochs— Bacon, S.N., Burke, R.M., Pezzopane, S.K., and Jayko, A.S., 2006, Last Glacial Sierra Nevada II: Science, v. 142, p. 382–385, doi: 10.1126/science.142. Maximum and Holocene lake levels of Owens Lake, eastern California, 3590.382-a. USA: Quaternary Science Reviews, v. 25, p. 1264–1282, doi: 10.1016/ Dorn, R.I., Turrin, B.D., Jull, A.J.T., Linick, T.W., and Donahue, D.J., 1987, j.quascirev.2005.10.014. Radiocarbon and cation-ratio ages for rock varnish on Tioga and Tahoe Beanland, S., and Clark, M.M., 1994, The Owens Valley Fault Zone, Eastern Cali- morainal boulders of Pine Creek, Nevada, California, and fornia, and Surface Rupture Associated with the 1872 Earthquake: U.S. their paleoclimatic implications: Quaternary Research, v. 28, p. 36–49, doi: Geological Survey Bulletin 1982, 29 p. 10.1016/0033-5894(87)90031-7. Benn, D.I., Owen, L.A., Finkel, R.C., and Clemmons, S., 2006, Pleistocene lake Duffield, W.A., and Smith, G.I., 1978a, Pleistocene history of volcanism and the outbursts floods and fan formation along the eastern Sierra Nevada, Cali- Owens River near Little Lake, California: U.S. Geological Survey Journal fornia: Implications for the interpretation of intermontane lacustrine of Research, v. 6, p. 395–408. records: Quaternary Science Reviews, v. 25, p. 2729–2748, doi: 10.1016/ Duffield, W.A., and Smith, G.I., 1978b, Pleistocene river erosion and intracanyon j.quascirev.2006.02.018. lava flows near Little Lake, Inyo County, California: California Geology, Benson, L.V., 1994, Carbonate deposition, Pyramid Lake subbasin, Nevada: 1. Se- v. 31, p. 81–89. quence of formation and altitudinal distribution of carbonate deposits (tufas): Firby, J.R., Sharpe, S.E., Whelan, J.F., Smith, G.R., and Spaulding, W.G., 1997, Palaeogeography, Palaeoclimatology, Palaeoecology, v. 109, p. 55–87, doi: Paleobiotic and isotopic analysis of mollusks, fish, and plants from core 10.1016/0031-0182(94)90118-X. OL-92, in Smith, G.I., and Bischoff, J.L., eds., An 800,000-Year Paleo- Bierman,P.R.,Gillespie,A.R.,andCaffee,M.W.,1995,Cosmogenicagesforearth- climatic Record from Core OL-92, Owens Lake, Southeast California: Geo- quake recurrence intervals and debris flow fan deposition, Owens Valley, logical Society of America Special Paper 317, p. 121–126. California: Science, v. 270, p. 447–450, doi: 10.1126/science.270.5235.447. Forester, R.M., Lowenstein, T.K., and Spencer, R.J., 2005, An ostracode based Bischoff, J.L., Fitts, J.P., and Fitzpatrick, J.A., 1997a, Responses of sediment geo- paleolimnologic and paleohydrologic history of Death Valley, 200 to 0 ka: chemistry to climate change in Owens Lake sediment: An 800-k.y. record of Geological Society of America Bulletin, v. 117, p. 1379–1386, doi: 10.1130/ saline/fresh cycles in core OL-92, in Smith, G.I., and Bischoff, J.L., eds., An B25637.1. 800,000-Year Paleoclimatic Record from Core OL-92, Owens Lake, South- Gale, H.S., 1914, Salines in the Owens, Searles, and Panamint Basins, Southeast- east California: Geological Society of America Special Paper 317, p. 37–48. ern California: U.S. Geological Survey Bulletin 580, p. 281–323. Bischoff, J.L., Stafford, T.W., Jr., and Rubin, M., 1997b, A time-depth scale for Gillespie, A.R., 1982, Quaternary Glaciation and Tectonism in the Southwestern Owens Lake sediments of core OL-92: Radiocarbon dates and constant Sierra Nevada, Inyo County, California [Ph.D. dissertation]: Pasadena, Cali- mass-accumulation: An 800-k.y. record of saline/fresh cycles in core OL-92, fornia Institute of Technology, 657 p. in Smith, G.I., and Bischoff, J.L., eds., An 800,000-Year Paleoclimatic Gillespie, A.R., Huneke, J.C., and Wasserburg, G.J., 1983, Eruption age of a Pleis- Record from Core OL-92, Owens Lake, Southeast California, Geological tocene basalt from 40Ar-39Ar analysis of