23
185 The Geological Society of America Special Paper 439 2008 Late Quaternary MIS 6–8 shoreline features of pluvial Owens Lake, Owens Valley, eastern California 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 Ecosystem Sciences, Reno, Nevada 89521, USA ABSTRACT The chronologic history of pluvial Owens Lake along the eastern Sierra Nevada in Owens Valley, California, has previously been reported forthe 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 20 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 Inyo Mountains and Coso Range east of the fault zone to as high as 1204 m west of the fault zone. This relict shoreline, referred to as the 1180 m shoreline, lies 20–40 m higher than the previously documented Last Glacial Maximum shoreline at 1160 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 60 ka and ca. 130 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 36 Cl model age of ca. 160 32 ka on reefal tufa at 1170 m elevation from the southeastern margin of the valley. This 36 Cl 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, Owens River system, deformed shorelines, marine isotope stage 6, MIS 6, Pleistocene lake, Great Basin, 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. on July 1, 2015 specialpapers.gsapubs.org Downloaded from

Late Quaternary MIS 6–8 shoreline features of pluvial ......Late Quaternary MIS 6–8 shoreline features of pluvial Owens Lake, ... Late Cenozoic Drainage History of the Southwestern

  • Upload
    others

  • View
    3

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Late Quaternary MIS 6–8 shoreline features of pluvial ......Late Quaternary MIS 6–8 shoreline features of pluvial Owens Lake, ... Late Cenozoic Drainage History of the Southwestern

185

The Geological Society of AmericaSpecial Paper 439

2008

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

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 Ecosystem Sciences, Reno, Nevada 89521, USA

ABSTRACT

The chronologic history of pluvial Owens Lake along the eastern Sierra Nevadain Owens Valley, California, has previously been reported for the interval of time fromca. 25 calibrated ka to the present. However, the age, distribution, and paleoclimaticcontext of higher-elevation shoreline features have not been formally documented. Wedescribe the location and characteristics of wave-formed erosional and depositionalfeatures, as well as fluvial strath terraces that grade into an older shoreline of pluvialOwens Lake. These pluvial-lacustrine features are described between the Olanchaarea to the south and Poverty Hills area to the north, and they appear to be verticallydeformed �20 � 4 m across the active oblique-dextral Owens Valley fault zone. Theyoccur at elevations from 1176 to 1182 m along the lower flanks of the Inyo Mountainsand Coso Range east of the fault zone to as high as �1204 m west of the fault zone.This relict shoreline, referred to as the 1180 m shoreline, lies �20–40 m higher thanthe previously documented Last Glacial Maximum shoreline at �1160 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, bracketthe age of the 1180 m shoreline to the time interval between ca. 340 � 60 ka and ca. 130 � 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 � 32 kaon reefal tufa at �1170 m elevation from the southeastern margin of the valley. This36Cl 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 OwensLake offers insight to the hydrologic balance along the east side of the southernSierra Nevada and will assist with future regional paleoclimatic models within thewestern Basin and Range.

Keywords: pluvial Owens Lake, Owens River system, deformed shorelines, marineisotope stage 6, MIS 6, Pleistocene lake, Great Basin, 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.

on July 1, 2015specialpapers.gsapubs.orgDownloaded from

Page 2: Late Quaternary MIS 6–8 shoreline features of pluvial ......Late Quaternary MIS 6–8 shoreline features of pluvial Owens Lake, ... Late Cenozoic Drainage History of the Southwestern

186 Jayko and Bacon

INTRODUCTION

Owens Valley is part of the Owens River system, which, dur-ing wet periods of the Pleistocene, extended from Mono Basin inthe north to Death Valley in the southeast (Gale, 1914; Black-welder, 1933; Smith et al., 1983; Jannik et al., 1991). ThroughoutHolocene time, the watershed has been hydrologically connectedalong only a third of its length—from Long Valley to Owens Lake(Fig. 1). The lower-elevation basins, including Indian Wells, Sear-les, Panamint, and Death Valleys, have unique lacustrine historiesthat are a consequence of the magnitude of glacial-pluvial events,which in part control the available water in the Owens River sur-face and groundwater systems. So, from an evolutionary perspec-tive, the duration and times of climate-controlled hydrologicconnectivity of the basins and their respective geochemistrieshave affected the distribution of various aquatic species (Hubbsand Miller, 1948; Miller, 1948; Firby et al., 1997; Reheis, 1999;Reheis et al., 2002a, 2002b).

In this study, we report the first age constraints and basin-wide documentation of shoreline features and fluvial strath ter-races of a pluvial Owens Lake at elevations between �1180 and1200 m, collectively referred to as the 1180 m shoreline. The1180 m shoreline is distinct from, and considerably older than,the Last Glacial Maximum (LGM) ca. 25 calibrated (cal) 14C kahighstand at �1160 m elevation, and it is also separate from theyounger, lower-elevation shorelines (Carver, 1970; Beanlandand Clark, 1994; Bacon et al., 2006; Orme and Orme, this vol-ume). Beanland and Clark’s (1994) paleoseismic study of theOwensValley fault zone noted the presence of an �1183 m strand-line along the northwest flank of the Coso Range. Other benchesand notches southeast of Olancha at approximately the sameelevation were also considered to be possible wave-formed plat-forms or river terraces (Beanland and Clark, 1994, their strand-line A; sites 1–3 herein; Fig. 1). However, they were uncertainwhether these sites represented offset from the �1160 m LGMstrandline or if they were older, and so they reported them as“�25 ka (?)” (Beanland and Clark, 1994, p. 6).

Tectonic Setting

Owens Valley lies adjacent to the southern part of the SierraNevada in eastern California (Fig. 1), and it is a 15–40-km-wide,200-km-long complex graben shaped by active transtensionaltectonism (e.g., Beanland and Clark, 1994; Savage and Lisowski,1995; Reheis and Dixon, 1996). The structural setting is impor-tant because of the high potential for deformation of shorelinesand spillway elevations by active tectonic processes. The valleyis bounded by the Sierra Nevada frontal faults on the west and theWhite and Inyo Mountains fault zones on the east (e.g., Pakiseret al., 1964; Bryant, 1984; Bacon et al., 2005; Slemmons et al.,2008). The valley floor is split longitudinally by the active, oblique-dextral Owens Valley fault zone, which last ruptured along an�120 km segment with a M 7.5–7.7 earthquake in A.D. 1872(e.g., Beanland and Clark, 1994; Slemmons et al., 2008) (Fig. 1).

The graben floor is underlain by at least four fault-bound struc-tural blocks with late Quaternary and/or Holocene displacement.

The valley is asymmetric from west to east due in part to rela-tive uplift of a horst block that lies west of the Owens Valley faultzone and east of the Sierra Nevada frontal faults, and in part to thedeposition of large, late Quaternary alluvial fans along the west-ern margin of the valley. The location of the Owens River mean-der belt reflects the position of the distal margins of these fans,which have deflected the river eastward toward the base of theInyo Mountains (Fig. 1).

During much of the middle and late Quaternary, a freshwaterlake occupied Owens Valley (Gale, 1914; Smith and Bischoff,1997). The current spillway or sill in Owens Valley at �1145 mnorth of Haiwee Reservoir (Gale, 1914) likely last overflowed ca.15.5 cal ka (Bacon et al., 2006) (Fig. 1). However, the sill droppedto that elevation during the late LGM following an event thateroded the berm, which confined the early LGM highstand at�1160 m.The location of the former 1180 m spillway is unknown.It may be eroded from the Haiwee area at the southernmost part ofOwens Valley, or it may lie further to the south at the southern endof Rose Valley, where erosional fluvial terraces formed in lateQuaternary (ca. 440–130 ka) basalt flows were noted by Duffieldand Smith (1978a, 1978b). Although an effort was made to estab-lish the location of this feature, it remains problematic.

Methods

Documentation of the 1180 m shoreline is based primarily onfield observations, crosscutting relations with dated volcanicrocks, and a cosmogenic 36Cl erosion age on tufa provided byFred Phillips (New Mexico Tech, 2007, personal commun.). Thespatial distribution of constructional and erosional lacustrine land-forms and erosional fluvial features associated with the 1180 mshoreline are described here. A description of each location isincluded in Table 1. Elevations estimated by global positioningsystem (GPS) were cross-checked with 1:24,000 topographicquadrangle maps and 30 m digital elevation models (DEMs).

A tufa sample was analyzed for 36Cl by standard procedures(Zreda, 1994), employing addition of a 35Cl spike and dissolu-tion of the carbonate sample in only nitric acid (F.M. Phillips,2007, personal commun.). Field and analytical results are givenin Table 2. The sample age was calculated using the spreadsheetprogram CHLOE (Phillips and Plummer, 1996) using the 36Clproduction parameters of Phillips et al. (2001) and the elevation-latitude scaling of Lal (1991). No correction for secular variationof Earth’s magnetic field was applied. The production parameteri-zation of Phillips et al. (1990) was largely calibrated using samplesfrom the same region, as well as using the same elevation-latitudescaling; thus, the age calculation should be relatively robust. Theage was calculated assuming that the dissolved material was onlyCaCO3 (i.e., 56 wt% CaO and 44 wt% CO2), given that dissolu-tion was only in nitric acid, and thus 36Cl should not have beenliberated from phases other than carbonate minerals (Phillips,2007, personal commun.).

on July 1, 2015specialpapers.gsapubs.orgDownloaded from

Page 3: Late Quaternary MIS 6–8 shoreline features of pluvial ......Late Quaternary MIS 6–8 shoreline features of pluvial Owens Lake, ... Late Cenozoic Drainage History of the Southwestern

Figure 1. Map of Owens Valley showing location of study areas and shoreline features of pluvial Owens Lake at elevationsbetween �1180 and 1200 m. Inset map shows the Owens River system and late Pleistocene lakes, the headwaters of theOwens 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 InyoMountain fault; SNFF—Sierra Nevada frontal fault; WMFZ—White Mountain fault zone.

on July 1, 2015specialpapers.gsapubs.orgDownloaded from

Page 4: Late Quaternary MIS 6–8 shoreline features of pluvial ......Late Quaternary MIS 6–8 shoreline features of pluvial Owens Lake, ... Late Cenozoic Drainage History of the Southwestern

TABLE 1. DESCRIPTION OF 1200 to 1180 METER SHORELINE FEATURES AND ASSOCIATED DEPOSITS IN OWENS VALLEY Siteno.

Elevation(m)

erutaeF noitpircseD gnitsaE gnihtroN

1 1180 4010297 413964 Wave-cut platform or paleoriver terraces of the Owens River of Beanland and Clark (1994).

Wave-cut platform; river terraces

2 1180 4018762 416422 Wave-cut notches into older alluvial fans that exhibit well-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).

Wave-cut notches

3 1179 ± 5 1170 ± 5

4024931 422364 Wave-cut notch and bluff; erosional feature formed into 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.

Wave-cut notch

4 1185 ± 5 4027460 424105 Embayment; remnant exposure of well-sorted, subrounded-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).

