Geological Society of America Bulletin, published online on 2 June 2014 as doi:10.1130/B31014.1

Rise and fall of late pluvial lakes in response to reduced evaporation and precipitation: Evidence from Lake Surprise,

Daniel E. Ibarra1,†, Anne E. Egger2, Karrie L. Weaver1, Caroline R. Harris1, and Kate Maher1 1Department of Geological and Environmental Sciences, Stanford University, Stanford, California 94305, USA 2Department of Geological Science, Central Washington University, Ellensburg, Washington 98926, USA

ABSTRACT ulations corroborate these fi ndings, simulat- using a new lacustrine paleoclimate record from ing an average precipitation increase of only Surprise Valley, California, we reconcile how Widespread late Pleistocene lake sys- 6.5% relative to modern, accompanied by a different factors, namely, the competing infl u- tems of the Basin and Range Province in- 28% decrease in total evaporation driven by ences of solar insolation and increased precipi- dicate substantially greater moisture avail- a 7 °C decrease in mean annual temperature. tation, infl uenced Pleistocene lakes in the Basin ability during glacial periods relative to LGM PMIP3 climate model simulations also and Range during the last deglaciation. Surprise modern times, but the climatic factors that suggest a seasonal decoupling of runoff and Valley was chosen because it is located in an drive changes in lake levels are poorly con- precipitation, with peak runoff shifting to important climatic transition between the more strained. To better constrain these climatic the late spring–early summer from the late arid Basin and Range Province and the wetter factors, we present a new lacustrine paleo- winter–early spring. Pacifi c Northwest. climate record and precipitation estimates Our coupled analyses suggest that moder- The geographic extent and temporal trends for Lake Surprise, a closed basin lake in ate lake levels during the LGM were a result of the latest Pleistocene lake levels suggest that northeastern California. We combine a de- of reduced evaporation driven by reduced orbital conditions and changes in atmospheric tailed analysis of lake hydrography and summer insolation and temperatures, not by circulation imposed wetter and/or cooler condi- constitutive relationships describing the increased precipitation. Reduced evapora- tions on the western United States. In the Basin water balance to determine the infl uence tion primed Basin and Range lake systems, and Range, the majority of lake highstands from of precipitation, evaporation, temperature, particularly smaller, isolated basins such as 31°N to 43°N occurred between 15 and 18 ka, and seasonal insolation on past lake levels. Surprise Valley, to respond rapidly to in- during Heinrich Stadial 1 (HS1, ca. 19–14.5 ka), At its maximum extent, during the last de- creased precipitation during late-Heinrich several thousand years after the Last Glacial glaciation, Lake Surprise covered 1366 km2 Stadial 1 (HS1). Post-LGM highstands were Maximum (LGM, ca. 26–19 ka; Benson et al., (36%) of the terminally draining Surprise potentially driven by increased rainfall dur- 1990; Adams and Wesnousky, 1998; García and Valley watershed. Using paired radiocarbon ing HS1 brought by latitudinally extensive Stokes, 2006; Munroe and Laabs, 2012; Lyle and 230Th-U analyses, we dated shoreline and strengthened midlatitude westerly storm et al., 2012). Prior to HS1, these lakes appear tufa deposits from wave-cut lake terraces in tracks, the effects of which are recorded in to have stood at moderate water levels during Surprise Valley, California, to determine the the region’s lacustrine and glacial records. much of the early marine oxygen isotope stage 2 hydrography of the most recent lake cycle. These results suggest that seasonal insolation (MIS 2, ca. 29–11 ka; e.g., Benson et al., 1995; This new lake hydrograph places the highest and reduced temperatures have been under- Wells et al., 2003; Bacon et al., 2006), and low lake level 176 m above the present-day playa investigated as long-term drivers of moisture to moderate levels through MIS 3 (ca. 57–29 at 15.19 ± 0.18 calibrated ka (14C age). This availability in the western United States. ka; e.g., Tackman, 1993; Phillips et al., 1994; signifi cantly postdates the Last Glacial Maxi- Reheis et al., 2012). The atmospheric mecha- mum (LGM), when Lake Surprise stood at INTRODUCTION nism driving these high lake levels is hypoth- only moderate levels, 65–99 m above modern esized to be midlatitude “dipping westerlies” playa, similar to nearby . The late Pleistocene landscape of the west- (Negrini, 2002), which reached as far south as To evaluate the climatic factors associated ern United States was characterized by vast Lake Elsinore, California (~34°N; Kirby et al., with lake-level changes, we use an oxygen iso- lake systems indicative of a hydrologic balance 2013) and Cave of the Bells, Arizona (~32°N; tope mass balance model combined with an dramatically different from the present (Fig. 1; Wagner et al., 2010). Despite evidence from late analysis of predictions from the Paleoclimate Mifflin and Wheat, 1979; Reheis, 1999a). Pleistocene lake records and other paleoclimate Model Intercomparison Project 3 (PMIP3) However, uncertainty in the timing of major archives, the temporal correspondence between climate model ensemble. Our isotope mass hydrologic changes has made it diffi cult to (Munroe and Laabs, 2012) or robust latitudinal balance model predicts minimal precipitation connect the observational record to well-dated trends in (Lyle et al., 2012) lake highstands and increases of only 2%–18% during the LGM climatic events. In addition, the precise con- stillstands during HS1 remain enigmatic. Fur- relative to modern, compared to an ~75% nection between lake levels and climate factors thermore, the mechanisms (e.g., reduced tem- increase in precipitation during the 15.19 ka has proven challenging to establish, because the peratures and lake surface evaporation, and/or highstand. LGM PMIP3 climate model sim- relationships among the physical and hydro- increased rainfall) that produced moderate lake logic controls on measured variables and past levels during the LGM, before the deglacial †E-mail: [email protected]. climatic states are unresolved. In this study, highstands, are not well understood. Knowledge

GSA Bulletin; Month/Month 2014; v. 1xx; no. X/X; p. 1–29; doi: 10.1130/B31014.1; 10 fi gures; 8 tables; Data Repository item 2014221.

For permission to copy, contact [email protected] 1 © 2014 Geological Society of America Geological Society of America Bulletin, published online on 2 June 2014 as doi:10.1130/B31014.1

Ibarra et al.

125°0'W 122°30'W 120°0'W 117°30'W 115°0'W the Pacifi c Northwest and the Basin and Range. Climate models of the LGM predict a more arid Wallowa

5°0'N Pacifi c Northwest and a relatively wet central 4 Nevada (Kim et al., 2008; Laîné et al., 2009; LI Braconnot et al., 2012). Paleoclimate records

FR MH from the north and west suggest reduced pre- cipitation and a mean annual temperature ~7 °C

N lower during the LGM (Worona and Whitlock, CB °30' WR 1995; Bradbury et al., 2004). To the southeast, 2 AL 4 UKL Lake Lahontan stood at moderate levels during the LGM (Benson et al., 1995), prior to a brief SV CL WL deglacial highstand at ca. 15.8 ka (Adams and LF Wesnousky, 1998). A detailed analysis of the Ruby Mtns LB Glacial Records DV transition zone between these regions can help 'N

°0 Pollen Record resolve the forcing mechanisms that could pro-

40 LL Late Pleistocene Lakes duce such disparate conditions. SL We use a “shore-based” approach (e.g., Red- Modern Lakes NV JL wine, 2003; García and Stokes, 2006; Kurth Extent of Major Glaciers RR LR CO CR et al., 2011; Munroe and Laabs, 2013) to quan- tify late Pleistocene lake levels in Surprise Val- 0 75 150 300 km 0'N ley, recorded in prominent wave-cut shorelines Tioga

37°3 along the steep valley walls (Egger and Miller, 2011; Irwin and Zimbelman, 2012). While Figure 1. Location map of late Pleistocene lakes (light blue), modern lakes (dark blue), lake sediment core studies can provide higher- and the extent of major Last Glacial Maximum (LGM) mountainous glaciation (gray) in resolution climate archives (e.g., Benson et al., the western United States (simplifi ed from Miffl in and Wheat, 1979; Reheis, 1999a; Ehlers 1990; Bischoff et al., 1997a, 1997b; Licciardi, et al., 2011). The black box delineates the extent of Figure 2A. Locations of additional paleo- 2001; Rosenbaum et al., 2012), records of climate archives compiled in Figure 10 and discussed within the text include glacier records shoreline ages document the history of lake (red triangles) and pollen records (red circle). Labeled lakes and pollen records are Alvord surface area, a measure of the balance of pre- Basin (AL), Lake (CR), Chewaucan Basin (CB), Columbus Lake (CO), Diamond cipitation and evaporation for a given basin Valley (DV), Fort Rock (FR), Jakes Lake (JL), (LB), Lake Clover (CL), (Miffl in and Wheat, 1979; Benson and Paillet, Lake (LF), Lake Lahontan (LL), Lake Russell (LR), Lake Surprise (SV), Little 1989; Currey, 1990; Reheis, 1999b; Munroe Lake (LI), Mahleur Lake (MH), Newark Valley (NV), Railroad Lake (RR), Spring Lake and Laabs, 2013). We incorporated stable iso- (SL), Upper Klamath Lake (UKL), Waring Lake (WL), and Lake Warner (WR). tope analysis into a simplifi ed hydrologic mass balance model based on lake surface area, which allows us to quantify the changes in of the spatial distribution of hydrologic shifts, ing insight into watershed-scale moisture avail- precipitation and evaporation during lake-level recorded by paleoclimate archives, is required ability driven by the climate system (Hostetler fl uctuations, and provides direct comparison to to resolve the underlying climatic drivers of pro- and Benson, 1994; Jones et al., 2007; Placzek climate model outputs. To obtain age-elevation found hydrologic change in the late Pleistocene. et al., 2011). constraints on lake level, we dated carbon- Most studies infer that Pleistocene lake levels While most research has focused on the large ate (tufa) deposits on wave-cut bedrock using record precipitation amounts driven by changes and complex lake systems of Lake Lahontan 230Th-U and radiocarbon (14C) geochronology. in midlatitude atmospheric circulation. Large and Lake Bonneville (Fig. 1; Benson et al., We used these new data to test climate model ranges of estimates for the extent of increased 1990; Oviatt et al., 1992; Adams and Wes- predictions for the LGM in a region where precipitation (80%–260% of modern), reduced nousky, 1998; Godsey et al., 2011; McGee et al., models diverge in their prediction of past rain- evaporation (12%–90% of modern), and 2012), small lakes—which record hydrologic fall magnitude and net change. decreased temperature (3–15 °C lower than conditions of smaller watersheds—are more Our new lake record and isotope mass balance modern) during the LGM have been calcu- responsive to climate fl uctuations and thus pro- calculations provide evidence that addresses lated using proxy data with modern analogs, vide higher-resolution paleoclimatic data (Gar- three key questions regarding the latest Pleisto- atmosphere-ocean global climate models cía and Stokes, 2006; Munroe and Laabs, 2013; cene lakes in the Basin and Range: (1) Can mod- (AOGCMs), and hydrologic, thermal evapora- Steinman et al., 2013). Thus, lakes with small erate LGM lake levels be explained by lower tion, and mass balance models (cf. Matsubara and watersheds, simple hypsometry, and short resi- evaporation rates due to decreasing summer Howard, 2009, their table 1). These broad ranges dence times should record perturbations in the solar insolation? (2) Is increased precipitation, are problematic for AOGCMs, which require hydrologic cycle more rapidly and with shorter driven by changes in atmospheric circulation, spatially resolved changes in the hydrologic lag times (Hendriks et al., 2012). required to explain post-LGM highstands dur- cycle for data-model intercomparison (e.g., One such small Pleistocene lake occupied ing late HS1? (3) Is there a temporal correspon- Braconnot et al., 2012; DiNezio and Tierney, Surprise Valley in northeastern California dence or robust latitudinal trend in lake high- 2013; Hargreaves et al., 2013). Stable isotope (Fig. 1). Surprise Valley is an ideal location for stands and stillstands during HS1 that records analysis of shoreline deposits can further con- testing paleoclimate models because it lies in changes in the position, strength, and/or charac- strain well-dated lake hydrographs by provid- the transition between two major climatic zones: ter of midlatitude atmospheric circulation?

2 Geological Society of America Bulletin, Month/Month 2014 Geological Society of America Bulletin, published online on 2 June 2014 as doi:10.1130/B31014.1

