<<

See discussions, stats, and author profiles for this publication at: https://www.researchgate.net/publication/262806093

Pluvial in the of the western —a view from the outcrop

Article in Science Reviews · August 2014 DOI: 10.1016/j.quascirev.2014.04.012

CITATIONS READS 28 172

4 authors, including:

Marith C. Reheis United States Geological Survey

119 PUBLICATIONS 3,485 CITATIONS

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

fun in retirement! View project

Tephrochronology Project View project

All content following this page was uploaded by Marith C. Reheis on 26 April 2016.

The user has requested enhancement of the downloaded file. All in-text references underlined in blue are added to the original document and are linked to publications on ResearchGate, letting you access and read them immediately. Quaternary Science Reviews 97 (2014) 33e57

Contents lists available at ScienceDirect

Quaternary Science Reviews

journal homepage: www.elsevier.com/locate/quascirev

Invited review Pluvial lakes in the Great Basin of the western United Statesda view from the outcrop

Marith C. Reheis a,*, Kenneth D. Adams b, Charles G. Oviatt c, Steven N. Bacon b a U.S. Geological Survey, MS-980, Federal Center Box 25046, Denver, CO 80225, USA b Research Institute, 2215 Raggio Parkway, Reno, NV 89512, USA c Department of Geology, Kansas State University, 108 Thompson Hall, Manhattan, KS 66506, USA article info abstract

Article history: Paleo-lakes in the western United States provide geomorphic and hydrologic records of climate and Received 1 March 2013 drainage-basin change at multiple time scales extending back to the . Recent reviews and studies Received in revised form of paleo- records have focused on interpretations of proxies in lake sediment cores from the northern 7 April 2014 and central parts of the Great Basin. In this review, emphasis is placed on equally important studies of Accepted 11 April 2014 lake history during the past w30 years that were derived from outcrop exposures and geomorphology, in Available online some cases combined with cores. Outcrop and core records have different strengths and weaknesses that must be recognized and exploited in the interpretation of paleohydrology and paleoclimate. Outcrops Keywords: and landforms can yield direct evidence of lake level, facies changes that record details of lake-level fl fl Great Basin uctuations, and geologic events such as catastrophic oods, drainage-basin changes, and isostatic Paleoclimate rebound. Cores can potentially yield continuous records when sampled in stable parts of lake basins and history can provide proxies for changes in lake level, water temperature and chemistry, and ecological conditions Quaternary in the surrounding landscape. However, proxies such as stable isotopes may be influenced by several Outcrop competing factors the relative effects of which may be difficult to assess, and interpretations may be confounded by geologic events within the drainage basin that were unrecorded or not recognized in a core. The best evidence for documenting absolute lake-level changes lies within the shore, nearshore, and deltaic sediments that were deposited across piedmonts and at the mouths of streams as lake level rose and fell. We review the different shorezone environments and resulting deposits used in such re- constructions and discuss potential estimation errors. Lake-level studies based on deposits and landforms have provided paleohydrologic records ranging from general changes during the past million years to centennial-scale details of fluctuations during the late and . Outcrop studies have documented the integration histories of several important drainage basins, including the Humboldt, Amargosa, Owens, and systems, that have evolved since the Miocene within the active tectonic setting of the Great Basin; these histories have influenced lake levels in terminal basins. Many pre-late Pleistocene lakes in the western Great Basin were significantly larger and record wetter conditions than the youngest lakes. Outcrop-based lake-level data provide important checks on core-based proxy interpretations; we discuss four such comparisons. In some cases, such as for Lakes Owens and Manix, outcrop and core data synthesis yields stronger and more complete records; in other cases, such as for Bonneville and Lahontan, conflicts point toward reconsideration of confounding factors in interpretation of core-based proxies. Published by Elsevier Ltd.

1. Introduction Russell (1885) and Gilbert (1890) in the late 19th century. The sediments and landforms of pluvial lakes found in these basins are Pluvial lakes in the western United States have been studied sensitive recorders of paleohydrologic conditions and provide in- intensively for over 125 years, beginning with the classic studies of formation on the magnitudes and rates of climatic change that influence lake level. Lake sediments also reveal sedimentary and geomorphic processes and can record geologic events fl * Corresponding author. such as drainage-basin changes, oods, deposition, E-mail address: [email protected] (M.C. Reheis). ground-rupturing earthquakes, and crustal rebound. From a http://dx.doi.org/10.1016/j.quascirev.2014.04.012 0277-3791/Published by Elsevier Ltd. 34 M.C. Reheis et al. / Quaternary Science Reviews 97 (2014) 33e57 biogeomorphic perspective, the distribution of many aquatic spe- A recent review of paleohydrologic and paleoclimatic in- cies in this region only makes sense when past hydrologic con- terpretations for the central and northern Great Basin focused on nections are understood. In addition, study of these basins provides core data (Benson, 2004), but outcrop-based work in the same area important paleoenvironmental context for archaeological sites and studies of lakes from the southern Great Basin (Fig. 1) have not found along the ancient lake margins. received similar attention. Our goal is to review notable advances Many different approaches have been used to decipher the made in the past few decades using outcrop studies of pluvial lake histories of pluvial basins. Periodic reviews of the state of paleo- deposits in the Great Basin and to provide examples of studies that lake research in the western U.S. have been published by yield significant insights on the interplay between Quaternary cli- Morrison (1965), Reeves (1968), Smith and Street-Perrott (1983), matic and geologic controls on lake records. Such insights include Benson and Thompson (1987), Currey (1990), Negrini (2002), documentation of: (1) many pre-late Pleistocene lakes in the Benson (2004), Orme (2008), and Grayson (2011). In general, in- western Great Basin that were significantly larger and record dividual lake studies published prior to the 1970s primarily relied wetter conditions than the late Pleistocene lakes (Reheis et al., on interpretation of sediments preserved in outcrop (e.g., Russell, 2002a; Redwine, 2003); (2) drainage-basin integration by lake 1885; Gilbert, 1890; Blackwelder and Ellsworth, 1936; Morrison, overflow (e.g., Reheis et al., 2002a; House et al., 2008); (3) 1964; Lajoie, 1968). After the 1970s, however, researchers drainage-basin changes caused by tectonic or volcanic damming, involved in deciphering the hydrologic and climatic records of triggering lake overflow and sometimes catastrophic floods (e.g., Pleistocene pluvial lakes in the western U.S. have increasingly used Bouchard et al., 1998; Reheis et al., 2002b; Carter et al., 2006); (4) cores collected from extant lakes including (Benson complex relations among basins and subbasins controlled by et al., 1997b, 2002, 2013), (Bradbury et al., 1989; changing threshold altitudes (e.g., Adams et al., 1999; Meek, 1999); Benson et al., 1991), (Davis, 1999), and Bear Lake and (5) well-constrained lake-level fluctuations during the late (Rosenbaum and Reynolds, 2004; Rosenbaum and Kaufman, 2009) Pleistocene and Holocene (e.g., Stine, 1990; Adams, 2003a, 2007; as well as from desiccated lakes such as (Oviatt Bacon et al., 2006). et al., 1999; Benson et al., 2011), (Smith et al., 1983), We use the term “pluvial lake basin” to mean a basin that has (Benson et al., 1996; Bischoff et al., 1997; Smith et al., recorded past lake levels higher than present that were usually 1997), (Lowenstein et al., 1999; Anderson and Wells, caused by periods of increased precipitation and (or) decreased 2003; Forester et al., 2005), and Lake Chewaucan (Cohen et al., evaporation (Reeves, 1968; Smith and Street-Perrott, 1983; Benson 2000; Negrini et al., 2000). and Thompson, 1987). For the region of study, the term is essen- Core and outcrop records have different strengths and weak- tially synonymous with “paleo-lake basin,” and we use the terms nesses in the interpretation of paleohydrology and paleoclimate interchangeably. Although most of these basins are presently (Cohen, 2003). Cores can potentially yield continuous records of closed (endorheic), others have spilled to or have been integrated lake history when sampled in the most stable parts of lake basins, with downstream closed basins. There are many other outcrop- are easy to sample in fine increments in laboratory settings, and can based lake-level records external to the hydrologic Great Basin, provide proxies for changes in lake level, water temperature and such as in (Tulare: Negrini et al., 2006), chemistry, and ecological conditions in the surrounding landscape (Estancia: Allen and Anderson, 2000; Anderson et al., 2002; Clo- using analyses of sedimentology, stable isotopes of oxygen and verdale: Krider, 1998), west Texas (King: Wilkins and Currey, 1997), carbon, elemental chemistry, and pollen and other microfossils. (Cochise: Waters, 1989; Kowler, 2010), and northern However, core studies are limited to small (typically <5 cm diam- Mexico (Babicora: Metcalfe et al., 2002; San Felipe: Roy et al., 2012), eter) views of lake sediments, and proxies such as stable isotopes but we restrict this review to the Great Basin proper. may be influenced by several competing factors whose relative effects may be difficult to assess. Many core records from lakes do 1.1. Settings of Paleo-Lake Basins not extend beyond the last w50 ka due to difficulties in coring. In addition, interpretation of cores may be confounded by geologic Most of the pluvial lake basins of the western U.S. are found events within the drainage basin that were unrecorded or not within the hydrologic Great Basin, an area of closed drainage recognized. For example, a drainage-basin change may be inter- encompassing about 520,000 km2 and covering large parts of preted from a core as evidence of climate change rather than , western , southeastern , and eastern California change in physiography. (Fig. 1). Virtually all of the individual basins encompassed within In contrast, outcrops and landforms can potentially yield long the hydrologic Great Basin were formed by strike-slip and exten- climate records, and provide direct evidence of lake level that re- sional faulting caused by interactions between the North American quires little interpretation. Where exposures are large, sediments and the Pacific Plates (Stewart, 1978). This shear and extension can reveal facies changes that provide details of lake-level fluctu- created the tectonic basins within which the pluvial lakes rose and ations, and record geologic events such as earthquakes, cata- fell. The size of the individual pluvial lakes through time depends strophic floods, drainage-basin changes, and isostatic rebound. For on the geometry and motion of basin-bounding faults and the some types of studies that employ physically based models of hy- amount of extension that a basin has undergone, as well as prox- drologic fluctuations or shoreline altitudes to infer isostatic imity to water sources. Rates of tectonic activity have also rebound, documentation of absolute changes in lake size (lake- controlled drainage integration histories, as described in Section 3. surface elevation, surface area, and volume) is necessary and is The two largest pluvial lakes, Bonneville and Lahontan, occupied more difficult to infer from core records. However, outcrops opposite sides of the Great Basin, were supplied by snowmelt-fed commonly yield more discontinuous records than cores and, due to perennial rivers, and integrated multiple tectonic basins when surface weathering, can be more difficult to sample and analyze for they were at their highest levels (Fig. 1). climate proxies. Outcrops may also be limited in geographic extent and not well located in lake basins. Ideally, research on a pluvial 2. Methods lake should incorporate and synthesize both core and outcrop studies for a more complete record, but this is rarely done Although many of the study methods for lake sediment in (although see Wells et al., 2003; Bacon et al., 2006; Negrini et al., outcrop are the same as for cores, the precise reconstruction of lake 2006; Reheis et al., 2012; Benson et al., 2013). level requires: (1) creation and preservation of surface features M.C. Reheis et al. / Quaternary Science Reviews 97 (2014) 33e57 35

Fig. 1. Overview map of the hydrologic Great Basin (black line) showing distribution of lakes and rivers during the late Pleistocene. Background is modern (1971e2000) mean annual precipitation (PRISM Climate Group, Oregon State University) overlain on hillshaded digital elevation model. Major pluvial lakes and rivers: AR, ; HR, Humboldt River; LB, Lake Bonneville; LL, ; MR, Mojave River; OR, system. Other lakes and features: Al, Alvord; Ch, Chewaucan; Col, Columbus; Coy, Coyote; FR, Fort Rock; Fr, Franklin; Ma, Manly; Ne, Newark; PL, Pyramid Lake; Ru, Russell; SCD, Smoke Creek Desert; Su, Surpise; Th, Thompson; Wa, Warner. Circled letters indicate cities: LA, Los Angeles; LV, Las Vegas; R, Reno; SLC, Lake City; W, Winnemucca. such as beach barriers that are easily modified or removed by features and landforms to lake surface, and (4) application of dating erosion or buried by deposition, (2) exposure of sedimentary se- techniques appropriate for outcrops. Dating methods and their quences by erosion, incision, or artificial means (e.g., trenching), (3) application to a wide variety of materials have been discussed in thorough understanding of the relationships between sedimentary detail elsewhere and most recently reviewed by Noller et al. (2000). 36 M.C. Reheis et al. / Quaternary Science Reviews 97 (2014) 33e57

Two methods that have recently been applied to date lacustrine as the age and elevation of soluble pack rat middens (e.g., sediments have less well-known pitfalls in their application, as Thompson et al., 1986; Bacon et al., 2006). follows: The shore zones of these large lakes were dynamic environ- Luminescence dating of lake deposits is subject to significant ments where particles from clay size to boulders were eroded, uncertainty arising from (1) the difficulty of estimating (variable) transported, and deposited in distinctive sedimentary packages and moisture content during burial (Forman et al., 2000), (2) the po- landforms, each of which have particular and often quantifiable tential for disequilibrium in the dose rate caused by migrating relationships to the surface of the lake in which they were depos- and post-depositional alteration (e.g., Reheis and ited. These different environments and deposits are discussed in Redwine, 2008), and (3) differential bleaching of sediments trans- order of their energetics, beginning with the highest energy de- ported and deposited by different mechanisms. The most accurate posits found at or above lake level and ending with sediments laid ages may be obtained from lake deposits in settings that receive down under calm conditions in deeper water. large amounts of aeolian sediment, such as sandy (Magee Pluvial basins are commonly surrounded by relatively steep et al., 2004; Long et al., 2012) or loessial regions (Forman et al., piedmonts covered by coalesced alluvial fans shed from large 2007; Zhang et al., 2011). mountain blocks. This setting typically provides abundant coarse Surface dating of lacustrine landforms by cosmogenic nuclide clastic sediment for transport by wave action along the shores of accumulation (Gosse and Phillips, 2001) is strongly affected by in- paleo lakes. Although wave action is the primary driver in creating heritance from previous exposure and the assumption that the shorelines, the diversity of shoreline types is largely a result of local sample has been continuously exposed and in the same orientation. conditions (Fig. 2). The controlling factors that dictate the type of Only a handful of studies have employed cosmogenic dating on lake shoreline feature formed along a particular shore include local deposits. Surface clast ages typically show widely scattered results slope, shoreline orientation, the amount and particle size distri- likely due to inheritance and (or) exhumation (e.g., Trull et al., 1995; bution of sediment supply, accommodation space, wind strength, Owen et al., 2007, 2011; Kong et al., 2011). Depth profiles, which can fetch length, water depth, and the length of time that lake level help account for prior exposure, require careful assessment and remains stable at a particular elevation (Forbes and Syvitski, 1994; modeling of the competing effects of surface erosion and inflation Adams and Wesnousky, 1998). through incorporation of dust (Machette et al., 2008; Kurth et al., In general, constructional beach features such as barriers and 2011). spits form on relatively low-gradient slopes (<4) with an abun- dant sediment supply, whereas erosional shoreline features form 2.1. Formation of the Lacustrine sedimentary record on relatively steep slopes (>6 )(Adams and Wesnousky, 1998). Field observations indicate that most “erosional” shorelines have a Reconstructing accurate lake-level curves requires detailed wedge-shaped gravel layer or thin lag of beach gravel that lies on knowledge of how the lake surface has fluctuated through time. the low-gradient, lakeward-sloping terrace tread at the base of the Although various proxies analyzed in cores can provide information shoreline angle (Russell, 1885; Gilbert, 1890; Adams and on whether a lake is rising or falling, only mud cracks, soils, or other Wesnousky, 1998). sedimentary evidence of shallow water or desiccation can place The largest waves traversing the surfaces of the pluvial lakes absolute bounds on lake-level changes inferred from cores. The probably ranged in height up to several meters with periods of 5e best evidence for documenting absolute lake-level change lies 7s(Adams, 2003b, 2004), depending on fetch. When these waves within the shore, nearshore, and deltaic sediments that were broke on the shore, the absolute height above still water level deposited across piedmonts and at the mouths of streams as lake reached by the upward rushing water probably ranged from one to level transgressed and regressed. Lake-level curves are constructed several meters, depending on beach slope. This range in runup from the ages and elevations of samples that were deposited above, heights is similar to the natural variability (3 m) expressed in the at or near, or below lake level (e.g., Benson and Thompson, 1987; heights of relatively closely spaced constructional shorelines Benson et al., 1990; Oviatt et al., 1992; Adams, 2007). However, formed above a stable lake level (Atwood, 1994; Adams and the most accurate and precise of these curves attempt to maximize Wesnousky, 1998). Therefore, the precision with which construc- the number of samples that reflect lake level at the time of depo- tional shoreline features can be used to reconstruct past lake levels sition, in addition to utilizing maximum lake-level constraints such is about 3 m along shores with a large fetch, but this value is