partially degassed xenoliths: Jour- Society of America Special Paper 317, p. 91–98. nal of Geophysical Research, v. 88, p. 4997–5008. Bischoff, J.L., Menking, K.M., Fitts, J.P., and Fitzpatrick, J.A., 1997c, Climatic Gillespie, A.R., Huneke, J.C., and Wasserburg, G.J., 1984, Eruption age of a oscillations 10,000–155,000 yr B.P. at Owens Lake, California, reflected in ϳ100,000-year-old basalt from 40Ar-39Ar analysis of partially degassed glacial rock flour abundance and lake salinity in core OL-92: Quaternary xenoliths: Journal of Geophysical Research, v. 89, p. 1033–1048. Research, v. 48, p. 313–325, doi: 10.1006/qres.1997.1933. Hooke, R.L., 1999, (?) shorelines in the eastern , Cali- Blackwelder, E., 1933, Lake Manly: An extinct lake of Death Valley: Geographi- fornia: Quaternary Research, v. 52, p. 328–336, doi: 10.1006/qres. 1999.2080. cal Review, v. 23, p. 464–471, doi: 10.2307/209632. Hooke, R.L., and Dorn, R.I., 1992, Segmentation of alluvial fans in Death Valley, Blair, T.C., 1999, Alluvial fan and catchment initiation by rock avalanching, California; New insights from surface-exposure dating and laboratory mod- Owens Valley, California: Geomorphology, v. 28, p. 201–221, doi: 10.1016/ eling: Earth Surface Processes and Landforms, v. 17, p. 557–574, doi: S0169-555X(98)00111-1. 10.1002/esp.3290170603. Blair, T.C., 2001, Outburst flood sedimentation on the proglacial Tuttle Canyon Hubbs, C.L., and Miller, R.R., 1948, The Great Basin with emphasis on glacial alluvial fan, Owens Valley, California, U.S.A.: Journal of Sedimentological and postglacial times: Bulletin of the University of Utah, v. 38, p. 19–166. Research, v. 71, p. 657–679. Jannik, N.O., Phillips, F.M., Smith, G.I., and Elmore, D., 1991, A 36Cl chronol- Bradbury, J.P., 1997, A diatom-based paleohydrologic record of climate change ogy of lacustrine sedimentation in the Pleistocene Owens River system: for the past 800 k.y. from Owens Lake, California, in Smith, G.I., and Geological Society of America Bulletin, v. 103, p. 1146–1159, doi: 10.1130/ Bischoff, J.L., eds., An 800,000-Year Paleoclimatic Record from Core OL- 0016-7606(1991)103Ͻ1146:ACCOLSϾ2.3.CO;2. 92, Owens Lake, Southeast California: Geological Society of America Spe- Jayko, A.S., 2008, Surficial Geology of the Darwin Hills 30Јϫ60Ј Quadrangle, cial Paper 317, p. 99–112. Inyo County, California: U.S. Geological Survey Map Series, scale Bryant, W.A., 1984, Evidence of recent faulting along the Owens Valley, Round 1:100,000. Valley, and White Mountains frontal fault zones, Inyo and Mono Counties, Jayko, A.S., Forester, R.M., Kaufman, D.S., Phillips, F.M., Yount, J.C., McGee- California: California Division of Mines and Geology Open-File Report hin, J., and Mahan, S.A., 2008, this volume, Late Pleistocene lakes, OFR-84–54SAC, scale 1:24,000. Panamint Valley, California, in Reheis, M.C., Hershler, R., and Miller, Bull, W.B., 1991, Geomorphic Responses to Climatic Change: New York, Oxford D.M., eds., Late Cenozoic Drainage History of the Southwestern Great University Press, 326 p. Basin and Lower Colorado River Region: Geologic and Biotic Perspec- Carter, C., 1997, Ostracodes in Owens Lake core OL-92: Alternation of saline and tives: Geological Society of America Special Paper 439, doi: 10.1130/ freshwater forms through time, in Smith, G.I., and Bischoff, J.L., eds., 2008.2439(07). An 800,000-Year Paleoclimatic Record from Core OL-92, Owens Lake, Kaufman, D.S., Porter, S.C., and Gillespie, A.R., 2004, Quaternary alpine glacia- Southeast California: Geological Society of America Special Paper 317, tion in Alaska, the Pacific Northwest, Sierra Nevada, and Hawaii, in Gille- p. 113–120. spie, A.R., Porter, S.C., and Atwater, B.F., eds., The Quaternary Period in the Downloaded from specialpapers.gsapubs.org on July 1, 2015