Lacustrine paleoclimate record from late Pleistocene Lake Surprise, California

BACKGROUND Basin through Buffalo Meadows during much one site, a radiocarbon age from a bone on the older and higher highstands similar to other second deepest excavated house fl oor was dated Geologic Setting and Previous Work lake systems (including Lahon- at 5640 ± 155 yr B.P. (radiocarbon age), the old- tan), likely recording progressive drying from est age obtained from all three sites (O’Connell Surprise Valley, located in northeastern Cali- the early middle to late Pleistocene (Reheis, and Inoway, 1994). fornia along the western margin of the Basin 1999a, 1999b; Reheis et al., 2002; Kurth et al., and Range Province (Fig. 1), is situated within 2011). However, fi eld and modeling studies Geochronology of Lake Shorelines a modern climatic transition from semiarid investigating surface areas, hydrog- coniferous forest in the west to arid sagebrush- raphy, and watershed runoff have assumed that Lake hydrograph assembly provides a fi rst- and grassland-dominated high-altitude desert the younger late Pleistocene Lake Surprise was order constraint on the climatic changes required in the east. The N-S–trending valley formed as inward draining, not connected to the Lahon- to sustain terminally draining Lake Surprise a result of extension since the mid-Miocene, tan system to the south or to Warner Valley to during the late Pleistocene. We targeted wave- and it is bounded by normal faults along the the northeast (Fig. 1; I.C. Russell, 1884; Hubbs cut lake terraces occurring in Surprise Valley range fronts of the Warner Range to west and and Miller, 1948; Miffl in and Wheat, 1979; using 230Th-U and radiocarbon geochronol- the smaller Hays Canyon Range to the east Benson and Paillet, 1989; Benson et al., 1995; ogy of tufas to provide absolute ages of sur- (Egger and Miller, 2011; Egger et al., 2014). Adams and Wesnousky, 1998; Sack, 2002; faces. Given recent advances in the precision Playas in the upper, middle, and lower sub- Matsu bara and Howard, 2009). of 230Th-U dating by multicollector–inductively basins occupy the valley fl oor (Fig. 2A). To the Late Pleistocene hydrographs have been coupled plasma–mass spectrometry (MC-ICP- south lies Duck Flat, a higher-elevation small produced for Great Basin lake systems includ- MS; Hernández-Mendiola et al., 2011; Shen et subbasin hydrologically connected to the Sur- ing Lahontan (e.g., Benson et al., 1995; Adams al., 2012; Cheng et al., 2013), and the establish- prise Valley watershed. et al., 2008), Franklin (Lillquist, 1994; Munroe ment of robust radiocarbon calibration curves Nearby Lake Lahontan and the Chewaucan and Laabs, 2011, 2013), Searles (Smith, 1984), (Reimer et al., 2013), these two methods pro- Basin (Fig. 1) have been the subject of detailed Bonneville (e.g., Oviatt et al. 1992), Russell vide complementary constraints on the age of study, but few studies have focused on Pleisto- (e.g., Zimmerman et al., 2011), and Chewaucan Pleistocene lacustrine carbonate. cene Lake Surprise, despite the recognition of (Licciardi, 2001) (Fig. 1). To date, no similar laterally continuous paleoshorelines throughout hydrographs exist for Lake Surprise, although 230Th-U Dating of Impure Lacustrine the valley and into Duck Flat (I.C. Russell, 1884; Personius et al. (2009) summarized constraints Carbonates C.J. Russell, 1927; Hubbs and Miller, 1948). on Pleistocene hydrography based on the distri- The 230Th-U dating method has been success- Erosional shoreline sequences in Surprise bution of lacustrine sediments deposited along fully applied to a range of terrestrial carbonate Valley are similar to those found in the Lahon- the western margin of the basin (Fig. 2A). materials, most notably to (e.g., tan and Bonne ville basins (e.g., Adams and Existing time-depth constraints for the latest Vacco et al., 2005; Oster et al., 2009; Wagner Wesnousky, 1998; Schofi eld et al., 2004; Felton Pleistocene lake cycle at Lake Surprise are con- et al., 2010; Asmerom et al., 2010; Polyak et et al., 2006; Jewell, 2007). On the east side of strained by the Trego Hot Spring tephra, with an al., 2012), pedogenic calcite/opal (e.g., Lud- Surprise Valley, wave-cut shoreline features are assigned radiocarbon age of 23.2 ± 0.3 calibrated wig and Paces, 2002; Sharp et al., 2003; Maher eroded into bedrock that consists of mid- to late ka from nearby Pyramid Lake (Benson et al., et al., 2007, 2014; Fletcher et al., 2010), trav- Cenozoic rhyolite, basalt, and tuff ( 1997), and archaeological remains ertine (e.g., Luo and Ku, 1991; Soligo et al., et al., 2006; Egger and Miller, 2011). Laminated (details in Table DR11). The Trego Hot Spring 2002), and a variety of lacustrine carbonates shoreline tufa on exposed bedrock is abundant tephra is exposed in two locations in the val- (e.g., Israelson et al., 1997; Haase-Schramm on several shorelines at elevations of 1420– ley at ~1378 m (Fig. 2A; Personius et al., 2009; et al., 2004; Placzek et al., 2006a, 2006b; Blard 1450 m (all elevations given as meters above Hedel, 1980, 1984). Hedel (1980) noted that the et al., 2011; McGee et al., 2012; Torfstein et al., sea level [asl]), but it is less common at higher tephra at both localities is deposited in fi ne sand 2013). Lacustrine carbonate associated with tufa and lower elevations (Fig. 2). Personius et al. and , likely associated with moderate to deep mounds and shoreline tufa deposits is common (2009) noted that the latest Pleistocene high- lake levels. Additionally, shallow-water deltaic in late Pleistocene lake systems in the western stand appears to be found throughout the valley deposits at 1475 and 1493 m postdate the Trego United States (e.g., Benson et al., 1995; Ku at ~1540 m. Additionally, two late Pleistocene Hot Spring deposit and likely represent deposi- et al., 1998; Felton et al., 2006; Godsey et al., highstand elevations have been proposed: tion between ca. 24 and 18 ka (Personius et al., 2011; Zimmerman et al., 2011) and elsewhere Miffl in and Wheat (1979) and Reheis (1999a) 2009). Personius et al. (2009) also reported (e.g., Moeyersons et al., 2006; Placzek et al., both reported highstand elevations of 1567 m, subaerial sediment deposition at 1475 m from 2006a, 2006b, 2011; Blard et al., 2011), making whereas Irwin and Zimbelman (2012) reported 13 to 1 ka (Fig. 2A). Archaeological evidence it useful for paleo–lake-level reconstructions. a highstand elevation of 1545 m. from three sites along the edges of the Holocene The 230Th-U dating method relies on the Analysis of modern topography by Irwin and playas (with a minimum elevation of 1355 m; fi rst three long-lived daughter products in the 238 Zimbelman (2012) suggests that the basin pour Fig. 2A) indicates that, since the middle Holo- U decay series. Uranium-238 (t1/2 = 4.46 point is 1621 m, i.e., signifi cantly higher than cene, the Modoc and Northern Paiute occu- b.y.; Jaffey et al., 1971) decays via the follow- any of the reported highstand elevations, indi- pied Surprise Valley as a seasonal winter home ing sequence: 238U → 234Th → 234Pa → 234U → 230 234 234 cating that Lake Surprise was a terminal lake. (James, 1983; O’Connell and Inoway , 1994). At Th. The Th (t1/2 = 24.1 d) and Pa (t1/2 = In support of that, Clawson et al. (1986) found 1.18 min) daughter isotopes are very short-lived that fl ows toward the center of the 1GSA Data Repository item 2014221, analytical compared to 238U, 234U, and 230Th. Samples can methods, discussion of the runoff coeffi cient assump- basin. C.J. Russell (1927) noted geomorphic tions, supporting fi gures, and tables, is available at thus be dated using the activity ratios (denoted evidence that suggests canyon-carving over- http://www.geosociety.org/pubs/ft2014.htm or by re- by parentheses) of the two long-lived daughter fl ow from Surprise Valley into the Lahontan quest to [email protected]. isotopes (234U/238U) and (230Th/238U). Due to the

Geological Society of America Bulletin, Month/Month 2014 3 Geological Society of America Bulletin, published online on 2 June 2014 as doi:10.1130/B31014.1

Ibarra et al.

120°20'W 120°0'W 119°40'W 120°4'W 120°3'W 120°4'W

Qc Water samples Qal Alluvium A B QalC Playa samples Qe Qpl Pleistocene lake sediments Weather Stations Qe Eolian deposits

Trego Hot Spring Qc Colluvium Tephra Locality Tufa Samples Archaeological Sites Upper Lake 41°36'N Paleoseismic Trench N 3'

w/ Lake Sediments 4 Qpl

C 1567m contour (earlier Qpl 41° Pleistocene HS)

41°40'N 1530m contour (deglaciation HS) B 1420 and 1440m contours (~LGM)

Middle Lake Watershed Qpl

Hays Canyon Range Qal 00.51Copyright:©0.25 2013 National km 00.51Cop0.25 right:© 2013 National km Geo ra hic Societ , i-cubed Geo ra hic ciet i-ube D N ' DE 1°20 4 Warner Mountains Warner Lower Lake

Qc E 41°26'N

Qc Qal Qal N ' Duck Flat °13 N 41 0' 1° 4

Buffalo Meadows

010205km 00.51Copyright:©0.25 2013 National km 00.510.25 km Geo ra hic Societ , i-cubed 41°25'N 119°59'W 119°58'W 119°58'W Figure 2. Location maps for samples collected as part of this study. Sample locations (latitude, longitude, and elevation) are listed in Table 1. (A) Detailed map of the Surprise Valley region including the calculated Lake Surprise watershed (black line) and outlines of Lake Sur- prise’s extent (green, red, and dark blue lines) at key periods, with water sample (blue circles), playa sample (red squares), and weather sta- tion (black stars) locations. Lake Surprise outlines were generated by contouring the merged digital elevation model (DEM) from the U.S. Geological Survey (USGS) National Elevation Data set using ArcGIS 10.1. For the Last Glacial Maximum (LGM) lake level (1420–1440 m), only 1440 m is shown on this panel. Only streams within the basin’s watershed are included, demonstrating that Lake Surprise was an inward-draining, closed basin lake system. Also shown are the locations of additional lake-level constraints from archaeological sites (yellow stars) (James, 1983; O’Connell and Inoway, 1994), Trego Hot Spring tephra localities (green diamonds; Hedel, 1980), and a paleoseismic trench with lacustrine sediments (gray square; Personius et al., 2009). Black boxes B, C, D, and E delineate the areas of additional panels. (B–E) Locations where carbonate samples were collected on exposed shoreline ridges (green circles). Group 1 tufa samples were collected from the prominent lower-elevation shorelines at 1419–1478 m (white circles); group 2 samples were collected from the less-prominent middle-elevation shorelines of 1509–1531 m (black circles); and group 3 samples were collected from higher-elevation shorelines >1542 m (black diamonds). Each sample locality includes the USGS 7.5′ quadrangles (1:24,000 scale), overlain with shaded relief and units adapted from Egger and Miller (2011) and Egger et al. (2014). HS—highstand.

4 Geological Society of America Bulletin, Month/Month 2014 Geological Society of America Bulletin, published online on 2 June 2014 as doi:10.1130/B31014.1

Lacustrine paleoclimate record from late Pleistocene Lake Surprise, California

TABLE 1. LOCATION OF SAMPLES Sample name Material type Sample Latitude Longitude Altitude* Type of analysis group no. (°N) (°W) (m) Accommodation zone shoreline (Fig. 2B) SVDI11-T2 Thick laminated tufa with porous 1 41.593662 120.070257 1453.5 U-Th (5 point isochrons [inner rind] + 5 subsamples [outer rind]), 14C, outer rind Sr/Ca (2 subsamples) and δ18O-δ13C (2 subsamples) SVDI11-T3 Thick laminated tufa 1 41.593293 120.070956 1437.7 U-Th (7 point isochron), 14C, Sr/Ca (2 subsamples) and δ18O-δ13C (2 subsamples) SVDI11-T4 Thin laminated tufa 1 41.592938 120.070912 1430.6 U-Th (7 point isochron), 14C, Sr/Ca (2 subsamples) and δ18O-δ13C (2 subsamples) SVDI11-T14 Thick laminated tufa 2 41.591140 120.052266 1478.4 U-Th (7 point isochron), 14C, Sr/Ca (2 subsamples) and δ18O-δ13C (2 subsamples) SVDI11-T18 Densely laminated tufa 3 41.596241 120.049039 1555.7 U-Th (7 subsamples), 14C, Sr/Ca (2 subsamples) and δ18O-δ13C Middle lake shoreline set (Fig. 2D) SVDI12-T1 Thick laminated tufa on 1 41.429916 119.975595 1419.5 U-Th (7 point isochron) and 14C, Sr/Ca, and δ18O-δ13C porous basalt 14 afutdetanimaL2T-21IDVS 5.1dnanorhcositniop7(hT-U5.9141595579.911619924.141 N HNO3 leach residue), C, Sr/Ca (3 replicates) and δ18O-δ13C (3 replicates) SVDI12-T3 Thin laminated tufa 1 41.429852 119.975167 1427.8 U-Th (2 subsamples), Sr/Ca (2 replicates), and δ18O-δ13C (2 replicates) SVDI12-T4 Thin laminated tufa 1 41.429829 119.974638 1439.0 U-Th (2 subsamples), Sr/Ca (2 replicates), and δ18O-δ13C (2 replicates) 14 SVDI12-T5 Poorly consolidated tufa on 1 41.429869 119.974454 1444.3 U-Th (7 point isochron and 1.5 N HNO3 leach residue), C, Sr/Ca porous basalt (2 replicates), and δ18O-δ13C (2 replicates) SVDI12-T7 Thin laminated tufa 1 41.428042 119.972521 1472.5 U-Th (1 sample), Sr/Ca and δ18O-δ13C SVDI12-T9 Thin laminated tufa 2 41.426983 119.970856 1508.9 U-Th (7 point isochron), 14C, Sr/Ca, and δ18O-δ13C 14 SVDI12-T10 Densely laminated tufa 2 41.426865 119.970569 1516.8 U-Th (7 point isochron and 1.5 N HNO3 leach residue), C, Sr/Ca (2 replicates), and δ18O-δ13C (2 replicates) SVDI12-T11 Poorly consolidated tufa 3 41.426084 119.969309 1554.9 U-Th (1 sample), Sr/Ca, and δ18O-δ13C SVDI12-T12 Thin laminated tufa 3 41.428400 119.967793 1576.9 U-Th (7 subsamples), 14C, Sr/Ca, and δ18O-δ13C Lower lake shoreline set (Fig. 2E) SVDI12-T13 Thick laminated tufa 1 41.217488 119.970070 1437.2 U-Th (7 point isochron), 14C, Sr/Ca, and δ18O-δ13C SVDI12-T14 Thin laminated tufa 2 41.219085 119.965077 1530.7 U-Th (7 point isochron), 14C, Sr/Ca, and δ18O-δ13C Upper lake shoreline set (Fig. 2C) 14 SVDI12-T15 Porous densely consolidated tufa 1 41.717964 120.070054 1433.1 U-Th (7 point isochron and 1.5 N HNO3 leach residue), C, Sr/Ca (2 replicates), and δ18O-δ13C (2 replicates) SVDI12-T17 Thin laminated tufa 3 41.717185 120.062345 1542.3 U-Th (1 sample), Sr/Ca, and δ18O-δ13C SVDI12-T18 Poorly consolidated tufa 3 41.713935 120.062480 1564.2 U-Th (2 subsamples), 14C, Sr/Ca, and δ18O-δ13C SVDI12T19 Poorly consolidated tufa on 3 41.711839 120.062596 1566.8 U-Th (1 sample), Sr/Ca, and δ18O-δ13C porous basalt Modern playa samples (Fig. 2A) SVDI12-C1 Playa carbonate - 41.532766 120.128760 - (234U/238U), (230Th/232Th), Sr/Ca, and δ18O-δ13C (2 replicates) (middle playa edge) SVDI12-P1 Playa surface sediment - 41.425504 119.987361 - 1:1 DI water extraction and NaOAc leach: (234U/238U) (middle playa edge) SVDI12-P2 Playa surface sediment - 41.532766 120.128760 - 1:1 DI water extraction and NaOAc leach: (234U/238U) (middle playa center) SVDI12-P3 Playa surface sediment - 41.602495 120.133613 - 1:1 DI water extraction and NaOAc leach: (234U/238U) (upper playa) SVDI12-P4 Playa surface sediment - 41.079306 119.890701 - 1:1 DI water extraction and NaOAc leach: (234U/238U) (Duck Flats) Modern water samples (Fig. 2A) SVDI12-WS1 Stream (Eagle Creek) - 41.427318 119.971265 - U concentration, (234U/238U), and Sr/Ca SVDI12-WS2 Stream (Emerson Creek) - 41.426983 119.970856 - U concentration, (234U/238U), and Sr/Ca SVDI12-WS3 Stream (Granger Creek) - 41.472795 120.188345 - U concentration, (234U/238U), and Sr/Ca SVDI12-WS4 Stream (Deep Creek) - 41.510626 120.215339 - U concentration, (234U/238U), and Sr/Ca SVDI12-WS5 Hot Spring (Boyd Spring) - 41.725518 120.082829 - U concentration, (234U/238U), and Sr/Ca SVDI12-WS6 Hot Spring (Seyferth Hot Springs) - 41.613584 120.106561 - U concentration and Sr/Ca SVDI12-WS7 Hot Spring (Leonard Hot Springs) - 41.598952 120.091965 - U concentration and Sr/Ca SVDI12-WS8 Hot Spring (Surprise Valley - 41.534183 120.079076 - U concentration and Sr/Ca Hot Springs) SVDI12-WS9 Groundwater (Lake City Well) - 41.634140 120.215041 - U concentration, (234U/238U), and Sr/Ca SVDI12-WS10 Stream (Mill Creek) - 41.642732 120.216200 - U concentration, (234U/238U), and Sr/Ca SVDI12-WS12 Stream (Fort Bidwell Creek) - 41.848853 120.152633 - U concentration, (234U/238U), and Sr/Ca SVDI12-WS13 Pond/Lake (Annie Lake) - 41.908525 120.108812 - U concentration, (234U/238U), and Sr/Ca SVDI12-WS15 Groundwater (Cockeral Ranch - 41.657510 120.219208 - U concentration, (234U/238U), and Sr/Ca Well #1) SVDI12-WS16 Groundwater (Cockeral Ranch - 41.649840 120.214560 - U concentration, (234U/238U), and Sr/Ca Well #2) SVDI12-WS17 Hot Spring (Lake City Hot Springs) - 41.666514 120.212447 - U concentration and Sr/Ca SVDI12-WS18 Hot Spring (Unnamed Hot Spring) - 41.218816 120.065092 - U concentration, (234U/238U), and Sr/Ca SVDI12-WS19 Stream (Lost Creek) - 41.084588 119.892403 - U concentration, (234U/238U), and Sr/Ca SVDI12-WS20 Groundwater (Eagleville Well) - 41.322792 120.117908 - U concentration, (234U/238U), and Sr/Ca SVDI12-WS21 Groundwater (Cedarville Well) - 41.530723 120.184478 - U concentration, (234U/238U), and Sr/Ca Note: DI—deionized water. Ac—acetate. *Tufa sample altitude determined by pinning latitude-longitude location to a 0.5-m-horizontal-resolution raster light detection and ranging (LiDAR) data set, reported as meters above sea level. Typical elevation error is ±0.1m (2σ).