Fig. 2. Shoreline features found around margins of pluvial lake basins. After Strahler and Strahler (1992) and Adams and Wesnousky (1998). M.C. Reheis et al. / Quaternary Science Reviews 97 (2014) 33e57 37 probably smaller along relatively protected shores or those with a stable. Base-level fall leads to rapid incision by the feeder streams, shorter fetch. For example, at Lake Tahoe, with fetch ranging from which can propagate many kilometers upstream (e.g., Adams, 2012) about 15 to 30 km, the elevations of the crests of historical and create outcrops for study. Most delta complexes on pluvial basin constructional shorelines were about 10e90 cm above maximum margins are Gilbert-type deltas where fluvial topsets abruptly tran- still-water level (Adams, 2004). The elevations of shoreline angles sition to deltaic foresets, that in turn tangentially transition to bot- associated with erosional shorelines vary by about 100 cm, but tomsets (Fig. 4)(Gilbert, 1885; Russell, 1885; Postma, 1990). The were found both above and below maximum still-water level. topset-foreset transition accurately reflects the lake-surface eleva- These observations agree with Gilbert (1890) who suggested that tion and the amplitude of the foresets indicates the depth of water shoreline angles approximate still water level whereas the crests of (accommodation space) into which the delta is building. The am- constructional features represent storm deposition. Similarly, bar- plitudes of foresets range in height from tens of meters for the largest rier crests were found to be about 1e3 m higher than adjacent deltas on the east side of Lake Bonneville (Milligan and Lemons, shoreline angles in the Lahontan basin (Adams and Wesnousky, 1998) to a few tens of centimeters for small microdeltas deposited 1998). into Walker Lake at times of low stream flow (Adams, 2007). The internal stratigraphy of constructional beach features can Within sandy delta complexes and along other sand-rich parts also provide detailed lake-level evidence. Along many gradually of the shore, wave ripples of various types and wave ripple lami- sloping shores with an abundant sediment supply, small beach nations are common and indicate shallow water deposition (Fig. 5). barriers of various types commonly enclose small (Fig. 3). The exact depth of the water in which these features form is Because of the generally coarse and pervious nature of shore de- sometimes unclear, but at Walker Lake, Adams (2007) was able to posits, water level in these lagoons accurately reflects the adjacent directly trace fluvial sediments (topsets) and beach deposits lake level (Adams and Wesnousky, 1998). As waves wash over the downslope to a transition into wave-rippled sands. Within this barrier crests, sediment and water are transported into the lagoons, low-gradient delta area, wave-rippled sands formed in water as forming small foresets that prograde into the lagoons analogous to shallow as 10 cm to water as deep as 3 m. Based on the modern topsets grading into foresets in a delta setting. The abrupt increase wind regime, the maximum deep-water wave height at Walker in the dip of these sediments has been referred to as a hingeline Lake would be about 2 m with five second periods, which corre- (Adams and Wesnousky, 1998) and is formed within a few centi- sponds to a wave base of about 20 m (Adams, 2007). Because of the meters of the surface of the , as observed along active relatively low gradient of the bed of Walker Lake in the delta area, shorelines in both Pyramid Lake and Lake Tahoe (Fig. 3). Hingelines however, large waves were probably restricted from entering this are commonly found in transverse exposures through coarse beach area, which means that most of the observed wave ripples were ridges and their elevations accurately reflect lake level at the time probably formed by more common small waves in relatively of their formation. shallow water (<3 m). Delta complexes contain readily decipherable lake-level infor- To infer the paleo-water depth of sedimentary features formed mation and commonly preserve stratigraphic sequences that indi- near and below wave base in the Lahontan basin, Adams (2010) first cate whether a lake was rising or falling at a particular time. Delta used the Wono and Trego Hot Springs layers to document complexes also commonly contain terrestrial organic matter useful lake levels at the times the tephra layers were deposited. Then, for radiocarbon dating (e.g., Born, 1972; Stine, 1990; Adams, 2007). sedimentary features formed from 5 to 90 m below the lake surface Because most pluvial lakes did not usually maintain constant water were examined in additional outcrops of these tephra layers at levels, the loci of delta deposition are commonly spread out for many lower elevations and therefore deeper water depths. Sandy sedi- kilometers along river valleys (e.g., Anderson and Link,1998; Adams, ments near wave base can display flaser or wavy bedding and 2007), forming long ribbon-like deltaic complexes with wedge- hummocky cross stratification (Fig. 5), indicative of storm condi- shaped accumulations in locations where lake level was relatively tions, and sediments below wave base range from massive silty

Fig. 3. Loop barrier formed in a period of less than 6 months on the west shore of Pyramid Lake after lake level rose by w2.75 m in of 1997. 38 M.C. Reheis et al. / Quaternary Science Reviews 97 (2014) 33e57

Fig. 4. Three views of Gilbert deltas. (A) Vertical air photo of Truckee building into Pyramid Lake in 1938 after lake level had dropped w17 m since 1891. Note that a small, barely emergent beach barrier formed near the topset-foreset transition. (B) Gilbert’s (1890) model of deltas where the topset-foreset contact closely represents lake level. Image has been mirrored so that foresets dip to left. (C) Microdelta sequence exposed along lower that formed during historical period, based on presence of a shotgun shell in sediments. Small shovel near center of photo is about 45 cm tall. Walker Lake is to left. clays to clays that sometimes contain a few thin, discontinuous of river water (e.g., Bouchard et al., 1998; Hart et al., 2004; McGee laminae of fine sand and or sand-sized ostracode tests (Adams, et al., 2012). 2010). Such laminae of sand-sized particles seem to be more common when an outcrop is located near the base of a steep slope, 3. Drainage basin changes and effects on pluvial lake records possibly reflecting suspension by wave activity and transport to the site by subaqueous gravity flows. Tectonic forces in extensional regimes generally act to create Various indicators of subaerial exposure, including desiccation internally drained basins separated by highlands or mountain cracks (Adams, 2007), channel features and other unconformities ranges (e.g., Peterson, 1981). Orographic focusing of precipitation on (Oviatt et al., 2005), soil development (Morrison, 1991), rooted uplands creates runoff that supports lakes or in endorheic terrestrial vegetation (Stine, 1990), and grounded blocks basins during periods of greater effective moisture. Drainage-basin (Stine, 1990), have been used to infer that lake level dropped below integration can be thought of as a result of competition between a particular elevation. Tiger beetle burrows superimposed on climate and tectonics, with climate-driven processes acting to fill a shallow-water lacustrine sediments also indicate subaerial expo- basin with sediment and tectonics acting to create more accom- sure because the larvae of these terrestrial creatures form vertical modation space for sediment. Well-known examples of drainage- burrows on moist sands immediately adjacent to lakes and streams system integration in the are the three (Fig. 6)(Smith and Bronson, 2003). The presence of these subaerial major river systems that can drain into to sustain Lake features within a stratigraphic sequence, coupled with the vertical Manly during pluvial periods (Knott et al., 2008): the long chain of arrangement of other sedimentary beds, can help determine overflowing lakes sustained by the Owens River from the west whether a lake was rising or falling (Stine, 1990; Oviatt et al., 1994; (reviewed by Phillips, 2008), the Amargosa River from the northeast Adams, 2007, 2010). (Menges, 2008), and the Mojave River from the southwest (Fig. 1; is a carbonate precipitate that forms below lake Meek, 1999; Enzel et al., 2003). Of these systems, only the Amargosa surface in pluvial basins. Consequently, the elevations and ages of River (in flood) now flows without interruption from its headwaters have long been used to infer lake-level fluctuations (Broecker to Death Valley. Another less well-studied example is the Humboldt and Orr, 1958; Broecker and Kaufman, 1965). In the Lahontan basin, River, a major eastern tributary of Lake Lahontan. We present Benson and colleagues have determined the formation sequence of several examples of drainage-basin integration documented in different types of tufas and their elevational distributions (Benson, recent studies and discuss the implications of such changes to 1994), constructed envelope lake-level curves (Benson et al., 1995, lacustrine climate records and migration of aquatic species. 2013), determined the sources of river water based on isotope values (Benson and Peterman, 1995), and correlated broad 3.1. Assembly of Humboldt River drainage trends in d18O and d13C from tufas with major features of the tufa- based lake-level curve (Benson et al., 1996, 2013). Similar studies The longest tributary of the Lahontan drainage basin is the have yielded insights on Bonneville lake-level history and sources Humboldt River, which presently flows about 400 km westward M.C. Reheis et al. / Quaternary Science Reviews 97 (2014) 33e57 39

Fig. 5. Views of wave-formed sedimentary features. (A) Wave ripples in shallow (w25 cm) water at Lake Tahoe. Paddle blade about 15 cm wide. (B) Wave ripples in sediment of Walker Lake. Some ripples formed in about 30 cm of water, near the foresetebottomset transition just above shovel handle. See Fig. 4C for a different view of same outcrop. (C) Wave ripples and laminae in Mullen Creek delta on west side of Pyramid Lake. Massive bed below prominent wave ripples is w50 cm thick. Bed directly overlying wave-rippled bed has undergone soft-sediment deformation. (D) Detail of wave ripple laminae. Note truncation of bundled, upward sweeping laminae. Trowel blade about 5 cm wide. (E) Hummocky cross stratification expressed in pebbly, sandy sediments at Smoke Creek in the Lahontan basin. Finer sand beds have been highlighted in white to emphasize structure. (F) Wavy (flaser) bedding delineated by relatively coarse sand beds (light colored) interbedded with silt-rich beds (dark gray) exposed in Pleistocene lacustrine deposits on the west side of Lake Tahoe. from near the NevadaeUtah border to terminate in the Humboldt of several basins formed during early Miocene extension, with final Sink (Figs. 1 and 7). The river carries about 38% of the total integration westward through the area of Carlin and Palisade can- streamflow of the Lahontan basin today (Benson and Paillet, 1989). yons, west of Elko, by about 9 Ma. Incision by this early stage of the During the last highstand of Lake Lahontan, the river entered the river removed substantial amounts of Miocene strata, carried the lake a short distance east of the town of Winnemucca, within a materials downstream towards Beowawe, and partially rejuvenated narrow valley cut through bedrock hills that divide the middle and the pre-early Miocene topography. The uppermost part of the lower reaches of the Humboldt. Studies during the past decade drainage, the Elko basin, was subsequently blocked during the show that the Humboldt River was assembled by fill-and-spill of , possibly due to a landslide at Carlin Canyon. This produced many previously endorheic basins from the middle Miocene to the a new, areally extensive lake in the Elko basin upstream of the middle Pleistocene, but this history has not previously been syn- canyon that filled the topography formed by earlier erosion. Tephra thesized and many details remain to be resolved. analyses from outcrops indicate that this lake persisted from at least The upper and middle reaches of the Humboldt River traverse at 2.5 to 2.2 Ma (Reheis et al., 2003), and drill hole data locally indicate least four subbasins and gather water from tributaries that drain as much as w275 m of lacustrine sediment (Wallace et al., 2008). other valleys to the north and south (Fig. 7). Some of these valleys are Numerous paired strath terraces of the Humboldt River are incised connected to the river by narrow bedrock canyons cut through into these lake deposits recording as much as 200 m of incision since intervening ranges and uplands, and incision has exposed lacustrine the early Pleistocene. Similarly, reconnaissance field work suggests and alluvial basin-fill deposits of Tertiary and Quaternary age. that the Reese River drainage, which enters the middle reach of the Wallace et al. (2008) showed that most of the upper Humboldt east Humboldt River at Battle Mountain, may also have been integrated of the town of Beowawe was assembled by progressive fill-and-spill during Pliocene (?) time (Wallace et al., 2008). 40 M.C. Reheis et al. / Quaternary Science Reviews 97 (2014) 33e57

Fig. 6. Filled tiger beetle burrows in a late Holocene tephra and associated wave-rippled sands at Walker Lake (Adams, 2007). The top of the burrows is coincident with a slight unconformity, which also indicates brief subaerial exposure. Inset shows photograph of a tiger beetle larva and vertical view of the top of burrows that they excavate.

The Humboldt River likely extended to a basin centered on the received runoff from the surrounding ranges as well as from present-day town of Battle Mountain by w9Ma(Fig. 7; Wallace Monitor and Antelope valleys (Reheis et al., 2002a). Lake Jonathan et al., 2008), and paleontological evidence suggests it may have overtopped its eastern threshold during a highstand that followed flowed west to the headwaters of the modern by deposition of the Rye Patch Dam ash bed at w670 ka, and the entire the early Pliocene (Repenning et al., 1995). If so, it may have been system became tributary to Lake Diamond (Davis, 1987; Reheis, dammed by flows or faulting near Iron Point by 4e5Ma 1999; Reheis et al., 2002a). The influx of water to this lake in turn (Wallace, 2005) and was not fully integrated with the ancestral caused it to overflow to the Humboldt River, although the threshold Lahontan basin to the west until as late as the early middle Pleis- has not yet been eroded low enough to allow unimpeded tocene. Reheis et al. (2002a) discussed sedimentary evidence for a throughflow except during pluvial periods (Mifflin and Wheat, lake in the middle Humboldt reach (proposed by Roberts, 1965), 1979). primarily on the basis of exposures in an open-pit gold mine. Su- The assembly of the Humboldt River from Miocene to middle perposed beach gravels record several highstands, but their alti- Pleistocene time (Fig. 7) has obvious implications for long-term tudes have been altered by uplift and thus are difficult to relate to climate records in the Lahontan basin, as well as regional biogeo- Lahontan deposits of similar age farther west. The lake likely per- graphic history (Hubbs et al., 1974; Hershler and Sada, 2002; Smith sisted from the late Pliocene to sometime after 760 ka, based on et al., 2002). Pliocene and early Pleistocene lacustrine deposits paleomagnetic data and lacustrine sediments containing the 2-Ma exposed in the main Lahontan basin southwest of Winnemucca Huckleberry Ridge, w1-Ma younger tuffs of Glass Mountain, and (Morrison, 1991; Reheis et al., 2002a) must record runoff from a 0.76-Ma Bishop ash beds. During middle Pleistocene highstands the much smaller area mainly fed by Sierran rivers (Fig. 1). Thus, lower lake may have been contiguous with the main body of Lake lake levels during this time than were attained in the middle to late Lahontan. Pleistocene do not necessarily indicate an effectively drier climate. A younger record of basin integration is preserved in several With respect to biogeography, development of a through-flowing valleys that are now tributary to the Humboldt River, including upper and middle Humboldt River would have permitted wide- , Monitor, Antelope, Kobeh, and Diamond valleys (Fig. 7). In spread distribution of aquatic and semi-aquatic species during the addition, Newark Valley may have discharged northward to the middle to late Miocene and Pliocene (Hubbs et al., 1974; Repenning Humboldt at least once during a middle (?) Pleistocene highstand et al., 1995; Smith et al., 2002). This distribution was likely inter- of pluvial Lake Newark (Redwine, 2003). Pine Valley contains a rupted during the Pliocene by events that separated the upper, well-documented record of Pliocene to middle Pleistocene lacus- middle, and lower reaches of the Humboldt, thus allowing ende- trine (Smith and Ketner, 1976; Gordon and Heller, mism to develop in formerly connected aquatic species as sug- 1993). Sometime after deposition of the 640-ka Creek B ash gested by Hershler and Sada (2002). bed, which lies near the top of the lake beds, the valley began draining into the Humboldt River. From the presence of a thick bed 3.2. Drainage-basin changes of Mono Lake of well-rounded beach gravel atop the lacustrine sequence, Reheis et al. (2002a) inferred that Pine Valley was breached when an Like the upper Humboldt River, Mono Lake (Lake Russell) exceptionally large lake spilled over a northern sill composed of experienced changing drainage-basin conditions that allowed it to Tertiary and sedimentary rocks at, or shortly after, 640 ka. Lake influence water levels in downstream basins at different times and Jonathan in Kobeh Valley, a Pliocene to middle Pleistocene lake, to provide a means of transfer for aquatic species (Hershler and M.C. Reheis et al. / Quaternary Science Reviews 97 (2014) 33e57 41

Fig. 7. Pluvial lakes of west-central Nevada, showing approximate times of integration of lake basins into Humboldt River drainage based on tephrochronology. Blue, late Pleis- tocene lakes; orange, older lake extent indicated by highest shoreline deposits; yellow, possible additional areas of older lakes now buried by younger alluvium. Red triangles are locations of beach deposits higher than late Pleistocene shorelines. Yellow arrows indicate hypothesized temporary drainage connections. Ancestral Humboldt River probably formed during late Miocene. During late Pliocene time, the river probably did not extend west of the “paleo-Lahontan drainage basin boundary” near Winnemucca, and probably had no connections to the upper Reese River, Pine, Diamond, or Kobeh valleys, or “Lake Elko” in the Elko basin (Reheis et al., 2002a; Wallace et al., 2008). The upper Humboldt above Iron Point was finally integrated with Lake Lahontan to the west sometime after 760 ka. Modified from Reheis et al. (2002a). 42 M.C. Reheis et al. / Quaternary Science Reviews 97 (2014) 33e57

Sada, 2002; Smith et al., 2002; Reheis et al., 2002b). Unlike the 1981), but probably became internally drained by the early Pleis- Humboldt, Mono basin lies in the belt in a zone of tocene. During highstands of pluvial Lake Russell, the lake first active transtensional shear (Stewart, 1988) and late Cenozoic discharged northward into a tributary of the Walker River (Fig. 8) volcanism (Hildreth, 2004). Thus, structural deepening and volca- that was episodically blocked by basaltic eruptions (Reheis et al., nism compete directly with sedimentary and climate controls on 2002b). Sometime after 1.3 Ma and possibly as late as 0.76 Ma, lake level in this basin. Mono basin lies on the divide between the faulting on the southeastern rim of the basin diverted discharges of Walker Lake arm of Lake Lahontan to the north and to Lake Russell highstands into Adobe Valley and thence to Owens the south (Fig. 8). The basin was relatively shallow and contained a Lake in southern Owens Valley. freshwater lake during the Pliocene, possibly with drainage con- Given the same climatic conditions, pluvial lakes in the Walker nections to the north or east (Gilbert et al., 1968; Miller and Smith, Lake basin during the early Pleistocene would have been larger, due