Embayment shoreline deposit

5 1180 ± 6 4027291 424143 Wave-cut notch formed into Miocene-Pliocene basalt of Coso Formation.

Wave-cut notch

6 1184 ± 4 4026825 423984 Abrasion platform or ravinement surface that has formed across metasedimentary rock.

Abrasion platform or ravinement surface

7 1182 ± 6 4026825 423927 Embayment; laterally continuous incised planar surface armored by a lag of sparse well-rounded gravel and subrounded-subangular basaltic cobble; surface underlain by rubified coarse sand.

Abrasion platform with gravel lag

8 1179 ± 4 4026802 423716 Tufa; large remnant exposure of tufa-cemented, well-sorted coarse sand with a gravelly to fine boulder lag at the surface; area of tufa fringes basaltic bedrock at edge of embayment.

Tufa

9 1179 ± 4 4026986 423573 Beach rock; moderately exposed and ~1 m thick, well-sorted coarse sand that is tufa-cemented; beach rock formed on shallow basaltic bedrock.

Beach rock

10 1169 ± 5 4027255 423493 Tufa mound; remnant exposure forming a topographic point that is primarily composed of ~1–2-m-thick tufa with silica veins.

Tufa mound

hcton tuc-evaW .tlasab otni demrof hcton tuc-evaW 548324 4208204 5 ± 5711 11

12 1181 ± 5 4027988 423855 Wave-cut notch between two embayments that is formed into basalt.

Wave-cut notch

13 1176 ± 3 4028230 424681 Abrasion platform; subrounded-rounded gravel, cobble, and tufa (pedogenic CaCO3?) lag at surface.

Abrasion platform

14 1181 ± 5 4028304 424733 Embayment; laterally continuous incised planar surface 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).

Abrasion platform

15 1175 ± 4 4028493 424602 Abrasion platform forming concordant surface that forms topographic point; surface has subrounded-rounded gravel and cobble lag.

Abrasion platform

16 1181 4040381 421739 6.7–4.3 Ma fanglomerate of Slate Canyon (Stone et al., 2004) truncated at 1181 m.

Truncated landform

hcton tuc-evaW .kcordeb etimolod otni demrof hcton tuc-evaW 418814 4503404 5 ± 9711 71

afuT .kcordeb etimolod derehtaew ro afuT 797814 7613404 3 ± 3811 81

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

20 1177 ± 3 4043260 418584 Pocket barrier; thin veneer of well-rounded bladed and spherical gravel composed of different lithologies; gravel deposit is situated within saddle.

Pocket barrier

21 1180 ± 5 4043360 418726 Pocket barrier; one well reirrab tekcoP .tsalc levarg dednuor-

22 1177 ± 5 4043340 418806 Abrasion platform; surface is incised and contains trace well-rounded gravel.

Abrasion platform

(continued)

on July 1, 2015specialpapers.gsapubs.orgDownloaded from

Page 5: Late Quaternary MIS 6–8 shoreline features of pluvial ......Late Quaternary MIS 6–8 shoreline features of pluvial Owens Lake, ... Late Cenozoic Drainage History of the Southwestern

TABLE 1. DESCRIPTION OF 1200 to 1180 METER SHORELINE FEATURES AND ASSOCIATED DEPOSITS IN OWENS VALLEY (continued)Siteno.

Elevation(m)

erutaeF noitpircseD gnitsaE gnihtroN

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. Slide material exhibits soil development; argillic horizon present (Blair, 1999).

Truncated landform

25 1180 ± 5 4073946 401078 Wave-cut abrasion platform formed into bedrock adjacent to Mazourka Canyon Road.

Wave-cut platform

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

Wave-cut notch and platform

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 into bedrock at base of range front.

Abrasion platform

29 1200 ± 6 4096225 389680 Higher and older strath surfaces formed on Red Mountain lava flow (340 ± 60 ka) that extend to north. Surfaces are incised relatively more deeply and dip to west.

Strath terraces

30 1195 ± 5 4096062 389700 Strath surfaces formed on Red Mountain lava flow (340 ± 60 ka) with local well-rounded granitic cobble clasts.

Strath terraces

31 1196 ± 5 4095984 389754 Strath surfaces formed on Red Mountain lava flow (340 ± 60 ka); no gravel.

Strath terraces

32 1197 ± 5 4095844 389564 Accumulation of spherical and well-rounded gravel and cobble consisting of granitic and metamorphic lithologies.

Deposit on strath terrace

33 1195 ± 5 4095871 389428 Strath surfaces formed on Red Mountain lava flow (340 ± 60 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.

Strath terraces

34 1186 ± 5 4091059 388094 Wave-cut notch into northern part of Armstrong Creek lava 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.

Wave-cut notch

35 1185 ± 5 4090721 388041 Wave-cut notch into northern part of Armstrong Creek lava flow (<300 ka).

Wave-cut notch

36 1192 ± 5 4087019 389101 Wave-cut notch and platform into Armstrong Creek lava flow (<300 ka).

Wave-cut notch and platform

37 1193 ± 5 4086966 388618 Wave-cut notch and platform into Armstrong Creek lava flow (<300 ka) with sparse ~2 cm rounded granitic clasts.

Wave-cut notch and platform

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 Armstrong Creek lava flow (<300 ka).

Shoreline deposit

40 1193 ± 6 4087138 388351 Wave-cut notches in Armstrong Creek lava flow (<300 ka) and associated deposit that contains very well-rounded, pea-sized gravel consisting mostly of volcanic lithologies.

Wave-cut notches/shoreline deposit

41 1204 4051096 403487 Strath terraces on bedrock at the mouth of Lone Pine Creek. 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.

Strath terraces

42 1199 4033141 406390 Degraded abrasion surface into bedrock at base of Sierra Nevada range front.

Abrasion surface

43 1182 4010355 410780 Erosional benches developed on old alluvial-fan deposits east of Sage Flat; trimline with gentle shoreline angle.

Abrasion surface and trimline

44 1173 4009510 411128 Tightly packed, clast-supported cobble and boulder deposit along east side of 1182 m benches, east of Sage Flat

Possible boulder beach berm

on July 1, 2015specialpapers.gsapubs.orgDownloaded from

Page 6: Late Quaternary MIS 6–8 shoreline features of pluvial ......Late Quaternary MIS 6–8 shoreline features of pluvial Owens Lake, ... Late Cenozoic Drainage History of the Southwestern

190 Jayko and Bacon

TABLE 2. FIELD AND ANALYTICAL DATA FOR 36Cl SAMPLE OL96-1 FROM SITE 10, CENTENNIAL FLAT EMBAYMENT

Variable Definition Units Value R36 Sample 36Cl/Cl ratio 36Cl/1015 Cl 46,073 ±R36 Sample 36Cl/Cl ratio uncertainty 36Cl/1015 Cl 1250 rb Bulk density g cm–3 2.65 ls Sample thickness cm 3 θavg

Water content cm3/cm3 0.1 9611 m velE

22983.63 N° taL71358.711 W° gnoL

Ssnow Snow shielding unitless 1.00 ST Total shielding unitless 0.999 Λf

Effective attenuation length g cm–2 170 LOI Loss on ignition wt% 42.39 Na2O wt% 0.04 MgO wt% 1.06 Al2O3 wt% 0.21 SiO2 wt% 0.93 P2O5 wt% 0.02 K2O wt% 0.03 CaO wt% 52.92 TiO2 wt% 0.01 MnO wt% 0.00 Fe2O3 wt% 0.06 Cl ppm 7.86 B ppm 14.00 Sm ppm 0.00 Gd ppm 5.00 U ppm 2.00 Th ppm 2.00 Msample Sample mass g 110.200 Mspike Mass 35Cl spike solution g 4.003 Cspike Concentration spike solution g g–3 0.997 S/S Analytical stable isotope ratio [35Cl/(35Cl + 37Cl)] 20.94 R/S Analytical 36Cl/Cl ratio 36Cl/1015 Cl 8120.00

1180 METER SHORELINE AND FLUVIAL FEATURESIN OWENS VALLEY

We document lacustrine and fluvial features at �1180 m ele-vation in Owens Valley and focus on four areas, two sites to theeast of the active Owens Valley fault zone and two to the west(Fig. 1). These sites are referred to as: (1) the Centennial Flatembayment, on the northeastern corner of the Coso Range; (2) theSwansea pocket barrier, in the southern Inyo Mountains rangefront near Swansea; (3) the southern end of the Big Pine volcanicfield in north-central Owens Valley, which includes fluvial ero-sional features on lava flows north of Taboose Creek and shore-line erosional features on lava flows near Sawmill Creek; and (4) the Alabama Hills strath terrace and wave-formed platformsnear Lone Pine. We also briefly describe other erosional shorelinefeatures observed in Owens Valley that are associated with the1180 m shoreline.

Depositional features associated with the shoreline includeremnant cobble and gravel beach deposits, sandy embaymentdeposits with a lag of gravel and tufa fragments, and calcareousdeposits in the form of tufa mounds with local secondary lami-nated silica, as well as sandy beach rock cemented with tufa.Erosional wave-generated landforms include trimlines, wave-

formed abrasion surfaces, platforms, and notches. Fluvial strathterraces are located near the shoreline features but at slightlyhigher elevation north of Independence. All these features arepresent on Paleozoic-Mesozoic bedrock or Miocene-Pliocenevolcanic and sedimentary rocks of the Coso Formation (e.g.,Stinson, 1977), on Miocene-Pliocene fanglomerate (e.g., Stoneet al., 2004) and late Pleistocene alluvium, or on late Pleistocenelava flows of the Big Pine volcanic field (e.g., Cox et al., 1963;Turrin and Gillespie, 1986, with localities and ages reported inConnor and Conway, 2000, Figure 7 therein; Dorn et al., 1987;Martel et al., 1987).

Dating of the 1180 m shoreline is based on crosscutting rela-tions of lacustrine features with dated lava flows of the Big Pinevolcanic field within the northern part of the valley, and from a36Cl exposure age on tufa at �1170 m elevation in the south (F.Phillips, New Mexico Tech, 2007, personal commun.). Recon-naissance observations of weathering also distinguish the clasticdeposits associated with the 1180 m shoreline from younger LGMclastic lacustrine deposits at �1160 m, based mainly on thedegree of weathering of the soil profile.

CENTENNIAL FLAT EMBAYMENT

Depositional and erosional remnants of the 1180 m shorelineare located in southern Owens Lake basin along the northwesternflank of the Coso Range near Centennial Flat (Figs. 1 and 2). Thisembayment contains erosional wave-formed platforms or abra-sion surfaces (Qab) and associated wave-formed notches andtrimlines that formed on massive, indurated late Cenozoic basaltand 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).These remnant wave-formed surfaces are distinctly flat-lying and typically terminate at abrupt breaks in slope that form trim-lines or notches (Fig. 4). Rare, erratic, and very well-rounded,cobble- and pebble-sized clasts of Mesozoic metasediment, in-cluding hornfelsic siltstone, mudstone, and intermediate meta-volcanic rock, occur just below the 1180 m trimlines and notches.These clasts also form a cobbly to pebbly lag on the surface ofsandy embayment deposits or on abrasion surfaces (sites 6 and 7;Table 1; Figs. 3A and 3B).