Geological Society of America Bulletin, Month/Month 2014 5 Geological Society of America Bulletin, published online on 2 June 2014 as doi:10.1130/B31014.1

Ibarra et al.

230 relatively short half-lives of Th (t1/2 = 75.4 removal and scavenging of hydrogenous Th by material, can constrain the reservoir effect or 234 230 k.y.) and U (t1/2 = 245 k.y.), Th-U geochro- absorption to particles (Anderson et al., 1982), identify anomalously young ages (e.g., Cassata nology has an upper age limit of ca. 500–800 ka as observed in ocean sediments (Bacon and et al., 2010; Zimmerman et al., 2011; Vazquez (Cheng et al., 2013). The equations for calculat- Anderson, 1982). In most cases, without correc- and Lidzbarski, 2012). ing the ingrowth of 230Th and the closed-system tion for initial 230Th, the measured (230Th/238U) decay of 234U are (Kaufman and Broecker, 1965; is too high, resulting in an age calculation METHODS Thurber et al., 1965; Neymark, 2011): that is erroneously old. Lake carbonates are likely to have two end-member sources of Th: Topographic Analyses ⎛ 234 ⎞ ⎛ 234 ⎞ U −−λλttU (1) hydrogenous Th in the primary carbonate ⎜ ⎟ =−()1 e 234 +⎜ ⎟ e 234 , (1) ⎝ 238U ⎠ ⎝ 238U ⎠ derived from water-soluble Th, and (2) detrital Calculation of Lake Surprise Basin Geometry t 0 Th incorporated within the carbonate as silici- Building on recent work by Irwin and clastic or organic particles (e.g., Lin et al., 1996; Zimbel man (2012), we used modern Surprise ⎛ 230 ⎞ ⎛ λ ⎞ ⎛ 234 ⎞ Th U −λ ttt−λ ⎜ ⎟ = ⎜ 230 ⎟ ⎜ ⎟ (e 234 − e 230 ) Blard et al., 2011; Torfstein et al., 2013). In Valley topography to calculate lake volume and ⎝ 238U ⎠ ⎝λλ− ⎠ ⎝ 238U ⎠ t 230 234 0 freshwater systems, low hydrogenous Th con- surface area at elevations spanning the modern

−−λλtt centrations are expected; however, some studies playa to 10 m above the proposed highstand ⎛ λλee230 − 234 ⎞ +,1+ 234 230 (2) of paleolake systems have observed substantial of Reheis (1999a). We analyzed digital eleva- ⎝⎜ λλ− ⎠⎟ 230 234 contribution from both Th end members (e.g., tion models (DEMs) from the U.S. Geological Lin et al., 1996; Haase-Schramm et al., 2004). Survey (USGS) National Elevation Data set λ λ where 230 and 234 are decay constants for Assuming that freshwater lake carbonates (NED) with ArcGIS 10.1 hydrology and 3-D 230Th and 234U, respectively, t is the time in include only the detrital end member, we will spatial analyst tools. Lake surface area and vol- 234 238 years since the mineral formed, and ( U/ U)t apply and assess two complementary strate- ume were calculated from the minimum playa 230 238 and ( Th/ U)t are the measured or corrected gies, single-sample correction and an isochron elevation to above the maximum lake terrace activity ratios. These equations are solved approach, to calculate the 230Th-U ages for elevation (1355–1577 m). To calculate vol- 234 238 ≥ for ( U/ U)0 and time (t). In closed-system coeval sets (n 5) of whole-rock dissolutions. ume, a triangulated irregular network (TIN) rocks and minerals older than ~106 yr, the topographic surface representation was created (234U/238U) and (230Th/238U) are equal to one, a of Lacustrine from the DEM and intercepted with lake surface condition referred to as “secular equilibrium.” Carbonates area. Lake volume calculations do not account The (234U/238U) of natural water is typically Radiocarbon geochronology has been exten- for postlake sediment infi ll and thus represent elevated (>1) due to the preferential release of sively applied to inorganic and organic carbon- minimum lake volumes. Empirical relation- the daughter 234Th nuclide across grain boundar- ates found as surfi cial deposits (e.g., tufa) and in ships, fi t using polynomial functions, were ies during the α-decay of 238U, a process known nearshore sediments to date Pleistocene shore- established to relate volume and surface area as α-recoil (Kigoshi, 1971; Fleischer, 1982; lines (e.g., Oviatt et al., 1992; Lin et al., 1996; to elevation (Fig. DR1 [see footnote 1]). The Oster et al., 2012). Under oxidizing conditions García and Stokes, 2006; Munroe and Laabs, inward-draining watershed area was calculated typical of surface waters, U is mobile as the 2013). Assuming a carbonate sample contains after smoothing the DEM with a 1 m vertical VI 2+ hexavalent uranyl ion [U O2] and is incorpo- carbon originally fi xed from the atmosphere, threshold. We assumed that modern contours rated into secondary minerals such as carbon- radiocarbon geochronology relies upon the and the basin topography accurately represent 14 14 ates (Reeder et al., 2001). In contrast, Th is decay of C (t1/2 = 5.73 k.y.) to N to calculate the watershed topography and lake hypsometry relatively insoluble under oxic conditions, and a radiocarbon (14C) age. Because the production of the late Pleistocene, with minimal isostatic ideal samples for 230Th-U dating have low initial and reservoir of 14C in the atmosphere have not rebound. The results of the Lake Surprise basin Th concentrations [e.g., (230Th/238U) = 0], such been constant with time, calibration data sets geometry analysis were also compared to work that all postdepositional 230Th accumulation is are required to correct the radiocarbon age of by previous authors (Miffl in and Wheat, 1979; derived from radioactive decay of 234U in the samples (Reimer et al., 2013). Beyond the Holo- Reheis, 1999a; Zimbelman et al., 2009; Irwin sample. However, by measuring the abundance cene, uncertainty in radiocarbon ages can be the and Zimbelman, 2012). 232 of Th (t1/2 ~14 b.y.), corrections can be made result of plateaus occurring in the radiocarbon Based on analyses of the high-resolution for any “detrital Th” (i.e., Th derived from other calibration curve during the last deglaciation topographic data set, we updated topographic processes aside from purely radioactive decay) and the short half-life of 14C (Trumbore, 2000; results from Irwin and Zimbelman (2012) in that may have been originally incorporated into Southon et al., 2012). In lacustrine carbonates, Table 2. The pour point elevation (1621 m) is the carbonate (Luo and Ku, 1991; Ludwig and two complications may occur. A reservoir effect >50 m higher than all proposed highstand ele- Titter ington, 1994; Ku, 2000; Neymark and due to the incorporation of signifi cant quanti- vations (Hubbs and Miller, 1948; Miffl in and Paces, 2000; Sharp et al., 2003; Paces et al., ties of “dead” carbon (i.e., low 14C/12C carbon Wheat, 1979; Reheis, 1999a; Personius et al., 2004; Fletcher et al., 2010). from weathering of carbonate bedrock) results 2009; Zimbelman et al., 2009; Irwin and Zim- In fact, the assumption of no initial 230Th is in a radiocarbon date that is up to thousands of belman, 2012) and the pre–MIS 2 and MIS 2 rarely valid for lake carbonates (Placzek et al., years older than the true age. Alternatively, in highstands studied here. Furthermore, all mod- 2006a, 2006b; Blard et al., 2011; Torfstein et poorly consolidated, porous, or impure carbon- ern streams drain terminally into the playa lakes al., 2013). In Mono Basin (Fig. 1, Lake Rus- ates, modern or younger atmospheric carbon (Fig. 2A). Groundwater contribution to Lake sell), unsupported dissolved Th, which may can contaminate the sample (Cassata et al., Surprise is assumed to be negligible because co precipi tate with lacustrine carbonates, can be 2010; Zimmerman et al., 2011, 2012). In an the –Pleistocene lacustrine sediments the result of high Th activity and complexation ideal situation, application of several absolute provide the primary groundwater storage capac- 2– 230 40 39 with CO3 (Anderson et al., 1982; Lin et al., dating methods, such as Th-U or Ar/ Ar, ity for modern-day Surprise Valley and are 1996; Zimmerman et al., 2012). This inhibits applied to coeval or depositionally equivalent recharged by nearshore lacustrine deposits and

6 Geological Society of America Bulletin, Month/Month 2014 Geological Society of America Bulletin, published online on 2 June 2014 as doi:10.1130/B31014.1

Lacustrine paleoclimate record from late Pleistocene Lake Surprise, California

# Holocene alluvial fans (Clawson et al., 1986).

†† The groundwater aquifer capacity is estimated Δ

(%) 3

77.1 to be 4.93 km (4 million acre-feet) to 122 m

precipitation below the valley fl oor (Clawson et al., 1986), which is only 2.9% of the lake volume at its lat-

†† est Pleistocene highstand of 1531 m (lake vol- ume = 164 km3; Table 2). 552.5 991.0**619.7** 10 (2 to 18)** ~75** (mm/yr) Estimated precipitation Field Methods †† § §

Δ Modern water and playa samples were col- (°C) lected in July 2012 to assess the geochemis-

temperature try of waters in the Surprise Valley watershed

O to better constrain the expected variability in 18 δ elemental and uranium isotopic composition of the late Pleistocene tufa samples. Modern water Tufa Tufa (‰, VPDB) samples were collected from seven streams, † fi ve groundwater wells, six hot springs, and one 95.0 75.0 35.0 small pond within the Surprise Valley watershed = 0.11 (HI)* index

met (Table 1; Fig. 2A). Water samples were fi ltered Hydrologic HI in the fi eld through 0.45 µm polyethylene fi lters into 1 L polyethylene bottles and acidifi ed with 1261 now water equivalent from Lake City (1929–1960), Cedarville (1894–2012), and high-purity nitric acid (HNO3) to ~2% HNO3. nd Wheat, 1979; Reheis, 1999a); values are from annual average of three

, 1979; Reheis, 1999a). Ten milliliters were aliquoted for concentration (m a.s.l.) elevation Pour point analyses (U, Th, Ca, and Sr), and 250–500 ml were dried down over ~2 wk in 50 mL Tefl on )

3 beakers for U isotope analysis. Four modern sur-

(km face sediment samples were collected from the

Lake volume playas (Table 1; Fig. 2A). One modern evaporite sample was collected from the middle playa. 86.2 5 09 Tufa Sample Collection 7 . . 2 2 Tufa samples were collected in August 2011 lake area and July 2012 from exposed bedrock on the Watershed area/ Watershed front edge or the crest of wave-cut terraces and

B horizontal shoreline benches, prominent on the ) A 7383 0404 00 2 east side of the valley due to the prevailing west- 8 3 (km

area, erly winds. Four sample localities were targeted Watershed Watershed adjacent to each of the three modern playa lakes

) (Figs. 2B–2E). While the absolute depth of tufa 2 0 1 0 1 3 7 formation relative to lake level is not certain 41 31 41 (km L

A or directly quantifi able, their association with Lake area, wave-cut shoreline features and biologic tex- tures (cf. Rouchy et al., 1996) indicates their TABLE 2. LAKE GEOMETRY, WATERSHED AREA, AND HYDROLOGIC INDICES FOR LAKE SURPRISE AREA, WATERSHED 2. LAKE GEOMETRY, TABLE 7651 5451 formation within the photic zone and likely near

1534 the lake surface and edge. Thus, the tufa sample

(m a.s.l.) elevations give minimum lake surface elevations at their time of deposition (e.g., Felton et al., 2006; Benson et al., 1996, 2011; Zimmerman enecotsielPet et al., 2012). Latitude and longitude of sample locations were recorded with a handheld global position- Lahontan 1567 1471 4040 2.75–2.34 ing system0.68 (GPS; Garmin Oregon 550t). Using a L)a9991(s a newly acquired light detection and ranging (LiDAR) data set, sample elevations were deter- mined by pinning the location coordinates to the ) 80 0.5-m-horizontal-resolution raster digital eleva- 0 is (meteorology derived) hydrologic index = runoff from tributary basin/(lake evaporation minus lake precipitation) (Miffl in a from tributary basin/(lake evaporation minus lake precipitation) (Miffl is (meteorology derived) hydrologic index = runoff 2(emr tion model derived from the LiDAR point cloud. ieheR met in and Calculated by Miffl in and Wheat (1979) assuming a 2.34 °C (5 °F) temperature reduction, 312 mm/yr precipitation, and an average hydrologic index of 0.68 in the south-central Oregon and extreme in and Wheat (1979) assuming a 2.34 °C (5 °F) temperature reduction, 312 mm/yr precipitation, an average Calculated by Miffl HI and Whitlock, 1995). Estimated from pollen assemblages at Little Lake, Oregon (Worona Percent change in precipitation, LGM minus modern. Modern precipitation is taken as 566 mm/yr, as the average of rainfall and s Percent change in precipitation, LGM minus modern. Modern precipitation is taken as 566 mm/yr, The elevation error associated with the elevation **See Table 8 for calculations. Table **See *HI is (topography derived) hydrologic index = lake surface area/tributary area (basin area minus lake area) (Miffl in and Wheat *HI is (topography derived) hydrologic index = lake surface area/tributary area (basin minus area) (Miffl : MIS—marine oxygen isotope stage; LGM—Last Glacial Maximum; VPDB—Vienna PeeDee Belemnite. Note : MIS—marine oxygen isotope stage; LGM—Last Glacial Maximum; VPDB—Vienna # †† (2012) † § Zimbelman (2009) Wheat (1979) northwestern Nevada region. Study Interpreted age Lake elevation This study MIS 2 (LGM) 1420 to 1440 851 to 977 3812 4.48 to 3.90 39 to 57 1621 0.29 to 0.34 –3.388 –7 Fort Bidwell (1911–2011) meteorological stations archived in the Western Regional Climate Center database (http://www.wrcc.dri.edu/summary/Climsmnca.html). meteorological stations archived in the Western Fort Bidwell (1911–2011) modern meteorological stations in Surprise Valley with lake evaporation assumed to be equivalent to pan evaporation. Average precipitation = 566 mm/yr, average pan evaporation = 905 mm/yr, and average average pan evaporation = 905 mm/yr, precipitation = 566 mm/yr, Average with lake evaporation assumed to be equivalent pan evaporation. modern meteorological stations in Surprise Valley in and Wheat (1979). estimated from graphs in Miffl Runoff = 38 mm/yr. runoff This study Modern meteorology This study Modern playa lakes 1355 to 1372 0 to 423 3812 >9.01 0 to 2.1 1621 0 to 0.12 This study MIS 4 or 6 1567 1499 3812 2.54 218 1621 0.65 This study MIS 2 (post-LGM) 1531 1366 3812 2.79 164 1621 0.56 –3.845 –5 Irwin and Zimbelman et al. Miffl O of the tufa samples, primarily collected from

Geological Society of America Bulletin, Month/Month 2014 7 Geological Society of America Bulletin, published online on 2 June 2014 as doi:10.1130/B31014.1

Ibarra et al.