Fig. 8. Pluvial lakes and drainages of the southwestern Great Basin. Approximate time represented is pluvial maximum period of OIS 6, when Owens River system (ORS) was last continuously connected from Owens Lake (O) through -Searles (CS) and Panamint (P) to Lake Manly (M) in Death Valley (Phillips, 2008), and may also have received overflow from Lake Russell (R; Reheis et al., 2002b) via Adobe Valley (AV). Dashed blue line shows overflow path from Lake Russell to the East Walker River (EWR) during the early Pleistocene. During early OIS 6, (T) probably filled and then drained into Lake Manly to integrate the upper and lower Amargosa River reaches (Morrison, 1999; Caskey et al., 2013). The Mojave River terminated in (Mx); (Mo) and the downstream connection to Lake Manly did not form until OIS 2 (Wells et al., 2003; Reheis and Redwine, 2008). Modified from Knott et al. (2008). AC, Afton Canyon; AR ¼ Amargosa River drainage; EM, Eagle Mountain; H, ; MR, Mojave River; SE, Swansea embayment. M.C. Reheis et al. / Quaternary Science Reviews 97 (2014) 33e57 43 to addition of water from the Mono basin, than could have been Valley on the north to Owens Lake basin on the south (Bacon et al., maintained after the basin discharge switched south to Owens 2006). Episodic earlier connections between Owens Lake and the Valleydand vice versa. Reheis et al. (2002a) showed that Walker lower basins probably played a significant role in migration and Lake highstands during oxygen-isotope stages (OIS) 6 and 2 were endemism of aquatic species (Hershler and Liu, 2008; Knott et al., somewhat lower than levels attained during the early Pleistocene, 2008). though this may in part have been due to threshold control be- tween Walker Lake basin and the main body of Lahontan until 3.4. Integration of basins along the Mojave River sometime after OIS 6 (King, 1993; Kurth et al., 2011). Aquatic spe- cies that became widely distributed along the eastern Sierra front The Mojave River flows through a region of the eastern Cali- during the late Miocene and Pliocene (Miller and Smith, 1981; fornia shear zone where faults have lower vertical displacement Hershler and Sada, 2002) were obviously divided into separate rates than those to the north. The progressive northeastward populations by development of an internally drained Mono basin, integration of the river by fill-and-spill, beginning in the Pliocene, but were later intermittently connected by the changing overflow has been reconstructed by several workers (Fig. 8)(Meek, 1999; Cox directions (Reheis et al., 2002b). et al., 2003; Wells et al., 2003; Reheis and Redwine, 2008). Lake Manix formed the terminus of the Mojave River for most of the 3.3. Integration of Amargosa River and Owens River period from w500 to 25 ka (Jefferson, 2003; Reheis et al., 2012), after which the river escaped eastward to form Lake Mojave. Wells The Amargosa River flows southward through Tecopa basin into et al. (2003) reconstructed the history of this lake and incision of its southern Death Valley (Fig. 8). This drainage provides an example threshold, allowing episodic discharge northward to Death Valley. of progressive filling and spilling of tectonically formed sedimen- Meek (1989) proposed that the failure of the Lake Manix tary basins (Menges, 2008). The headwaters basin of the Amargosa threshold might have triggered initially rapid, even catastrophic, formed during the late Miocene, and eventually drained or was incision of the threshold to the level of the former lake-basin floor captured into the sometime before the middle to form Afton Canyon as Lake Mojave filled in the next basin Pleistocene. The timing of the next integration event around Eagle downstream. Wells et al. (2003) and Enzel et al. (2003) rejected Mountain and into Tecopa basin is not well constrained. Morrison that hypothesis on the grounds that many strath terraces within the (1999) inferred a highstand of Lake Tecopa early in OIS 6 as well canyon record progressive downcutting. However, Reheis and as during OIS 16. However, Menges (2008) used projections of Redwine (2008) showed that initially rapid incision at about 25 alluvial-fan slopes and relative fan-surface ages to suggest that the ka could not be rejected on the basis of soil development. Several Amargosa-Tecopa divide at Eagle Mountain was not breached until studies of soils formed on alluvial-fan and beach deposits in the sometime after 150e100 ka. Preliminary cosmogenic ages suggest a have shown that soil properties can be used to somewhat older minimum age of w200 ka and post-incision differentiate between deposits of late and early Holocene and late terrace ages that are less than 50 ka (C. Menges, 2011, oral com- Pleistocene ages (e.g., McFadden et al., 1989; Reheis et al., 1989; mun.). Regardless, it is difficult to understand how substantial lakes McDonald et al., 1995). Yet, soils developed on the inset strath in OIS 16 and 6 could be sustained without input from the Amar- terraces showed no trends in age-related properties with height gosa River; research on this issue is ongoing. The final integration above the river as would be expected if downcutting had been event through the divide between Tecopa basin and southern Death gradual. In addition, soil properties on the strath terraces could not Valley probably also occurred during OIS 6, on the basis of relative be differentiated from those of soils formed on the last-highstand ages of pre- and post-incision terraces (Morrison, 1999; Menges, beach barriers (Reheis and Redwine, 2008). 2008) and the presence of a large Lake Manly in Death Valley Recent mapping demonstrated an internal integration event beginning around 185 ka fed initially by water with Amargosa River within the Manix basin that did trigger a catastrophic flood (Reheis composition (Forester et al., 2005; Caskey et al., 2013). et al., 2012). The boundary between the western and Afton sub- The Owens River system, in contrast, is an example of compe- basins of Lake Manix lay on the south flank of Buwalda Ridge, tition between active tectonics and climate-driven sedimentation; composed of sheared Pliocene fanglomerate along the left-lateral its history was recently reviewed by Phillips (2008). This system, Manix (Miller et al., 2011b)(Fig. 9). From about 500 to 190 which lies entirely within the tectonically active Walker Lane and ka, Lake Manix was confined to its western subbasins and fluctu- eastern California shear zone, drains the southern ated in response to climate change and perhaps to brief diversions and WhiteeInyo Mountains in east-central California. During of the river into Harper Lake basin (Meek, 1999). During this time pluvial periods, the Owens River system extended from Mono basin about 24 m of sediment accumulated near the confluence of the on the north to Death Valley in the southeast dropping a total Mojave River and Manix Wash, gradually filling the accommoda- elevation of w2275 m (Gale, 1914; Blackwelder, 1933; Smith and tion space. Street-Perrott, 1983; Reheis et al., 2002b)(Fig. 8) to form a chain When the lake filled at the onset of OIS 6, the Buwalda Ridge of lakes occupying one or more successively lower-elevation basins threshold failed and triggered a catastrophic flood into the previ- (Jayko and Bacon, 2008; Orme and Orme, 2008). All of the ously endorheic Afton subbasin. The evidence for this flood consists component basins are fault-bounded or lie atop strike-slip faults of several meters of chaotically bedded blocks (as much as 2 m and most are actively subsiding; thus, the elevations of the across) of semi-indurated green lake sediment derived from the thresholds between these basins has changed through time, in part basin upstream mixed with blocks of brown playa mud. These controlling individual lake levels. The lower-elevation basins, deposits are overlain by and interfinger upstream with fluvial sand including Indian Wells/China, Searles, Panamint, and Death Valleys, and gravel that also contains clasts of reworked lake mud and have unique lacustrine histories as a consequence of the frequency grades up into finer-grained deposits that contain a 184-ka tephra. and magnitude of glacial-pluvial climatic events, which in addition Upstream in the older subbasins, lake deposits that underlie this to tectonics have controlled the duration and discharge of stream tephra show evidence for erosion and slope instability caused by flow within the Owens River system (Phillips, 2008 and references rapid draining of the lake, including gravels containing rip-up clasts therein; Jayko et al., 2008; Smith, 2009; Grayson, 2011). Since the of lake mud that truncate thinly bedded lacustrine silts, and beds of PleistoceneeHolocene transition, the Owens River system has been convoluted gravel and mud. Subsequent lake-level changes affected hydrologically limited to a third of its upstream length from Long both subbasins, but much of the upstream subbasin was only 44 M.C. Reheis et al. / Quaternary Science Reviews 97 (2014) 33e57

Fig. 9. Aerial photograph of Lake Manix basin. Solid and dashed black lines are important Quaternary faults. CMF, Cave Mountain fault; DLF, Dolores Lake fault; MF, Manix fault; PF, Pisgah fault; TCAF, Tin Can Alley fault. Modified from figures 2 and 3 in Reheis et al. (2007). submerged during relatively high lake levels. Thus, sediment re- Oneida divide, fed by local streams (not the Bear River), had cords in part reflect a geomorphic event (basin integration) rather commenced by 100 ka. Their data did not indicate when Oneida than climatic conditions (Reheis et al., 2012). Narrows was deeply incised, but the gorge was present by the time Lake Bonneville rose high enough to flood into Thatcher basin 3.5. Bonneville: changing inputs of the Bear River (w20,000 cal yr B.P.). A basaltic tephra that is present in both cores and outcrops in Gilbert (1890,p.219e220) noted that river diversions in the basin provides a clue to when the upper Bear River was southern (Fig. 10), including diversion of the Bear River into permanently diverted into the Bonneville basin. The Hansel Valley the Bonneville basin, would have a significant effect on the water ash (Miller et al., 2008) is found at the base of the Bonneville budget of the lake. Bright (1963) conducted studies in the Soda sediments in outcrops as high as w1340 m above sea level (masl) Springs/Gem Valley/Lake Thatcher area of southern Idaho, and and in sediment cores from the floor of (GSL) at demonstrated that the Bear River had been diverted by basaltic altitudes as low as w1260 masl. In GSL cores, sediments of a eruptions during the late Pleistocene. Lake Thatcher, a mid- to late- shallow saline-to-brackish lake directly underlie the Hansel Valley Pleistocene lake that was fed for part of its history by the upper ash. For the ash to be present at the base of the Lake Bonneville Bear River, began discharging across Oneida divide into the Bon- section at altitudes within a vertical range of w80 m, the lake must neville basin about 27,000 14C yr B.P. (w32,000 cal yr B.P.). Bright have risen very rapidly, and the cause of that abrupt rise is more (1963) thought that Oneida Narrows was rapidly incised at that likely to be related to a sharp increase in river input rather than to time, after which the Bear River became a permanent tributary to an increase in the ratio of precipitation to evaporation in the the Bonneville basin. drainage basin. A likely explanation for this is the diversion of the Research by Bouchard et al. (1998) using strontium isotopes in Bear River into the Bonneville basin at about the time of the mollusks collected from outcrops added complexity to the story. eruption of the Hansel Valley ash. Radiocarbon ages associated with They concluded that (1) the Bear River had entered the Thatcher the ash in cores from Great Salt Lake suggest an approximate age of basin by w140 ka; (2) the river did not enter the Thatcher basin 25,000 14C yr B.P. (w30,000 cal yr B.P.) (Thompson et al., 1990) (see between w140 and w80 ka, but was present after w50 ka; and (3) Section 4.3). This age is similar to that proposed by Bright (1963) for overflow from Lake Thatcher to the Bonneville basin across the the incision of Oneida Narrows. M.C. Reheis et al. / Quaternary Science Reviews 97 (2014) 33e57 45

tephra layers that indicate these large lakes existed during OIS 16 (Reheis et al., 2002a), consistent with marine oxygen-isotope re- cords (Lisiecki and Raymo, 2005) and glacial deposits (Mickelson and Colgan, 2003; Pierce, 2003; Dahms, 2004) showing that this glaciation was one of the most extensive in the past 2 Ma. High- stands in some of these basins likely triggered overflow and inci- sion that integrated endorheic basins with Lake Lahontan (Section 3.1; Fig. 7). Below these highest shorelines, sequences of progressively lower shorelines are often preserved (Fig. 11; Reheis et al., 2002a; Redwine, 2003), with the lowest highstand representing the OIS 2 pluvial maximum (e.g., Lillquist, 1994; Adams and Wesnousky, 1998; Munroe and Laabs, 2013b). These intermediate shorelines have been difficult to date. However, 36Cl cosmogenic profile ages and supporting UeTh and 14C ages of lake tufa have recently been published for the Walker Lake arm of Lake Lahontan, Lake Co- lumbuseRennie to the south, and Lake Newark to the east (Kurth et al., 2011). Remarkably, all three basins contain shorelines at or

Fig. 10. Map of Lake Bonneville. TB ¼ Thatcher basin; OD ¼ Oneida divide; RRP ¼ . Large gray arrow in indicates general flow path from Sevier basin into Great Salt Lake basin. Modified from Currey (1982).

4. Lake-level records of paleoclimate

Lake level directly reflects the balance between precipitation (runoff) and evaporation and thus effective moisture (Benson and Thompson, 1987). Sedimentologic and geomorphic indicators of lake level can be accurately measured across broad areas and pre- served for millennia, thus allowing reconstruction of long paleo- climate records provided tectonic effects and isostatic rebound can be assessed or corrected. In this section, we review examples of paleoclimate constrained by outcrop studies of the Great Basin pluvial lakes and compare them where possible with core records. Fig. 11. Comparison of shoreline ages and altitudes in Walker Lake subbasin of Lake 4.1. Older shorelines of Lakes Lahontan, Newark, and Columbus Lahontan and Lakes Columbus and Newark measured using 36Cl, U-series, and 14C. Some error bars are smaller than symbol size. Long arrow on Walker Lake plot rep- resents minimum age of 1400-m shoreline based on 36Cl profile age; maximum age Across the western Great Basin, many basins preserve shorelines less than w760 ka based on presence of reworked Bishop ash in beach deposits. Gray of lakes that reached higher levels than those attained during OIS 2. bars show approximate times of glacial . Note general increase in The highest known lake deposits in these basins locally contain age with altitude. Modified from Kurth et al. (2011). 46 M.C. Reheis et al. / Quaternary Science Reviews 97 (2014) 33e57 above the OIS 2 highstands that appear to have formed during OIS was deposited. To address these inconsistencies, Benson et al. 4. Lake Bonneville sediments record buried nearshore deposits of (1997b) also studied several outcrops in the area. At Squaw Creek, that approximate age (Oviatt and McCoy, 1992). Cool, relatively Benson et al. (1997b) reinterpreted the sedimentary package con- deep-water conditions during OIS 4 in southern Oregon and adja- taining the THS tephra in the downstream outcrop as fluvial de- cent California are inferred from core proxy records from Lake posits not graded to lake level. This reinterpretation was based, in Chewaucan (Cohen et al., 2000; Negrini et al., 2000) and Lake part, on sedimentary features associated with the tephra but also Modoc (Bradbury, 1991), and from outcrop for Lake Malheur on sedimentological analysis of THS and Wono tephra outcrops on (Dugas, 1998), but their shoreline altitudes are not known. The next the east shore of Pyramid Lake and at Agency Bridge, as well as of higher shorelines were formed during OIS 6, and there is at least core PLC92B. one older intermediate shoreline in two of the basins. Such records Outcrops of the THS and Wono tephra layers studied by Davis might be interpreted to indicate uplift of shorelines on footwall (1983) and Benson et al. (1997b) represent only a small fraction of blocks, and indeed this has been documented on the west side of the number of outcrops containing one or both of these tephra layers Walker Lake (Kurth et al., 2011, and references therein; Fig. 11). in the Lahontan basin. Adams (2010) documented the sedimentology However, uplift cannot explain stepped sequences of shorelines and stratigraphy of nine more sites where the THS and (or) Wono that are preserved on the eastern, relatively undeformed margin of tephra layers crop out at a range of elevations, and reexamined the Walker Lake and those that rim the basins of lakes Newark and Squaw Creek sites of Davis (1983) and the Agency Bridge site (Davis, Columbus-Rennie. Kurth et al. (2011) interpreted these results to 1978; Benson et al. (1997b). At Squaw Creek, Adams (2010) indicate a regional decrease in lake levels and thus a decrease in the concurred with the interpretation of Davis (1983) that the THS effective wetness of pluvial periods since the early middle Pleis- tephra was deposited within deltaic topsets at an elevation of tocene. It is unclear whether this is a “real” climate signal or 1254.3 m, which closely approximated lake level at that time. This whether a reduction in effective moisture may be related to other interpretation is supported by the presence of the THS tephra within factors such as an increasing rain-shadow effect from uplift of lacustrine beds at five other sites in western subbasins of Lake ranges to the west (Reheis, 1999). Lahontan ranging in altitude from 1165 to 1229 m. At Agency Bridge, both the Wono and THS beds were deposited within lacustrine sed- 4.2. Lahontan: comparison of interpretations from core and outcrop iments at an altitude of 1208e1209 m (Adams, 2010), or about 30 m records above where Benson et al. (1997b) claimed lake level was at that time. However, Benson et al. (2013) revised their previous interpretation to When working in any pluvial basin, it is important to keep in suggest that the lake-surface elevation at THS time “could have been mind that each basin only has one history, regardless of how subtle no greater than 1208e1220 m, and may have been significantly and complex that history may be. Therefore, when there are lower.” Given the occurrence of the THS layer in lacustrine beds well different interpretations of lake state at a particular time or in above this altitude range, including at the Squaw Creek site, their trends through time, one or both interpretations must be wrong. revision is still about 35 m too low. Benson et al. (2013) also agreed Differences in interpretations can stem from a variety of causes that with Adams (2010) that lake level was near 1217 m in the western include inappropriate age samples, poorly constrained age models, subbasins of Lahontan when the Wono tephra was deposited. and the use of proxies whose response to changing hydrologic The differences in lake-surface altitudes discussed above, which conditions are not fully understood. in turn correspond to large differences in lake areas and volumes, Lake-level interpretations for the Lahontan basin can be made are not a function of imprecise or inadequate dating because the using extensive outcrop data as well as data from multiple cores THS and Wono tephra layers explicitly tie all of these sites together. collected from Pyramid Lake. Comparing core and outcrop records Instead, the differences in interpretations may in part stem from in Lahontan is facilitated by the presence of multiple volcanic the different proxies used to imply lake state. Fig.12 shows the tufa- tephra layers spanning a wide age range that have been deposited based lake-level curve of Benson et al. (2013) along with oxygen in various sedimentary environments from deep water to shore isotope data from core PLC92B. Superimposed on the tufa-based (Davis, 1978, 1982, 1983, 1985; Sarna-Wojcicki and Davis, 1991; curve is an alternative interpretation of lake-level changes for the Sarna-Wojcicki et al., 1991; Benson et al., 1997b, 2013; Adams, 2007, period 35e10 cal ka based on ages and elevations of tephra layers 2010; Adams et al., 2008). Ideally, each tephra represents a precise and organic carbon samples found in shore or nearshore environ- chronostratigraphic marker that can confidently be used to corre- ments. For the alternative curve, the lake-surface elevations for the late sediment layers between isolated outcrops and core sites over a Mount St. Helens M (MSHM), THS, and Wono tephra layers are from wide range of elevations. Adams (2010), whereas the lake-surface elevation for the Timber Davis (1983) used two outcrops of the Trego Hot Springs (THS) Lake bed is inferred from its elevation contained in beach sands tephra along Squaw Creek in the northern Smoke Creek Desert to exposed in a trench at the north end of Winnemucca Dry Lake determine that Lake Lahontan was rising through an elevation (Adams et al., 2008). The ages of the THS, Wono, and Timber Lake between 1256 and 1260 m when this tephra was deposited. In the beds are from Benson et al. (2013). upstream outcrop, the tephra was contained in fluvial deposits, but The lake-level interpretations derived from these different in the outcrop about 500 m downstream and a few meters lower in datasets are broadly similar, except for the period from 25 to 35 cal elevation, the tephra was deposited in a deltaic sedimentary ka, spanning the time when the THS tephra was deposited. The package graded to the surface of the lake (Davis, 1983). Subse- outcrop-based curve was drawn with smooth, relatively gently quently, Benson et al. (1997b) noted that Davis’ (1983) interpreta- sloping lines during this period because of limited data. In reality, tion of lake level conflicted with their d18O record from core PLC92B the rate of lake-level change may have been much more rapid, from the north end of Pyramid Lake. In the core, the THS tephra is similar to rates of change associated with the highstand between 15 associated with a maximum in d18O, which to Benson et al. (1997b) and 16 cal ka, which is shown by all three curves (Fig. 12). If the meant that Pyramid Lake was probably shallow, approximately the lake-level rise marked by the THS tephra resembled more of a spike size it is today (1177 m). The slightly older Wono tephra is also than a slow rise, then the disagreement between the three curves associated with relatively heavy d18O values in PLC92B, which led to would be even less. It is almost a certainty that with more data the their interpretation that Pyramid Lake was spilling across Emerson shape of the outcrop-based curve will continue to change, partic- Pass (w1200 m) and into the Smoke Creek Desert when that tephra ularly during the period prior to the highstand. M.C. Reheis et al. / Quaternary Science Reviews 97 (2014) 33e57 47

that give inherently more reliable results. For some carbonate samples, for example, the effects of contamination with young carbon or from a radiocarbon reservoir have not been evaluated. Radiocarbon ages for wood or charcoal samples are generally more reliable than ages for mollusk shells or tufa (e.g., Taylor et al., 1992; Trumbore, 2000). In Fig. 13, radiocarbon ages for wood or charcoal samples from buried soils directly beneath deposits of Lake Bon- neville place limits on the age of lake transgression at different altitudes. Radiocarbon ages for shells or tufa that are older than the youngest wood ages at that altitude are less likely to be correct than the wood ages. The older shell ages may reflect unrecognized reservoir effects in the lake at that location and altitude. Lake Bonneville occupied a hydrographically closed basin dur- ing its transgressive and regressive phases, and therefore lake level was not stable during those time periods. Even during the over- flowing phase of the lake, during which the Bonneville flood occurred and the Provo shoreline was formed, lake level was not constant (Miller et al., 2013). From outcrop studies, Miller et al. (2013) suggested that the Bonneville flood was older than previ- ously thought, and that the Red Rock Pass overflow threshold was progressively raised (probably because of landslide movement) through Provo-shoreline time. The highest-altitude Provo-shore- line barriers are the youngest components of the shoreline. Conclusions from study of Fig. 13 are listed below as a set of hypotheses that should be tested by further work. Core in- Fig. 12. (A) Comparison of tufa-based lake-level curve (black line) for Lake Lahontan simplified from Fig. 12 of Benson et al. (2013) with lake levels defined from tephra terpretations could be useful in tests of these hypotheses, if out- outcrops (blue stars) and radiocarbon ages (red dots) of organic carbon from shore crops and cores are reliably correlated (the measured age of environments (blue line). THS is Trego Hot Springs tephra and MSHM is Mt. St. Helens potentially correlative outcrop and core sediment is a necessary but M tephra. All tephra ages are from Benson et al. (2013). Radiocarbon ages for the not sufficient criterion for reliable correlations because all age es- revised curve are from Born (1972), Adams and Wesnousky (1998), Briggs et al. (2005), timates have uncertainties, some of which are large). In- Adams et al. (2008), and Adams (2010) and plotted at the center of their 2-sigma calibrated age ranges. (B) Oxygen isotope record from core PLC92B collected from terpretations of core proxies should be consistent with the Pyramid Lake (based on data in Benson et al., 2013).