The most extensive depositional feature at the embaymentconsists of a 3–4-m-thick, massive and moderately sorted, coarsearkosic sand deposit (unit Qem; Figs. 2 and 3B). The Qem depositis widely distributed; it extends for nearly 1 km, and it ranges inelevation from 1165 to 1182 m. The northwestern half of theembayment is truncated by LGM and younger gravelly beachridges (Qg) below �1160 m (Beanland and Clark, 1994; Baconet al., 2006; Orme and Orme, this volume).

The Qem deposits grade to the same elevation near an iso-lated, 3–5-m-thick outcrop of whitish-gray, bedded silt and finesand with a gravelly lag of tufa fragments mapped as unit Qm inthe western margin of the embayment (site 8; Table 1; Figs. 2and 3C). The lithologic association of the Qm sediment is typicalof a shallow-water to wetlands depositional environment linked

on July 1, 2015specialpapers.gsapubs.orgDownloaded from

Page 7: Late Quaternary MIS 6–8 shoreline features of pluvial ......Late Quaternary MIS 6–8 shoreline features of pluvial Owens Lake, ... Late Cenozoic Drainage History of the Southwestern

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

to groundwater discharge. Outcrops of beach rock cemented by tufa, gravelly tufa lag, and isolated well-rounded pebbles andgravel clasts lie along a trimline that crosses a bedrock slope at�1179 m. These features lie along the southwestern margin ofthe embayment between sites 8 and 9 (Table 1; Figs. 2 and 3C). The in situ tufa beach rock and fragmental lag on bedrock

surfaces range in thickness from �0.5 m at the �1179 m trim-line and notches to �3–4 m downslope at 1169 m near out-crops of tufa mounds. These massive mounds consist of frac-tured and bedded tufa with secondary siliceous laminations asmuch as 2–6 mm thick that infill dissolution voids (site 10; Table 1; Figs. 2 and 3D).

Figure 2. Map showing location of Centennial Flat embayment along the northeast corner of the Coso Range. Numbered site locations (stars) are describedin text and Table 1. CFF—Centennial Flat fault, modified from Jayko (2008). LGM—Last Glacial Maximum; MIS—marine oxygen isotope stage.

on July 1, 2015specialpapers.gsapubs.orgDownloaded from

Page 8: Late Quaternary MIS 6–8 shoreline features of pluvial ......Late Quaternary MIS 6–8 shoreline features of pluvial Owens Lake, ... Late Cenozoic Drainage History of the Southwestern

Figure 3. Photographs showing shoreline features and embayment deposits at the Centennial Flat embayment (see Figs. 1and 2). (A) Abrasion platform in foreground and wetland/embayment deposit in background; person for scale, view lookingwest. (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 ofthe 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 erosionsince the tufa formed. Dashed lines represent age uncertainties calculated from the 1� analytical standard deviation. Thearrows indicate the ages associated with the estimated limits on total tufa erosion.

on July 1, 2015specialpapers.gsapubs.orgDownloaded from

Page 9: Late Quaternary MIS 6–8 shoreline features of pluvial ......Late Quaternary MIS 6–8 shoreline features of pluvial Owens Lake, ... Late Cenozoic Drainage History of the Southwestern

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) Abrasionplatforms and wave-formed notches looking west from site 13 to site 11. (B) Well-developed abrasion surface consisting of wave-formed notch andabrasion platform on a local bedrock headland; view looking west.

Relative Age of DepositsThe relative degree of weathering, in the form of rubification

and accumulation of pedogenic clay, is used to differentiate thehigher sandy deposits of the embayment (Qem) at �1180 m fromthe lower LGM and younger gravelly beach deposits (Qg)between 1145 and 1160 m. From a reconnaissance-level investi-gation of the soil profile of these deposits, the moderately devel-oped soils associated with the 1180 m shoreline exhibit colors thatare brownish yellow (10YR 6/6) within soils developed fromarkosic parent material. Furthermore, an increase in clay is appar-ent in the upper part of the profiles, which also exhibit no dis-cernible carbonate accumulation. In contrast, the LGM andyounger deposits are similar in color to the parent material (i.e.,no rubification) and lack apparent increase in clay content, butthey do have an incipient development of pedogenic carbonatethat forms thin, discontinuous coatings under clasts. Locally, dis-continuous outcrops of beach rock are cemented by tufa carbon-ate (as opposed to pedogenic carbonate).

Cosmogenic 36Cl Age EstimateIn 1996, Fred M. Phillips in the company of Antony and

Amalie Jo Orme sampled the tufa mound at site 10 in the Centen-nial Flat embayment for chlorine-36 (36Cl) surface exposure dat-ing. The amount of 36Cl production in this sample was influencedby spallation of calcium. As a result, the calculated age is sensi-tive to surface erosion, as is typical of spallation-dominatednuclides such as 10Be. Interpretation of the cosmogenic nuclideage therefore requires careful evaluation of the total erosion of thetufa, which proceeds at a relatively high rate due to the high sol-ubility of calcite (Phillips, 2007, personal commun.). The tufawas present as scattered lag and remnant tufa heads projecting�20 cm above the soil surface (Fig. 3D).

The tufa mound sampled was close to the highest shorelinedeposits, may have been reef-forming, and is therefore interpretedto be a nearshore tufa deposit rather than a deep-water tufa tower.Similar shoreline tufa deposits at Mono Lake, Panamint Valley,and Lake Lahontan have thicknesses around 0.5–3.5 m (e.g., Ben-son, 1994). The 36Cl exposure age of the tufa is estimated assum-ing that 30–75 cm of tufa has been removed by chemical andphysical erosion. The total erosion limits correspond to limitingerosion rates of 2.2 mm/k.y. (for 30 cm erosion) and 4.1 mm/k.y.(for 75 cm erosion), which are assumed to be constant (Phillips,2007, personal commun.). Figure 3E shows the calculated 36Clexposure age of the tufa as a function of total erosion depth fromTable 2. The lower and upper limits of the estimated erosion yieldage limits between 192 and 127 ka, respectively; therefore, we getan average age of ca. 160 � 32 ka. Application of the alternative36Cl production rate of Stone et al. (1996) for spallation from cal-cium would give a somewhat older age of ca. 218 � 33 ka(Phillips, 2007, personal commun.). These results, although fromonly a single sample from tufa sampled at �1170 m (site 10), areconsistent with a marine oxygen isotope stage (MIS) 6 age. Thiselevation lies below the highstand elevation for this locality. Thesampled tufa is continuously exposed up to the beach deposits andrelicts of tufa at 1180 m. We infer that the tufa at this elevationformed a reef that was established as the lake was rising, as wehave not found evidence for a lake highstand that was stableenough to erode shoreline features in bedrock at 1170 m elevationelsewhere in the basin.

Lacustrine SettingThe Centennial Flat embayment of pluvial Owens Lake

was open to the northwest and contained at least two islands, aswell as beach cliffs or erosional escarpments distinguished by

on July 1, 2015specialpapers.gsapubs.orgDownloaded from

Page 10: Late Quaternary MIS 6–8 shoreline features of pluvial ......Late Quaternary MIS 6–8 shoreline features of pluvial Owens Lake, ... Late Cenozoic Drainage History of the Southwestern

194 Jayko and Bacon

trimlines, notches, and abrasion surfaces formed on bedrock. Thewidespread accumulation of sand and silty sediment characteris-tic of both beach and marsh or wetlands deposits in the embay-ment means they likely formed in a bay head and (or) bay-sidebeach depositional environment similar to that described byAdams and Wesnousky (1998) for the Jessup embayment of LakeLahontan in Nevada. The deposition of clastic sediment anddevelopment of erosional features within the embayment likelywas influenced by large wave energy and an unobstructed fetchdirected from the north.

Since the development of the Centennial Flat embayment,younger LGM beach ridges and barriers associated with the�1160 m water level formed below cliffed headlands and trun-cated the older embayment deposits. The Centennial Flat fault,which is a northwest-trending and possibly oblique-dextral fault

with down-to-the-west displacement, is located west of the pre-served embayment (Figs. 1 and 2). Relative uplift on the east sideof the Centennial Flat fault apparently preserved the lacustrinerecord within the embayment. The LGM and younger beachdeposits are also well preserved on the upthrown side of the Cen-tennial Flat fault, whereas to the west of the fault, they are mostlyburied by alluvium or eroded.

Swansea Pocket Barrier

The best example of a beach deposit associated with the 1180 m shoreline is located within the foothills of the southernInyo Mountains near the historic site of Swansea, and it is re-ferred to as the Swansea pocket barrier (Figs. 1 and 5). At thissite, 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, atthe Swansea pocket barrier (see Fig. 1 for location). LGM—Last Glacial Maximum.

on July 1, 2015specialpapers.gsapubs.orgDownloaded from

Page 11: Late Quaternary MIS 6–8 shoreline features of pluvial ......Late Quaternary MIS 6–8 shoreline features of pluvial Owens Lake, ... Late Cenozoic Drainage History of the Southwestern

Late Quaternary MIS 6–8 shoreline features of pluvial Owens Lake 195

abrasion surface at �1177 m between two bedrock knobs com-posed of Paleozoic dolomite (site 20; Table 1; Figs. 5 and 6A).Well-rounded, discoidal clasts lie on a thin veneer of �2–4-cm-wide fragmented and angular tufa at this site (site 20; Table 1;Figs. 5 and 6B). The tufa appears to have been brecciated in situby postdepositional mechanical weathering. The Qg1 depositalso rests on bedrock benches (Qab) that terminate against wave-cut notches (e.g., site 17; Table 1; Figs. 5 and 6C). A trimline ad-jacent to a wave-formed notch locally separates a hillslope withrough microtopography and abundant colluvium from a notice-ably much smoother and more sparsely colluviated hillslopebelow (Figs. 5, 6A, and 6C).

Pinkish tufa covers several abrasion surfaces on bedrockslopes south of the Qg1 deposit between 1177 and 1179 m eleva-tion (sites 17–19, 23; Table 1; Figs. 5 and 6A). The tufa is lami-

nated with 2–4-mm-wide secondary silica, which partly fills andcements fractured dolomite bedrock. This pinkish tufa contrastswith the whitish or grayish white pedogenic carbonate observedelsewhere in the valley. Most of the abrasion surfaces are denudedof beach gravel and associated tufa, but a few rare, well-roundedbeach cobbles and pebbles were found on bedrock surfaces to theeast (sites 21, 22; Table 1; Fig. 5).