234 238 238 232 exposed wave-cut shorelines, is ±10 cm. The (normalized by either U or U). The mea- by the y-intercept of the ( U/ Th)M versus 230 238 230 232 location and elevation of all samples are pre- sured ( Th/ U) is corrected using the equation ( Th/ Th)M Rosholt isochron or the slope of 232 238 230 238 sented in Table 1 and Figure 2. from Kaufman (1993): the ( Th/ U)M versus ( Th/ U)M Osmond isochrons (Lin et al., 1996; Blard et al., 2011). Analytical Procedures ⎛ 230 ⎞ ⎛230 ⎞ ⎛232 ⎞ ⎛ 230 ⎞ Various methods have been investigated to con- Th Th Th Th −λ t ⎜ ⎟ = ⎜ ⎟ − ⎜ ⎟ ⎜ ⎟ ()e 230 , (3) struct sample pair isochrons. Initially, studies ⎝ 238U ⎠ ⎝ 238U⎠ ⎝ 238U⎠ ⎝ 232TTh⎠ The laboratory work carried out to deter- AMM 0 relied on leachate/residue (L/R) and leachate/ mine the Lake Surprise hydrograph during the leachate (L/L) methods (e.g., Kaufman, 1971; last glacial cycle included the application of where the A denotes the authigenic or sup- Schwarcz and Latham, 1989) to produce mea- 230Th-U geochronology, stable isotopic analy- ported (230Th/238U), M denotes the measured surement pairs used to construct isochrons. ses, elemental analyses, and accelerator mass isotope ratios (230Th/238U) and (233Th/238U), 0 However, leachate schemes have been found to spectrometer (AMS) radiocarbon (14C) geo- denotes the assumed initial (230Th/232Th), and preferentially solubilize U or Th. Nevertheless, if chronology on tufa samples. To support the t is the apparent time calculated before correc- the majority of U is in the authigenic carbonate 230Th-U geochronology, we also analyzed the tion for initial Th (cf. Kaufman, 1993; Israelson phase, these methods may be acceptable (e.g., Ku elemental and U isotopic composition of mod- et al., 1997; Placzek et al., 2006a, 2006b). Once and Liang, 1984; Schwarcz and Latham, 1989; ern playa sediments (using leachate methods of the measured ratio is corrected for initial Th, a Przyblowicz et al., 1991; Kaufman, 1993). Alter- Tessier et al., 1979; Maher et al., 2003, 2006; corrected age is calculated by substituting of natively, to produce consistent mixing relation- 230 238 Oster et al., 2012) and modern water samples. ( Th/ U)A into Equation 1. The magnitude and ships between the authigenic and detrital compo- Detailed analytical methods are provided in the error associated with this correction are a func- nents, a total sample dissolution (TSD) method, Data Repository (see footnote 1). tion of the concentration of 232Th in a sample, involving full digestion of suites (n ≥ 3) of coeval 230 232 U and Th isotopic compositions were ana- the assumed ( Th/ Th)0, and the error on the samples (Bischoff and Fitzpatrick, 1991; Luo and 230 232 lyzed at the ICP-MS–thermal ionization mass assumed ( Th/ Th)0. The error on the assumed Ku, 1991), has been used in recent studies (e.g., 230 232 spectrometry (TIMS) facility at Stanford Uni- ( Th/ Th)0 is typically the greatest source of Hall and Henderson, 2001; Soligo et al., 2002; 230 232 versity, using a Nu Instruments Plasma high- error on the calculated age when ( Th/ Th)M Garnett et al., 2004; Haase-Schramm et al., 2004; resolution MC-ICP-MS. Column chemistry was < 10. Although some authors have assumed that Blard et al., 2011; Torfstein et al., 2013). Here we 230 232 performed using standard element specifi c ion ( Th/ Th)0 is unity (e.g., Sylvestre et al., 1999) used the TSD method for all subsamples. We also exchange chromatography methods (Luo et al., or equal to upper continental crust (e.g., Ludwig measured four residues from a 1.5 N HNO3 leach 1997; Potter et al., 2005; Stirling et al., 2007). and Titterington, 1994; Polyak and Asmeron, of separate subsamples with suffi cient sample Isotopic standards CRM 145, SRM 4321A, 2001), it has been observed that in terrestrial material. IRMM-036, and OU Th “U,” and multiple aqueous environments, including paleolakes, Several methodologies have been proposed 230 232 dissolution aliquots of spiked rock references ( Th/ Th)0 can vary from 0.5 to 4.2 (cf. to construct statistically signifi cant linear regres- BZVV, BCR-2, and TML were measured during Placzek et al., 2006a, their table 5). Thus, for a sions of the paired isochrons. Statistical rigor is analytical sessions. The long-term averages of given system, the most accurate constraint on required to produce meaningful and representa- 230 232 230 the isotopic standards and rock reference stan- ( Th/ Th)0 is via modern lake carbonate mea- tive Th-U ages while accounting for both ana- dards (Fig. DR2; Table DR2 [see footnote 1]) surements (e.g., Lin et al., 1996; Israelson et al., lytical uncertainties in MC-ICP-MS or TIMS agree with previously measured literature values 1997) or an isochron approach (e.g., Placzek measurements, as well as geologic scatter. Luo (Turner et al., 1997, 2001; Thomas et al., 1999; et al., 2006a; Blard et al., 2011; Torfstein et and Ku (1991) proposed a methodology using a Raptis et al., 1998; Cheng et al., 2000; Shen al., 2013), as described next. We calculated simple least-squares fi t with counting statistics, et al., 2002; Amelin and Back, 2006; Placzek single-sample ages for all subsamples and error- which does not consider error-weighting or error et al., 2006a; Sims et al., 2008; Neymark, 2011; weighted averages for each set of subsamples. correlation (in x and y errors), noting that doing Oster et al., 2012; Torfstein et al., 2013). Isochron method. A complementary approach so would require more than three or four coeval to single-sample corrections is an isochron samples, which is often unpractical given sample U-Series Age Determination approach. Isochrons allow for the calculation size and laboratory method limitations (cf. Ku, To calculate detrital Th–corrected 230Th-U of the (230Th/238U) and (234U/238U) of the pure 2000). Ludwig and Titterington (1994) proposed ages, we relied upon both the isochron authigenic carbonate from sets of coeval sam- a 3-D simultaneous solution of the Osmond approach, involving suites of coeval samples, ples, without necessarily needing to constrain isochrons using maximum-likelihood estima- 230 232 and single-sample detrital Th correction. From ( Th/ Th)0. Two mathematically equiva- tion (MLE), allowing for the projection of the 21 tufa samples, we processed a total of 111 lent pairs of two-dimensional (2-D) isochrons detritus-free end member onto a traditional 230Th- subsamples. This included 13 sets of 7 coeval sharing one common axis or a simultaneous 234U-238U evolution plot. This method is imple- samples, two sets of 5 coeval samples from the three-dimensional (3-D) solution can yield mented in Isoplot (Ludwig, 2012). Finally, recent 230 238 234 238 inner and outer rind of a thick lower-elevation ( Th/ U)A and ( U/ U)A. The 2-D iso- studies (e.g., Hall and Henderson, 2001; Soligo et 230 238 234 238 sample (SVDI11-T2), and 10 supporting single chrons yield ( Th/ U)A and ( U/ U)A via al., 2002; Garnett et al., 2004; Blard et al., 2011) 238 232 or duplicate samples. Additionally, to support the slopes of Rosholt isochrons [( U/ Th)M vs. have used error-weighted 2-D linear fi ts to the 230 232 238 232 234 232 the construction of isochrons, four 1.5 N HNO3 ( Th/ Th)M and ( U/ Th)M vs. ( U/ Th)M] individual Rosholt or Osmond isochrons using leach residues and one modern carbonate sam- or the y-intercepts of Osmond-type isochrons methods originally developed by York (1968) and 232 238 230 238 232 238 ple were also analyzed. [( Th/ U)M vs. ( Th/ U)M and ( Th/ U)M implemented in Isoplot (Ludwig, 2012). 234 238 Single-sample correction. For a single (sub-) vs. ( U/ U)M] (Osmond et al., 1970; Rosholt, For the suites of coeval samples, to calculate 230 234 230 238 234 238 sample, the Th daughter (supported by U 1976; Luo and Ku, 1991; Bischoff and Fitzpatrick, ( Th/ U)A and ( U/ U)A, and calculate a and 238U) can be determined given an assumed 1991; Ludwig and Titterington, 1994). Further- detrital Th–corrected isochron age, we evalu- initial (230Th/232Th) and 232Th concentration more, the detrital (230Th/232Th) can be determined ated both solutions to separate 2-D isochrons

8 Geological Society of America Bulletin, Month/Month 2014 Geological Society of America Bulletin, published online on 2 June 2014 as doi:10.1130/B31014.1

Lacustrine paleoclimate record from late Pleistocene Lake Surprise, California

(e.g., Hall and Henderson, 2001; Soligo et al., for comparison to paleoclimate records, we A positive covariance with a rating of weak, 2002; Garnett et al., 2004; Blard et al., 2011) report the percent change in precipitation, total moderate, or strong following the method of and the simultaneous 3-D solution of the evaporation, and runoff simulated by the LGM Davis et al. (2009) is shown in Table 3. For all Osmond isochrons using MLE (Ludwig and Tit- experiments minus preindustrial control experi- samples <1531 m, both δ18O-δ13C and δ18O- terington, 1994). Both methods calculate nearly ments for each climate model in the PMIP3 Sr/Ca are moderately correlated (Figs. 3C and identical absolute ages but differ in the calcu- ensemble. For relative humidity and mean annual 3D). For the lowest-elevation samples of LGM 230 238 234 238 δ18 δ13 lated error on ( Th/ U)A and ( U/ U)A. For temperature, we report the absolute change. age, O- C are strongly correlated; how- the 2-D solutions, to best illustrate spread along ever, δ18O-Sr/Ca are not correlated, likely due the isochrones, we use the slopes of Rosholt RESULTS to minimal spread in measured Sr/Ca ratios. 238 232 230 232 δ18 δ13 isochrons [( U/ Th)M vs. ( Th/ Th)M and For middle-elevation samples, O- C are 238 232 234 232 δ18 ( U/ Th)M vs. ( U/ Th)M]. While construct- In the following sections, we present our moderately correlated, and O-Sr/Ca are very ing isochrones, some subsamples were rejected, results and examine the reliability of the geo- strongly correlated (Figs. 3A and 3B). Values of including residues, to minimize the error on chronologic approaches, spatial relationships, δ18O, δ13C, and Sr/Ca from the highest-elevation 238 232 230 232 the ( U/ Th)M versus ( Th/ Th)M isochron and the U-series and stable isotope systematics. samples suggest potential alteration of the sam- slope and produce a mean square weighted We then construct a late Pleistocene hydrograph ples due to recrystallization or pedogenic over- deviation (MSWD) closest to ~1 (cf. Hall and for Lake Surprise. For the purposes of discus- printing (Table 3; Fig. 3). Henderson, 2001; Garnett et al., 2004). In an sion, the tufa samples are split into three groups The covariance of δ18O-δ13C and δ18O-Sr/Ca effort to maximize the precision of our 230Th-U by elevation: (1) samples from prominent lower- can be used to evaluate the role of evaporation ages, we attempted to construct isochrons using elevation shorelines at 1419–1472 m; (2) sam- in lake systems (Müller et al., 1972; Eugster and 5–7 subsamples, i.e., more than previous stud- ples from the less-prominent middle-elevation Kelts, 1983; Talbot, 1990; Li and Ku, 1997; Davis ies that have used isochron approaches to date shorelines of 1478–1531 m; and (3) samples et al., 2009; Chamberlain et al., 2013). Terminal impure carbonates and evaporites (e.g., Luo and from higher-elevation shorelines >1542 m basin lakes have comparatively long residence Ku, 1991; Blard et al., 2011). (Figs. 2B–2E). times, resulting in the evaporative enrichment of 18O in the lake water and preferential outgassing 12 Climate Model Output Analysis Evaluating Stable Isotope, Sr/Ca, and of C-rich CO2 from the system (Talbot, 1990). U-Series Systematics If a lake is hydrologically closed, and evaporation To calculate changes in precipitation and is high, then δ18O and δ13C will covary (Talbot, evapotranspiration predicted by AOGCMs for Stable Isotopic and Elemental Signatures of 1990; Li and Ku, 1997; Davis et al., 2009). Simi- the western United States, we computed aver- Tufas and Modern Samples larly, Sr/Ca ratios have been used to assess evapo- age monthly climatologies using the monthly Measurements of δ18O, δ13C, and Sr/Ca in rative effects in lake systems (Müller et al., 1972; precipitation, total evaporation, runoff, relative paleoshoreline carbonates and modern waters Davis et al., 2009). Müller et al. (1972) showed humidity, and mean annual temperature out- and sediment are shown in Figure 3. All paleo- that for highly evaporative lakes, Sr/Ca in car- puts for the LGM (21 ka boundary conditions) shoreline samples from <1531 m have δ18O bonates is proportionally correlated to Sr and Ca and preindustrial control experiments com- values of −2.50‰ to −4.51‰, and δ13C values concentrations in lake water. With a KSr less than piled by the Paleoclimate Model Intercompari- of 2.68‰–4.10‰; the modern playa carbon- 1 (Gabitov and Watson, 2006), Sr is not removed son Project 3 (PMIP3; Braconnot et al., 2012; ate has δ18O = −4.05‰ and δ13C = −0.11‰. as effi ciently from lake water as Ca, which is http:// pmip3 .lsce .ipsl .fr/). The nine models we High-elevation samples have much lower δ18O taken up by precipitating carbonates; thus, Sr included are NCAR-CCSM4 (Gent et al., 2011; values of −13.13‰ to −9.33‰, and lower δ13C is concentrated when the net evaporative fl ux is Brady et al., 2013), CNRM-CM5 (Voldoire values of −6.32‰ to −1.72‰. Others report high relative to lake volume (Eugster and Kelts, et al., 2013), FGOALS-g2 and IPSL-CM5A-LR a large, −14.4‰, seasonal variation in mod- 1983; Davis et al., 2009). This correlation is not (Kageyama et al., 2013a, 2013b), MRI-CGCM3 ern lake-water δ18O from the middle playa of expected with an open-system lake because Sr (Yukimoto et al., 2012), MPI-ESM-P, GISS- −9.5‰–4.9‰ and minimal variation in creek is fl ushed from the system. Surprise Valley tufa E2-R, COSMOS-ASO, and MIROC-ESM waters, from −15.0‰ to −13.7‰ (Ingraham and samples demonstrate a moderate to high δ18O- (Sueyoshi et al., 2013); all were used because Taylor, 1989; Sladek et al., 2004; Table DR3 δ13C and δ18O-Sr/Ca covariance among sample of their inclusion into the PMIP3/CMIP5 data- [see footnote 1]). groups (Table 3; Fig. 3). base (Coupled Model Intercomparison Project 5; If the range of measured modern water Sr/Ca The extent to which the Surprise Valley lake http:// cmip-pcmdi .llnl .gov /cmip5/). Monthly refl ects the range of late Pleistocene lake-water system was a terminal basin in the past deter- and annually summed precipitation, total evapo- Sr/Ca (Tables 3 and 4), and assuming a range mines the utility of lake shoreline ages and other ration, and runoff anomalies (LGM minus pre- of values for the partition coeffi cient (KSr) of geochemical data as indicators of past climate. industrial control) were calculated for Surprise Based on previous studies of lacustrine carbon- (/Sr Ca )calcite Valley using bilinear interpolation for both Sr ==−ates and the relatively robust trends observed K solution 012.. 035 climate model experiments. Similarly, annu- (/Sr Ca ) here, we propose that the δ18O-δ13C-Sr/Ca sys- ally averaged relative humidity and tempera- tem refl ects the fact that the latest Pleistocene ture anomalies were calculated. Not all output (Gabitov and Watson, 2006), the Sr/Ca mea- Lake Surprise was indeed a closed, inward- variables are uniformly archived in the PMIP3 sured in the tufa samples (Table 3) is at the high draining pluvial lake system. database. COSMOS-ASO does not report run- end of the range of expected values. The high- off values, and COSMOS-ASO, FGOALS-g2, est-elevation samples have a lower Sr/Ca (aver- U-Series Measurements of Modern Samples MIROC-ESM, and MPI-ESM-P do not report age = 0.86 ± 0.66 mmol/mol) compared to the The 19 water samples and 5 modern playa relative humidity values. No bias correction was samples from elevations <1531 m (average = samples (see Fig. 2A for locations) demonstrate applied to the climate models outputs. Instead, 1.16 ± 0.30 mmol/mol). the range of expected variability in U concentra-