4.3. Bonneville: lake level records and rebound effects

Gilbert (1890) noted that the major shorelines of Lake Bonne- ville (Bonneville and Provo shorelines) were higher in the center of the basin than they were at the margins, and he reasoned that the most likely cause of this fact was that the shorelines had been uplifted in response to the removal of the water load as Lake Bonneville evaporated. Following Gilbert’s work, isostatically rebounded Bonneville shorelines have been a valuable source of data for determining the elasticity of the crust and viscosity of the upper mantle in the eastern Great Basin (e.g., Crittenden, 1963; Bills et al., 1994). Currey (1982) and Currey et al. (1984) mapped the major shorelines of Lake Bonneville using aerial photographs and 1:24,000-scale topographic maps, plus field checking in some cases. Repeated tests of Currey’s mapping of the Bonneville and Provo shorelines, over many years and throughout the basin (e.g., Oviatt, 1989; Miller et al., 2012), have demonstrated that his data set is accurate and reproducible, although the precision of both the horizontal and vertical measurements of shoreline altitudes can Fig. 13. Diagram showing calibrated radiocarbon ages, basaltic ashes, and proposed now be improved and additional data points on the major shore- lake-level changes in the Bonneville basin during last w32,000 yr. Altitudes adjusted lines can be obtained using GPS and high-resolution digital- for differential rebound (see text). Basaltic ashes: HV, Hansel Valley; PE, ; elevation models. PB, Pahvant Butte; TH, Tabernacle Hill. U1, U2, U3 are unnamed transgressive-phase Since the 1950s, over 1000 radiocarbon ages have been obtained oscillations. Swan Lake age, from sediment cores in Swan Lake basin (Rosario et al., 2013), plotted separately because it is for organic materials formed in equilibrium for samples collected from sites in the Bonneville basin, mostly with the atmosphere, and may be the only such age available from the regressive from outcrops. Compilations and reviews (Scott et al., 1983; Currey phase of the lake. It is younger than the youngest overflow at Red Rock Pass. Dashed and Oviatt, 1985; Oviatt et al., 1992) have demonstrated that most dark-blue and dashed red lines are from Miller et al. (2012) and show alternative in- of the ages from outcrop samples are not useful for reconstructing terpretations of lake-level change during development of Provo shoreline. Dashed gray line shows interpretation of lake-level change during Bonneville lake cycle. Sources of lake history. In many cases, the relationship between the sample information: Scott et al., 1983; Currey and Oviatt, 1985; Oviatt and Nash, 1989; Oviatt fi and lake level has not been suf ciently established. In other cases et al., 1992; Oviatt, 1997; Oviatt et al., 2003, 2005; Madsen et al., in press; Oviatt and the measured age of a sample conflicts with other ages for materials Nash, in press. 48 M.C. Reheis et al. / Quaternary Science Reviews 97 (2014) 33e57 hypothetical constraints in this list until integrated tests call for the hypotheses to be rejected.

1. Lake Bonneville rose rapidly from low altitudes, similar to those of modern GSL, at about 30,000 cal yr B.P., probably caused by diversion of the upper Bear River into the Bon- neville basin (see discussion above under Section 3.5). 2. The lake fluctuated at moderate altitudes in the closed basin until about 25,000 cal yr B.P. (Oviatt et al., 1992). 3. The lake rose rapidly in the closed basin between about 25,000 and 21,000 cal yr B.P. (Oviatt et al., 1992; Oviatt and Nash, 2014). 4. From w21,000 to w18,000 cal yr B.P. the rate of transgression decreased (Oviatt et al., 1992) probably because of an in- crease in groundwater outflow from the basin as the lake reached higher altitudes and began to approach topographic divides composed of unconsolidated alluvium, and (or) because of a change in climate (decrease in precipitation, or increase in evaporation in the basin). 5. The Bonneville flood occurred as soon as the lake reached its highest altitude and began to overflow across the alluvial-fan dam just north of Red Rock Pass (Gilbert, 1890, p. 175) after 18,400 cal yr B.P.; a round-number estimate of the flood age is 18,000 cal yr B.P. (the flood age is constrained by charcoal ages from a pre-Bonneville paleosol 7 m below the Bonne- ville shoreline, and a number of tufa and gastropod ages from the Provo shoreline e refer to Fig. 13). 6. The age of the end of Provo-shoreline development is probably between 16,500 and 15,000 cal yr B.P. (Godsey et al., 2011; Miller et al., 2013); ages and dynamics of the Provo shoreline are currently under study. 7. There is no outcrop evidence for a “mid-Provo regression” Fig. 14. Comparison of the Lake Bonneville lake-level chronology (A), which was (Godsey et al., 2011). derived from radiocarbon ages of samples collected from outcrops plus limited infor- 8. The final regression, as a closed-basin lake after formation of mation from cores (simplified from Fig. 13), with (B) total inorganic carbon (TIC) and fi the Provo shoreline, was rapid (Oviatt, 2014). interpretations of the Blue Lake core (BL04-4; data from Benson et al., 2011, gure 2 and supplementary data). A. G ¼ Gilbert episode; P ¼ Provo shoreline; B ¼ Bonneville 9. The Gilbert episode was rapid and occurred at about shoreline; TO ¼ transgressive-phase oscillations; S ¼ Stansbury oscillation. B. Inter- 11,600 cal yr B.P., when the lake rose to about 1295 m from an preted time intervals (Benson et al., 2011): GI ¼ Gilbert; PR ¼ Provo; HS ¼ highstand; average low level similar to that of modern GSL; it is unlikely ST ¼ Stansbury. Alternative interpretations of the TIC curve (G? ¼ Gilbert episode; that the Gilbert lake reached altitudes as high as Currey’s eP? ¼ end Provo; BF? ¼ Bonneville flood) based on ostracode in samples from the BL04-4 core observed by Oviatt in 2013, observations of the BL04-4 core and core (1982) Gilbert shoreline in the central and western parts of photos by Oviatt in 2005 and in 2013, and observations of outcrops and cores in the the basin (Oviatt, 2014). Bonneville basin by Oviatt and colleagues, some of which have been published (e.g., 10. Wetlands began forming on the basin floor after the lake had Oviatt, 1987, 1997; Oviatt et al., 1994, 2005). evaporated, created by discharge of the large volume of groundwater stored in piedmont and mountain aquifers when Lake Bonneville was high (Oviatt et al., 2003, 2005; margin of the Bonneville basin. Though there is broad agreement in Oviatt, 2012). lake level as interpreted from the core vs. outcrop, important de- tails are in conflict and these conflicts suggest alternatives to the In some cases, core and outcrop information together provide published core-proxy interpretations. In the following discussion, “ ” “ ” insights that cannot be derived from just one or the other. An A refers to Fig. 14A (outcrop data) and B to Fig. 14B (core data). example of this is the age and altitude limits of the Hansel Valley ash (see section 3.5). Three 14C ages on bulk organic carbon and one 4.3.1. Beginning of Bonneville on pollen collected near the ash in two sediment cores from the A. Lake Bonneville began to rise abruptly and rapidly about Great Salt Lake suggest an approximate age of 25,000 14C yr B.P. 30 cal ka (based on depositional settings of the dated Hansel Valley (w30,000 cal yr B.P.; Thompson et al., 1990). This age places the ash bed; see discussion above under Section 3.5). B. In core BL04-4 eruption of the Hansel Valley ash at a time that does not conflict total inorganic carbon (TIC) varies considerably in the presumed with the lowest-altitude basal-stratigraphic wood age in the Bon- time interval of the initial rise of Lake Bonneville, but there is no neville sequence (Fig. 13). Outcrops plus cores have provided an major and consistent shift in TIC near this depth in the core. estimate of the altitudinal range and upper limit of ash occurrences. A comparison of the Bonneville shoreline chronology (Fig. 13) 4.3.2. Transgressive-phase oscillations with the interpretations of the Blue Lake core BL04-4 (Benson et al., A. Three major transgressive-phase oscillations (TO) have been 2011) provides an example of how interpretations of a core that identified in outcrops (Oviatt, 1997), and other oscillations probably were not fully integrated with information from outcrops can lead occurred in the closed-basin lake. Their potential correlation with to an incomplete or inaccurate reconstruction of lake history iceberg-rafting events in the North Atlantic Ocean, as suggested by (Fig. 14). Benson et al. (2011) used paleomagnetic secular variation Oviatt (1997), should be examined closely as better age control (PSV) ages to interpret core BL04-4 from Blue Lake on the western becomes available in both areas. B. Post-Stansbury transgressive- M.C. Reheis et al. / Quaternary Science Reviews 97 (2014) 33e57 49 phase lake-level oscillations cannot be determined from the TIC 4.4. Late Quaternary history of Owens Lake: core and outcrop record in BL04-4, because the core record is incomplete and TIC and comparisons other proxies in the Blue Lake core seem to be insensitive to these lake-level changes. Owens Lake has been the site of numerous stratigraphic core investigations, with one of the earliest by Smith and Pratt (1957). 4.3.3. Stansbury oscillation Core OL-92 revealed a hydrologic history extending back A. There is unequivocal outcrop evidence of two lake-level 800,000 yr (Smith and Bischoff, 1997). More detailed studies on OL- fluctuations on the order of 45 m, referred to collectively as the 92 and additional cores have documented the timing of Quaternary Stansbury oscillation (Oviatt et al., 1990). B. The Stansbury oscilla- hydrologic changes and mountain glaciation during the past d18 d13 tion (ST) is marked in the PSV time interval that is approximately 155,000 yr based on proxy indicators including: O and C, correlated with the calibrated age of the oscillation as determined cation exchange capacity, total inorganic and organic carbon, from outcrop, although the TIC data are interpreted to indicate a mineralogy, magnetic susceptibility, and sedimentology (Newton, longer period of lower lake level. 1991; Benson et al., 1996, 1997a,b; 2002; Bischoff et al., 1997; Menking et al., 1997; Smith and Bischoff, 1997; references therein; Li et al., 2000; Smoot et al., 2000; Bischoff and Cummins, 4.3.4. Highstand 2001), and biological proxy climatic indicators such as fossil di- A. The Bonneville flood probably occurred at about 18 cal ka atoms, ostracodes, and pollen (e.g., Bradbury, 1997; Carter, 1997; immediately after the lake transgressed to the low point on the Forester, 2000; Mensing, 2001). Three different and shallower basin divide and began to overflow to the drainage sediment cores were also used to characterize the Holocene stra- basin. Outcrops and shoreline geomorphology suggest there tigraphy to make paleoclimatic inferences using lithology, miner- probably was no prolonged period of overflow prior to the flood alogy, magnetic properties, and (Newton, 1991). Another (see hypothesis 5 in the list above). B. The beginning of the high- investigation of Benson et al. (2002) used changes in d18O and TIC stand is placed at the abrupt increase in TIC at 18.59 ka. Although values to suggest that the western Great Basin experienced five Benson et al. (2011) did not recognize it, the abrupt increase in TIC distinct intervals of climatic variability in the Holocene. The past at the reported PSV age of 18.59 ka in BL04-4, may actually mark the 1000 yr was investigated on interannual to decadal time scales to Bonneville flood (BF?, Fig. 14A). If so, the estimated PSV age is older infer the frequency and duration of lake-level fluctuations and than the charcoal ages from outcrops that indicate the flood was regional climatic variability (Li et al., 2000; Smoot et al., 2000). younger than 18.4 ka. The existence of the flood event (one of the largest documented paleo-floods in ) and a deter- 4.4.1. Outcrop records of Owens Lake mination of its age can only be deciphered from outcrop observa- Recent study of a previously undated, older and higher shoreline tions, although if the event is correctly identified in cores its age can (1180 m shoreline) in Owens Valley provided geomorphic and be verified and it can be used as a valuable stratigraphic tool, stratigraphic evidence of a large pluvial lake of OIS 6 age, based on analogous to a well-characterized . cross-cutting relations with dated lava flows of the Big Pine vol- canic field and 36Cl model age of 160 32 ka on tufa (Jayko and 4.3.5. Provo Bacon, 2008). The more recent late Pleistocene (OIS 2) to early fl A. The Bonneville ood marks the beginning of the period of Holocene stratigraphic and geomorphic history of Owens Lake fl over ow at Red Rock Pass and the beginning of formation of the basin (Fig. 8) has been investigated in several studies (Lubetkin and Provo shoreline. The end of Provo time may have been at about Clark, 1988; Beanland and Clark,1994; Bacon et al., 2006; Bacon and 15 cal ka or earlier (Godsey et al., 2011; Miller et al., 2013). There is Pezzopane, 2007; Orme and Orme, 2008). Numerous 14C dates on no outcrop evidence for a mid-Provo lake-level oscillation (Godsey charcoal, carbonized wood, shell, and tufa, in addition to teph- et al., 2011) as called for by Benson et al. (2011).B.Benson et al. rochronology, provide age control on gravelly and sandy shoreline (2011) thought Provo-shoreline formation was represented by the and nearshore deposits in the basin. There are two distinct Last- “ ” time interval marked by PR in the Blue Lake core. An alternative Glacial highstand shorelines in southern Owens Valley that consist hypothetical interpretation, which should be tested with further of well-developed gravelly beach ridges, wave-formed platforms work on outcrops and cores, is that the end of Provo time is marked cut into bedrock and alluvium, and sandy fan deltas at elevations of fi “ ” by the rst abrupt increase in TIC at eP? . The overlying interval of about 1160 m and 1145 m (Fig. 15). These two shorelines mark the slightly lower TIC values, overlain by a second abrupt TIC increase, times of overflows and subsequent downcutting across the sill from might represent a brief post-Provo oscillation followed by rapid 1160 m to the present elevation of 1145 m; their ages are con- lake-level decline. strained by 14C dates on tufa and shell of w25,000 cal yr B.P. and w15,500 cal yr B.P., respectively (Lubetkin and Clark, 1988; Bacon 4.3.6. Gilbert episode et al., 2006). Owens Lake during the PleistoceneeHolocene tran- A. The Gilbert episode occurred at about 11.6 cal ka and probably sition (w15,500 to 11,500 cal yr B.P.) is characterized as having an lasted for a very short time, less than the two-sigma age range of a overall regression consisting of extreme lake-level oscillations be- typical radiocarbon age (Oviatt, 2014). B. Benson et al. (2011) tween the elevations of w110 0e1139 m based on stratigraphy and assigned the Gilbert time period (GI) to a relatively long period of 14C dates on charcoal, shells, and tufa (Koehler, 1995; Bacon et al., low TIC that corresponded with the time period as 2006; Bacon and Pezzopane, 2007; Orme and Orme, 2008). defined in Greenland ice cores. However, during this time interval Sequence stratigraphic analysis and 14C dates on charcoal and outcrops indicate the lake in the Bonneville basin was low- tufa plus tephrochronology on beach ridge deposits provide evi- dprobably within a few meters of the average altitude of modern dence of an early Holocene transgression and multiple oscillations Great Salt Lakedand the Gilbert episode had not yet occurred of Owens Lake at several locations (Fig. 15; Bacon et al., 2006; Bacon (Oviatt, 2014). The abrupt rise in TIC marked by “G?” may represent and Pezzopane, 2007). This transgression began at an elevation of the Gilbert episode at the Blue Lake site, possibly caused by an w1100 m around 11,500 cal yr B.P. and appears to have attained a influx of Sevier-basin water and/or by a large increase in organic brief highstand up to w1135 m before dropping and stabilizing to productivity, and consequent carbonate precipitation, in the form well-developed shoreline features near an elevation of 1120 m shallow water of the Gilbert lake at this site (Oviatt, 2014). at w8000 cal yr B.P. Orme and Orme’s (2008) detailed investigation 50 ..Rhi ta./Qaenr cec eiw 7(04 33 (2014) 97 Reviews Science Quaternary / al. et Reheis M.C. e 57