The Qg1 deposit and associated Qab surfaces are locallyburied by undifferentiated post–15 cal ka alluvial sediment andcolluvium (Qa2), and they are dissected by active alluvial washes(Qa3) along the southeastern edge of the deposit. Bedrock cliffstruncate the western edge of the deposit. The latest Quaternary(�15 cal ka) lake plain and beach ridges (Carver, 1970; Bacon et al., 2006; Orme and Orme, this volume) lie �40 m below thecliffs at elevations below �1140 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 notchin Figure 5 and is about 50 m � 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.

on July 1, 2015specialpapers.gsapubs.orgDownloaded from

Page 12: Late Quaternary MIS 6–8 shoreline features of pluvial ......Late Quaternary MIS 6–8 shoreline features of pluvial Owens Lake, ... Late Cenozoic Drainage History of the Southwestern

196 Jayko and Bacon

Lacustrine SettingThe morphologic setting of the remnant gravel deposit and

associated wave-formed abrasion features at this locality appearto be similar to relatively protected pocket barriers formed withinpluvial lake systems (e.g., Russell, 1885; Adams and Wesnousky,1998). The Swansea pocket barrier was open to the southwestthrough a saddle of bedrock (Fig. 5) and appears to have formedthe western margin of a relatively steep-sided and deep embay-ment that is now filled with latest Quaternary alluvial fans andcolluvial slopes.

Sites near the Big Pine Volcanic Field, Northern Owens Valley

Lacustrine and fluvial features observed in the northern partof the study area include an extensive upland area dominated byfluvial strath surfaces that are overlain by lag cobbles and sandydeposits near Taboose Creek. This upland area is succeeded to thesouth by a remnant wave-formed shoreline platform and over-lying sandy beach deposits near Sawmill Creek. The valley floor

between these creeks, as well as near Aberdeen, ranges in eleva-tion from at least 1170 m south of the Tinemaha Narrows to atleast 1160 m near Independence west of the Owens Valley faultzone within the active meander belt of the Owens River (Qfl)(Fig. 1). Erosional shoreline features lie above the valley floor at�1180 m east of the Owens Valley fault zone along the InyoMountains, whereas erosional shoreline and fluvial features arefound within the southern Big Pine volcanic field at �1185–1200 m west of the fault zone (sites 27–41; Table 1; Fig. 7).

Volcanic rocks within the southern and central parts of theBig Pine volcanic field provide age control for the associatedlacustrine and fluvial geomorphic features (Fig. 1). Basalt lavaflows in the volcanic field have yielded K/Ar and 40Ar-39Ar datesthat range from 1.2 Ma to 280 ka (for Qv1) and an age of 130 �50 ka for Qv3, whereas the intermediate unit Qv2 is undated (Coxet al., 1963; Gillespie, 1982; Gillespie et al., 1983, 1984; Turrinand Gillespie, 1986, see compiled ages in Connor and Conway,2000; Dorn et al., 1987; Martel et al., 1987) (Fig. 7). The relativedegree of weathering and surface morphology on the Qv1 andQv2 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.

on July 1, 2015specialpapers.gsapubs.orgDownloaded from

Page 13: Late Quaternary MIS 6–8 shoreline features of pluvial ......Late Quaternary MIS 6–8 shoreline features of pluvial Owens Lake, ... Late Cenozoic Drainage History of the Southwestern

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

The Qv2 lava flows exhibit relatively smooth surfaces that aremoderately dissected (Figs. 8A and 9A), whereas the younger180–80 ka (Qv3) flow (Dorn et al., 1987) tends to have a roughand blocky surface with irregular meter-scale microtopography, amuch fresher, darker appearance, and a thinner accumulation ofoverlying eolian sand and silt. We use the dated basalt flows tohelp constrain the relative age of inset strath terraces and wave-formed platforms in the following sections.

Taboose Creek Strath TerracesTwo distinct erosional surfaces are developed north of

Taboose Creek along the southern margin of the Red Mountainlava flow west of the Owens Valley fault zone (Qv1; ca. 340 �60 ka; Turrin and Gillespie, 1986, see map Figure 7 in Connor andConway, 2000; Dorn et al., 1987) (Fig. 7). The lower erosionalsurface ranges in elevation from 1195 to 1197 m, slopes gently to

the east, and is dissected by well-developed, 3- to 4-m-deep chan-nels that trend east toward the Owens River (sites 31–34; Figs. 9Aand 9B). North of the lower surface, there is a �3-m-high south-facing slope resembling a terrace riser that separates the lower andupper surfaces. The upper surface lies at �1200 m and exhibitswestward-trending channels as much as 6–8 m deep (sites 30–34;Table 1; Figs. 7, 9A, and 9B). The erosional surfaces are �10–15 mhigher than Quaternary alluvium (Qa) to the south on active fansurfaces emanating from the Sierra Nevada. The relatively sub-dued ridges or knobs appear to be the higher portions of arcuatepressure ridges of the flow surface that have been beveled tonearly flat benches. These pressure ridges are evident on aerialphotographs and satellite imagery, and they become progres-sively less preserved and more eroded to the south (Fig. 8A).

Isolated ridges on the 1195–1197 m surface exhibit a well-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 orthophotoquad showing location of sites described in text and Table 1. (B) Photo showing locations of fluvial strath terraces developed on late Pleistocene basaltflows near Tinemaha Narrows. (C) Photo showing isolated subrounded and rounded clasts of a fluvial lag deposit, as well as angular basalt brecciatedclasts on a strath terrace. The brecciation and fragmental character of the basalt are secondary and from weathering of more massive basalt flows.

on July 1, 2015specialpapers.gsapubs.orgDownloaded from

Page 14: Late Quaternary MIS 6–8 shoreline features of pluvial ......Late Quaternary MIS 6–8 shoreline features of pluvial Owens Lake, ... Late Cenozoic Drainage History of the Southwestern

198 Jayko and Bacon

brecciated basalt clasts form a patchy mosaic of roughly equant-shaped 20–40-cm-wide fragments lying between larger basalticoutcrops. The breccia clasts were formed in situ from weatheringand disaggregation of the basalt flows and have not been trans-ported significant distances. There are well-rounded cobbles andboulders, usually observed as isolated clasts rather than clusters,which are coated by an iron oxide stain, above and slightlydownslope of the erosion surfaces (site 34; Figs. 8A and 8C).Preservation of the iron oxide coating implies that the clasts arereworked out of an older, local deposit, and have not been trans-ported a great distance. The lithologies of observed clasts are gran-odioritic and metasedimentary, which are characteristic of SierraNevada plutonic rocks and roof pendants. Clasts composed ofBishop Tuff lithology are ubiquitous in middle and late Quater-nary Owens River and Owens Lake deposits. This lithology is

absent from these surfaces, suggesting the gravel may have beendeposited by local streams, such as Taboose Creek, flowing eastfrom the Sierra Nevada rather than by the Owens River.

Relative Age Constraints. The strath terraces were formedduring times when sedimentation rates were relatively low com-pared to vertical uplift rates. These straths, cut on the ca. 340 �60 ka flows of Red Mountain, have subsequently been incised bychannels. The degree of weathering on basaltic flow surfacesdescribed earlier and the development of weathering rinds ongravel and cobble clasts support an age older than ca. 25 cal ka.Furthermore, LGM outwash and younger alluvial fans of the steeppiedmont that lie east and south (?) of the site are graded lowerthan the terraces, providing additional evidence for an older age.

Depositional Setting. The two planar and gently east- andwest-sloping erosional surfaces are developed across the southern

Figure 9. Site localities and abrasion surfaces developed on basalt younger than 300 ka (Qv2) north of Sawmill Creek. (A) Portion of digitalorthophoto quad showing location of sites described in text and Table 1. (B) Photo showing locations of wave-formed surface developed on latePleistocene 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 othergenerally coarser, angular basalt clasts are secondary and from weathering of larger basalt clasts.

on July 1, 2015specialpapers.gsapubs.orgDownloaded from

Page 15: Late Quaternary MIS 6–8 shoreline features of pluvial ......Late Quaternary MIS 6–8 shoreline features of pluvial Owens Lake, ... Late Cenozoic Drainage History of the Southwestern

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

Red Mountain lava flow. The occurrence of erratic cobbles andgravel to small boulders in conjunction with these surfaces indi-cates that they formed by fluvial rather than lacustrine processes.Therefore, we interpret these erosion surfaces as fluvial terraces.The fluvial terraces are on the upthrown block of the Owens Val-ley fault zone and are �1.5 km west of the mapped trace (Fig. 7).Near these terraces, the Owens Valley fault zone steps to the eastand changes to a more northerly strike, with an implied increasein the vertical component of slip. It appears that the terraces wereinitially cut during two different phases and were subsequentlyincised by a later cutting phase that formed the channels, whichlikely were influenced by relative uplift associated with theOwens Valley fault zone. As a result, these planar surfaces erodedon lava flows are interpreted as strath terraces (e.g., Bull, 1991).These strath terraces are graded to the south, to the Sawmill Creekshoreline, and are used as a datum to constrain the elevation of alake level below �1195–1197 m.

Sawmill Creek ShorelineA planar and flat erosional surface on the southern margin of

the Qv2 flow is locally exposed under a thin and discontinuousveneer of alluvium between the elevations of 1192 and 1193 m(sites 37–41; Table 1; Figs. 7 and 9A). In this area, erosional featuresare characterized as isolated benches in the form of flat platforms,as well as more continuous platforms that intersect low-reliefknobs with significant slope breaks that appear as notches (sites37 and 38; Fig. 9B). At one site, a relatively small lag (�40 m2)of very well-rounded, roughly equant, 3–5-mm-sized basalt peb-bles is preserved against a well-developed wave-formed notch(site 41; Table 1; Figs. 9B and 9C). The well-rounded, size-sorted,equant nature of the basalt pebbles, the close proximity to theirsource rock, and the absence of detrital clasts of the Bishop Tuff,indicative of an Owens River fluvial deposit, suggest that thebasaltic pebbles could have been deposited in a beach (shoreline)environment.

A �1-m-thick deposit of well-rounded and moderatelysorted coarse pebbly sand underlies the well-rounded basaltic lagpebbles. This sand exhibits considerable rubification (7.5YR 5/6)but has a weakly developed soil, indicated by a lack of soil struc-ture, that appears to be currently influenced by active bioturbationfrom rodents. The well-rounded and weathered clasts of thisdeposit contrast with very light-colored, arkosic, fresh sandy andpebbly alluvium on the alluvial-fan surface only 2–3 m lower.Based on the small fraction of arkosic sediment within the beachdeposit, it appears that there has been little recent eolian or allu-vial contribution at the site.

Age Constraints. The beach deposits overlie a bench or abra-sion platform that lies along the southern margin of an undated lavaflow (Qv2) west of the Owens Valley fault zone (Fig. 7). This plat-form is well developed along the southern margin of the Qv2 flowbut is not discernible toward the axis of the flow, except at the north-ern flow margin, where a portion of the adjacent Qv1 lava flow isexposed (Fig. 7). Notches interpreted as either fluvial or shorelinefeatures at �1185–1186 also occur on the Qv1 flow to the north

(sites 35 and 36; Table 1; Figs. 1 and 7). Compared to the surfaceof the 340 � 60 ka (Qv1) Red Mountain lava flow (Dorn et al.,1987), the Qv2 flow appears to be younger, has pressure ridges thatare moderately to well-preserved, and is not as dissected by alluvialchannels. Compared to the surface of the ca. 130 � 50 ka (Qv3)lava flow to the north (Dorn et al., 1987), the Qv2 flow is moreweathered, has more muted, meter-scale microtopography, and ismore dissected. Therefore, based on relative degrees of surfaceweathering characteristics, Qv2 is intermediate in age (Fig. 8A).