206 Jayko and Bacon

United States: Elsevier, Netherlands, Developments in Quaternary Science, America Bulletin, v. 114, p. 991–1006, doi: 10.1130/0016-7606(2002)114 v. 1, p. 77–103. Ͻ0991:DRIMBDϾ2.0.CO;2. Ku, T.L., Luo, S., Lowenstein, T.K., Li, J., and Spencer, R.J., 1998, U-series Reheis, M.C., Sarna-Wojcicki, A.M., Reynolds, R.L., Repenning, C.A., and Mif- chronology of lacustrine deposits in Death Valley, California: Quaternary flin, M.D., 2002b, Pliocene to middle Pleistocene lakes in the western Great Research, v. 50, p. 261–275, doi: 10.1006/qres.1998.1995. Basin: Ages and connections, in Hershler, R., Madsen, D.B., and Currey, Lal, D., 1991, Cosmic-ray labeling of erosion surfaces; in situ production rates and D.R., eds., Great Basin Aquatic Systems History: Smithsonian Contribu- erosion models: Earth and Planetary Science Letters, v. 104, p. 424–439, tions to the Earth Sciences, v. 33, p. 53–108. doi: 10.1016/0012-821X(91)90220-C. Russell, I.C., 1885, Geological History of Lake Lahontan, a Quaternary Lake in Lowenstein, T.K., Li, J., Brown, C., Roberts, S.M., Ku, T.L., Luo, S., andYang, W., Northwestern, Nevada: U.S. Geological Survey Monograph 11, 288 p. 1999, 200 k.y. paleoclimate record from Death Valley core: Geology, Russell, I.C., 1889, Quaternary History of Mono Valley, California: Washington, v. 27, p. 3–6, doi: 10.1130/0091-7613(1999)027Ͻ0003:KYPRFDϾ2.3.CO;2. D.C., Eighth Annual Report of the U.S. Geological Survey, p. 267–394. Lubetkin, L.K.C., and Clark, M.M., 1988, Late Quaternary activity along the Lone Savage, J.C., and Lisowski, M., 1995, Strain accumulation in Owens Valley: Seis- Pine fault, eastern California: Geological Society of America Bulletin, v. 100, mological Society of America Bulletin, v. 85, p. 151–158. p. 755–766, doi: 10.1130/0016-7606(1988)100Ͻ0755:LQAATLϾ2.3.CO;2. Slemmons, D.B., Vittori, E., Jayko, A.S., Carver, G.A., and Bacon, S.N., 2008, Martel, S.J., Harrison, T.M., and Gillespie, A.R., 1987, Late Quaternary vertical Quaternary Fault and Lineament Map of Owens Valley, Inyo County, East- displacement rate across the Fish Springs fault, Owens Valley fault zone, ern California: Geological Society of America Map and Chart Series California: Quaternary Research, v. 27, p. 113–129, doi: 10.1016/0033-5894 MCH096, 25 p. (in press). (87)90071-8. Smith, G.I., and Bischoff, J.L., 1997, Core OL-92 from Owens Lake: Project ratio- Menking, K.M., 1997, Climatic signals in clay mineralogy and grain-size varia- nale, geologic setting, drilling procedures, and summary, in Smith, G.I., and tions in Owens Lake core OL-92, in Smith, G.I., and Bischoff, J.L., eds., An Bischoff, J.L., eds., An 800,000-Year Paleoclimatic Record from Core OL- 800,000-Year Paleoclimatic Record from Core OL-92, Owens Lake, South- 92, Owens Lake, Southeast California: Geological Society of America Spe- east California: Geological Society of America Special Paper 317, p. 25–36. cial Paper 317, p. 1–8. Miller, R.R., 1946, Correlation between fish distribution and Pleistocene hydrog- Smith, G.I., Barczak, V.J., Moulton, G.F., and Liddicoat, J.C., 1983, Core KM-3, raphy in eastern California and southwestern Nevada, with a map of the a Surface-to-Bedrock Record of Late Cenozoic Sedimentation in Searles Pleistocene waters: The Journal of Geology, v. 54, p. 43–53. Valley, California: U.S. Geological Survey Professional Paper 1256, p. 1–24. Miller, R.R., 1948, The Cyprinodont Fishes of the Death Valley System of East- Smith, G.R., Dowling, T.E., Gobalet, K.W., Lugaski, D.K., Shiozawa, D.K., and ern California and Southwestern Nevada: University of Michigan Museum Evans, R.P., 2002, Biogeography and timing of evolutionary events among of Zoology Miscellaneous Publication 68, 155 p. Great Basin fishes, in Hershler, R., Madsen, D.B., and