Fig. 15. Revised lake-level curve of Owens Lake between w27,000 cal yr B.P. and present. Lake-level curve was previously developed from compilation of 45 14C dates, tephra correlation and paleomagnetic secular variations from 12 studies of lacustrine and fluvial-deltaic stratigraphy, packrat middens, and hydrologic proxy indicators from sediment core studies in Owens Lake basin and Owens Valley (Bacon et al., 2006; references therein). Revised curve adds results of recent investigations of late Holocene shoreline and aeolian depositional history of Owens Lake basin that directly date or bracket the age of four previously undated shorelines at elevations of 1108 m, 1103 m, 1101 m, and 1099 m (Bacon et al., 2013). Also shown are periods of relatively wet, variable, and dry conditions inferred from core proxy indicators of Benson et al. (1996, 1997a; 2002) for w27,000e3100 cal yr B.P. and of Li et al. (2000) for w1000e 135 yr B.P. Positions of present Owens Lake sill at 1145 m and historical highstand (w1880 A.D.) at 1096 m also shown. M.C. Reheis et al. / Quaternary Science Reviews 97 (2014) 33e57 51 failed to identify this early Holocene transgression, perhaps outcrops, from which the relative direction of lake-level oscillations because their numerous 14C dates from shells were from samples was determined (Bacon et al., 2006; Bacon and Pezzopane, 2007), obtained from stratigraphic sections in lower exposures of beach are largely based on the fundamentals of sequence stratigraphy. gravels. The stratigraphic proxy indicators commonly used in sediment core The late Holocene shoreline record in Owens Lake basin has studies at Owens Lake to indicate shallow to dry lake level include recently been refined by detailed geomorphic mapping and lacus- frosted (etched) quartz grains, coarse sand, oolites, and rubified trine and aeolian stratigraphic studies on previously undated de- clay with prismatic soil structure, all of which are indicators of posits (Bacon et al., 2013)(Fig. 15). Four well-developed late shallow to dry lake conditions near the depocenter of the basin Holocene (pre-historical) shorelines surround Owens Lake playa at (Benson et al., 1996, 1997a; Smith et al., 1997). The timing of change elevations of w1108, 1103, 1101, and 1099 m in the form of wave-cut in cores is based on 14C dates on organic-rich sediment (humates). notches, gravelly and sandy beach ridges, and sandy delta bars. In addition to the sediment hiatuses, the variation of d18O and d13C The 1108 m shoreline was previously inferred by Bacon et al. values and other geochemical indicators may also indicate changes (2006) as late Holocene in age (<4300 cal yr B.P.) with an age (Benson et al., 1996, 1997a, 2002; Li et al., 2000), thus potentially closer to w3500 cal yr B.P., based on relative degree of preservation, constraining the direction (though not the amount) of lake-level geomorphic and topographic cross-cutting relations, and regional oscillations in the absence of outcrop data. paleoclimatic proxy evidence. Investigations of interbedded sandy and gravelly deposits within a well-developed beach ridge at 4.4.3. Revised lake-level curve for Owens Lake w1108 m in the Swansea embayment yielded two new infrared The lake-level curve for Owens Lake presented in this review stimulated luminescence (IRSL) dates of 3620 260 and revises the Bacon et al. (2006) curve (Fig. 15). The revised curve, 3490 290 yr B.P. (Bacon et al., 2013). The late Holocene beach based on all the types of records discussed in the previous section, ridge cross-cuts older PleistoceneeHolocene shoreline deposits now includes the results from four previously undated late Holo- studied by Orme and Orme (2008) and was constructed at the base cene shoreline features and vegetated (Bacon et al., 2013), as of early Holocene lake plains (Bacon et al., 2006). Soil-geomorphic well as the timing and duration of additional proxy indicators of characteristics of the beach ridge confirm the IRSL ages and wet and dry conditions from sediment cores during the entire demonstrate that late Holocene water levels reached an elevation Holocene (Li et al., 2000; Benson et al., 2002). The revised lake-level as high as 1108 m at Owens Lake. curve shows a regression of Owens Lake from its latest Pleistocene The lower shoreline at an elevation of 1103 m has a direct age highstands at altitudes of w1160 and 1145 m between w27,000 and constraint from natural exposures through a well-developed beach 15,300 cal yr B.P., respectively, and an overall drop in water level of ridge at Carroll Creek on the western margin of the basin. Here, w45 m by w11,400 cal yr B.P. detrital charcoal preserved in sandy beach rock at the crest of a Between 18,000 and 11,000 cal yr B.P., six w1000e2000-yr beach ridge yielded a 14C date of 870 40 yr B.P. (910e700 cal yr high-frequency lake-level oscillations occurred with changes in B.P.). Erosional shorelines below 1103 m were dated by employing elevations up to 60 m as reflected by the integrated core indicators sediment sampling from hand-augered boreholes at the crests of of “dry” conditions with the age of dated shoreline and nearshore two parallel and vegetated liner ridges near the Swansea deposits. At the Last Glacial termination, a lowstand at embayment that produced five IRSL ages in morphostratigraphic w11,200 cal yr B.P. was followed by a transgression to a short-lived order. The outer dune ridge was sampled to a depth of 4 m and early Holocene highstand near 1135 m and a subsequent drop to yielded IRSL ages that indicate two episodes of sand accumulation form shoreline features at 1120 m at w7860e7650 cal yr B.P. The at 760 90 and 410 40 yr B.P. The inner dune ridge was sampled early Holocene highstand documented by geomorphic data is to a depth of 2.5 m and produced two IRSL ages. Alluvium at the supported by d18O and TIC values in sediment cores of Benson et al. base of the section yielded an age of 5000 210 yr B.P. and the age (2002), who concluded that relatively wet conditions prevailed of overlying aeolian deposits indicate sand accumulation at from 10,000e8000 cal yr B.P. (Fig. 15). 300 30 yr B.P. The 14C age of the 1103-m beach ridge is supported Core data indicate that the lake lowered w30 m to near desic- by the oldest IRSL age of 820 30 yr B.P. from overlying aeolian cation levels between w6500 and 4400 cal yr B.P. (Fig. 15) based on deposits that bury its associated lake plain. The ages of the 1101 and oolite and sand deposition recording shallow and oscillating lakes 1099 m shorelines are bracketed to between 300 30 and in the basin at 8000e6500 cal yr B.P. (Benson et al., 1997a,b; Smith 400 30 yr B.P., based on cross-cutting relations with laterally et al., 1997) and d18O and TIC values suggesting drought conditions continuous wave-cut notches that only occur on the outer dune at 6500e3800 cal yr B.P. (Benson et al., 2002). Outcrop records ridge (Bacon et al., 2013). The historical maximum lake level of support extremely low lake levels at this time based on an IRSL age Owens Lake reached w1096 m around 1880 A.D., which similarly of 5000 210 yr B.P. from alluvial fan sediment sampled at an eroded into the base of the inner (300 30 yr B.P.) dune ridge elevation of w1099 m (Bacon et al., 2013). This period of shallow to forming erosional shoreline features in the area. near-desiccation water levels was followed by a w25-m rise in lake level to an elevation of 1108 m. The lake stabilized at this elevation 4.4.2. Integration of sediment core and outcrop records at Owens and constructed a gravelly beach ridge at 3540 340 IRSL yr B.P. Lake: Late Pleistocene to Holocene This direct age is supported by proxy indicators that indicate the At least 10 significant lake-level oscillations and many more onset of wet conditions and a “deep” lake at w3100 200 cal yr B.P. minor fluctuations are distinguished in the latest Quaternary lake- based on a shift in d18O values and stabilizing of TIC and magnetic level history of Owens Lake by the integration of sediment core data susceptibility in core records (Benson et al., 2002)(Fig. 15). sets with outcrop records (Fig. 15; Bacon et al., 2006). The lake-level Geomorphic and sediment core records for the interval curve of Owens Lake from w27,000 cal yr B.P. to present was w3000e1000 yr B.P. are currently being investigated. The newly developed from compilation of 45 14C dates combined with tephra obtained age constraints on the 1103, 1101, and 1099 m shorelines, correlation and paleomagnetic secular variations from 12 in- discussed above, are in general agreement with the core record of Li vestigations of the lacustrine and fluvial-deltaic stratigraphy, et al. (2000) that indicates an effectively dry climate between paleoenvironmental indicators in packrat middens, and hydrologic w1065 and 795 yr B.P. corresponding to the Medieval Climatic proxy indicators from sediment core studies throughout Owens Anomaly during which Owens Lake approached playa conditions. Lake basin and Owens Valley. The stratigraphic analyses of They also inferred relatively wet climatic conditions from w795 to 52 M.C. Reheis et al. / Quaternary Science Reviews 97 (2014) 33e57

535 yr B.P. coeval with the 1103 m shoreline dated at w800 cal yr or subunits through cuts that are perpendicular to the B.P. Li et al. (2000) also suggest that regional climate became colder shoreline, the upper and lower altitude limits of each subunit were with Owens Lake having frequently oscillating lake levels between defined in a particular area. The records from each area were then w535 and 255 yr B.P., which corresponds to the time of the Little correlated using 14C ages to assemble a coherent lake-level record. and construction of the 1101 and 1099 m shorelines Additional constraints on lake level are provided by 14C-dated lake (Fig. 15). Since w135 yr B.P. [w1880 A.D.], the lake level has steadily deposits in the Coyote subbasin to the north (Fig. 9; Miller et al., dropped from its historic highstand at an elevation of w1096 m due 2011a). Lake Manix can overflow into this subbasin when lake to strong anthropogenic impacts in Owens Valley (e.g., Li et al., level exceeds 541 masl, the altitude of the bedrock threshold be- 2000). tween them. Lake Manix, which was probably shallow due to In summary, the late Quaternary lake-level record of Owens previous sediment infilling, achieved multiple highstands during Lake demonstrates that integration of sediment core datasets and OIS 3 and early OIS 2. A preliminary lake-level curve, constrained outcrop records is possible and yields a combined result that is at by > 60 calibrated 14C ages on freshwater mollusk shells, indicates sufficient resolution to test paleoclimatic models with a high de- rapid fluctuations between 45 and 25 ka and as many as 8 high- gree of certainty. Furthermore, the refined Holocene portion of the stands within 10 m of the 543-m upper threshold for Lake Manix lake-level record of Owens Lake correlates well with the sediment- (Reheis et al., 2010). and outcrop-based records of Lake Elsinore, Little Bear Lake, and Tulare Lake in coastal southern and south-central California (Kirby 4.6. Other lake-level records in the Great Basin et al., 2012), indicating that similar episodes of climate change are also reflected along the southcentral-eastern Sierra Nevada A review of lake-level records from outcrop would be incom- Mountains. plete without a summary of the many studies published on other pluvial lakes in the region. Many of these have focused on recon- 4.5. Lake Manix: integration of core and outcrop records structing lake-level fluctuations during OIS 2 and are based on only a few dates, but a few are more comprehensive or extend to older Combined outcrop and core studies have yielded a paleohy- records. In some basins, reservoir effects complicate the interpre- drologic record for Lake Manix (Fig. 9)(Jefferson, 2003; Reheis and tation of radiocarbon ages, and the dated sediments are often not Redwine, 2008; Reheis et al., 2010, 2012). As described in section from highstand positions or well-understood stratigraphic con- 3.4, the lake was greatly influenced by an intra-basin integration texts. These lake-level records are being used to test competing event that occurred early in OIS 6. Lake levels were also controlled hypotheses on the source(s) of moisture that fed lake highstands by another shallow intra-basin threshold between the main lake following the : north-south migration of basin and the Coyote subbasin to the north (Meek, 1994; Miller westerlies (e.g., Antevs, 1948; Benson and Thompson, 1987; et al., 2009) as well as possible episodic diversions to Harper Lake Bartlein et al., 1998) vs. northward transport of tropical moisture upstream of the Manix basin (Fig. 7; Meek, 1999; Cox et al., 2003; (Lyle et al., 2012). Garcia et al., 2014). Thus, detailed mapping in addition to coring Few ages from nearshore deposits are available to constrain was required to shed light on the possible effects of these di- highstands of Oregon pluvial lakes (Fig. 1). Negrini (2002) synthe- versions on the lake-level record. sized available ages for Lakes Chewaucan and Fort Rock to infer that The 45-m Manix core, obtained from the older part of the lake the lakes were at moderate levels during late OIS 3, dropped to low basin, records the initial influx of water from upstream at about 450 levels, then rose gradually beginning around 25 cal ka to highstands ka (Reheis et al., 2012). Sedimentology, stable isotope, and ostra- at about 20e18 cal ka (Allison, 1982); a lower highstand occurred at code faunal analyses indicate that lakes were present repeatedly about 13.8 cal ka (Licciardi, 2001). For Lake Alvord, outcrop data and consistently from OIS 12 through OIS 2, including during indicate a deep lake at and prior to ca 14 cal ka that was probably interstadial OIS 3 and at least parts of OIS 5, 7, and 9; discharging eastward into Lake Coyote, triggering at least one the core site was dry during OIS 11. These results are interpreted to catastrophic flood that temporarily connected the Alvord basin indicate that precipitation and runoff responded to atmospheric with a tributary of the Owyhee River (Carter et al., 2006). circulation patterns that were different from those thought to Across northern and central California and Nevada, several lake control lake levels in the central and northern Great Basin, which records have recently been obtained. A lake-level curve for Lake were generally dry or shallow and saline during interglacials (e.g., Surprise (Fig. 1), based on U-series and 14C dating of tufa collected Smith et al., 1997; Oviatt et al., 1999; Forester et al., 2005). The from wavecut scarps, indicates that this lake stood at moderate ostracode-based stable isotopes yielded a d18O record dominated levels from w27 to 20 cal ka, then rose gradually to a highstand at by evaporation effects in this low-desert lake. Relatively shallow 15 ka (Ibarra et al., in press). Munroe and Laabs (2013a) produced a lakes, deltas, and mudflat conditions are associated with the most lake-level curve for Lake Franklin (Fig. 7) based on numerous 14C positive d18O values at the core site during OIS 6-2 (Reheis et al., ages, including previously published dates from Lillquist (1994). 2012). These conditions were triggered by the failure of the Franklin’s record is similar to that of Surprise but has an earlier Buwalda Ridge threshold, which allowed Lake Manix to increase its highstand with a double peak at about 17 and 16 cal ka. Lakes surface area by expanding into the deeper Afton subbasin, and later Clover and Waring, to the east, have yielded a few highstand ages reflect the encroachment of the Mojave River delta from the west. ranging from about 17 to 19.5 cal ka (Garcia and Stokes, 2006; The history of Lake Manix during OIS 3-2 is derived entirely Munroe and Laabs, 2013b). Farther south, a lake-level curve for from mapping, measured sections, and dating of freshwater clams Lake Newark indicates a highstand at about 16.9 cal ka (Redwine, that today live exclusively in shallow water, generally less than 2 m 2003; Reheis et al., 2003), and nearby Lake Jakes has a similar deep (Berger and Meek, 1992). The highest beach deposits at and highstand age (Garcia and Stokes, 2006). Lake Diamond was high below 543 m altitude consist of multiple transgressiveeregressive and overflowing from about 26 to 16.5 cal ka; its hydrologic packages of lacustrine sediment that are typically separated in maximum is unknown because of threshold incision during this outcrop by layers of clasts coated with tufa deposited during period (Tackman, 1993). Lake Columbus apparently reached its transgressions, and locally by intervening thin (commonly <1m highstand somewhat earlier, at about 18 cal ka (Kurth et al., 2011). thick) alluvial-fan deposits with buried soils (Reheis and Redwine, Lake Russell, in the Mono basin (Fig. 8), has a longer lake-level 2008; Reheis and Miller, 2010). By physically tracing these packages record that extends at least into OIS 3 (Benson et al., 1990, 2003) M.C. Reheis et al. / Quaternary Science Reviews 97 (2014) 33e57 53 and likely into OIS 4 (Zimmerman et al., 2006, 2011); the chro- youngest lakes (e.g., Reheis et al., 2002a; Redwine, 2003; Kurth nology is currently being debated. Older shorelines are also pre- et al., 2011), as well as complex relations among basins and sub- served (Reheis et al., 2002b). The OIS 2 highstand occurred 16e basins controlled by changing threshold altitudes (e.g., Adams et al., 15 cal ka (Lajoie, 1968; Benson et al., 1998). 1999; Meek, 1999). Lake-level records from outcrop provide critical Lakes in southern California and Nevada (Fig. 8) lie in low- controls on interpretation of lake level from hydrologic proxies in altitude desert basins and are more influenced by evaporation cores, and can yield evidence of rapid, well-constrained lake-level than the northern Great Basin lakes. Lakes Searles and Panamint fluctuations during the late Pleistocene and Holocene (e.g., Adams, were fed by progressive lake spillover and do not contain complete 2003a, 2007; Bacon et al., 2006; Miller et al., 2012). Conflicts be- hydrologic records. The Searles record was studied by G.I. Smith tween outcrop and core records may point to inconsistencies in (1979, 2009), combining core and outcrop data with 14C and Ue dating, age models, or interpretation of hydrologic proxies and Th ages (Lin et al., 1998) and summarized by Phillips (2008) in the provide direction to further investigations. The careful combination context of the Owens River drainage system. Searles Lake had of outcrop- and core-based data can yield more powerful insights multiple fluctuations from low to moderate levels from w37 to on paleohydrology and paleoclimate than can be achieved 27 cal ka, followed by a sharp rise and overflow to Panamint from separately. 27 to 19.5 cal ka. By 18 ka, the lake had dropped to very low levels; then it spiked again to overflow between 16 and 15 cal ka. Lake References Panamint, farther down the chain of lakes, was also at a relatively high level w34e27 cal ka and again at about 22e20 cal ka based on Adams, K.D., 2003a. Age and paleoclimatic significance of late Holocene lakes in the 14 , NV, USA. Quat. Res. 60, 294e306. C ages on tufa and shells (Jayko et al., 2008). Lake Mojave, which Adams, K.D., 2003b. Estimating paleowind strength from beach deposits. Sedi- began to fill at around 25 ka (Reheis and Redwine, 2008), has a mentology 50, 565e577. combined outcrop and core-based lake-level record (Wells et al., Adams, K.D., 2004. Shorezone Erosion at Lake Tahoe: Historical Aspects, Processes, and Stochastic Modeling. Reno, NV, Desert Research Institute Final Technical 2003). This record is interpreted to indicate two major full-lake Report to the U.S. Bureau of Reclamation and Tahoe Regional Planning Agency, phases w22e20 and 17e13.5 cal ka, during which Lake Mojave 94 p. episodically discharged northward toward Death Valley. Core Adams, K.D., 2007. Late Holocene sedimentary environments and lake-level fluc- e sedimentology indicates that before, after, and between these two tuations at Walker Lake, Nevada, USA. Geol. Soc. Amer. Bull. 119, 126 139. Adams, K.D., 2010. Lake levels and sedimentary environments during deposition of phases, the lake fluctuated from high water to desiccation (Wells the Trego Hot Springs and Wono in the Lake Lahontan basin, Nevada, et al., 2003). USA. Quat. Res. 73, 118e129. Adams, K.D., 2012. Response of the to lowering base level at Pyramid Lake, Nevada, based on historical air photos and LiDAR data. Geosphere 8, 607e 5. Conclusions 627. Adams, K.D., Wesnousky, S.G., 1998. Shoreline processes and the age of the Lake The focus of this paper was to review outcrop-based studies of Lahontan highstand in the Jessup embayment, Nevada. Geol. Soc. Amer. Bull. 110, 1318e1332. pluvial lake deposits in the Great Basin that have been published in Adams, K.D., Wesnousky, S.G., Bills, B.G., 1999. Isostatic rebound, active faulting, and the last few decades. In addition, we summarized techniques best potential geomorphic effects in the Lake Lahontan basin, Nevada and California. suited to the study of lake-sediment outcrops, particularly with Geol. Soc. Amer. Bull. 111, 1739e1756. fi Adams, K.D., Goebel, Ted, Graf, Kelly, Smith, G.M., Camp, A.J., Briggs, R.W., respect to lake level. These studies have yielded signi cant insights Rhode, David, 2008. Late Pleistocene and early Holocene lake-level fluctuations on the interplay between Quaternary climatic and geologic controls in the Lahontan Basin, Nevada: implications for the distribution of archaeo- on lake records, and are critical to the accurate interpretation of logical sites. Geoarch 23, 608e643. Allen, B.D., Anderson, R.Y., 2000. A continuous, high-resolution record of late climate and lake-level proxies measured in cores of extant and Pleistocene climate variability from the Estancia basin, New Mexico. Geol. Soc. desiccated lakes. Amer. Bull. 112, 1444e1458. Outcrop studies of paleo-lake deposits have allowed recon- Allison, I.S., 1982. Geology of Pluvial Lake Chewaucan, Lake County, Oregon. In: struction of drainage-basin integration of important fluvial systems Oregon State Monographs, Studies in Geology, vol. 11, p. 79. Anderson, D.E., Wells, S.G., 2003. Latest Pleistocene highstands in Death Valley, in various tectonic settings in the Great Basin. Recent studies allow California. In: Enzel, Y., Wells, S.G., Lancaster, N. (Eds.), Paleoenvironments and a broad outline of the post-Miocene assembly of the Humboldt Paleohydrology of the Mojave and Southern Great Basin Deserts, Geol. Sci. e River (Reheis et al., 2002a, 2003; Wallace et al., 2008), one of the Amer. Spec. Paper 368, pp. 115 128. Anderson, R.Y., Allen, B.D., Menking, K.M., 2002. Geomorphic expression of abrupt most important drainages in the region, but much work remains to climate change in southwestern North America at the glacial termination. Quat. be done. Drainage-switching of pluvial Lake Russell (Mono Lake) Res. 57, 371e381. significantly affected lake levels in both Walker Lake and Owens Anderson, S.A., Link, P.K., 1998. Lake Bonneville sequence stratigraphy, Pleistocene Bear River Delta, Cache Valley, Idaho. In: Pitman, J.K., Carroll, A.R. (Eds.), Lake at different times and likely controlled the migration of Modern and Ancient Lake Systems: Utah Geol. Assoc. Guidebook 26, pp. 91e aquatic species (Reheis et al., 2002b). Integration of the Amargosa 104. River (Menges, 2008) and the episodic integration of the Owens Antevs, E., 1948. The Great Basin with emphasis on glacial and postglacial times, Part III: climatic Changes and pre-White Man. Bull. Univ. Utah Biological Series, River system certainly impacted lake levels in the downstream 168e191. basins throughout Pleistocene time (e.g., Morrison, 1999; Knott Atwood, G., 1994. Geomorphology applied to flooding problems of closed-basin et al., 2008; Phillips, 2008). Arrival of the Bear River in Lake Bon- lakes e specifically Great-Salt-Lake, Utah. Geomorphology 10, 197e219. Bacon, S.N., Pezzopane, S.K., 2007. A 25,000-year record of earthquakes on the neville profoundly affected lake level as well as chemical compo- Owens Valley fault near Lone Pine, California: implications for recurrence in- sition of lake waters (Bouchard et al., 1998). A subbasin-integration tervals, slip rates, and segmentation models. Geol. Soc. Amer. Bull. 119, 823e event in the Manix basin, first documented in outcrop, strongly 847. affected the paleohydrologic status of the lake as recorded in core Bacon, S.N., Burke, R.M., Pezzopane, S.K., Jayko, A.S., 2006. Last glacial maximum and Holocene lake levels of Owens Lake, eastern California, USA. Quat. Sci. Rev. proxies (Reheis et al., 2012). Thus, outcrop-based drainage inte- 25, 1264e1282. gration records are important in interpreting paleohydrologic Bacon, S.N., Lancaster, N., Stine, S., Rhodes, E.J., Holder, G.A.M., 2013. Refined late conditions in long-lived paleo-lake basins. Holocene lake-level history of Owens Lake, east-central California. Geol. Soc. Amer. Abstr. Prog. 45 (7), 552. Outcrop-based studies are the only means of accurate recon- Bartlein, P.J., Anderson, K.H., Anderson, P.M., Edwards, M.E., Mock, C.J., Thompson, R.S., struction of lake highstands; both outcrops and cores can provide Webb, R.S., Webb, T.I.I.I., Whitlock, C., 1998. Paleoclimate simulations for North evidence of lowstands. On a coarse scale, such studies have docu- America over the past 21,000 years: features of the simulated climate and com- parisons with paleoenvironmental data. Quat. Sci. Rev. 17, 549e585. mented many pre-late Pleistocene lakes in the western Great Basin Beanland, S., Clark, M.M., 1994. The Owens Valley fault zone, eastern California, and that were significantly larger and record wetter conditions than the surface rupture associated with the 1872 earthquake. U.S. Geol. Surv. Bull. 1982. 54 M.C. Reheis et al. / Quaternary Science Reviews 97 (2014) 33e57