Depositional Setting. The numerous breaks in slope at theintersection with flat platforms along the southern margin of thelava flow, combined with a deposit we interpret as beach pebblesand sand, suggest that these features are wave-formed notches orangles. Development of these features likely was influenced bylarge wave energy and unobstructed fetch directed from thesouth-southeast. Although the evidence is not robust for widelydistributed beach deposits associated with the �1192–1193 mplatforms in the area, we interpret these features as a shoreline thatappears to be older than adjacent post-LGM alluvium.

Alabama Hills Strath Terraces

In the vicinity of Lone Pine, there are several planar erosionalfeatures formed on indurated Mesozoic metasedimentary rocksbetween the elevations of 1200 and 1204 m along the easternAlabama Hills range front (Fig. 1). These erosional features arediscontinuous and occur in the form of bedrock benches andnotches. In addition, a relatively flat and well-exposed fluvial cut-and-fill terrace on bedrock was observed at the mouth of LonePine Creek (site 42; Table 1; Fig. 1).

The terrace cut on bedrock is overlain by a fill of clast-sup-ported, well-rounded, boulder and cobble conglomerate at itsnorthern edge along the Whitney Portal road west of Lone Pine(Fig. 10A). The relatively fresh and loose conglomerate was ini-tially mapped by Blair (2001) and later by Benn et al. (2006) aspossible evidence for catastrophic outburst floods from failure ofmoraine-damned lakes within the steep eastern range front of theSierra Nevada. The same deposit is mapped and dated nearby asoutwash/alluvium by Bierman et al. (1995), who reported cosmo-genic 10Be surface ages from boulders that were later recalculatedwith up-to-date production rates by Bacon and Pezzopane (2007)to between ca. 26 and 11 ka. These 10Be surface ages are cor-roborated by additional age constraints of the same fan betweenca. 25 and 12 cal ka by Lubetkin and Clark (1988) and Beanlandand Clark (1994), based on relative weathering characteristicsand crosscutting relations of 14C-dated lithoid tufa deposits northof Lone Pine Creek. In contrast, further upstream, the cut terraceappears to be overlain by a sequence of interbedded gravellychannel bed deposits and massive sandy and gravelly to boulderydeposits that are reddened or deeply weathered (Blair, 2001). Theweathered deposits are in sharp contact with and overlain by theloose and fresh alluvium dated as old as ca. 26 ka.

This terrace is interpreted as a fluvial strath terrace be-cause it is cut into indurated bedrock and it is located on the

on July 1, 2015specialpapers.gsapubs.orgDownloaded from

Page 16: Late Quaternary MIS 6–8 shoreline features of pluvial ......Late Quaternary MIS 6–8 shoreline features of pluvial Owens Lake, ... Late Cenozoic Drainage History of the Southwestern

Figure 10. Site localities in southern Owens Valley. (A) Photo showing strath terrace at �1200 m elevation developed onMesozoic bedrock. Strath is overlain by grayish weathered Last Glacial Maximum (LGM) glacial outwash at this localityand by both the LGM and older reddish weathered outwash deposit upstream. (B) Photo showing wave-formed notches at1160, 1180, and �1300 m elevation on granitic bedrock headlands of a promontory called Haystack. (C) View looking easttoward Coso Range from �1182 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 SierraNevada on the left showing benches at 1182 m elevation. (E) Photo showing tightly packed, clast-supported cobble-boulderdeposit (interpreted as a possible berm) adjacent to the benches at �1177 m elevation. Field notebook (�10 cm � 17 cm)for scale near the center of the photo.

on July 1, 2015specialpapers.gsapubs.orgDownloaded from

Page 17: Late Quaternary MIS 6–8 shoreline features of pluvial ......Late Quaternary MIS 6–8 shoreline features of pluvial Owens Lake, ... Late Cenozoic Drainage History of the Southwestern

Late Quaternary MIS 6–8 shoreline features of pluvial Owens Lake 201

upthrown block of the Owens Valley fault zone with a mappedtrace �100 m to the east (Fig. 1). The surface of the strath ter-race projects out to the east above the valley floor and �40–44 m above the LGM shoreline at �1160 m. The Lone PineCreek appears to have subsequently cut a channel in bedrockthrough the strath terrace down to �1160 m, which may reflectbase-level changes of pluvial Owens Lake as it dropped from�1160 m to �1145 m between ca. 25 and 15.5 cal ka. Becausethe strath is locally overlain by very weathered deposits, thestrath terrace must have been cut well before deposition of theyounger ca. 25–12 ka gravels. We interpret the old, pre-LGMstrath terrace and other benches along the range front of theAlabama Hills as evidence of a nearby shoreline and adjacentstreams entering a lake near Lone Pine that had a maximum ele-vation of �1200 m. This elevation is similar to the elevation offeatures documented at the southern Big Pine volcanic field onthe west side of the Owens Valley fault zone.

Erosional Wave-Formed Features in Owens Valley

Several erosional shoreline features are recorded at localitiesalong the range front of the Inyo Mountains at �1180 m in eleva-tion and along the Sierra Nevada at �1200 m (sites 16, 25, and43; Table 1; Fig. 1). These sites preserve wave-formed featuresconsisting of abrasion platforms and notches in resistant bedrock,and truncated landforms composed of unconsolidated Quaternaryand Tertiary alluvial deposits.

Bedrock NotchesNotches cut on quartz monzonite are located at Haystack

(Kern Knob) along the Inyo Mountains directly east of LonePine (site 24; Table 1; Figs. 1 and 10B). Three distinct notchesoccur at this site at elevations of 1160 and 1180 m, as well as ahigher one at �1300 m (Fig. 10B). During lake levels above�1120 m, the area of Haystack formed a headland that pro-truded out from the range front. South of Lone Pine along theSierra Nevada range front at Cottonwood Creek, there is a rem-nant bench on bedrock resembling a wave-formed platform at�1199 m elevation (site 43; Table 1; Fig. 1). There is a largegully (as much as 20 m deep) below this bench that exposes asequence of strongly indurated fanglomerate overlain by a suc-cession of weathered alluvium.

Sage Flat BenchesErosional features in the form of benches occur at �1182 m

elevation. They are inset within old alluvial-fan deposits east ofSage Flat in southern Owens Valley (site 43, Table 1; Figs. 1 and10C). The benches are relatively smooth and terminate against atrimline that forms what appears to be a gentle to moderate shore-line angle (Fig. 10D). The alluvial-fan deposits that underlie thebenches are locally exposed in railroad cuts and are composedmostly of debris-flow deposits consisting of matrix-supportedsands and silts with minor gravel. In addition, there are less preva-

lent exposures of tightly packed, clast-supported cobble andboulder deposits. These clast-supported deposits may representboulder berms that were formed by reworking of the older fanmaterial along a shoreline (site 44, Table 1; Fig. 10E). If so, thesegravelly, clast-supported deposits are the best example of clasticsediment associated with the 1180 m shoreline on the west side of the valley.

Abrasion Platforms—Inyo Range FrontRemnant erosional landforms that appear to be abrasion

platforms formed by wave erosion are cut on bedrock along thelower range front of the Inyo Mountains between Independenceand just south of Tinemaha Narrows (sites 27–29; Table 1; Figs. 1and 7). These relict abrasion surfaces gently slope basinward andoccur on relatively narrow bedrock noses and isolated knobs thatprotrude from the range front at elevations between �1180 and1184 m (sites 26–29; Table 1; Figs. 1 and 11). Most of these sur-faces are devoid of surficial sediment and occur on Paleozoicmetasedimentary or Mesozoic granitic bedrock. These landformsterminate toward the range front at distinct breaks in slope asrelatively well-preserved, wave-formed notches (sites 27 and 29;Fig. 11), or at saddles between the steeper slopes of the range front(sites 26 and 28; Fig. 11). The morphologic setting of the shore-line platforms indicates that they were developed on headlandsand islands similar to shoreline features in Mono Lake basin (e.g.,Russell, 1889; Reheis et al., 2002a) and Lake Lahontan basin(e.g., Russell, 1885; Adams and Wesnousky, 1998). Developmentof these features likely was influenced by a large fetch and waveenergy directed from the south.

Truncated LandformsSouth of the Swansea pocket barrier locality, along the range

front of the southern Inyo Mountains, there is a relict and heavilydissected alluvial-fan complex composed of late Tertiary (6.7–4.3 Ma) alluvium and interbedded volcanic tuff (Stone et al., 2004)(site 16; Table 1; Figs. 1 and 12A). The base of the fan is abruptlytruncated, forming a nearly linear slope break with distinct facetsat �1180 m elevation (Fig. 12A). We infer that the slope break rep-resents a relict shoreline angle formed by the 1180 m shoreline.

An additional landform that exhibits truncation at �1180 mis the North Long John Canyon slide complex, which is locatedalong the Inyo Mountains range front east of Lone Pine (Blair,1999) (site 26; Table 1; Fig. 1). We infer that the abrupt truncationof this landform represents a shoreline angle, as it is similar inelevation and morphology to other features described along theeastern and southern margins of the valley. Blair (1999) inter-preted the �25 million m3 slide complex as early Holocene (?)age. However, based on a moderately developed soil that in-cludes a Bt horizon (Blair, 1999), rubification, as well as a mutedsurface morphology, we interpret this landslide as older thanLGM and possibly equivalent in age to the �MIS 6–aged glacialoutwash/ alluvial fans mapped in the Fish Springs area by Zehfusset al. (2001).

on July 1, 2015specialpapers.gsapubs.orgDownloaded from

Page 18: Late Quaternary MIS 6–8 shoreline features of pluvial ......Late Quaternary MIS 6–8 shoreline features of pluvial Owens Lake, ... Late Cenozoic Drainage History of the Southwestern

202 Jayko and Bacon

Figure 11. Photos of localities where noticeable platforms are developed on bedded, inclined Paleozoic metasedimentaryrock at 1180 m elevation east of the Owens Valley fault zone along the base of the Inyo Range. We interpret the platforms tobe 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 �1184 m on bedrock atsite 27. (C) Surface at �1181 m at site 28. (D) Surface at �1183 m on bedrock at site 29.

DISCUSSION

Age of the 1180 m Shoreline

The age of the pluvial Owens Lake that formed the �1180 mshoreline is constrained to be younger than the 340 � 60 ka RedMountain flow and older than the 130 � 50 ka basaltic lava flowsof the Big Pine volcanic field, based on crosscutting relations ofthe Sawmill Creek shoreline, Taboose Creek strath surfaces, andthe lack of features on the younger lava flow. Thus, the shorelinedeveloped sometime between MIS 8 and 6, which occurred dur-ing the intervals of 280–260 ka and 185–130 ka, respectively(Winograd et al., 2006). The 36Cl model age of ca. 160 � 32 ka(CHLOE production parameters; Phillips et al., 2001) or 218 �33 ka (calcium spallation rate of Stone et al., 1996) also supportsthe MIS 8–6 age as determined from crosscutting relationsbetween dated lava flows.