Geological Society of America Bulletin, Month/Month 2014 9 Geological Society of America Bulletin, published online on 2 June 2014 as doi:10.1130/B31014.1

Ibarra et al. Sr/Ca average C replicate 13 δ C 13 δ average O replicate 18 δ O 8 C 16 18 δ 13 .3 δ average )* 2 064.3– r O C 18 13 δ O- δ 18 δ Covariance ( 4 03.18.72 )* 2 3.574 –3.998 –0.095 –4.109 –0.131 Sr/Ca r ), rated using no correlation, weak, moderate, and strong based on 2 (mmol/mol) (‰, VPDB) (‰, VPDB) (‰, VPDB) (‰, VPDB) r Sample group covariation and averages O-Sr/Ca Strong Moderate 18 Moderate Moderate Moderate Strong δ (m) 4 No correlation Strong 1 Altitude Covariance ( TABLE 3. STABLE ISOTOPE AND ELEMENTAL ANALYSES AND ELEMENTAL ISOTOPE 3. STABLE TABLE 1 group no. a3T-2 1 IDVS : VPDB—Vienna PeeDee Belemnite. Note : VPDB—Vienna cient ( *Pearson product-moment coeffi (middle playa edge) Sample group 1 0.06 0.61 –3.48 ± 0.72 3.56 ± 0.51 1.14 ± 0.18 SVDI12-T1SVDI12-T2c 1 1 1419.5 1.064 1419.5 1.214 –3.696 –3.124 3.599 3.637 Sample name Sample SVDI11-T2-2SVDI11-T3-1aSVDI11-T3-2SVDI11-T4-1a 1 1SVDI11-T4-1bSVDI11-T14-1a 1 1SVDI11-T14-1c 1453.5 1437.7 1SVDI11-T18-1c 2 1437.7 2 1.157 1430.6 1.123 3 1430.6 1478.4 1.144 1.110 1478.4 –3.385 –4.045 1.144 1555.7 1.517 –2.988 1.561 –3.238 3.472 1.250 3.282 –3.225 –2.501 3.302 –2.851 3.789 –12.559 3.690 4.103 3.598 –6.142 SVDI12-T13SVDI12-T14 1SVDI12-T15 2SVDI12-T17 1437.2 1 1530.7SVDI12T19 1.201 3 0.878 1433.1 3 –3.343 1542.3 1.126 –3.846 1566.8 3.728 0.439 –3.089 3.737 0.667 –13.134 –3.387 3.764 –3.844 –11.041 –5.565 3.751 –6.321 –13.240 3.702 –11.057 –5.604 –6.402 Davis et al. (2009). SVDI12-T2b 1 1419.5 1.291 –3.472 3.632 Sample group 3 0.42 0.61 ± 2.87 –11.30 –4.80 ± 4.15 0.86 ± 0.66 SVDI11-T2-1 1 1453.5 1.070SVDI11-T18-1e 3 –3.123 1555.7 3.866 0.690 SVDI12-T15SVDI12-T18 1 3 1433.1 1564.2 1.070 0.645 –3.129 –11.548 3.726 –6.219 –11.789 –6.266 SVDI12-T4bSVDI12-T5bSVDI12-T7 1 1 1439.0SVDI12-T11 1 1444.3 1.039 1472.5 3 1.093 –3.511 1.140 1554.9 –4.509 3.555 –3.463 1.236 2.679 3.654 –9.871 –2.032 –3.789 3.629 SVDI12-T2a 1 1419.5 1.273 –3.491 3.504 All samples <1531 m 0.36 0.47 –3.46 ± 0.78 3.59 ± 0.49 1.16 ± 0.30 Sample group 2 0.92 0.40 –3.33 ± 0.95 3.69 ± 0.39 1.21 ± 0.54 SVDI12-C1-1 SVDI12-T3bSVDI12-T4aSVDI12-T5a 1 1SVDI12-T9 1427.8 1 1439.0SVDI12-T10b 1.267 1444.3 1.026 2 2 –3.457 0.989 –3.827 1508.9 1516.8 3.621 –3.888 1.084 3.532 1.093 3.205 –3.587 –3.538 3.609 3.514 –3.465 3.696 SVDI12-T10aSVDI12-T12 2 1516.8 3 1.100 1576.9 –3.552 1.115 3.596 –9.332 –1.716 –9.434 –1.730 Middle playa shoreline set Accommodation zone shoreline set Lower playa shoreline set Upper playa shoreline set Modern carbonate sample O 2 4 6 8 10 12 14 2.5 3.0 3.5 4.0 4.5 18 C for each C for C and δ 13 13 ) were calcu- ) were 2 C systematics =0.47 13 2 C (B and D) for C (B and D) for O- δ r =0.40 13 18 2 r O- δ (‰) (‰) 18 O- δ δ 18 2024 VPDB VPDB cients ( r cients C C 13 13 O and δ δ δ 18 4 δ Groups 2 and 3 Correlation =0.61 2 =0.61 r 2 r 6 2.5 3.0 3.5 4.0 4.5 O (A and C) δ O (A 1.75 18 Group 3 =0.06 2 r =0.92 2 r Group 2 =0.36 2 r =0.42 2 r Sr/Ca (mmol/mol) Sr/Ca (mmol/mol) Group 1 0.50 0.75 1.00 1.25 1.50 0.75 1.00 1.25 1.50 1.75 AB CD 2 4 6 8

10 12 14

2.5 3.0 3.5 4.0 4.5

VPDB VPDB VPDB

(‰) O (‰) O

δ δ

18 18 lated for linear least squares regressions for Sr/Ca- for regressions least squares linear lated for sample group and for the combination of sample groups 1 and 2. Combining the combination of sample groups and for sample group Sr/Ca- 1 and 2 (A–B) demonstrates similar sample groups tufa samples grouped by elevation. Sample are grouped using the same symbol- grouped by elevation. Sample are tufa samples grouped coeffi 2B–2E. Pearson product-moment ogy as in Figures Figure 3. Correlation plots of Sr/Ca- δ 3. Correlation Figure signatures (C–D), suggesting potential recrystallization and/or different source source different and/or (C–D), suggesting potential recrystallization signatures PeeDee Belemnite. VPDB—Vienna origins. water among all samples <1531 m. Group 3 samples have very different δ 3 samples have very different among all samples <1531 m. Group

10 Geological Society of America Bulletin, Month/Month 2014 Geological Society of America Bulletin, published online on 2 June 2014 as doi:10.1130/B31014.1

Lacustrine paleoclimate record from late Pleistocene Lake Surprise, California

TABLE 4. Sr/Ca, [238U], AND (234U/238U) OF MODERN WATER AND PLAYA SAMPLES Sample epytlairetaM (234U/238U) [238U] (ng/mL) [Sr] [Ca] Sr/Ca name (ng/mL) (µg/mL) (mmol/mol) Modern water samples )keerCelgaE(maertS1SW-21IDVS 804.1407.8897.621200.0±0570.0700.0±933.2 )keerCnosremE(maertS2SW-21IDVS 0.0±648.1 734.2227.31790.373700.0±9301.050 )keerCregnarG(maertS3SW-21IDVS 742.1153.42083.664210.0±6374.0600.0±992.2 )keerCpeeD(maertS4SW-21IDVS 953.3036.41834.7013200.0±2180.0600.0±063.2 )gnirpSdyoB(gnirpStoH5SW-21IDVS 663.3301.2674.519600.0±2552.0600.0±630.2 SVDI12-WS6 Hot Spring (Seyferth Hot Springs)* - 0.0045 ± 0.0002 548.531 29.163 8.604 SVDI12-WS7 Hot Spring (Leonard Hot Springs)* - 0.0031 ± 0.0001 236.911 13.973 7.755 SVDI12-WS8 Hot Spring (Surprise Valley Hot Springs)* - 0.0951 ± 0.0028 260.084 19.437 6.120 SVDI12-WS9 Groundwater (Lake City Well) 2.137 ± 0.006 0.3692 ± 0.0105 181.739 29.890 2.781 M(maertS01SW-21IDVS )keerClli 319.2623.21305.879000.0±2620.0800.0±540.2 SVDI12-WS12 Stream (Fort Bidwell Creek) 2.399 ± 0.007 0.1137 ± 0.0035 33.688 7.382 2.087 )ekaLeinnA(ekaL/dnoP31SW-21IDVS 018.1776.7073.039000.0±0630.0500.0±874.1 SVDI12-WS15 Groundwater (Cockeral Ranch Well #1) 2.130 ± 0.006 0.4574 ± 0.0129 51.796 12.764 1.856 SVDI12-WS16 Groundwater (Cockeral Ranch Well #2) 1.508 ± 0.004 0.9190 ± 0.0535 133.410 25.135 2.428 SVDI12-WS17 Hot Spring (Lake City Hot Springs)* - 0.0047 ± 0.0003 400.516 22.214 8.247 SVDI12-WS18 Hot Spring (Unnamed Hot Spring) 1.391 ± 0.004 0.0995 ± 0.0029 3.978 2.093 0.869 )keerCtsoL(maertS91SW-21IDVS 12118.2014861.0±5035.2500.0±048.1 851.2887. SVDI12-WS20 Groundwater (Eagleville Well) 1.797 ± 0.005 0.7240 ± 0.0657 176.296 36.048 2.237 SVDI12-WS21 Groundwater (Cedarville Well) 3.508 ± 0.040 0.9628 ± 0.0815 152.034 29.660 2.345 Modern playa samples SVDI12-C1 Modern playa carbonate (middle playa edge) 1.441 ± 0.005 Playa Pore Waters SVDI12-P1 1:1 DI water extraction (middle playa edge) 1.455 ± 0.005 SVDI12-P2 1:1 DI water extraction (middle playa center) 1.640 ± 0.006 SVDI12-P3 1:1 DI water extraction (upper playa) 1.595 ± 0.006 SVDI12-P4 1:1 DI water extraction (Duck Flats) 1.591 ± 0.006 Playa authigenic carbonate fraction SVDI12-P1 NaOAc carbonate leach (middle playa edge) 1.327 ± 0.004 SVDI12-P2 NaOAc carbonate leach (middle playa center) 1.542 ± 0.006 SVDI12-P3 NaOAc carbonate leach (upper playa) 1.513 ± 0.004 SVDI12-P4 NaOAc carbonate leach (Duck Flats) 1.571 ± 0.004 Note: DI—deionized water. Ac—acetate. *(234U/238U) not measured due to low U concentrations. tions and (234U/238U) within the Surprise Valley Rosholt isochrons and 3-D Osmond isochrons lying data points, all Rosholt and Osmond ages watershed (Table 4). The average measured U for suites of coeval samples. For all suites of are concordant with nearly identical calculated concentrations in surface waters and ground- 5+ coeval samples, calculated Rosholt isochron 230Th-U ages (Fig. 5A). Differences in the age waters from Surprise Valley are 0.430 ± 1.722 ages are reported in Table 5. One important error between the two methods are the result of ng/mL and 0.687 ± 1.722 ng/mL, respectively. assumption is that these isochrons constitute the way in which analytical uncertainty and geo- Hot spring waters have much lower U concen- mixing relationships between the authigenic logic scatter are handled (cf. Ludwig and Titter- trations, ranging from 0.0031 to 0.2552 ng/mL, carbonate end member and a Th end member ington, 1994; Ku, 2000). To be consistent with suggesting that hot springs contribute minimally with a distinct Th isotope composition, either 230Th-U isochron ages reported in recent litera- to the U budget of modern lake and ground- detrital or hydrogenous Th. Hydrogenous 232Th ture (e.g., Hall and Henderson, 2001; Soligo water. Surface-water (234U/238U) measurements, in modern surface waters was less than 6 pg/mL et al., 2002; Garnett et al., 2004; Blard et al., draining from primarily basaltic bedrock, range (see Data Repository methods [see footnote 1]), 2011), hereafter we will refer to the 230Th-U from 1.478 to 2.360. The (234U/238U) values which suggests that detrital Th is the only sub- ages from the Rosholt isochrons. from groundwater and hot spring waters dem- stantial Th isotope end member for Lake Sur- For all 111 subsamples, we calculated single- onstrate a greater range of variability from 1.391 prise carbonates and is likely introduced into the sample detrital Th–corrected ages. For samples to 3.508. Modern playa pore waters, authigenic tufa via incorporation of siliciclastic or organic with robust 2-D Rosholt isochrons (n = 68), we 238 232 carbonate fractions, and the modern playa car- particles containing Th. used the y intercept of the ( U/ Th)M versus 230 232 230 232 bonate all fall within the lower end of the range To calculate the most robust isochron fi t ( Th/ Th)M plot to calculate the ( Th/ Th)0 of observed water sample variability (Table 4). that minimizes the error on the calculated value used in Equation 3 (Torfstein et al., 2013). 230 238 230 ( Th/ U)A and the eventual Th-U age, For the remaining samples (n = 43), an error- Geochronology some subsamples were excluded from the error- weighted mean of the isochron intercepts and weighted linear regressions (Table DR4 [see the modern carbonate (230Th/232Th) measure- 230Th-U Geochronology footnote 1]). For three samples, no robust 2-D ment was used to calculate a basin average 230 232 230 232 The results of the U and Th isotope measure- Rosholt or 3-D Osmond isochrons could be cal- ( Th/ Th)0 [( Th/ Th)0 = 1.327 ± 0.096, ments, and the 230Th-U age calculations are pre- culated. Additionally, two samples produced no error-weighted mean, 2σ] (Table DR5; Fig. DR6 sented in Table 5 and Table DR4 (see footnote 1). robust 3-D Osmond isochron solutions because [see footnote 1]). Because the distribution of 230 230 232 Figure 4 and Figures DR3–DR5 (see footnote 1) the calculated Th-U age error was greater than errors on the ( Th/ Th)0 isochron intercepts show 230Th-U age calculations via pairs of 2-D the corrected 230Th-U age. After excluding out- is nonuniform and varies between samples (Fig.