Benson, L., 1994. Carbonate deposition, Pyramid Lake subbasin, Nevada; 1, Sequence Broecker, W.S., Kaufman, Aaron, 1965. Radiocarbon chronology of Lake Lahontan of formation and elevational distribution of carbonate deposits (tufas). Palae- and Lake Bonneville II, Great Basin. Geol. Soc. Amer. Bull. 76, 537e566. ogeogr., Palaeoclim. Palaeoecol. 109, 55e87. Broecker, W.S., Orr, P.C., 1958. Radiocarbon chronology of Lake Lahontan and Lake Benson, L., 2004. Western Lakes. In: Gillespie, A.R., Porter, S.C., Atwater, B.F. (Eds.), Bonneville. Geol. Soc. Amer. Bull. 69, 1009e1032. The Quaternary Period in the United States. Elsevier, New York, pp. 185e204. Bright, R.C., 1963. Pleistocene Lakes Thatcher and Bonneville, Southeastern Idaho Benson, L.V., Thompson, R.S., 1987. The physical record of lakes in the Great Basin. (Ph.D. thesis). Univ. Minn., Minneapolis. In: Ruddiman, W.F., Wright Jr., H.E. (Eds.), North America and Adjacent Oceans Carter, C., 1997. Ostracodes in Owens Lake core OL-92: alternation of saline and during the Last Deglaciation. Geological Society of America, Boulder, CO, United freshwater forms through time. In: Smith, G.I., Bischoff, J.L. (Eds.), An 800,000- States, pp. 241e260. year Paleoclimatic Record from Core OL-92, Owens Lake, Southeast California, Benson, L.V., Paillet, F.L., 1989. The use of total lake-surface area as an indicator of Geol. Soc. Amer. Spec. Pap. 317, pp. 113e120. climatic change: examples from the Lahontan basin. Quat. Res. 32, 262e275. Carter, D.T., Ely, L.L., O’Connor, J.E., Fenton, C.R., 2006. Late Pleistocene outburst Benson, L., Peterman, Z., 1995. Carbonate deposition, Pyramid Lake subbasin, flooding from pluvial Lake Alvord into the Owyhee River, Oregon. Geo- Nevada: 3. The use of 87Sr values in carbonate deposits (tufas) to determine the morphology 75, 346e367. hydrologic state of paleolake systems. Palaeogeogr., Palaeoclim. Palaeoecol. 119, Caskey, S.J., Reheis, M.C., Paces, J.B., 2013. New shoreline ages and OIS-6 lake-level 201e213. history of Lake Manly, Death Valley. Geol. Soc. Amer. Abst. Prog. 45 (4). Benson, L.V., Currey, D.R., Dorn, R.I., Lajoie, K.R., Oviatt, C.G., Robinson, S.W., Cohen, A.S., 2003. Paleolimnology, the History and Evolution of Lake Systems. Smith, G.I., Stine, S., 1990. Chronology of expansion and contraction of four Oxford University Press, New York. Great Basin lake systems during the past 35,000 years. Palaeogeogr., Palaeoclim. Cohen, A.S., Palacios, M., Negrini, R.M., Wigand, P.E., Erbes, D., 2000. A high reso- Palaeoecol. 78, 241e286. lution continental paleoclimate record for the middle-late Pleistocene from Benson, L.V., Meyers, P.A., Spencer, R.J., 1991. Change in the size of Walker Lake Summer Lake, Oregon, II: sedimentology, paleontology, and geochemistry. during the past 5000 years. Palaeogeogr., Palaeoclim. Palaeoecol. 81, 189e214. J. Paleolimnol. 24, 151e182. Benson, L., Kashgarian, M., Rubin, M., 1995. Carbonate deposition, Pyramid Lake Cox, B.F., Hillhouse, J.W., Owen, L.A., 2003. Pliocene and Pleistocene evolution of the subbasin, Nevada: 2. Lake levels and polar positions reconstructed Mojave River, and associated tectonic development of the Transverse Ranges from radiocarbon ages and elevations of carbonates (tufas) deposited in the and Mojave Desert, based on borehole stratigraphy studies and mapping of Lahontan basin. Palaeogeogr., Palaeoclim. Palaeoecol. 117, 1e30. landforms and sediments near Victorville, California. In: Enzel, Y., Wells, S.G., Benson, L.V., Burdett, J.W., Kashgarian, M., Lund, S.P., Phillips, F.M., Rye, R.O., 1996. Lancaster, N. (Eds.), Paleoenvironments and Paleohydrology of the Mojave and Climatic and hydrologic oscillations in the Owens Lake basin and the adjacent Southern Great Basin Deserts, Geol. Soc. Amer. Spec. Paper 368, pp. 1e42. Sierra Nevada, California. Science 274, 746e749. Crittenden, M.D., 1963. New data on the isostatic deformation of Lake Bonneville. Benson, L., Burdett, J., Lund, S., Kashgarian, M., Mensing, S., 1997a. Nearly syn- U.S. Geol. Surv. Prof. Pap. 454-E. chronous climate change in the Northern Hemisphere during the last glacial Currey, D.R., 1982. Lake Bonneville: Selected features of relevance to neotectonic termination. Nature 388, 263e265. analysis. U.S. Geol. Surv. Open-File Rep. 82e1070. Benson, L.V., Smoot, J.P., Kashgarian, M., Sarna-Wojcicki, A.M., Burdett, J.W., 1997b. Currey, D.R., 1990. Quaternary palaeolakes in the evolution of semidesert basins, Radiocarbon ages and environments of deposition of the Wono and Trego Hot with special emphasis on Lake Bonneville and the Great Basin, U.S.A. Palae- Springs tephra layers in the Pyramid Lake subbasin, Nevada. Quat. Res. 47, 251e ogeogr., Palaeoclim. Palaeoecol. 76, 189e214. 260. Currey, D.R., Oviatt, C.G., 1985. Durations, average rates, and probable causes of Lake Benson, L.V., Lund, S.P., Burdett, J.W., Kashgarian, M., Rose, T.P., Smoot, J.P., Bonneville expansions, stillstands, and contractions during the last deep-lake Schwartz, M., 1998. Correlation of late-Pleistocene lake-level oscillations in cycle, 32,000 to 10,000 years ago. In: Kay, P.A., Diaz, H.F. (Eds.), Problems of Mono Lake, California, with North Atlantic climate events. Quat. Res. 49, 1e10. and Prospects for Predicting Great Salt Lake Levels. , Univ. Utah Benson, L., Kashgarian, M., Rye, R., Lund, S., Paillet, F., Smoot, J., Kester, C., Center for Public Affairs and Administration, pp. 9e24. Mensing, S., Meko, D., Lindstrom, S., 2002. Holocene multidecadal and multi- Currey, D.R., Atwood, G., Mabey, D.R., 1984. Major levels of Great Salt Lake and Lake centennial droughts affecting northern California and Nevada. Quat. Sci. Rev. 21, Bonneville: Utah Geol. Miner. Surv. Map 73. 659e682. Dahms, D.E., 2004. Glacial limits in the middle and southern Rocky Mountains, Benson, L., Lund, S., Negrini, R.M., Linsley, B., Zic, M., 2003. Response of North U.S.A., south of the Yellowstone Ice Cap. In: Ehlers, J., Gibbard, P.L. (Eds.), American Great Basin lakes to Dansgaard-Oeschger oscillations. Quat. Sci. Rev. Quaternary Glaciationseextent and Chronology, Part II. Elsevier, Ltd., New York, 22, 2239e2251. pp. 275e288. Benson, L.V., Lund, S.P., Smoot, J.P., Rhode, D.E., Spencer, R.J., Verosub, K.L., Davis, J.O., 1978. Tephrochronology of the Lake Lahontan area, Nevada and Cali- Louderback, L.A., Johnson, C.A., Rye, R.O., Negrini, R.M., 2011. The rise and fall of fornia. Nev. Arch. Res. Paper 7, 137. Lake Bonneville between 45 and 10.5ka. Quat. Int. 235, 57e69. Davis, J.O., 1982. Bits and pieces: the last 35,000 years in the Lahontan area. In: Benson, L.V., Smoot, J.P., Lund, S.P., Mensing, S.A., Foit Jr., F.F., Rye, R.O., 2013. Insights Madsen, D.B., O’Connell, J.F. (Eds.), Man and Environment in the Great Basin. from a synthesis of old and new climate-proxy data from the Pyramid and Society for American Archeology, pp. 53e75. basins for the period 48 to 11.5 cal ka. Quat. Int. 310, 62e82. Davis, J.O., 1983. Level of Lake Lahontan during deposition of the Trego Hot Springs Berger, R., Meek, N., 1992. Radiocarbon dating of Anodonta in the Mojave River tephra about 23,400 years ago. Quat. Res. 19, 312e324. basin. Radiocarbon 34, 578e584. Davis, J.O., 1985. Correlation of late Quaternary tephra layers in a long pluvial Bills, B.G., Currey, D.R., Marshall, G.A., 1994. Viscosity estimates for the crust and sequence near Summer Lake. Or. Quat. Res. 23, 38e53. upper mantle from patterns of lacustrine shoreline deformation in the eastern Davis, J.O., 1987. Tephrochronology of middle Pleistocene lake sediments in Kobeh Great Basin. J. Geophys. Res. 99, 22,059e22,086. Valley, Nevada: implications for neotectonics and drainage history of the cen- Bischoff, J.L., Cummins, K., 2001. of the Sierra Nevada (79,000e tral Great Basin. Geol. Soc. Amer. Abstr. Progr 19 (6), 370. 15,000 yr B.P.) as recorded by rock flour in sediments of Owens Lake, California. Davis, O.K., 1999. Pollen analysis of a late-glacial and Holocene sediment core from Quat. Res. 55, 14e24. Mono Lake, Mono County, California. Quat. Res. 52, 243e249. Bischoff, J.L., Menking, K.M., Fitts, J.P., Fitzpatrick, J.A., 1997. Climatic oscillations Dugas, D.P., 1998. Late Quaternary variations in the level of paleo-Lake Malheur, 10,000e155,000 yr B.P. at Owens Lake, California reflected in glacial rock flour eastern Oregon. Quat. Res. 50, 276e282. abundance and lake salinity in core OL-92. Quat. Res. 48, 313e325. Enzel, Y., Wells, S.G., Lancaster, N., 2003. Late Pleistocene lakes along the Mojave Blackwelder, E., 1933. Lake Manly: an extinct lake of Death Valley. Geogr. Rev. 23, River, southeast California. In: Enzel, Y., Wells, S.G., Lancaster, N. (Eds.), Paleo- 464e471. environments and Paleohydrology of the Mojave and Southern Great Basin Blackwelder, E., Ellsworth, E.W., 1936. Pleistocene lakes of the Afton basin, Cali- Deserts, Geol. Soc. Amer. Spec. Paper 368, pp. 61e77. fornia. Amer. J. Sci. 231, 453e463. Forbes, D.L., Syvitski, J.P.M., 1994. Paraglacial Coasts. In: Carter, R.W.G., Born, S.M., 1972. Late Quaternary History, Deltaic Sedimentation, and Mudlump Woodroffe, C.D. (Eds.), Coastal evolution: Late Quaternary Shoreline Mor- Formation at Pyramid Lake, Nevada. Center for Water Resources, Desert phodynamics. Cambridge University Press, Cambridge, Great Britain, Research Institute, Reno. pp. 373e424. Bouchard, D.P., Kaufman, D.S., Hochberg, A., Quade, J., 1998. Quaternary history of Forester, R.M., 2000. An ostracode record of Holocene climate change from Owens the Thatcher Basin, Idaho, reconstructed from the 87Sr/86Sr and amino-acid Lake, California. In: Jayko, A.S., Miller, C.I. (Eds.), Impacts of Climate Change on composition of lacustrine fossils: implications for the diversion of the Bear Landscapes of the Eastern Sierra Nevada and Western Great Basin, pp. 7e8. U.S. River into the Bonneville Basin. Palaeogeogr. Palaeoclim. Palaeoecol. 141, 95e Geol. Surv. Open-File Rept, 01-202. 114. Forester, R.M., Lowenstein, T.K., Spencer, R.J., 2005. An ostracode based paleo- Bradbury, J.P., 1991. The late Cenozoic stratigraphy and paleolimnology of limnologic and paleohydrologic history of Death Valley: 200 to 0 ka. Geol. Soc. Tule Lake, Siskiyou Co. California. J. Paleolimnol. 6, 205e255. Amer. Bull. 117, 1379e1386. Bradbury, J.P., 1997. A diatom-based paleohydrologic record of climate change for Forman, S.L., Pierson, J., Lepper, K., 2000. Luminescence geochronology. In: the past 800 k.y. from Owens Lake, California. In: Smith, G.I., Bischoff, J.L. (Eds.), Noller, J.S., Sowers, J.M., Lettis, W.R. (Eds.), Quaternary Geochronology. AGU An 800,000-year Paleoclimatic Record from Core OL-92, Owens Lake, southeast Reference Shelf. American Geophysical Union, Washington, DC, pp. 157e176. California, Geol. Soc. Amer. Spec. Pap. 317, pp. 99e112. Forman, S.L., Pierson, J., Gomez, J., Brigham-Grette, J., Nowaczyk, N.R., Melles, M., Bradbury, J.P., Forester, R.M., Thompson, R.S., 1989. Late Quaternary paleolimnology 2007. Luminescence geochronology for sediments from Lake El’gygytgyn, of Walker Lake, Nevada. J. Paleolimnol. 1, 249e267. northeast Siberia, Russia: constraining the timing of paleoenvironmental events Briggs, R.W., Wesnousky, S.G., Adams, K.D., 2005. Late Pleistocene and late Holocene for the past 200 ka. J. Paleolimnol. 37, 77e88. lake highstands in the Pyramid Lake subbasin of Lake Lahontan, Nevada, USA. Gale, H.S., 1914. Salines in the Owens, Searles, and Panamint basins, southeastern Quat. Res. 64, 257e263. California. U.S. Geol. Surv. Bull. 580-L, 251e323. M.C. Reheis et al. / Quaternary Science Reviews 97 (2014) 33e57 55