The degree of soil development in the form of rubification,presence of pedogenic clay, and weathering rinds provide a mea-

sure of relative age. Soils and weathered clasts found within sedi-ment deposited by the 1180 m shoreline at the Centennial Flatembayment and sediment deposited prior to, or coeval with, theshoreline at Taboose and Sawmill Creeks, at the Alabama Hillsstrath terraces, and on the North Long John Canyon slide complexappear to have similar weathering characteristics. These depositsalso have soil development similar to that described on alluvialfans south of Big Pine at Fish Springs, where cosmogenic 10Beand 26Al dates on boulders yielded ages that range from 171 �21 ka to 136 � 17 ka (Zehfuss et al., 2001).

An additional age estimate for the 1180 m shoreline may bedetermined by dividing the amount of measured displacementbetween shoreline features across the Owens Valley fault zone bythe average vertical slip rate of �0.12 m/k.y. on the Owens Val-ley fault zone at Lone Pine (from Bacon and Pezzopane, 2007) to solve for time. From this calculation, the time necessary toaccommodate �20 � 4 m of apparent vertical deformation acrossthe Owens Valley fault zone ranges from �200 to 130 k.y. Thisindependent age estimate of the 1180 m shoreline is within the

on July 1, 2015specialpapers.gsapubs.orgDownloaded from

Page 19: Late Quaternary MIS 6–8 shoreline features of pluvial ......Late Quaternary MIS 6–8 shoreline features of pluvial Owens Lake, ... Late Cenozoic Drainage History of the Southwestern

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

Figure 12. (A) Aerial photograph showing abrupt truncation of Tertiaryalluvial fans along the southern Inyo Range at �1180 m elevation (site 16;Fig. 1). (B) Aerial photograph showing truncation of the Long John Can-yon fan slide of Blair (1999) along the southern Inyo Range at �1180 melevation (site 25; Fig. 1).

range of constraining ages from dated lava flows, and it is simi-lar to the 36Cl model age on tufa.

Correlation to Sierran Glacial Chronologies

The penultimate glaciation in the Sierra Nevada, referred toas the Tahoe glaciation, had two major advances, which are re-ferred to as the younger Tahoe (Tahoe II) and older Tahoe (Tahoe I)(e.g., Kaufman et al., 2004). The younger Tahoe advance is esti-mated at ca. 50–42 ka from 36Cl cosmogenic exposure ages onboulders from moraine crests at Bloody Canyon in Mono Basin(Phillips et al., 2001; Kaufman et al., 2004; corrected from ca. 66–55 ka reported in Phillips et al., 1990). Boulders on olderTahoe moraines range between 218 and 133 ka in age (Phillips et al., 1990; Kaufman et al., 2004) (Fig. 1). Along the range to the south, the younger Tahoe is constrained between 119 ka and53 ka and is therefore most likely MIS 4 in age (Gillespie, 1982;Gillespie et al., 1984). Near Sawmill Canyon, located west of the

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

Owens Lake Core OL-92 during MIS 6–8

Depths to MIS 6–8 horizons in the Owens Lake core OL-92have been identified by paleoclimate proxies including lithologic,chemical, mineralogic, geophysical, and paleontological data(Smith and Bischoff, 1997). The cycling between closed- andopen-lake systems defined both by paleoecology and mineralogyof lacustrine sediment refines the duration and magnitude of thecycles between interglaciations and glaciations. The Owens Lakecore reached a depth of 323 m, penetrating the Bishop tuff (ca.760 ka) at �304–309 m depth, and it terminated near the Brunhes-Matayama magnetic reversal (ca. 783 ka) at �310–320 m depth(Smith and Bischoff, 1997). Sediment correlated with MIS 6ranges from �80–92 m depth, whereas sediment correspondingto MIS 8 ranges from �111–131 m depth. The absence of fishremains in the core between �96 and 112 m depth combined withthe 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 strongerand longer duration of cold and freshwater during MIS 6 than dur-ing MIS 8 (Bradbury, 1997; Carter, 1997).

In addition to biologic proxy indicators, the total inorganiccontent (TIC) and cation exchange capacity (CEC) of the clayfraction 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 incombination with depth-age estimates in the core, full glacial con-ditions in Owens Valley began at ca. 190 ka (MIS 6/7) and endedat ca. 120 ka (MIS 5/6) (Bischoff et al., 1997a). Other proxy indi-cators in core OL-92, such as the presence of rock flour (a glacialerosional by-product) and salinity, support the TIC and CEC vari-ations. These additional proxies also show that between ca. 260and 220 ka, there were three short-duration cold intervals, withthe coldest around 220 ka (Bischoff et al., 1997c; Menking 1997),which is also strongly expressed in the Devils Hole and sea-surfacetemperature paleoclimate records (Winograd et al., 2006).

Importance for Aquatic Species

The most recent overflow of the Owens River system viaSearles and Panamint Valleys to its terminus in Death Valleyprobably last occurred during MIS 6 or older time (Jayko et al.,this volume). U-series tufa dates from the highstand Blackweldershoreline in Death Valley range in age from 216 to 120 ka (Hooke

on July 1, 2015specialpapers.gsapubs.orgDownloaded from

Page 20: Late Quaternary MIS 6–8 shoreline features of pluvial ......Late Quaternary MIS 6–8 shoreline features of pluvial Owens Lake, ... Late Cenozoic Drainage History of the Southwestern

204 Jayko and Bacon

and Dorn, 1992; Ku et al., 1998, Hooke, 1999). Sediment coredata from near Badwater in Death Valley indicate a large lakestand at MIS 6 and a smaller lake at MIS 2 but mainly ground-water-supported wetlands during MIS 4 (Lowenstein et al., 1999;Forester et al., 2005). Results from Pleistocene pluvial studies atMono Basin indicate that Lake Russell may have spilled south-ward through Adobe Valley to Owens River during MIS 6 (Put-nam, 1947, personal commun. to Hubbs, in Hubbs and Miller,1948) and (or) MIS 4 (Reheis et al., 2002a). Thus, during MIS 6(older Tahoe), there may have been a connection from Mono LakeBasin to Death Valley, as initially speculated by Blackwelder(1933). Following MIS 6 time, water from the Owens River sys-tem made it no farther than Searles Valley during MIS 4 (Jayko et al., this volume) and as far downstream as Panamint Valley dur-ing MIS 2 (Smith et al., 1983; Jayko et al., this volume). Fish inthe Owens River system have clearly been isolated from the othersouthern and eastern populations in the Mojave and Amargosadrainage basins for at least �120 k.y., since the waning of MIS 6(Jayko et al., this volume).

Previous paleontologic and geomorphic work has shown thatthe Owens River system was linked to the Lahontan watershed inthe north (Russell, 1889; Miller, 1946; Hubbs and Miller, 1948;Smith et al., 2002; Reheis et al., 2002a). Fish remains includingCatostomus (sucker) and Siphateles (tui chub) from OL-92 andolder sediments in the region indicate that the Lahontan fluvialconnection is older than the Bishop Tuff, dated at ca. 760 ka(Firby et al., 1997; Miller, 1946; Hubbs and Miller, 1948). Also,Prosopium (whitefish) and Oncorrhynchus (trout) at the base ofOL-92 core between ca. 730 and 695 ka suggest colder conditionsat this time (Firby et al., 1997). These cold-water Salmonidae(trout) species clearly persisted following the eruption of theLong Valley caldera (Firby et al., 1997), which deposited theBishop Tuff, but they apparently could not survive subsequentenvironmental and (or) tectonic changes, which may haveincluded eruptive activity in the Big Pine volcanic field or disrup-tion of the southward drainage from the Mono Basin watershed(cf. Russell, 1889).

SUMMARY AND CONCLUSIONS

Deformed shoreline features and strath terraces graded to theshorelines at elevations between �1180 and 1200 m are youngerthan 340 � 60 ka and older than 130 � 50 ka, based on crosscut-ting relations with dated lava flows of the Big Pine volcanic field.Therefore, this phase of pluvial Owens Lake existed within thetime interval between MIS 8 and MIS 6. Additional factors refinethe age of the pluvial lake to approximately MIS 6 time. Theseinclude: (1) a 36Cl model age of 160 � 32 ka on tufa from south-ern Owens Valley (F. Phillips, 2007, personal commun.; Table 2;Fig. 3E); and (2) an age based on a vertical slip rate on the OwensValley fault zone at Lone Pine in combination with the apparentvertical displacement of deformed shorelines across the OwensValley fault zone. Such a correlation is consistent with the rela-tive duration and intensity of MIS 6 as indicated by proxy indica-

tors in core OL-92 and the timing of the older Tahoe (Tahoe I)glaciation in the southern Sierra Nevada. This pluvial OwensLake was much larger and deeper than preceding younger lakesin Owens Valley. It had a minimum area of at least �880 km2,based on the map distribution of described features in Owens Val-ley, but the exact location of the spillway has yet to be resolved,so the full extent of the lake is unknown. Such a large, fresh lakeduring MIS 6 is consistent with environmental interpretations offaunal and mineralogic indicators identified in core OL-92 (Brad-bury, 1997; Carter, 1997; Bischoff et al., 1997c; Menking, 1997).

Landscape evolution and rates of geomorphic processes canbe quantified where there is sufficient age control on importantlandscape-modifying events, such as lake-level fluctuationswithin large lake systems during glacial-pluvial cycles. Recogni-tion and documentation of climate-induced landscape changeprovide a baseline for studies that constrain other processes suchas erosion, fault-slip rates, and water balance within a region. Thisstudy has documented the distribution of shoreline and fluvialfeatures that outline the highstand of a large ancestral pluvial lakeof probable MIS 6 age that had not previously been documented.As with many large ancestral pluvial lakes in the western UnitedStates, which are typically named after pioneers of their field, wepropose to name the MIS 6 pluvial Owens Lake as Lake GI, inrecognition of George I. Smith’s investigations in Owens Valley,as well as other pluvial lakes within the Owens River system, overthe past 50 yr.

ACKNOWLEDGMENTS

We owe a special thanks to Fred Phillips for graciously pro-viding the 36Cl date from site 10 at 1170 m elevation, pro bono,when he learned about the results of this study. We also thank AlanGillespie for directing us to published sources of some of the dataso often cited elsewhere in the Turrin and Gillespie (1986)abstract. We thank Wally Woolfenden, U.S. Forest Service; AlanGillespie, University of Washington; and Scott Stine, CaliforniaState University–Hayward for constructive and insightful reviewsthat helped strengthen the manuscript. We additionally thank MarithReheis for thorough editing and helpful comments. We acknowl-edge Silvio Pezzopane for engaging discussions on the 1180 mshoreline in Owens Valley. S. Bacon acknowledges Eric McDonald,Desert Research Institute, for partial support (Army Research OfficeContract No. DAAD19-03-1-0159). We also gratefully acknowl-edge lively discussions with many Quaternary geologists who pio-neered the modern lacustrine studies in the region during PacificCell–Friends of the Pleistocene field trips between 1996 and 2002.