Geological Society of America Bulletin, Month/Month 2014 11 Geological Society of America Bulletin, published online on 2 June 2014 as doi:10.1130/B31014.1

Ibarra et al. ) 260.0±728.176.2±52.81060.0±587.1730.0±972.0 750.0±175.136.1±96.62450.0±925.1410.0±733.0 98 36 6 3 650.0±706.164.1±44.91450.0±575.1610.0±062.0 670.0±2 7 7 6 12.0 initial 0 0 0.0±616.180.2±48.81780.0±485.1220.0±452.0 3 . . . 0±1 0±7 0±1 U) continued ±710.214.9± ( 238 6 ± 0.017 U/ 39. 85. 38 1 6.168.5 .105. 234 ( 178 148. .0±8 1± 4± ± 58 10 44.6 2 4 8.11212.0±389.155 . . . 6 0 2 (ka cal BP) 1260.0±8 2370.0 16 2 Corrected Age Robust Fit) 2x Rosholt Isochron Derived TSD Method Detrital Corrected Age (Most TSD Method Detrital Corrected 270.0±76 † 5 3.0 U) Th) isochron Th) isochron Th) isochron 238 ± ± § 232 232 232 U/ 4 3 U/ U/ U/ 8 5 0 234 8.10 5 8 5 .1660.0 .1 .1 238 238 238 8 6 Th) vs ( Th) vs ( Th) vs ( 10 10 50. 1 232 232 232 U)* ( .0±962 .0± .0± 0± 238 Th/ Th/ Th/ ± 230 230 230 343. 7 4 0 Th/ (tlohsoRtsuboroN (tlohsoR (tlohsoRtsuboroN113.0±90 72 91 02.0 230 . . . ( 0 0 0 0 ts Activity Ratios (Most Robust Fit) uboroN ------5 5 5 7 7 4 4 5 7 4 0.289 ± 0.014 1.563 ± 0.064 21.93 ± 1.53 1.599 ± 0.066 Samples No. of Sub- TSD Rosholt Isochron Carbonate 650.0±216 670.0±21 460.0±13 740.0±475.141.5±25.524 6 712 280.0±766.126.01±79.62670. 0 6 31. 70 12 initial .0 .0 . 0 0±785. U) ±621. ±909. ±710. 238 U/ 4 6 8 . .1 .1 234 1 . 1 1 2 2 191.8 no 92.8 n rhcosI)hT or 5 0 1 4 6 4 hcos 3 8 1 4 8 7 . . .9 . 5 . . 7±61.8106 5±44. 2±3 1±5 9±8 2±48. ±2 ± i ) 48.1 h 8 8.11 8 1 . . . T 6 0 5 6 0 232 22 29 16 23 227 0157 232 U/ TSD Method Detrital Corrected Age (All TSD Method Detrital Corrected U/ Corrected Age (ka) ( 238 Sub-Samples) 2x Rosholt Isochron Derived 238 2 † 21. 70. 50. 02. 40. 12.0± 0 0 0 .0 .0 .0 Th) vs ( U) Th) vs ( 0 0 0 0 0 0 232 238 ± ± ± ± ± ± ± ± ± 232 U/ 98 6 76 45 46 4 8 403.1362.1±638.0 7 3 Th/ 68.1 75.1 16.1 89.1 35.1 Th/ 234 7 5.1 5 0.2 230 . . 230 1 1 (t l o hs TABLE 5. ISOCHRON DERIVED ACTIVITY RATIOS AND DETRITAL THORIUM CORRECTED AGES THORIUM CORRECTED AND DETRITAL RATIOS ACTIVITY 5. ISOCHRON DERIVED TABLE 40 81 9 45 8 5 5 66 5 oR 20.0±1 90.0±092.0 21.0±063.0 50.0±523.0 51.0±702.0 0.0±4 1 0.0±3 1.0±5 U)* ( . ts 0 238 ubo ± 8 Th/ r 7 7 72. 5 4 2 3 2.0 3 230 oN . . . ( 0 0 0 0 Activity Ratios (All Sub-Samples) TSD Rosholt Isochron Carbonate ------7 5 No Robust Rosholt ( 57 ± 0.051 0.283 ± 0.049 1.561 21.52 ± 4.34 1.597 ± 0.051 7 7 7 7 7 7 7 7 7 Samples 7 + residue 0.306 ± 0.120 ± 0.060 1.611 22.57 ± 9.81 1.651 ± 0.064 5 + residue 0.310 ± 0.050 ± 0.064 1.615 22.87 ± 4.17 1.656 ± 0.067 7 + residue ± 0.015 0.311 1.624 ± 0.016 22.78 ± 1.23 1.666 ± 0.017 7 + residue ± 0.015 0.311 1.624 ± 0.016 22.78 ± 1.23 1.66 No. of Sub- 7 (no residue) ± 0.175 0.321 1.557 ± 0.113 ± 15.16 24.82 ± 0.121 1.598 (no residue) 5 0.335 ± 0.086 1.544 ± 0.181 ± 8.31 26.27 1.586 ± 0.190 7 (no residue) ± 0.035 0.371 1.522 ± 0.135 29.94 ± 4.42 1.568 ± 0.014 7 (no residue) ± 0.035 0.371 1.522 ± 0.135 29.94 ± 4.42 1.568 ± 0.014 3 0 7 2 41T- 5 9 8 81T-11IDVS 11 41T 1 1 3 7T- 9T- 4 1 1 1 1 1 1 3 4 T-2 T T-2 T-21 T-21 T-2 T-21 T-2 T-2 T-21 T-21 T- T- - - 2 2 2 2 11 11 11IDVS 1IDVS 1IDVS 1IDV 1ID 1IDVS 1IDVS 1ID 1IDV 1ID I I I I I I I DVS DVS DVS DVS DVS DVS DVS VS VS VS Lower Lake Shoreline Set Sample name Middle Lake Shoreline Set S SVDI11-T2 Outer Rind SVDI11-T2 SVDI12-T5** Upper Lake Shoreline Set SVDI11-T2 Inner Rind SVDI11-T2 Accomodation Zone Shoreline Set SVDI12-T5** SVDI12-T2** S SVDI12-T2**

12 Geological Society of America Bulletin, Month/Month 2014 Geological Society of America Bulletin, published online on 2 June 2014 as doi:10.1130/B31014.1

Lacustrine paleoclimate record from late Pleistocene Lake Surprise, California egAnobracoidaR 81 42.0±49.41 32.0±74.91 03.0±31.12 5 72.0±42.12 12.0±48.02 2 . . 0±91.51 0±22.12

(ka cal.) 8.58 ± 0.07 10.69 ± 0.11 14.53 ± 0.35 14.53 12.70 ± 0.06 12.70 25.31 ± 0.30 16.00 ± 0.18 19.22 ± 0.23 22.13 ± 0.23 22.13 Corrected Age gvAdethgieW-rorrEnoitcerroCelpmaSelgniS 610.0 920.0±6 800.0±7 810 310.0±755.1 230.0±935.1 110.0±685.1 initial . 0 U) ± ± 238 48 79 U/ 4 5 8.1 7 5.1 4 .1 .1 234 ( 1.504 ± 0.193 1.237 ± 0.003 ) # 74.0±54.71 78.1±48.91 41 21.1±17.91 31.1±03.12 68.0±01.62 8 8 . . continued 2±89.81 0±45.81 11.78 ± 2.8511.78 1.840 ± 0.021 11.41 ± 2.0111.41 1.682 ± 0.247 70.37 ± 1.55 1.196 ± 0.003 18.33 ± 1.82 1.527 ± 0.104 19.80 ± 2.00 16.67 ± 6.57 1.543 ± 0.011 17.28 ± 1.61 1.610 ± 0.011 15.92 ± 6.4415.92 1.128 ± 0.004 27.49 ± 2.87 29.21 ± 8.45 1.197 ± 0.024 24.70 ± 2.25 1.177 ± 0.003 15.41 ± 5.85 1.261 ± 0.009 21.47 ± 0.9421.47 1.609 ± 0.003 24.84 ± 0.92 1.544 ± 0.003 24.91 ± 3.1024.91 1.507 ± 0.039 1 Corrected Age (ka) Age Corrected 290.0±9 040.0±506.1 440.0±1 570.0 430.0±536.1 640.0 620.0±275.1 initial U) ± ± 238 99 548. U/ 2 8 9.1 5 5 . . 234 1 1 1 )noitu 4 7.1 5 8 6 6 1.2±5.62 loSD- . .3 . . . 2±8 2±1 3±5 4±6 ±9 ± 5.61 .02 . . . . 3d 7 91 81 6 1 2 n o msO E 4 09 5 5 430.0±306.11 930.0±275.1440.0±062.0 420.0±035.13 LM(e 4 4 7 0 0 0 0 U) Corrected Age (ka) ( . . . . 0±308 0±1 0±6 0±5 238 gA U/ d 8 8 5 234 e 5 8 5 tc . . . . 15 1 1 18 err oCl a t 93 120 ir 3 4 2 3 t 0.0±4 0. 0. 0.0±7 0.0±533 U) ( eDD . 0 0±86 0 238 ± ± Th/ 37 24 S Th). 7 4 T 230 Th). 2 2 2 2 3 ( . .0 . . . . 232 TABLE 5. ISOCHRON DERIVED ACTIVITY RATIOS AND DETRITAL THORIUM CORRECTED AGES ( THORIUM CORRECTED AND DETRITAL RATIOS ACTIVITY 5. ISOCHRON DERIVED TABLE 232 0 0 0 0 0 Th/ U/ 230 234 ------Th) vs ( 4 0.285 ± 0.026 1.560 ± 0.029 21.7 ± 2.2 ± 0.031 1.595 7 5 5 5 7 4 4 Th) vs ( 232 232 Samples U/ U/ 7 + residue 0.310 ± 0.057 ± 0.055 1.600 ± 5.2 23.2 1.640 ± 0.054 7 + residue 0.312 ± 0.014 1.624 ± 0.019 23.0 ± 1.3 1.665 ± 0.018 No. of Sub- No Robust Osmond Isochron No Robust Osmond Isochron No robust Osmond Isochron No robust Osmond Isochron 7 (no residue) ± 0.088 0.326 ± 0.130 1.520 ± 8.4 26.0 ± 0.140 1.560 7 (no residue) 0.363 ± 0.039 1.530 ± 0.150 29.0 ± 3.9 1.580 ± 0.160 238 238 3 0 71T- 41T- 51T- 8 9 11 1 1 1 7T- 1 3T-21IDVS 1 4T- 4T 3T T-21 T-2 T-2 T-21 T-21 T-21 - - 2 2 2 2 2 1-T14 11IDVS 11IDVS 1IDVS 1ID 1IDV 1I 1IDVS 1IDVS 1IDVS I I I I DVS D DVS DVS DVS Slope of Rosholt Isochron ( t Rosholt isochrons. DR4 (text footnote 1) for samples omitted from robust fi Table See DR4 (text footnotes 1) for samples omitted from error-weighted mean. Table See **Rosholt and Osmond isochron ages calculated with without residue. † § # *Slope of Rosholt Isochron ( VS VS SVDI12-T9 S Middle Lake Shoreline Set SVDI12-T5** SVDI11-T2 Outer RindSVDI11-T2 No Robust Osmond Isochron Sample name Upper Lake Shoreline Set SVDI12-T5** Accomodation Zone Shoreline Set SVDI12-T12 Lower Lake Shoreline Set SVDI12-T2** SVDI11-T18 SVDI12-T2** SVDI11-T2 Inner Rind SVDI11-T2 SVDI1

Geological Society of America Bulletin, Month/Month 2014 13 Geological Society of America Bulletin, published online on 2 June 2014 as doi:10.1130/B31014.1

Ibarra et al.