Garcia, A.F., Stokes, M., 2006. Late Pleistocene highstand and recession of a small, Kurth, G., Phillips, F.M., Reheis, M.C., Redwine, J.L., Paces, J.B., 2011. Cosmogenic high-altitude pluvial lake, Jakes Valley, central Great Basin, USA. Quat. Res. 65, nuclide and uranium-series dating of old, high shorelines in the western Great 179e186. Basin, USA. Geol. Soc. Amer. Bull. 123, 744e768. Garcia, A.L., Knott, J.R., Mahan, S.A., Bright, J., 2014. Geochronology and paleo- Lajoie, K.R., 1968. Late Quaternary Stratigraphic and Geologic History of Mono Basin, environment of pluvial Harper Lake, Mojave Desert, California, USA. Quat. Res. Eastern California (Ph.D. dissertation). Univ. California at Berkeley. 81, 305e317. Li, H., Bischoff, J.L., Ku, T., Lund, S.P., Stott, L.D., 2000. Climate variability in east- Gilbert, C.M., Christensen, M.N., Al-Rawi, Y., Lajoie, K.R., 1968. Structural and volcanic central California during the past 1000 years reflected by high-resolution history of Mono Basin, California-Nevada. Geol. Soc. Amer. Mem. 16, 275e329. geochemical and isotopic records from Owens Lake sediments. Quat. Res. 54, Gilbert, G.K., 1885. The topographic features of lake shores. U.S. Geol. Surv. Ann. Rep. 187e197. 5, 75e123. Licciardi, J.M., 2001. Chronology of latest Pleistocene lake-level fluctuations in the Gilbert, G.K., 1890. Lake Bonneville. In: U.S. Geol. Surv. Monograph 1. pluvial Lake Chewaucan basin, Oregon, USA. J. Quat. Sci. 16, 545e553. Godsey, H.S., Oviatt, C.G., Miller, D.M., Chan, M.A., 2011. Stratigraphy and chronology Lillquist, K.D., 1994. Late Quaternary Lake Franklin: Lacustrine chronology, coastal of offshore to nearshore deposits associated with the Provo shoreline, Pleisto- geomorphology, and hydro-isostatic deflection in Ruby Valley and northern cene Lake Bonneville, Utah. Palaeogeogr. Palaeoclim. Palaeoecol. 310, 442e450. Butte Valley, Nevada (M.S. thesis). University of Utah, Salt Lake City. Gordon, I., Heller, P.L., 1993. Evaluating major controls on basinal stratigraphy, Pine Lin, J.C., Broecker, W.S., Hemming, S.R., Hajdas, I., Anderson, R.F., Smith, G.I., Valley, Nevada: implications for syntectonic deposition. Geol. Soc. Amer. Bull. Kelley, M., Bonani, G., 1998. A reassessment of U-Th and 14C ages for late-glacial 105, 47e55. high-frequency hydrological events at Searles Lake, California. Quat. Res. 49, Gosse, J.C., Phillips, F.M., 2001. Terrestrial in situ cosmogenic nuclides: theory and 11e23. application. Quat. Sci. Rev. 20, 1475e1560. Lisiecki, L.E., Raymo, M.E., 2005. A PlioceneePleistocene stack of 57 globally Grayson, D.K., 2011. The Great Basin, a Natural Prehistory. Univ. California Press, distributed benthic d18Orecords. Paleoceanography 20, 17. Berkeley. Long, H., Lai, Z., Fuchs, M., Zhang, J., Li, Y., 2012. Timing of late Quaternary palae- Hart, W.S., Quade, J., Madsen, D.b., Kaufman, D.S., Oviatt, C.G., 2004. The 87Sr/86Sr olake evolution in Tengger Desert of northern China and its possible forcing ratios of lacustrine carbonates and lake-level history of the Bonneville paleo- mechanisms. Glob. Planet. Change 92e93, 119e129. lake system. Geol. Soc. Amer. Bull. 116, 1107e1119. Lowenstein, T.K., Li, J., Brown, C., Roberts, S.M., Ku, T.-L., Luo, S., Yang, W., 1999. 200 Hershler, R., Liu, H.-P., 2008. Ancient vicariance and recent dispersal of springsnails k.y. paleoclimate record from Death Valley salt core. Geology 27, 3e6. (Hydrobiidae: Pyrgulopsis) in the Death Valley system, California-Nevada. In: Lubetkin, L.K.C., Clark, M.M., 1988. Late Quaternary activity along the Lone Pine Reheis, M.C., Hershler, R., Miller, D.M. (Eds.), Late Cenozoic Drainage History of fault, eastern California. Geol. Soc. Amer. Bull. 100, 755e766. the Southwestern Great Basin and Lower Region: Geologic and Lyle, M., Heusser, L., Ravelo, C., Yamamoto, M., Barron, J., Diffenbaugh, N.S., Herbert, T., biotic perspectives, Geol. Soc. Amer. Spec. Paper 439, pp. 91e101. Andreasen, D., 2012. Out of the tropics: the Pacific, Great Basin lakes, and late Hershler, R., Sada, D.W., 2002. Biogeography of Great Basin aquatic snails of the Pleistocene water cycle in the western United States. Science 337, 1629e1633. genus Pyrgulopsis. In: Hershler, R., Madsen, D.B., Currey, D.R. (Eds.), Great Basin Machette, M.N., Slate, J.L., Phillips, F.M., 2008. Terrestrial cosmogenic-nuclide dating Aquatic Systems History. Smithsonian Contributions to the Earth Sciences 33, of alluvial fans in Death Valley, California. U.S. Geol. Surv. Prof. Pap. 1755. pp. 255e276. Madsen, D.B., Oviatt, C.G., Young, D.C., Page, D., 2014. Chapter 3: Old River Bed delta Hildreth, W., 2004. Volcanological perspectives on Long Valley, Moun- geomorphology and chronology. In: Madsen, D.B., Schmitt, D.N., Page, D. (Eds.), tain, and Mono Craters; several contiguous but discrete systems. J. Volc. Geo- The Palaeoarchaic Occupation of the Old River Bed Delta. Univ. Utah Anthropol. therm. Res. 136, 169e198. Pap (in press). House, P.K., Pearthree, P.A., Perkins, M.E., 2008. Stratigraphic evidence for the role of Magee, J.W., Miller, G.H., Spooner, N.A., Questiaux, D., 2004. Continuous 150 k.y. lake spillover in the inception of the lower Colorado River in southern Nevada monsoon record from Lake Eyre, Australia: insolation-forcing implications and and western Arizona. In: Reheis, M.C., Hershler, R., Miller, D.M. (Eds.), Late unexpected Holocene failure. Geology 32, 885e888. Cenozoic Drainage History of the Southwestern Great Basin and Lower Colorado McDonald, E.V., McFadden, L.D., Wells, S.G., 1995. The relative influences of climate River Region: Geologic and biotic perspectives, Geol. Soc. Amer. Spec. Paper change, desert dust, and lithologic control on soil-geomorphic processes on 439, pp. 335e353. alluvial fans, Mojave Desert, California. Summ. Results. San. Bernardino Co. Mus. Hubbs, C.L., Miller, R.R., Hubbs, L.C., 1974. Hydrographic history and relict fishes of Assoc. Quart. 42, 35e42. the north-central Great Basin. Calif. Acad. Sci. Mem. 7. McFadden, L.D., Ritter, J.B., Wells, S.G., 1989. Use of multiparameter relative- Ibarra, D.E., Egger, A.E., Weaver, K.L., Harris, C.R., Maher, K., Rise and fall of late age methods for age estimation and correlation of alluvial-fan surfaces Pleistocene pluvial lakes in response to reduced evaporation and precipitation: on a desert piedmont, eastern Mojave Desert, California. Quat. Res. 32, evidence from Lake Surprise, California. Geol. Soc. Amer. Bull. (in press). 276e290. Jayko, A.S., Bacon, S.N., 2008. Late Quaternary MIS 6-8 shoreline features of pluvial McGee, D., Quade, J., Edwards, R.L., Broecker, W.S., Cheng, H., Reiners, P.W., Owens Lake, Owens Valley, eastern California. In: Reheis, M.C., Hershler, R., Evenson, N., 2012. Lacustrine cave carbonates: novel archives of paleohydro- Miller, D.M. (Eds.), Late Cenozoic drainage history of the southwestern Great logic change in the Bonneville Basin (Utah, USA). Earth Planet. Sci. Lett. 352e Basin and lower Colorado River Region: geologic and biotic perspectives, Geol. 352, 182e194. Soc. Amer. Spec. Paper 439, pp. 185e206. Meek, N., 1989. Geomorphic and hydrologic implications of the rapid incision of Jayko, A.S., Forester, R.M., Kaufman, D.S., Phillips, F.M., Yount, J.C., McGeehin, J., Afton Canyon, Mojave Desert, California. Geology 17, 7e10. Mahan, S.A., 2008. Late Pleistocene lakes and wetlands, , Inyo Meek, N., 1994. The stratigraphy and geomorphology of Coyote basin. In: County, California. In: Reheis, M.C., Hershler, R., Miller, D.M. (Eds.), Late Ceno- Reynolds, J. (Ed.), Calico, Coyote Basin & Lake Havasu Giants. San Bernardino Co. zoic Drainage History of the Southwestern Great Basin and Lower Colorado Mus. Q., pp. 5e13. River Region: Geologic and biotic perspectives, Geol. Soc. Amer. Spec. Paper Meek, N., 1999. New discoveries about the Late Wisconsinan history of the Mojave 439, pp. 151e184. River system. In: Reynolds, R.E., Reynolds, J. (Eds.), Tracks along the Mojave: a Jefferson, G.T., 2003. Stratigraphy and paleontology of the middle to late Pleistocene Field Guide from Cajon Pass to the Calico Mountains and Coyote Lake. San Manix Formation, and paleoenvironments of the central Mojave River, southern Bernardino Co. Mus. Q, pp. 113e117. California. In: Enzel, Y., Wells, S.G. (Eds.), Paleoenvironments and Paleohy- Menges, C.M., 2008. Multistage late Cenozoic evolution of the Amargosa River drology of the Mojave and Southern Great Basin Deserts, Geol. Soc. Amer. Spec. drainage, southwestern Nevada and eastern California. In: Reheis, M.C., Paper 368, pp. 43e60. Hershler, R., Miller, D.M. (Eds.), Late Cenozoic Drainage History of the South- King, G.Q., 1993. Late Quaternary history of the lower Walker River and its impli- western Great Basin and Lower Colorado River Region: Geologic and biotic cations for the Lahontan paleolake system. Phys. Geogr. 14, 81e96. perspectives, Geol. Soc. Amer. Spec. Paper 439, pp. 39e90. Kirby, M.E., Zimmerman, S.R.H., Patterson, W.P., Rivera, J.J., 2012. A 9170-year record Menking, K.M., Bischoff, J.L., Fitzpatrick, J.A., Burdette, J.W., Rye, R.O., 1997. Climatic/ of decadal-to-multi-centennial scale pluvial episodes from the coastal South- hydrologic oscillations since 155,000 yr B.P. at Owens Lake, California, reflected west United States: a role for atmospheric rivers? Quat. Sci. Rev. 46, 57e65. in abundance and stable isotope composition of sediment carbonate lake Knott, J.R., Machette, M.N., Klinger, R.E., Sarna-Wojcicki, A.M., Liddicoat, J.C., salinity. Quat. Res. 48, 58e68. Tinsley III, J.C., David, B.T., Ebbs, V.M., 2008. Reconstructing late Pliocene to Mensing, S.A., 2001. Late-glacial and early Holocene vegetation and climate change middle Pleistocene Death Valley lakes and river systems as a test of pupfish near Owens Lake, eastern California. Quat. Res. 55, 57e65. (Cyprinodontidae) dispersal hypotheses. In: Reheis, M.C., Hershler, R., Metcalfe, S., Say, A., Black, S., McCulloch, R., O’Hara, S., 2002. Wet conditions during Miller, D.M. (Eds.), Late Cenozoic Drainage History of the Southwestern Great the last glaciation in the Chihuahuan Desert, Alta Babicora basin, Mexico. Quat. Basin and Lower Colorado River Region: Geologic and biotic perspectives, Geol. Res. 57, 91e101. Soc. Amer. Spec. Paper 439, pp. 1e26. Mickelson, D.M., Colgan, P.M., 2003. The southern . In: Koehler, P.A., 1995. Late Pleistocene mollusks from the Dolomite Site, Owens Valley, Gillespie, A.R., Porter, S.C., Atwater, B.F. (Eds.), The Quaternary Period in the California. Veliger 38, 312e318. United States, Developments in Quaternary Science 1. Elsevier, Netherlands, Kong, P., Naa, C., Brown, R., Fabel, D., Freeman, S., Xiaoa, W., Wanga, Y., 2011. pp. 1e16. Cosmogenic 10Be and 26Al dating of paleolake shorelines in Tibet. J. Asian Earth Mifflin, M.D., Wheat, M.M., 1979. Pluvial lakes and estimated pluvial climates of Sci. 41, 263e273. Nevada. Nev. Bur. Mines Geol. Bull. 94. Kowler, A., 2010. Paleohydrologic and paleoclimatic implications of late Pleistocene Miller, D.M., Dudash, S.L., McGeehin, J.P., 2011a. Chronology of pluvial Lake Coyote, shoreline deposits in the southern Basin and Range Province. Geol. Soc. Amer. California, and implications for 25 to 10-ka Mojave River paleohydrology. Abst. Prog. 42 (5), 33. PACLIM 2009 Abstracts. Quat. Int. 235, 116. Krider, P.R., 1998. Paleoclimatic significance of late Quaternary lacustrine and al- Miller, D.M., Oviatt, C.G., Nash, B.P., 2008. Late Pleistocene Hansel Valley basaltic luvial stratigraphy, Animas Valley, New Mexico. Quat. Res. 50, 283e289. ash, northern Lake Bonneville, Utah, USA. Quat. Int. 178, 238e245. 56 M.C. Reheis et al. / Quaternary Science Reviews 97 (2014) 33e57