REFERENCES CITED

Adams, K.D., and Wesnousky, S.G., 1998, Shoreline processes and the age of theLake Lahontan highstand in the Jessup embayment, Nevada: GeologicalSociety of America Bulletin, v. 110, p. 1318–1332, doi: 10.1130/0016-7606(1998)110�1318:SPATAO�2.3.CO;2.

Bacon, S.N., and Pezzopane, S.K., 2007, A 25,000-year record of earthquakes onthe Owens Valley fault near Lone Pine, California: Implications for recur-

on July 1, 2015specialpapers.gsapubs.orgDownloaded from

Page 21: Late Quaternary MIS 6–8 shoreline features of pluvial ......Late Quaternary MIS 6–8 shoreline features of pluvial Owens Lake, ... Late Cenozoic Drainage History of the Southwestern

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

rence intervals, slip rates, and segmentation models: Geological Society ofAmerica Bulletin, v. 119, p. 823–847, doi: 10.1130/B25879.1, doi: 10.1130/B25879.1.

Bacon, S.N., Jayko, A.S., and McGeehin, J.P., 2005, Holocene and latest Pleis-tocene oblique dextral faulting on the southern Inyo Mountains fault, OwensLake basin, California: Bulletin of the Seismological Society of America, v. 95, p. 2472–2485, doi: 10.1785/0120040228.

Bacon, S.N., Burke, R.M., Pezzopane, S.K., and Jayko, A.S., 2006, Last GlacialMaximum and Holocene lake levels of Owens Lake, eastern California,USA: Quaternary Science Reviews, v. 25, p. 1264–1282, doi: 10.1016/j.quascirev.2005.10.014.

Beanland, S., and Clark, M.M., 1994, The Owens Valley Fault Zone, Eastern Cali-fornia, and Surface Rupture Associated with the 1872 Earthquake: U.S.Geological Survey Bulletin 1982, 29 p.

Benn, D.I., Owen, L.A., Finkel, R.C., and Clemmons, S., 2006, Pleistocene lakeoutbursts floods and fan formation along the eastern Sierra Nevada, Cali-fornia: Implications for the interpretation of intermontane lacustrinerecords: Quaternary Science Reviews, v. 25, p. 2729–2748, doi: 10.1016/j.quascirev.2006.02.018.

Benson, L.V., 1994, Carbonate deposition, Pyramid Lake subbasin, Nevada: 1. Se-quence of formation and altitudinal distribution of carbonate deposits (tufas):Palaeogeography, Palaeoclimatology, Palaeoecology, v. 109, p. 55–87, doi:10.1016/0031-0182(94)90118-X.

Bierman,P.R.,Gillespie,A.R.,andCaffee,M.W.,1995,Cosmogenicagesforearth-quake recurrence intervals and debris flow fan deposition, Owens Valley,California: Science, v. 270, p. 447–450, doi: 10.1126/science.270.5235.447.

Bischoff, J.L., Fitts, J.P., and Fitzpatrick, J.A., 1997a, Responses of sediment geo-chemistry to climate change in Owens Lake sediment: An 800-k.y. record ofsaline/fresh cycles in core OL-92, in Smith, G.I., and Bischoff, J.L., eds., An800,000-Year Paleoclimatic Record from Core OL-92, Owens Lake, South-east California: Geological Society of America Special Paper 317, p. 37–48.

Bischoff, J.L., Stafford, T.W., Jr., and Rubin, M., 1997b, A time-depth scale forOwens Lake sediments of core OL-92: Radiocarbon dates and constantmass-accumulation: An 800-k.y. record of saline/fresh cycles in core OL-92,in Smith, G.I., and Bischoff, J.L., eds., An 800,000-Year PaleoclimaticRecord from Core OL-92, Owens Lake, Southeast California, GeologicalSociety of America Special Paper 317, p. 91–98.

Bischoff, J.L., Menking, K.M., Fitts, J.P., and Fitzpatrick, J.A., 1997c, Climaticoscillations 10,000–155,000 yr B.P. at Owens Lake, California, reflected inglacial rock flour abundance and lake salinity in core OL-92: QuaternaryResearch, v. 48, p. 313–325, doi: 10.1006/qres.1997.1933.

Blackwelder, E., 1933, Lake Manly: An extinct lake of Death Valley: Geographi-cal Review, v. 23, p. 464–471, doi: 10.2307/209632.

Blair, T.C., 1999, Alluvial fan and catchment initiation by rock avalanching,Owens Valley, California: Geomorphology, v. 28, p. 201–221, doi: 10.1016/S0169-555X(98)00111-1.

Blair, T.C., 2001, Outburst flood sedimentation on the proglacial Tuttle Canyonalluvial fan, Owens Valley, California, U.S.A.: Journal of SedimentologicalResearch, v. 71, p. 657–679.

Bradbury, J.P., 1997, A diatom-based paleohydrologic record of climate changefor the past 800 k.y. from Owens Lake, California, in Smith, G.I., andBischoff, J.L., eds., An 800,000-Year Paleoclimatic Record from Core OL-92, Owens Lake, Southeast California: Geological Society of America Spe-cial Paper 317, p. 99–112.

Bryant, W.A., 1984, Evidence of recent faulting along the Owens Valley, RoundValley, and White Mountains frontal fault zones, Inyo and Mono Counties,California: California Division of Mines and Geology Open-File ReportOFR-84–54SAC, scale 1:24,000.

Bull, W.B., 1991, Geomorphic Responses to Climatic Change: New York, OxfordUniversity Press, 326 p.

Carter, C., 1997, Ostracodes in Owens Lake core OL-92: Alternation of saline andfreshwater forms through time, in Smith, G.I., and Bischoff, J.L., eds., An 800,000-Year Paleoclimatic Record from Core OL-92, Owens Lake,Southeast California: Geological Society of America Special Paper 317, p. 113–120.

Carver, G.A., 1970, Quaternary Tectonism and Surface Faulting in Owens LakeBasin, California: Reno, University of Nevada, MacKay School of MinesTechnical Report AT2, 103 p.

Connor, C.B., and Conway, F.M., 2000, Basaltic volcanic fields, in Sigurdsson,H., et al., eds., Encyclopedia of Volcanoes: San Diego, California, Acade-mic Press, p. 331–344.

Cox, A., Doell, R.R., and Dalrymple, G.B., 1963, Geomagnetic polarity epochs—Sierra Nevada II: Science, v. 142, p. 382–385, doi: 10.1126/science.142.3590.382-a.

Dorn, R.I., Turrin, B.D., Jull, A.J.T., Linick, T.W., and Donahue, D.J., 1987,Radiocarbon and cation-ratio ages for rock varnish on Tioga and Tahoemorainal boulders of Pine Creek, eastern Sierra Nevada, California, andtheir paleoclimatic implications: Quaternary Research, v. 28, p. 36–49, doi:10.1016/0033-5894(87)90031-7.

Duffield, W.A., and Smith, G.I., 1978a, Pleistocene history of volcanism and theOwens River near Little Lake, California: U.S. Geological Survey Journalof Research, v. 6, p. 395–408.

Duffield, W.A., and Smith, G.I., 1978b, Pleistocene river erosion and intracanyonlava flows near Little Lake, Inyo County, California: California Geology, v. 31, p. 81–89.

Firby, J.R., Sharpe, S.E., Whelan, J.F., Smith, G.R., and Spaulding, W.G., 1997,Paleobiotic and isotopic analysis of mollusks, fish, and plants from core OL-92, in Smith, G.I., and Bischoff, J.L., eds., An 800,000-Year Paleo-climatic Record from Core OL-92, Owens Lake, Southeast California: Geo-logical Society of America Special Paper 317, p. 121–126.

Forester, R.M., Lowenstein, T.K., and Spencer, R.J., 2005, An ostracode basedpaleolimnologic and paleohydrologic history of Death Valley, 200 to 0 ka:Geological Society of America Bulletin, v. 117, p. 1379–1386, doi: 10.1130/B25637.1.

Gale, H.S., 1914, Salines in the Owens, Searles, and Panamint Basins, Southeast-ern California: U.S. Geological Survey Bulletin 580, p. 281–323.

Gillespie, A.R., 1982, Quaternary Glaciation and Tectonism in the SouthwesternSierra Nevada, Inyo County, California [Ph.D. dissertation]: Pasadena, Cali-fornia Institute of Technology, 657 p.

Gillespie, A.R., Huneke, J.C., and Wasserburg, G.J., 1983, Eruption age of a Pleis-tocene basalt from 40Ar-39Ar analysis of partially degassed xenoliths: Jour-nal of Geophysical Research, v. 88, p. 4997–5008.

Gillespie, A.R., Huneke, J.C., and Wasserburg, G.J., 1984, Eruption age of a�100,000-year-old basalt from 40Ar-39Ar analysis of partially degassedxenoliths: Journal of Geophysical Research, v. 89, p. 1033–1048.

Hooke, R.L., 1999, Lake Manly(?) shorelines in the eastern Mojave Desert, Cali-fornia: Quaternary Research, v. 52, p. 328–336, doi: 10.1006/qres. 1999.2080.

Hooke, R.L., and Dorn, R.I., 1992, Segmentation of alluvial fans in Death Valley,California; New insights from surface-exposure dating and laboratory mod-eling: Earth Surface Processes and Landforms, v. 17, p. 557–574, doi:10.1002/esp.3290170603.

Hubbs, C.L., and Miller, R.R., 1948, The Great Basin with emphasis on glacialand postglacial times: Bulletin of the University of Utah, v. 38, p. 19–166.

Jannik, N.O., Phillips, F.M., Smith, G.I., and Elmore, D., 1991, A 36Cl chronol-ogy of lacustrine sedimentation in the Pleistocene Owens River system:Geological Society of America Bulletin, v. 103, p. 1146–1159, doi: 10.1130/0016-7606(1991)103�1146:ACCOLS�2.3.CO;2.

Jayko, A.S., 2008, Surficial Geology of the Darwin Hills 30� � 60� Quadrangle,Inyo County, California: U.S. Geological Survey Map Series, scale1:100,000.

Jayko, A.S., Forester, R.M., Kaufman, D.S., Phillips, F.M., Yount, J.C., McGee-hin, J., and Mahan, S.A., 2008, this volume, Late Pleistocene lakes,Panamint Valley, California, in Reheis, M.C., Hershler, R., and Miller,D.M., eds., Late Cenozoic Drainage History of the Southwestern GreatBasin and Lower Colorado River Region: Geologic and Biotic Perspec-tives: Geological Society of America Special Paper 439, doi: 10.1130/2008.2439(07).

Kaufman, D.S., Porter, S.C., and Gillespie, A.R., 2004, Quaternary alpine glacia-tion in Alaska, the Pacific Northwest, Sierra Nevada, and Hawaii, in Gille-spie, A.R., Porter, S.C., and Atwater, B.F., eds., The Quaternary Period in the

on July 1, 2015specialpapers.gsapubs.orgDownloaded from

Page 22: Late Quaternary MIS 6–8 shoreline features of pluvial ......Late Quaternary MIS 6–8 shoreline features of pluvial Owens Lake, ... Late Cenozoic Drainage History of the Southwestern

206 Jayko and Bacon

United States: Elsevier, Netherlands, Developments in Quaternary Science,v. 1, p. 77–103.