Currey, D.R., eds., Orme, A.J., and Orme, A.R., 2008, this volume, Late Pleistocene climate and tec- Great Basin Aquatic Systems History: Smithsonian Contributions to the tonic impacts on Owens Lake levels and linkages, in Reheis, M.C., Hersh- Earth Sciences, v. 33, p. 175–234. ler, R., and Miller, D.M., eds., Late Cenozoic Drainage History of the Stinson, M.C., 1977, Geology of the Keeler 15Ј Quadrangle, Inyo County, Califor- Southwestern Great Basin and Lower Colorado River Region: Geologic and nia: California Division of Mines and Geology Map Sheet 38, scale 1:62,500. Biotic Perspectives: Geological Society of America Special Paper 439, doi: Stone, J.O., Allan, G.L., Fifield, L.K., and Cresswell, R.G., 1996, Cosmogenic 10.1130/2008.2439(09). chlorine-36 from calcium spallation: Geochimica et Cosmochimica Acta, Pakiser, L.C., Kane, M.F., and Jackson, W.H., 1964, Structural Geology and Vol- v. 60, p. 679–692, doi: 10.1016/0016-7037(95)00429-7. canism of Owens Valley Region, California—A Geophysical Study: U.S. Stone, P., Dunne, G.C., Conrad, J.E., Swanson, B.J., Stevens, C.H., and Valin, Geological Survey Professional Paper 438, 65 p. V.C., 2004, Geologic Map of the Cerro Gordo Peak 7.5Ј Quadrangle, Inyo Phillips, F.M., and Plummer, M.A., 1996, CHLOE: A program for interpreting County, California: U.S. Geological Survey Geologic Scientific Investiga- in-situ cosmogenic nuclide data for surface exposure dating and erosion tion Map I-2851, scale 1:24,000. studies, in Jull, A.J.T., 7th International Conference Accelerator Mass Spec- Turrin, B., and Gillespie, A., 1986, K/Ar ages of basaltic volcanism of the Big Pine trometry Abstracts: Radiocarbon, v. 38, p. 98–99. volcanic field, California: Implications for glacial stratigraphy and neotec- Phillips, F.M., Zreda, M.G., Smith, S.S., Elmore, D., Kubik, P.W., and Sharma, P., tonics of the Sierra Nevada: Geological Society of America Abstracts with 1990, Cosmogenic chlorine-36 chronology for glacial deposits at Bloody Programs, v. 18, no. 6, p. 777. Canyon, eastern Sierra Nevada: Science, v. 248, p. 1529–1532, doi: Winograd, I.J., Landwehr, J.M., Coplen, T.B., Sharp, W.D., Riggs, A.C., Ludwig, 10.1126/science.248.4962.1529. K.R., and Kolesar, P.T., 2006, Devils Hole, Nevada, b18O record extended Phillips, F.M., Stone, W.D., and Fabryka-Martin, J.T., 2001, An improved to the mid-Holocene: Quaternary Research, v. 66, p. 202–212, doi: 10.1016/ approach to calculating low-energy cosmic-ray neutron fluxes near the j.yqres.2006.06.003. land/atmosphere interface: Chemical Geology, v. 175, p. 689–701, doi: Zehfuss, P.H., Bierman, P.R., Gillespie, A.R., Burke, R.M., and Caffe, M.W., 10.1016/S0009-2541(00)00329-6. 2001, Slip rates on the Fish Springs fault, Owens Valley, California, de- Reheis, M., 1999, Highest pluvial-lake shorelines and Pleistocene climate of the duced from cosmogenic 10Be and 26Al and soil development on fans: Geo- western Great Basin: Quaternary Research, v. 52, p. 196–205, doi: 10.1006/ logical Society of America Bulletin, v. 113, p. 241–255, doi: 10.1130/ qres.1999.2064. 0016-7606(2001)113Ͻ0241:SROTFSϾ2.0.CO;2. Reheis, M.C., and Dixon, T.H., 1996, Kinematics of the Eastern California shear Zreda, M.G., 1994, Development and Calibration of the Cosmogenic 36Cl Surface zone: Evidence of slip transfer from Owens and Saline Valley fault zones Exposure Dating Method and Its Application to the Chronology of Late to fault zone: Geology, v. 24, p. 339–342, doi: Quaternary Glaciations [Ph.D. thesis]: Socorro, New Mexico Institute of 10.1130/0091-7613(1996)024Ͻ0339:KOTECSϾ2.3.CO;2. Mining and Technology, 318 p. Reheis, M.C., Stine, S., and Sarna-Wojcicki, A.M., 2002a, Drainage reversals in Mono Basin during the late Pliocene and Pleistocene: Geological Society of MANUSCRIPT ACCEPTED BY THE SOCIETY 17 JULY 2007