Miller, D.M., Dudash, S.L., McGeehin, J.P., 2009. Chronology of pluvial Lake Coyote, Oviatt, C.G., Currey, D.R., Miller, D.M., 1990. Age and paleoclimatic significance of California, and implications for 25 to 10-ka Mojave River paleohydrology. In: the Stansbury shoreline of Lake Bonneville, northeastern Great Basin. Quat. Res. PACLIM 2009 Abstracts. 33, 291e395. Miller, D.M., Clark, D.L., Wells, M.L., Oviatt, C.G., Felger, T.J., Todd, V.R., 2012. Progress Oviatt, C.G., Currey, D.R., Sack, D., 1992. Radiocarbon chronology of Lake Bonneville, report geologic map of the Grouse Creek 30’ 60’ quadrangle and Utah part of eastern Great Basin, USA. Palaeogeogr. Palaeoclim. Palaeoecol. 99, 225e241. the Jackpot 30’ 60’ quadrangle, Box Elder County, Utah, and Cassia County, Oviatt, C.G., Habiger, G., Hay, J., 1994. Variation in the composition of Lake Bonne- Idaho. Utah Geol. Surv. Open-File Rep. 598. ville marl: a potential key to lake-level fluctuations and paleoclimate. Miller, D.M., Oviatt, C.G., McGeehin, J.P., 2013. Stratigraphy and chronology of Provo J. Paleolimnol. 11, 19e30. shoreline deposits and lake-level implications, late Pleistocene Lake Bonneville, Oviatt, C.G., Thompson, R.S., Kaufman, D.S., Bright, J., Forester, R.M., 1999. Reinter- eastern Great Basin, USA. Boreas 42, 342e361. pretation of the Burmeister Core, Bonneville Basin. Utah. Quat. Res. 52, 180e184. Miller, D.M., Reheis, M.C., Wan, E., Wahl, D.B., Olson, H., 2011b. Pliocene and early Oviatt, C.G., Madsen, D.M., Schmitt, D.N., 2003. Late Pleistocene and early Holocene Pleistocene paleo-geography of the Coyote Lake and Alvord Mountain area, rivers and wetlands in the Bonneville basin of western North America. Quat. Mojave Desert, California. In: Reynolds, R.E. (Ed.), The Incredible Shrinking Res. 60, 200e210. Pliocene: 2011 Desert Symposium. Calif State Univ. Fullerton, pp. 53e67. http:// Oviatt, C.G., Miller, D.M., McGeehin, J.P., Zachary, C., Mahan, S., 2005. The Younger biology.fullerton.edu/dsc/pdf/2011shrinkingpliocene.pdf. Dryas phase of Great Salt Lake, Utah, USA. Palaeogeogr. Palaeoclim. Palaeoecol. Miller, R.R., Smith, G.R., 1981. Distribution and Evolution of Chasmistes (Pisces: 219, 263e284. Catostomidae) in Western North America. In: Occasional Papers of the Museum Owen, L.A., Bright, J., Finkel, R.C., Jaiswal, M.K., Kaufman, D.S., Mahan, S., Radtke, U., of Zoology, University of Michigan 696, pp. 1e46. Schneider, J.S., Sharp, W., Singhvi, A.K., Warren, C.N., 2007. Numerical dating of Milligan, M.R., Lemons, D.R., 1998. A sequence stratigraphic overview of sandy and a late Quaternary -shoreline complex at the northern end of gravelly lacustrine deltas deposited along the eastern margin of late Pleistocene playa, Mojave Desert, California: a comparison of the applicability of radio- Lake Bonneville, northern Utah and southern Idaho. In: Pittman, J.K., carbon, luminescence, terrestrial cosmogenic nuclide, electron spin resonance, Carroll, A.R. (Eds.), Modern and Ancient Lake Systems, Utah Geol. Assoc. U-series and amino acid racemization methods. Quat. Int. 166, 87e110. Guidebook, vol. 26. New Problems and Perspectives, Salt Lake City, pp. 105e129. Owen, L.A., Frankel, K.L., Knott, J.R., Reynhout, S., Finkel, R.C., Dolan, J.F., Lee, J., 2011. Morrison, R.B., 1964. Lake Lahontan: geology of southern Carson Desert, Nevada. Beryllium-10 terrestrial cosmogenic nuclide surface exposure dating of Qua- U.S. Geol. Surv. Prof. Pap. 401. ternary landforms in Death Valley. Geomorphology 125, 541e557. Morrison, R.B., 1965. Quaternary geology of the Great Basin. In: Wright Jr., H.E., Peterson, F.F., 1981. Landforms of the Basin & Range Province, defined for soil sur- Frey, D.C. (Eds.), The Quaternary of the United States. Princeton University Press, vey. Nev. Agric. Exp. Stn. Tech. Bull. 28. Princeton, New Jersey, pp. 265e285. Phillips, F.M., 2008. Geological and hydrological history of the paleo-Owens River Morrison, R.B., 1991. Quaternary stratigraphic, hydrologic, and climatic history of drainage since the late Miocene. In: Reheis, M.C., Hershler, R., Miller, D.M. (Eds.), the Great Basin, with emphasis on Lakes Lahontan, Bonneville, and Tecopa. In: Late Cenozoic Drainage History of the Southwestern Great Basin and Lower Morrison, R.B. (Ed.), Quaternary Nonglacial Geology: Conterminous U.S. The Colorado River Region: Geologic and biotic perspectives, Geol. Soc. Amer. Spec. Geology of North America, K-2. Geol. Soc. Amer., Boulder, Colorado, pp. 283e Paper 439, pp. 115e150. 320. Pierce, K.L., 2003. Pleistocene glaciations of the Rocky Mountains. In: Gillespie, A.R., Morrison, R.B., 1999. Lake Tecopa: Quaternary geology of Tecopa Valley, California, a Porter, S.C., Atwater, B.F. (Eds.), The Quaternary Period in the United States, multimillion-year record and its relevance to the proposed nuclear-waste re- Developments in Quaternary Science 1. Elsevier, Netherlands, pp. 63e76. pository at , Nevada. In: Wright, L.A., Troxel, B.W. (Eds.), Postma, G., 1990. Depositional architecture and facies of river and fan deltas: a Cenozoic Basins of the Death Valley Region, Geol. Soc. Amer. Spec. Paper 333, synthesis. In: Colella, A., Prior, D.B. (Eds.), Coarse-grained Deltas. Blackwell pp. 301e344. Scientific Publications, Oxford, pp. 13e27. Munroe, J.S., Laabs, B.J.C., 2013a. Latest pleistocene history of pluvial Lake Franklin, Redwine, J.L., 2003. The Quaternary Pluvial History and Paleoclimate Implications of northeastern Nevada, USA. Geol. Soc. Amer. Bull. 125, 322e342. Newark Valley, East-central Nevada; Derived from Mapping and Interpretation Munroe, J.S., Laabs, B.J.C., 2013b. Temporal correspondence between pluvial lake of Surficial Units and Geomorphic Features (M.S. thesis). Humboldt State Uni- highstands in the southwestern US and Heinrich Event 1. J. Quat. Sci. 28, 49e versity, Arcata, 358 p. 58. Reeves Jr., C.C., 1968. Introduction to Paleolimnology. Elsevier, Amsterdam. Negrini, R.M., 2002. Pluvial lake sizes in the northwestern Great Basin throughout Reheis, M.C., 1999. Highest pluvial-lake shorelines and Pleistocene climate of the the Quaternary period. In: Hershler, R., Madsen, D.B., Currey, D. (Eds.), Great western Great Basin. Quat. Res. 52, 196e205. Basin Aquatic Systems History: Smithsonian Contributions to the Earth Sci- Reheis, M.C., Redwine, J.L., 2008. Lake Manix shorelines and Afton Canyon terraces: ences, vol. 33. Smithsonian Institution Press, Washington, D.C, pp. 11e52. implications for incision of Afton Canyon. In: Reheis, M.C., Hershler, R., Negrini, R.M., Erbes, D.B., Faber, K., Herrera, A.M., Roberts, A.P., Cohen, A.S., Miller, D.M. (Eds.), Late Cenozoic drainage history of the southwestern Great Wigand, P.E., Foit, F.F.J., 2000. A paleoclimate record for the past 250,000 years Basin and lower Colorado River region: Geologic and biotic perspectives, Geol. from Summer Lake, Oregon, USA: I. Chronology and magnetic proxies for lake Soc. Amer. Spec. Paper 439, pp. 227e260. level. J. Paleolimnol. 24, 125e149. Reheis, M.C., Harden, J.W., McFadden, L.D., Shroba, R.R., 1989. Development rates of Negrini, R.M., Wigand, P.E., Draucker, S., Gobalet, K., Gardner, J.K., Sutton, M.Q., late Quaternary soils, Silver Lake playa, California. Soil. Sci. Soc. Amer. Jour 53, Yohe, R.M., 2006. The Rambla highstand shoreline and the Holocene lake-level 1127e1140. history of Tulare Lake, California, USA. Quat. Sci. Rev. 25, 1599e1618. Reheis, M.C., Sarna-Wojcicki, A.M., Reynolds, R.L., Repenning, C.A., Mifflin, M.D., Newton, M.S., 1991. Holocene Stratigraphy and Magnetostratigraphy of Owens and 2002a. Pliocene to Middle Pleistocene lakes in the Western Great Basin: ages Mono Lakes, Eastern California (Unpublished Ph.D. dissertation). Univ. So. Calif., and connections. In: Hershler, R., Madsen, D.B., Currey, D.R. (Eds.), Great Basin Los Angeles. Aquatic Systems History. Smithsonian Institution Press, Washington, D.C, Noller, J.S., Sowers, J.M., Lettis, W.R., 2000. Quaternary Geochronology: Methods pp. 53e108. and Applications. American Geophysical Union, Washinton, DC. Reheis, M.C., Stine, S., Sarna-Wojcicki, A.M., 2002b. Drainage reversals in Mono Orme, A.R., 2008. Pleistocene pluvial lakes of the American West: a short history of Basin during the late Pliocene and Pleistocene. Geol. Soc. Amer. Bull. 114, 991e research. In: Geological Society, London, Spec. Pub. 301, pp. 51e78. 1006. Orme, A.J., Orme, A.R., 2008. Late Pleistocene climate and tectonic impacts on Reheis, M., Redwine, J., Adams, K., Stine, S., Parker, K., Negrini, R., Burke, R., Kurth, G., Owens Lake levels and linkages. In: Reheis, M.C., Hershler, R., Miller, D.M. (Eds.), McGeehin, J., Paces, J., Phillips, F., Sarna-Wojcicki, A., Smoot, J., 2003. Pliocene to Late Cenozoic drainage history of the southwestern Great Basin and lower Holocene lakes in the western Great Basin: new perspectives on paleoclimate, Colorado River region: Geologic and Biotic Perspectives, Geol. Soc. Amer. Spec. landscape dynamics, tectonics, and paleodistribution of aquatic species. In: Paper 439, pp. 207e225. Easterbrook, D.J. (Ed.), Quaternary Geology of the United States (INQUA 2003 Oviatt, C.G., 1987. Lake Bonneville stratigraphy at the Old River Bed, Utah. Am. J. Sci. Field Guide Volume). Desert Research Institute, Reno, pp. 155e194. 287, 383e398. Reheis, M.C., Miller, D.M., Redwine, J.L., 2007. Quaternary stratigraphy, drainage- Oviatt, C.G., 1989. Quaternary geology of part of the Sevier Desert, Millard County, basin development, and geomorphology of the Lake Manix basin, Mojave Utah. Utah Geol. Min. Surv. Spec. Stud. 70. Desert. U.S. Geol. Surv. Open-File Report 2007e1281. Oviatt, C.G., 1997. Lake Bonneville fluctuations and global climate change. Geology Reheis, M.C., Miller, D.M., 2010. Environments of nearshore lacustrine deposition in 25, 155e158. the Pleistocene Lake Manix basin, south-central California. In: Reynolds, R.E., Oviatt, C.G., 2012. Sapropels in the Great Salt Lake basin: indicators of massive Miller, D.M. (Eds.), Overboard in the Mojave: 20 Million Years of Lakes and groundwater-discharge events. In: Am. Geophys. Union Ann. Meet. abst. PP31A- Wetlands: 2010 Desert Symposium. Calif State Univ, Fullerton, pp. 24e37. 2013. http://biology.fullerton.edu/dsc/pdf/2010Overboard.pdf. Oviatt, C.G., 2014. The Gilbert episode in the Great Salt Lake basin, UT. Utah Geol. Reheis, M.C., Miller, D.M., McGeehin, J.P., 2010. Late Pleistocene shoreline fluctua- Surv. Misc. Publ. 14-3. tions of Lake Manix, Mojave Desert: Paleoclimate implications. Geol. Soc. Amer. Oviatt, C.G., McCoy, W.D., 1992. Early Wisconsin lakes and in the Great Abst. Prog. 42 (4), 53. Basin, U.S.A. In: Clark, P.U., Lea, P.D. (Eds.), The Last Interglacialeglacial Tran- Reheis, M.C., Bright, J., Lund, S.P., Miller, D.M., Skipp, G., Fleck, R.J., 2012. A half- sition in North America, Geol. Soc. Amer. Spec. Paper 270, pp. 279e287. million-year record of paleoclimate from the Lake Manix core, Mojave Desert, Oviatt, C.G., Nash, W.P., 1989. Basaltic volcanic ash and volcanic eruptions in the California. Palaeogeogr., Palaeoclim. Palaeoecol. 365e366, 11e37. Bonneville basin, Utah. Geol. Soc. Am. Bull. 101, 292e303. Repenning, C.A., Weasma, T.R., Scott, G.R., 1995. The early Pleistocene (latest Oviatt, C.G., Nash, B.P., 2014. The Pony Express basaltic ash: a stratigraphic marker Blancan-earliest Irvingtonian) Froman Ferry and history of the Glenns in Lake Bonneville sediments, Utah. Utah Geol. Surv. Misc. Publ. 14-1, 10. Ferry Formation, southwestern Idaho. U.S. Geol. Surv. Bull. 2105. M.C. Reheis et al. / Quaternary Science Reviews 97 (2014) 33e57 57

Roberts, R.J., 1965. Stratigraphy and structure of the Antler peak quadrangle, Smoot, J.P., Litwin, R.J., Bischoff, J.L., Lund, S.P., 2000. Sedimentary record of the 1872 Humboldt and Lander Counties, Nevada. U.S. Geol. Surv. Prof. Pap. 459A, 1e93. earthquake and “tsunami” at Owens Lake, southeast California. J. Sed. Geol. 135, Rosario, J., Miller, D., Wahl, D., Anderson, L., Oviatt, J., Tripati, A., Presnetova, L., 241e254. McGeehin, J., 2013. Sedimentary evolution of Swan Lake and its paleoclimate Stewart, J.H., 1978. Basin and Range structure in western North America, a review. implications during deglacial and Holocene time, southeastern Idaho. In: In: Smith, R.B., Eaton, G.P. (Eds.), Cenozoic Tectonics and Regional Geophysics of PACLIM 2013 Abstracts. the Western Cordillera, Geol. Soc. Amer. Memoir 152, pp. 1e32. Rosenbaum, J.G., Reynolds, R.L., 2004. Record of late Pleistocene glaciation and Stewart, J.H., 1988. Tectonics of the Walker Lane belt, western Great Basin: deglaciation in the southern : II. Flux of glacial flour in a sedi- and Cenozoic deformation in a zone of shear. In: Ernst, W.G. (Ed.), Meta- ment core from Upper Klamath Lake, Oregon. J. Paleolimnol. 31, 235e252. morphism and Crustal Evolution of the Western United States. Prentice Hall, Rosenbaum, J.G., Kaufman, D.S., 2009. Paleoenvironments of Bear Lake, Utah and Englewood Cliffs, New Jersey, pp. 683e713. Idaho, and its catchment. Geol. Soc. Amer. Spec. Pap. 450. Stine, S., 1990. Late Holocene fluctuations of Mono Lake, eastern Calfornia. Palae- Roy, P.D., Caballero, M., Lozano, S., Morton, O., Lozano, R., Jonathan, M.P., ogeogr., Palaeoclim. Palaeoecol. 78, 333e381. Sánchez, J.L., Macías, M.C., 2012. Provenance of sediments deposited at paleo- Strahler, A.H., Strahler, A.N., 1992. Modern Physical Geography, fourth ed. John lake San Felipe, western Sonora Desert: implications to regimes of summer and Wiley and Sons, New York. winter precipitation during last 50 cal kyr BP. J. Arid Envir 81, 47e58. Tackman, G., 1993. Middle and Late Pleistocene Hydrologic History of Diamond Russell, I.C., 1885. Geological history of Lake Lahontan, a Quaternary lake of Valley, Eureka and Elka Counties, Nevada, with Climatic and Isostatic Implica- northwestern Nevada. U.S. Geol. Surv. Monogr. 11, 288. tions (M.S. thesis). University of Utah, Salt Lake City, 192 pp. Sarna-Wojcicki, A.M., Davis, J.O., 1991. Quaternary tephrochronology. In: Taylor, R.E., Long, A., Kra, R., 1992. Radiocarbon after four decades: an interdisci- Morrison, R.B. (Ed.), Quaternary Nonglacial Geology: Conterminous U.S. The plinary perspective. Springer-Verlag, New York. Geology of North America, K-2. Geol. Soc. Amer., Boulder, Colorado, pp. 93e116. Thompson, R.S., Benson, L., Hattori, E.M., 1986. A revised chronology for the last Sarna-Wojcicki, A.M., Lajoie, K.R., Meyer, C.E., Adam, D.P., Rieck, H.J., 1991. Teph- Pleistocene lake cycle in the central Lahontan Basin. Quat. Res. 25, 1e9. rochronologic correlation of upper Neogene sediments along the Pacific margin, Thompson, R.S., Toolin, L.J., Forester, R.M., Spencer, R.J., 1990. Accelerator-mass conterminous United States. In: Morrison, R.B. (Ed.), Quaternary Nonglacial spectrometer (AMS) radiocarbon dating of Pleistocene lake sediments in the Geology, Conterminous U.S. The Geology of North America, K-2. Geol. Soc. Great Basin. Palaeogeogr., Palaeoclim. Palaeoecol 78, 301e313. Amer., Boulder, Colorado, pp. 117e140. Trull, T.W., Brown, E.T., Marty, B., Raisbeck, G.M., Yiou, F., 1995. Cosmogenic 10Be and Scott, W.E., McCoy, W.D., Shroba, R.R., Rubin, M., 1983. Reinterpretation of the 3He accumulation in Pleistocene beach terraces in Death Valley, California, exposed record of the last two cycles of Lake Bonneville, western United States. U.S.A.: implications for cosmic-ray exposure dating of young surfaces in hot Quat. Res. 20, 261e285. climates. Chem. Geol. 119, 191e207. Smith, C.R., Bronson, L.R., 2003. Distribution, preferences and seasonality of Trumbore, S.E., 2000. Radiocarbon geochronology. In: Noller, J.S., Sowers, J.M., tiger beetles, Genus Cicindela (Coleoptera: Cicindelidae), in Surprise Valley, Lettis, W.R. (Eds.), Quaternary Geochronology, Methods and Applications. Modoc County, California. Cicindela 35, 1e22. American Geophysical Union Reference Shelf 4, Washington, DC. Smith, G.I., 1979. Subsurface stratigraphy and geochemistry of Late Quaternary Wallace, A.R., 2005. Sedimentation, faulting, and erosion in the Carlin basin, , Searles Lake, California. U.S. Geol. Surv. Prof. Pap. 1043. northeastern Nevada, and implications for exploration and ground Smith, G.I., 2009. Late Cenozoic geology and lacustrine history of Searles Valley, water resources. In: Rhoden, H.N., Steininger, R.C., Vikre, P.G. (Eds.), Geological Inyo and San Bernardino Counties, California. U.S. Geol. Surv. Prof. Pap. 1727. Society of Nevada Symposium 2005: Window to the World. Geol. Soc. Nevada, Smith, G.I., Pratt, W.P., 1957. Core logs from Owens, China, Searles, and Panamint Reno, pp. 147e159. basins, California. U.S. Geol. Surv. Bull. 1045-A. Wallace, A.R., Perkins, M.E., Fleck, R.J., 2008. Late Cenozoic paleogeographic evo- Smith, G.I., Street-Perrott, F.A., 1983. Pluvial lakes of the western United States. In: lution of northeastern Nevada: evidence from the sedimentary basins. Geo- Porter, S.C. (Ed.), The Late Pleistocene. Univ. Minn. Press, Minneapolis, pp. 190e sphere 4, 36e74. 212. Waters, M.R., 1989. Late Quaternary lacustrine history and paleoclimatic signifi- Smith, G.I., Bischoff, J.L., 1997. Core OL-92 from Owens Lake: project rationale, cance of pluvial Lake Cochise, southeastern Arizona. Quat. Res. 21, 1e11. geologic setting, drilling procedures, and summary. In: Smith, G.I., Bischoff, J.L. Wells, S.G., Brown, W.J., Enzel, Y., Anderson, R.Y., McFadden, L.D., 2003. Late Qua- (Eds.), An 800,000-year Paleoclimatic Record from Core OL-92, Owens Lake, ternary geology and paleohydrology of pluvial Lake Mojave, southern Califor- southeast California, Geol. Soc. Amer. Spec. Pap. 317, pp. 1e8. nia. In: Enzel, Y., Wells, S.G., Lancaster, N. (Eds.), Paleoenvironments and Smith, G.I., Barczak, V.J., Moulton, G., Liddicoat, J.C., 1983. Core KM-3, a surface to Paleohydrology of the Mojave and Southern Great Basin Deserts, Geol Soc. bedrock record of late Cenozoic sedimentation in Searles Valley, California. U.S. Amer. Spec. Paper 368, pp. 79e114. Geol. Surv. Prof. Pap. 1256. Wilkins, D.E., Currey, D.R., 1997. Timing and extent of late Quaternary paleolakes in Smith, G.I., Bischoff, J.L., Bradbury, J.P., 1997. Synthesis of the paleoclimatic record the Trans-Pecos closed basin, west Texas and south-central New Mexico. Quat. from Owens Lake Core OL-92. In: Smith, G.I., Bischoff, J.L. (Eds.), An 800,000- Res. 47, 306e315. year Paleoclimatic Record from Core OL-92, Owens Lake, southeast California, Zhang, J., Jia, Y., Lai, Z., Long, H., Yang, L., 2011. Holocene evolution of Huangqihai Geol. Soc. Amer. Spec. Pap. 317, pp. 143e160. Lake in semi-arid northern China based on sedimentology and luminescence Smith, G.R., Dowling, T.E., Gobalet, K.W., Lugaski, T., Shiozawa, D.K., Evans, R.P., dating. Holocene 21, 1261e1268. 2002. Biogeography and timing of evolutionary events among Great Basin Zimmerman, S.R.H., Hemming, S.R., Kent, D.V., Searle, S.Y., 2006. Revised chronol- fishes. In: Hershler, R., Madsen, D.B., Currey, D.R. (Eds.), Great Basin Aquatic ogy for late Pleistocene Mono Lake sediments based on paleointensity corre- Systems History, Smithsonian Contributions to the Earth Sciences 33. Smith- lation to the global reference curve. Earth Planet. Sci. Lett. 252, 94e107. sonian Institution Press, Washington, DC, pp. 175e234. Zimmerman, S.R.H., Hemming, S.R., Hemming, N.G., Tomascak, P.B., Pearl, C., 2011. Smith, J.F.J., Ketner, K.B., 1976. Stratigraphy of post-Paleozoic rocks and summary of High-resolution chemostratigraphic record of late Pleistocene lake-level vari- resources in the CarlinePinon Range area, Nevada. U.S. Geol. Surv. Prof. Pap. 867-B. ability, Mono Lake, California. Geol. Soc. Amer. Bull. 123, 2320e2334.

View publication stats