Ku, T.L., Luo, S., Lowenstein, T.K., Li, J., and Spencer, R.J., 1998, U-serieschronology of lacustrine deposits in Death Valley, California: QuaternaryResearch, v. 50, p. 261–275, doi: 10.1006/qres.1998.1995.

Lal, D., 1991, Cosmic-ray labeling of erosion surfaces; in situ production rates anderosion models: Earth and Planetary Science Letters, v. 104, p. 424–439,doi: 10.1016/0012-821X(91)90220-C.

Lowenstein, T.K., Li, J., Brown, C., Roberts, S.M., Ku, T.L., Luo, S., andYang, W.,1999, 200 k.y. paleoclimate record from Death Valley salt core: Geology,v.27,p.3–6,doi:10.1130/0091-7613(1999)027�0003:KYPRFD�2.3.CO;2.

Lubetkin, L.K.C., and Clark, M.M., 1988, Late Quaternary activity along the LonePine fault, eastern California: Geological Society of America Bulletin, v. 100,p. 755–766, doi: 10.1130/0016-7606(1988)100�0755:LQAATL�2.3.CO;2.

Martel, S.J., Harrison, T.M., and Gillespie, A.R., 1987, Late Quaternary verticaldisplacement rate across the Fish Springs fault, Owens Valley fault zone,California: Quaternary Research, v. 27, p. 113–129, doi: 10.1016/0033-5894(87)90071-8.

Menking, K.M., 1997, Climatic signals in clay mineralogy and grain-size varia-tions in Owens Lake core OL-92, in Smith, G.I., and Bischoff, J.L., eds., An800,000-Year Paleoclimatic Record from Core OL-92, Owens Lake, South-east California: Geological Society of America Special Paper 317, p. 25–36.

Miller, R.R., 1946, Correlation between fish distribution and Pleistocene hydrog-raphy in eastern California and southwestern Nevada, with a map of thePleistocene waters: The Journal of Geology, v. 54, p. 43–53.

Miller, R.R., 1948, The Cyprinodont Fishes of the Death Valley System of East-ern California and Southwestern Nevada: University of Michigan Museumof Zoology Miscellaneous Publication 68, 155 p.

Orme, A.J., and Orme, A.R., 2008, this volume, Late Pleistocene climate and tec-tonic impacts on Owens Lake levels and linkages, in Reheis, M.C., Hersh-ler, R., and Miller, D.M., eds., Late Cenozoic Drainage History of theSouthwestern Great Basin and Lower Colorado River Region: Geologic andBiotic Perspectives: Geological Society of America Special Paper 439, doi:10.1130/2008.2439(09).

Pakiser, L.C., Kane, M.F., and Jackson, W.H., 1964, Structural Geology and Vol-canism of Owens Valley Region, California—A Geophysical Study: U.S.Geological Survey Professional Paper 438, 65 p.

Phillips, F.M., and Plummer, M.A., 1996, CHLOE: A program for interpreting in-situ cosmogenic nuclide data for surface exposure dating and erosionstudies, in Jull, A.J.T., 7th International Conference Accelerator Mass Spec-trometry Abstracts: Radiocarbon, v. 38, p. 98–99.

Phillips, F.M., Zreda, M.G., Smith, S.S., Elmore, D., Kubik, P.W., and Sharma, P.,1990, Cosmogenic chlorine-36 chronology for glacial deposits at BloodyCanyon, eastern Sierra Nevada: Science, v. 248, p. 1529–1532, doi:10.1126/science.248.4962.1529.

Phillips, F.M., Stone, W.D., and Fabryka-Martin, J.T., 2001, An improvedapproach to calculating low-energy cosmic-ray neutron fluxes near theland/atmosphere interface: Chemical Geology, v. 175, p. 689–701, doi:10.1016/S0009-2541(00)00329-6.

Reheis, M., 1999, Highest pluvial-lake shorelines and Pleistocene climate of thewestern Great Basin: Quaternary Research, v. 52, p. 196–205, doi: 10.1006/qres.1999.2064.

Reheis, M.C., and Dixon, T.H., 1996, Kinematics of the Eastern California shearzone: Evidence of slip transfer from Owens and Saline Valley fault zonesto Fish Lake Valley fault zone: Geology, v. 24, p. 339–342, doi:10.1130/0091-7613(1996)024�0339:KOTECS�2.3.CO;2.

Reheis, M.C., Stine, S., and Sarna-Wojcicki, A.M., 2002a, Drainage reversals inMono Basin during the late Pliocene and Pleistocene: Geological Society of

America Bulletin, v. 114, p. 991–1006, doi: 10.1130/0016-7606(2002)114�0991:DRIMBD�2.0.CO;2.

Reheis, M.C., Sarna-Wojcicki, A.M., Reynolds, R.L., Repenning, C.A., and Mif-flin, M.D., 2002b, Pliocene to middle Pleistocene lakes in the western GreatBasin: Ages and connections, in Hershler, R., Madsen, D.B., and Currey,D.R., eds., Great Basin Aquatic Systems History: Smithsonian Contribu-tions to the Earth Sciences, v. 33, p. 53–108.

Russell, I.C., 1885, Geological History of Lake Lahontan, a Quaternary Lake inNorthwestern, Nevada: U.S. Geological Survey Monograph 11, 288 p.

Russell, I.C., 1889, Quaternary History of Mono Valley, California: Washington,D.C., Eighth Annual Report of the U.S. Geological Survey, p. 267–394.

Savage, J.C., and Lisowski, M., 1995, Strain accumulation in Owens Valley: Seis-mological Society of America Bulletin, v. 85, p. 151–158.

Slemmons, D.B., Vittori, E., Jayko, A.S., Carver, G.A., and Bacon, S.N., 2008,Quaternary Fault and Lineament Map of Owens Valley, Inyo County, East-ern California: Geological Society of America Map and Chart SeriesMCH096, 25 p. (in press).

Smith, G.I., and Bischoff, J.L., 1997, Core OL-92 from Owens Lake: Project ratio-nale, geologic setting, drilling procedures, and summary, in Smith, G.I., andBischoff, J.L., eds., An 800,000-Year Paleoclimatic Record from Core OL-92, Owens Lake, Southeast California: Geological Society of America Spe-cial Paper 317, p. 1–8.

Smith, G.I., Barczak, V.J., Moulton, G.F., and Liddicoat, J.C., 1983, Core KM-3,a Surface-to-Bedrock Record of Late Cenozoic Sedimentation in SearlesValley, California: U.S. Geological Survey Professional Paper 1256, p. 1–24.

Smith, G.R., Dowling, T.E., Gobalet, K.W., Lugaski, D.K., Shiozawa, D.K., andEvans, R.P., 2002, Biogeography and timing of evolutionary events amongGreat Basin fishes, in Hershler, R., Madsen, D.B., and Currey, D.R., eds.,Great Basin Aquatic Systems History: Smithsonian Contributions to theEarth Sciences, v. 33, p. 175–234.

Stinson, M.C., 1977, Geology of the Keeler 15� Quadrangle, Inyo County, Califor-nia: California Division of Mines and Geology Map Sheet 38, scale 1:62,500.

Stone, J.O., Allan, G.L., Fifield, L.K., and Cresswell, R.G., 1996, Cosmogenicchlorine-36 from calcium spallation: Geochimica et Cosmochimica Acta, v. 60, p. 679–692, doi: 10.1016/0016-7037(95)00429-7.

Stone, P., Dunne, G.C., Conrad, J.E., Swanson, B.J., Stevens, C.H., and Valin,V.C., 2004, Geologic Map of the Cerro Gordo Peak 7.5� Quadrangle, InyoCounty, California: U.S. Geological Survey Geologic Scientific Investiga-tion Map I-2851, scale 1:24,000.

Turrin, B., and Gillespie, A., 1986, K/Ar ages of basaltic volcanism of the Big Pinevolcanic field, California: Implications for glacial stratigraphy and neotec-tonics of the Sierra Nevada: Geological Society of America Abstracts withPrograms, v. 18, no. 6, p. 777.

Winograd, I.J., Landwehr, J.M., Coplen, T.B., Sharp, W.D., Riggs, A.C., Ludwig,K.R., and Kolesar, P.T., 2006, Devils Hole, Nevada, 18O record extendedto the mid-Holocene: Quaternary Research, v. 66, p. 202–212, doi: 10.1016/j.yqres.2006.06.003.

Zehfuss, P.H., Bierman, P.R., Gillespie, A.R., Burke, R.M., and Caffe, M.W.,2001, Slip rates on the Fish Springs fault, Owens Valley, California, de-duced from cosmogenic 10Be and 26Al and soil development on fans: Geo-logical Society of America Bulletin, v. 113, p. 241–255, doi: 10.1130/0016-7606(2001)113�0241:SROTFS�2.0.CO;2.

Zreda, M.G., 1994, Development and Calibration of the Cosmogenic 36Cl SurfaceExposure Dating Method and Its Application to the Chronology of LateQuaternary Glaciations [Ph.D. thesis]: Socorro, New Mexico Institute ofMining and Technology, 318 p.

MANUSCRIPT ACCEPTED BY THE SOCIETY 17 JULY 2007

Printed in the USA

on July 1, 2015specialpapers.gsapubs.orgDownloaded from

Page 23: Late Quaternary MIS 6–8 shoreline features of pluvial ......Late Quaternary MIS 6–8 shoreline features of pluvial Owens Lake, ... Late Cenozoic Drainage History of the Southwestern

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

  A.S. Jayko and S.N. Bacon  eastern California

8 shoreline features of pluvial Owens Lake, Owens Valley,−Late Quaternary MIS 6  Geological Society of America Special Papers

  E-mail alerting services

  article

to receive free e-mail alerts when new articles cite thiswww.gsapubs.org/cgi/alertsclick

  Subscribe

  Papers

to subscribe to Geological Society of America Specialwww.gsapubs.org/subscriptionsclick

  Permission request

  to contact GSA.www.geosociety.org/pubs/copyrt.htm#gsaclick

reflect official positions of the Society.of their race, citizenship, gender, religion, or political viewpoint. Opinions presented in this publication do notthis and other forums for the presentation of diverse opinions and positions by scientists worldwide, regardless

providestheir organization's Web site providing the posting includes a reference to the article's full citation. GSA file may not be posted to any Web site, but authors may post the abstracts only of their articles on their own orcopies of items in GSA's journals for noncommercial use in classrooms to further education and science. This use a single figure, a single table, and/or a brief paragraph of text in subsequent works and to make unlimitedemployment. Individual scientists are hereby granted permission, without fees or further requests to GSA, to Copyright not claimed on content prepared wholly by U.S. government employees within scope of their

Notes

Geological Society of America

on July 1, 2015specialpapers.gsapubs.orgDownloaded from