Printed in the USA Downloaded from specialpapers.gsapubs.org on July 1, 2015

Geological Society of America Special Papers

Late Quaternary MIS 6−8 shoreline features of pluvial Owens Lake, Owens Valley, eastern California

A.S. Jayko and S.N. Bacon

Geological Society of America Special Papers 2008;439; 185-206 doi:10.1130/2008.2439(08)

E-mail alerting services click www.gsapubs.org/cgi/alerts to receive free e-mail alerts when new articles cite this article

Subscribe click www.gsapubs.org/subscriptions to subscribe to Geological Society of America Special Papers

Permission request click www.geosociety.org/pubs/copyrt.htm#gsa to contact GSA.

Copyright not claimed on content prepared wholly by U.S. government employees within scope of their employment. Individual scientists are hereby granted permission, without fees or further requests to GSA, to use a single figure, a single table, and/or a brief paragraph of text in subsequent works and to make unlimited copies of items in GSA's journals for noncommercial use in classrooms to further education and science. This file may not be posted to any Web site, but authors may post the abstracts only of their articles on their own or their organization's Web site providing the posting includes a reference to the article's full citation. GSA provides this and other forums for the presentation of diverse opinions and positions by scientists worldwide, regardless of their race, citizenship, gender, religion, or political viewpoint. Opinions presented in this publication do not reflect official positions of the Society.

Notes

Geological Society of America