4 230 238 12 234 238 Slope = 0.311 ± 0.015 (2σ) = ( Th/ U)authigenic Slope = 1.624 ± 0.016 (2σ) = ( U/ U)authigenic 230 232 Intercept = 0.904 ± 0.096 = ( Th/ Th)initial Intercept = -0.829 ± 0.080 MSWD = 3.7 MSWD = 0.49 10 3 Rosholt Isochron Age: 22.8 ± 1.2 (234U/238U) = 1.666 ± 0.017 initial 8 Th) Th) 232

2 232 6 Th/ U/ 230 234 ( ( 4

1 Residue

2 Residue

A - “Rosholt” Isochron B - “Rosholt” Isochron 0 0 0 2 4 6 8 0 2 4 6 8 (238U/232Th) (238U/232Th)

Rosholt Isochrons 1.8 230Th-U Age = 23.0 ± 1.3 ka 234 238 Measured Isotope Ratios (2σ) ( U/ U)initial = 1.665 ± 0.018 MSWD = 17 Error-weighted Regression 1.6 Error Envelope (2σ) U)

Osmond 3-D Isochron 238 1.4 U/

Measured Isotope Ratios (2σ) 234 ( Residue Projected Data Points (2σ) 1.2 Isochron Intercept (2σ) C - “Osmond” 3-D Isochron Isochron 1.0 0.0 0.2 0.4 0.6 0.8 (230Th/238U)

Figure 4. Example of Rosholt and Osmond isochron plots calculating 230Th-U ages (in ka) for SVDI12-T2. All error ellipses and error enve- lopes are 2σ. (A–B) Rosholt isochrons of (230Th/232Th) vs. (238U/232Th) and (234U/232Th) vs. (238U/232Th) (measured values), with error-weighted 230 238 234 238 linear regressions (see text for details), calculate the ( Th/ U)authigenic and ( U/ U)authigenic, based on the slopes of the linear regressions. The 230 232 238 232 230 232 intercept of ( Th/ Th) vs. ( U/ Th) is the ( Th/ Th)initial value of the suite of coeval samples. Samples not included in linear regressions are gray ellipses. (C) Projected representation of a three-dimensional (3-D) Osmond isochron calculation of 230Th-U ages for coeval suites of samples (see text for details) onto a 230Th-234U-238U evolution plot. Black ellipses are measured values, and white ellipses are projected values 230 234 238 based on the projected isochron (gray line). Gray ellipses are the calculated Th-U age (95% confi dence) and ( U/ U)0 of the sample. MSWD—mean square of weighted deviates.

DR6 [see footnote 1]), we chose to use an error- ple detrital Th–corrected ages for suites of sub- with robust isochrons, all samples are concor- weighted mean and standard deviation, rather samples (n = 2–7) were calculated using Isoplot dant or nearly concordant (Fig. 5A). On aver- than an unweighted mean, to represent the 3.75 (Ludwig, 2012). age, the error-weighted mean single-sample 230 232 basin average ( Th/ Th)0 for samples without A comparison of the error-weighted means of ages are 0.17 k.y. younger than the Rosholt ages, robust isochrons or a suffi cient number of sub- the single-sample detrital Th–corrected 230Th-U with no apparent trend with elevation (Fig. 5B). samples. The calculated value is similar to other ages to the Rosholt isochron 230Th-U ages (con- One outlier is SVDI12-T1, which has a much Great Basin paleolakes (Lin et al., 1996, 1998; sisting of samples from the two lower-elevation older Rosholt isochron age (26.44 ± 5.86 ka Fig. DR6 [see footnote 1]). All age calculations groups) indicates the utility of the two comple- compared to 18.54 ± 0.88 ka). The older age is 230 232 and the error-weighted means of the single-sam- mentary age calculation methods. For samples a result of a low ( Th/ Th)0 isochron intercept

14 Geological Society of America Bulletin, Month/Month 2014 Geological Society of America Bulletin, published online on 2 June 2014 as doi:10.1130/B31014.1

Lacustrine paleoclimate record from late Pleistocene Lake Surprise, California

32 AB32

28 28

24 24

20 20

16 16 Group 1 Group 2 Th-U “Rosholt” Isochron Age (ka) Th-U “Rosholt” Isochron Age (ka) 12 Group 3 12 230 230 1:1 Line Rejected 8 8 8 121620242832 8 121620242832 230Th-U 3-D “Osmond” Isochron Age (ka) 230Th-U SS Age (ka)

32 CD32

28 28

24 24

20 20

16 Th-U SS Age (ka BP) 16 230

Th-U “Rosholt” Isochron Age (ka) 12 12 230

8 8 8 12 16 20 24 8 12 16 20 24 14C Age (ka) 14C Age (ka) Figure 5. Comparison among 230Th-U age calculations, and between paired calibrated radiocarbon (14C) ages and 230Th-U ages. Samples are grouped by the same symbology as in Figures 2 and 3. Samples rejected due to apparent open-system behavior have gray error ellipses (see text for details). All error ellipses are 2σ. (A) Rosholt isochron-based 230Th-U ages compared to three-dimensional (3-D) Osmond isochron- based 230Th-U ages. (B) Rosholt isochron-based 230Th-U ages compared to error-weighted averages of single-sample (SS) detrital-corrected 230Th-U ages. (C) Rosholt isochron-based 230Th-U ages compared to radiocarbon ages. (D) Error-weighted averages of single-sample (SS) detrital-corrected 230Th-U ages compared to radiocarbon ages.

238 232 and steeper slope in the ( U/ Th)M versus bonate, and playa samples and form three groups the 15 samples have radiocarbon ages that inter- 230 232 ( Th/ Th)M isochron. Due to this anoma- (Fig. 6): (1) the lower-elevation samples of ca. sect the IntCal13 calibration curve (Reimer 230 232 234 238 lously low ( Th/ Th)0 value, the basin average 27–18 ka, which cluster around a ( U/ U)0 et al., 2013) at multiple ages, resulting in two 230 232 ( Th/ Th)0 was used to calculate the single- of 1.497–1.682; (2) the middle-elevation, post- or more statistically probable calibrated ages. 234 238 sample ages for SVDI12-T1. SVDI12-T2, from LGM samples with elevated ( U/ U)0 from For all samples, the ages discussed hereafter are the same elevation (1419.5 m), yielded nearly 1.784 to 1.846; and (3) the highest-elevation the calibrated average ages of the tufa samples 234 238 concordant ages of 22.78 ± 1.23 ka (Rosholt samples, with ( U/ U)0 between 1.128 and taken from the age range with the highest rela- age) and 24.84 ± 0.92 ka (single-sample age). 1.261, below the minimum of modern values tive probability based on the area under the cali- 234 238 The calculated ( U/ U)0 values of the tufa (Figs. 2B–2E and 6). brated age probability distribution. For all but samples (Table 5; Table DR4 [see footnote 1]), one sample (SVDI12-T5-b), this area was >88% which should refl ect the (234U/238U) value of the Radiocarbon Geochronology of the total area under the curve (Table 6). water precipitating carbonate, demonstrate less The radiocarbon geochronology results and A hydrograph using only the radiocarbon variation in (234U/238U) than modern water, car- calibrations are presented in Table 6. Two of ages indicates that 12 of the 15 radiocarbon ages

Geological Society of America Bulletin, Month/Month 2014 15 Geological Society of America Bulletin, published online on 2 June 2014 as doi:10.1130/B31014.1

Ibarra et al.

2.6 Group 1 “Rosholt” Isochron Age AB234 238 and ( / U)initial Group 2 “Rosholt” Isochron Age and (234/238U) initial 230 Range of Single Sample Detrital Th Figure 6. Comparison of Th-U age vs. 234 238 2.2 234 238 ( U/ U) calculated using Equation 2 and Corrected Ages and ( U/ U)initial 0 the measured modern values of (234U/238U). Modern Water and

Playa ( Only sample groups 1 and 2 are plotted.

initial 234 238 (A) Rosholt isochron-derived ( U/ U)0 U)

234 are shown as ellipses, with the range val- 238

1.8 U/

U/ ues for the single-sample detrital Th–cor- 238 234 (

U) rected samples shaded in gray. All errors are 2σ. (B) Box and whisker plots show the median and interquartile ranges of modern (234U/238U) and the rejected single-sample 1.4 234 238 Rejected Sample Minimum modern ( U/ U) 234 238 ( ages’ ( U/ U)0 from the group 3 samples. 234 Group 3 U/ 238 U) initial 1.0 10 14 18 22 26 30 34 Age (ka)

TABLE 6. NEW RADIOCARBON AGES FOR LAKE SURPRISE Sample name Laboratory Altitude 14C age Calibrated Relative area Median Calibrated age number (m) (yr ± 1σ) age range under distribution age (ka cal. ± 2σ)*,§ (cal. yr ± 2σ)* (cal. yr)† IntCal13 Accommodation zone shoreline set SVDI11-T2-1 inner rind Beta –342012 1453.5 15,930 ± 70 18,988–19,457 (1.000) 19,212 19.22 ± 0.23 SVDI11-T3-2 Beta –342013 1437.7 17,580 ± 70 20,978–21,511 (1.000) 21,247 21.24 ± 0.27 SVDI11-T4-1b Beta –340102 1430.6 17,280 ± 60 20,631–21,052 (1.000) 20,836 20.84 ± 0.21 SVDI11-T14-1c Beta –342014 1478.4 10,790 ± 50 12,644–12,758 (1.000) 12,708 12.70 ± 0.06 SVDI11-T18-1c Beta –340103 1555.7 13,310 ± 40 15,820–16,187 (1.000) 16,010 16.00 ± 0.18 Middle lake shoreline set SVDI12-T1-a Beta –340104 1419.5 1,7560 ± 60 20,971–21,464 (1.000) 21,217 21.22 ± 0.25 SVDI12-T2-b Beta –342015 1419.5 18,270 ± 70 21,899–22,355 (1.000) 22,140 22.13 ± 0.23 SVDI12-T5-b Beta –342016 1444.3 9470 ± 40 10,581–10,794 (0.888) 10,713 10.69 ± 0.11 )100.0(758,01–558,01 00.0±68.01 )250.0(700,11–369,01 20.0±99.01 )950.0(560,11–120,11 20.0±40.11 SVDI12-T9-1 Beta –340105 1508.9 12,420 ± 50 14,183–14,875 (1.000) 14,515 14.53 ± 0.35 SVDI12-T10-b Beta –342017 1516.8 12,600 ± 50 14,693–15,178 (1.000) 14,964 14.94 ± 0.24 SVDI12-T12-1 Beta –340106 1576.9 7810 ± 40 8459–8498 (0.045) 8587 8.48 ± 0.02 8509–8656 (0.929) 8.58 ± 0.07 8968–1768 )920.0( 10.0±86.8 Lower lake shoreline set SVDI12-T13 Beta –342018 1437.2 17,490 ± 90 20,836–21,432 (1.000) 21,127 21.13 ± 0.30 SVDI12-T14 Beta –342019 1530.7 12,750 ± 50 15,006–15,364 (1.000) 15,192 15.19 ± 0.18 Upper lake shoreline set SVDI12-T15-b Beta –342020 1433.1 16,150 ± 70 19,248–19,699 (1.000) 19,491 19.47 ± 0.23 SVDI12-T18 Beta –342021 1564.2 20,970 ± 110 25,016–25,611 (1.000) 25,322 25.31 ± 0.30 *Calibrated using the Calib 7.0 program with IntCal13 (Stuiver and Reimer, 1993; Reimer et al., 2013). †Median age calculated using the Calib 7.0 program. §Bolded ages are the preferred ages with the highest probability and are used for comparison to the 230Th-U ages (Table 5; Fig. 5).

16 Geological Society of America Bulletin, Month/Month 2014 Geological Society of America Bulletin, published online on 2 June 2014 as doi:10.1130/B31014.1

Lacustrine paleoclimate record from late Pleistocene Lake Surprise, California

YD B-A HS1 LGM MIS 1 / Holocene MIS 2 MIS 3 A Basin Pour Point

1600

Cooks Canyon subaerial sediments Cooks Canyon lacustrine sediments (Personius et al., 2009) (Personius et al., 2009) 1500

Elevation (m) Trego Hot Spring tephra moderate to deep water Playa Prehistoric Settlements deposition (Hedel, 1980; 1984) 1400 (~0.5 to 6 ka)

Modern Playa Minimum

Calculated Precipitation Increase 80 B (Isotope Model) 14C Ages “Rosholt” 230Th-U Ages 60 SS Detrital Th Corrected Ages No Fill - Rejected Datapoint 40 14C Derived Hydrograph Age Envelope

20

Precipitation (% Modern) 0

0 5 10 15 20 25 30 35 Age (ka) Figure 7. (A) Hydrograph for the latest pluvial cycle of Lake Surprise. The gray error envelope bounds all sample ages with elevations <1531 m; the gray line relies on radiocarbon ages only. Dashed lines are tentative correlations. Other constraints include moderate- to deep-water deposition of the Trego Hot Spring tephra at 1378 m (Hedel, 1980, 1984), Cooks Canyon deltaic lacustrine and subaerial sediments deposited at 1493 m (Personius et al., 2009), and playa prehistoric settlements (O’Connell and Inoway, 1994). See Figure 2A for the location of the additional lake- level constraints. (B) Calculated precipitation increase (% modern) from the isotope mass balance model. See Tables 7 and 8 for assumptions and calculations. MIS—marine oxygen isotope stage; YD—; B-A—Bølling- Allerød; HS1—Heinrich Stadial 1; LGM—Last Glacial Maximum. record the transgression and regression of Lake localities, which have discordant ages of 8.58 radiocarbon dates are unlikely to be affected Surprise (Fig. 7A). Two samples from the low- ± 0.07, 16.00 ± 0.16, and 25.31 ± 0.30 ka cal. by dead carbon incorporation. For the samples est shoreline of the middle playa set are not con- Of these three samples, the lowest-elevation dated in this study, with ages three to fi ve times cordant but have very similar ages of 21.22 ± sample recorded an age of 16.00 ± 0.16 ka cal., the radiocarbon half-life of 5.73 k.y., modern 0.25 and 22.13 ± 0.23 ka cal. Similarly, two which is potentially compatible with lower-ele- contamination of 5% can impact the calculated nearly identical-elevation samples from differ- vation samples. age by as much as 7 k.y. (cf. Cassata et al., 2010, ent sampling localities have concordant ages fi g. 4). Open-system behavior in the uranium of 21.24 ± 0.27 and 20.13 ± 0.03 ka cal. With Comparison of Radiocarbon and 230Th-U system, typically a result of U-loss due to the the exception of one anomalously young age, Geochronology preferential solubility of U relative to Th, results the lowest-elevation samples record transgres- Other studies attempting to reconcile radio- in an anomalously high (230Th/238U) and an older sion from 1419.5 to 1453.5 m and have LGM carbon and 230Th-U ages in lacustrine carbon- apparent age (cf. Kurth et al., 2011). ages ranging from 19.22 ± 0.23 to 22.13 ± 0.23 ates, particularly tufa samples, have considered Eleven group 1 and 2 samples dated using ka cal. Four tufa samples at middle elevations the limitations of both methods (Lin et al., 1996; both radiocarbon and 230Th-U demonstrate (1530.7–1478.4 m) and the anomalous low- Placzek et al., 2006a; Kurth et al., 2011; Zim- acceptable agreement (Figs. 5C and 5D). Seven elevation sample (1444.3 m) record consistent merman et al., 2012). The Surprise Valley water- of the 11 sample ages (Rosholt 230Th-U ages) are gradual regression from an apparent highstand shed, unlike many other watersheds containing concordant between the two methods, and the at 15.19 ± 0.18 to 10.69 ± 0.11 ka cal. Out- late Pleistocene lakes, drains primarily basaltic remaining four 230Th-U ages are on average 3.19 liers to this apparent lake cycle are three of the bedrock and contains no substantial carbon- k.y. older (Fig. 5C). Previous studies on other highest-elevation samples at different sample ate sediments (Egger and Miller, 2011). Thus, paleolakes have also observed that a majority of

Geological Society of America Bulletin, Month/Month 2014 17 Geological Society of America Bulletin, published online on 2 June 2014 as doi:10.1130/B31014.1

Ibarra et al. samples dated using both methods, while falling pluvial lake cycles; however, both samples (28.08 ± 4.44 ka), as originally suspected by the 234 238 close to the 1:1 concordant line, have slightly recorded ( U/ U)0 less than the range of mod- authors, due to potential contamination. How- older 230Th-U ages (e.g., Kaufman and Broecker, ern water and playa samples. A single-sample ever, the younger age of 18.40 ± 3.17 ka of the 1965; Placzek et al., 2006a; Blard et al., 2011). age from a group 3 sample at 1542.3 m yielded deltaic sediments, using a weighted mean of the Given these constraints, for samples dated by an early MIS 4 age of 70.37 ± 1.55 ka. Addi- optically stimulated luminescence (OSL) and both methods, we consider the radiocarbon ages tionally, one subsample of SVDI12-T18b at IRSL ages (Personius et al., 2009), agrees well with discordant 230Th-U ages as minimum ages 1564.2 m recorded an early MIS 5 age of 115.92 with the constructed lake-level curve. Finally, and the 230Th-U ages as maximum ages. ± 6.44 ka. This age is questionable, as the other three of four subaerial ages from Personius et al. The only signifi cant outlier from the lower- subsample yields an incalculable age, presum- (2009) are much younger than the proposed lake elevation group 1 and 2 samples is SVDI12-T5. ably due to U loss. Uranium loss is evidenced level at that elevation (1475 m); however, the This tufa sample, formed on vesicular basalt by a much lower U concentration relative to the oldest subaerial age, 13.38 ± 2.64 ka, is within bedrock, has a signifi cantly younger radiocarbon 115.92 ± 6.44 ka subsample (417 vs. 2544 ppb), error of the plotted radiocarbon- and 230Th-U– age of 10.69 ± 0.11 ka cal., compared to concor- and an anomalously high (230Th/238U) of 2.277. based hydrographs during the deglacial lake- dant 230Th-U ages of 22.87 ± 4.17 ka (Rosholt This open-system behavior is further refl ected in level recession. All other previous constraints on age) and 24.91 ± 3.10 ka (single-sample age). a discordant radiocarbon age of 25.31 ± 0.30 ka Lake Surprise lake level are consistent with the This suggests either an insuffi cient detrital Th cal. While inconclusive, these two samples may constructed hydrograph plotted in Figure 7A. correction or modern contamination of the record a higher, older lake cycle(s) in Surprise carbonate sample. The only concordant group Valley, similar to those found in other Basin and DISCUSSION 3 sample age is the single-sample average of Range lake systems, corresponding to MIS 4 or SVDI11-T18, with a radiocarbon age of 16.00 ± MIS 6 (Reheis, 1999b; Kurth et al., 2011). In the following discussion, we examine the 0.18 ka cal. and a 230Th-U age of 15.41 ± 5.85 ka; new Lake Surprise hydrograph of Figure 7A however, the single-sample ages range from 6.30 Lake Surprise Hydrograph and explore the implications for western United to 20.74 ka. We reject both the radiocarbon and States during the late Pleisto- 230Th-U ages from this sample based on the δ18O, To construct the Lake Surprise hydrograph, cene. Our key results are: (1) The Lake Surprise δ13 234 238 C, Sr/Ca, and ( U/ U)0 values, which col- we used only ages obtained from lower and watershed remained a closed, inward-draining lectively suggest open-system behavior due to middle elevations and plotted calibrated radio- basin throughout the last glacial cycle with recrystallization or pedogenic overprinting. carbon ages, single-sample 230Th-U ages, and lateral correspondence found among samples Taking the 11 accepted radiocarbon ages as Rosholt isochron 230Th-U ages against the from the four shoreline localities based on topo- the “true” age of the samples also allows for a LiDAR-derived sample elevations (Fig. 7A). graphic analysis, δ18O-δ13C-Sr/Ca covariance, comparison between the ages calculated using We connected radiocarbon ages assuming that and concordant geochronology; (2) the new lake isochron and single-sample methods (Fig. 5B). a given sample records the minimum lake-water hydrograph is in agreement with previous lake- Single-sample 230Th-U ages are on average 0.80 level (e.g., Felton et al., 2006; Benson et al., level constraints and places the highest lake level k.y. older than the radiocarbon ages and are 1996). For samples of statistically equivalent ~176 m above modern playa at 15.19 ± 0.18 ka closer than the Rosholt 230Th-U isochron ages, ages, the sample at higher elevation was taken cal. and during the LGM ~80 m above modern which are 1.18 k.y. older. The single-sample as the likely lake level, rather than recording playa; and (3) multiple lines of evidence reveal detrital correction method only corrects the higher-frequency fl uctuations. In addition to that samples from the highest shorelines are (230Th/238U) ratio (using Eq. 3), whereas Rosholt the radiocarbon hydrograph, we constructed an likely from older, higher lake cycles and were 230Th-U isochrons also correct the (234U/238U) age envelope of all available ages from samples infl uenced by variable amounts of open-system of the authigenic carbonate (cf. Ku, 2000). <1531 m (Fig. 7A). Based on the radiogenic exchange or pedogenic overprinting. Using these Although the correction of the (234U/238U) of and stable isotopic evidence presented here, we results, we modify an existing δ18O isotope mass the authigenic carbonate is negligible at the have concluded that Lake Surprise lake levels balance model by including basin geometry to young age range of these samples, we consider reached 1531 m (a water depth of 176 m) at assess hydrologic controls on Lake Surprise lake the Rosholt 230Th-U isochron ages as the most 15.19 ± 0.18 ka cal. (radiocarbon age) during levels and compare the calculated changes in accurate 230Th-U ages of the samples, because the latest Pleistocene (Fig. 7A). This highstand precipitation (Fig. 7B) to climate models from this method corrects both the U and Th ratios age is supported by a concordant, but signifi - the PMIP3 climate model ensemble (Fig. 8). of the authigenic carbonate for detritus. Addi- cantly older, 230Th-U isochron age of 18.25 ± tionally, depending on the (232Th/238U) ratio of 2.67 ka (Rosholt age). At higher elevations, one Hydrologic Controls on Lake Surprise a given subsample, the age error of the single- sample (SVDI11-T18) suggests that Lake Sur- Lake Levels sample error-weighted 230Th-U ages will vary prise may have briefl y reached 1555 m at 16.00 signifi cantly. Because we observe concor- ± 0.18 ka cal. (radiocarbon age), although stable A description of the changes in atmospheric 234 238 dance between the Rosholt isochron and error- isotope, Sr/Ca, and ( U/ U)0 evidence sug- circulation patterns and the seasonal role of weighted averages of the single-sample 230Th-U gests that this age may be in error due to calcite solar insolation in regional climate during the ages (Fig. 5B), we consider the single-sample recrystallization causing open-system behavior. last deglaciation requires quantitative predic- 230Th-U ages as valid supporting ages in the Our results correlate well with the limited tions of meteorologic and hydrologic processes. hydrograph construction. previous studies (Table DR1 [see footnote 1]) We incorporate the Lake Surprise hydrograph, and help to clarify some ambiguous relation- the stable isotope analyses, and the topographic Older Pre-LGM Ages from High Shorelines ships. An infrared-stimulated luminescence analysis of the basin geometry into two models Two samples from the highest shoreline (IRSL) age on feldspar grains constraining the that enable us to estimate past changes in pre- elevations and the Upper Lake shoreline set older age range of the deltaic deposits from cipitation and evaporation: hydrologic index (Fig. 2) record ages that may be from past Personius et al. (2009) appears to be too old and isotope mass balance.

18 Geological Society of America Bulletin, Month/Month 2014 Geological Society of America Bulletin, published online on 2 June 2014 as doi:10.1130/B31014.1

Lacustrine paleoclimate record from late Pleistocene Lake Surprise, California

0 MIS 6) highstand (1567 m) (Table 2). Modern ABweather station data ([P – ET ]/[E – P ]) and 100 T T L L modern maximum lake levels (AL/AT) indicate -10 that the modern HI equals 0.11 and 0.12, respec- tively. Quantifying these variables in the past, 80 particularly tributary evapotranspiration (ETT), requires calibration using modern meteorologi- -20 cal observations (Miffl in and Wheat, 1979). Dur- 60 ing the LGM stillstands, HI was almost 300% of modern, with values of 0.29–0.34. At the post- -30 LGM deglacial highstand, the HI was 0.56. The 40 calculated HI values are consistent with other nearby lake systems evaluated by Miffl in and Wheat (1979) in south-central Oregon and north- -40 20 western Nevada and provide a useful framework with which to constrain changes in precipitation,

Precipitation (% Modern) evaporation, and temperature during the LGM.

Lake Evaporation (% Modern) -50 0 Isotope Mass Balance Calculations The stable isotope mass balance provides a complementary method to the HI calculations -20 -60 for determining basin average changes in pre- cipitation. We modifi ed the isotope mass bal- ance model of Jones et al. (2007), which was developed for similar steady-state midlatitude lake systems in Turkey (Jones et al., 2007; Jones Surprise Valley Surprise Valley and Imbers, 2010). We then incorporated the PMIP 3 Ensemble (Isotope Model)Surprise Valley (Isotope Model) Lake Lahontan Lake Lahontan basin geometry (i.e., lake surface area and tribu- Mifflin and Wheat (1979)Lake LahontanGreat Basin Mifflin and Wheat (1979)Lake LahontanGreat Basin Hostetler and Benson (1990) (Total Evaporation) Hostetler and Benson (1990) tary area), derived from the ArcGIS analysis, Matsubara and Howard (2009) Matsubara and Howard (2009) PMIP 3 Ensemble Surprise Valley to calculate changes in precipitation, relative to modern, at the LGM and post-LGM highstand. Figure 8. Last Glacial Maximum (LGM) isotope mass balance–calculated precipitation The time-varying (t) water balance and iso- (A) and lake evaporation (B), and the PMIP3 climate model ensemble predictions for Lake topic mass balance for a lake, assuming ground- Surprise. The values derived by this study are compared to literature values derived using a water fl ux across the sediment-water interface is mass balance model (Miffl in and Wheat, 1979), a thermal evaporation model (Hostetler and negligible, can be described by two equations. Benson, 1990) for Lake Lahontan, and a hydrologic model (Matsubara and Howard, 2009) The fi rst describes the change in lake volume for the Great Basin. Individual climate model predictions are presented as black dots and a (V ; modifi ed from Jones et al., 2007; Jones and box-and-whisker plot (Table DR6 [see text footnote 1]). PMIP3 climate model output is total L Imbers, 2010; Steinman et al., 2013): evaporation, calculated from the surface latent heat fl ux. Isotope mass balance calculations from all LGM samples are presented as a box-and-whisker plot (Table 8). Lake evapora- dV L =+−QQQ, (5) tion is calculated using the equations of Linacre (1992) and Jones et al. (2007). See Figure dt pre DR7 (text footnote 1) for the isotope calculation’s sensitivity to relative humidity, assumed change in temperature, average wind speed, and input δ18O (runoff and precipitation). The where Q is the input and output fl uxes, and sub- black and white bars indicate the range of predicted values by the literature. scripts correspond to on-lake precipitation (p), runoff (r), and lake surface evaporation (e). The isotopic mass balance is similarly described as Hydrologic Index Calculations A A R PET− (modifi ed from Benson and Paillet, 2002; Jones HI ==L L = T = TT, (4) The “pluvial hydrologic index” (HI), origi- − − − et al., 2007; Jones and Imbers, 2010; Doebbert AT AABLEPLLEPLL nally derived by Miffl in and Wheat (1979), is a et al., 2010; Steinman et al., 2013): measure of the ratio of lake surface area to tribu- δ18 × dOV()LL 18 tary area given an equilibrium (steady-state) where AL, AB, and AT are the areas of the lake, =×+()δ OQ dt pp lake level, whereby moisture into the lake equals basin, and tributary (AT = AB – AL), P is the aver- δδ18 ×−18 × moisture out of the lake. It has been used in vari- age on-lake (subscript L) and tributary (subscript ()(),OQrr OQ ee (6) ous forms (e.g., Hostetler and Benson, 1990; T) precipitation, EL is the gross lake evaporation, δ18 Reheis, 1999b) to solve for the ratio of runoff to ETT is the average tributary evapotranspiration, where subscripts also apply to the O of the evaporation using paleohydrologic constraints and RT is the combined surface and groundwater input and output fl uxes. We simplifi ed Equation derived from related (e.g., wave-cut runoff, where RT = PT – ETT. We computed the 5 by combining precipitation and runoff into the shorelines, , strandlines). Based only HI from AL/AT (1) the range of LGM stillstands total surface and subsurface inputs (Qw) multi- δ18 on mass balance, the dimensionless hydrologic (1420–1440 m), (2) the post-LGM highstand plied by the isotopic composition, Ow. Apply- index (HI) is (Miffl in and Wheat, 1979): (1531 m), and (3) the pre–MIS 2 (possibly ing the chain rule to Equation 6 yields:

Geological Society of America Bulletin, Month/Month 2014 19 Geological Society of America Bulletin, published online on 2 June 2014 as doi:10.1130/B31014.1

Ibarra et al.

dOδ18 dV E using the equations of Benson and White Despite increasing the potential unknowns, L +=δ18 L L VL OL δ18 dt dt (1994) for Oe and the evaporation model of the isotope model allows for an explicit solu- E δδ18 ×−18 × Linacre (1992) for L (see Table 7 for details). tion to a steady-state isotope mass balance for ()()OQww OQ ee. (7) δ18 The value for OL was derived from the mea- the lake input fl uxes (Eq. 9) while also consid-

sured tufa samples (Table 3) and converted ering basin geometry (AL and AT). Additionally, By substituting Equation 5 into Equation 7 using the temperature-dependent water-calcite we can account for temperature and evaporation and rearranging, we fi nd an expression for time- fractionation factor (Table 7; Kim and O’Neil, processes while solving for the basin average δ18 varying changes in OL: 1997). Given the calculation for the total water precipitation. input fl ux, Q , P was calculated according δ18 w L dOL 18 18 V =−×−((δδOOQ ) ) to mass balance (Qw = [PL × AL] + [RT × AT]) Calculating Late Pleistocene Precipitation L dt wLw assuming that RT is 0.15 × PL. Integrated water- Amounts for Lake Surprise δδ18−× 18 ((OOQeLe ) ). (8) shed modeling of the Lake Estancia basin, New To calculate precipitation changes during the Mexico, during the LGM, assuming a colder LGM and the post-LGM highstand, we applied

Since Qe = ALEL, and assuming that for a and wetter climate, suggests that annual runoff the hydrologic index (Eq. 4) and our isotope given lake level, recorded in the isotopic com- (RT) was 15% of precipitation (PL) (Menking et mass balance model (Eqs. 5–9). Our isotope δ18 position and altitude of a tufa sample, d OL/dt al., 2004). No stream gauges have been histori- mass balance model also incorporates lake

= 0, solving for Qw yields an expression for the cally recorded in Surprise Valley, but analysis surface area within a basin of fi xed total area, volumetric fl ux of moisture into the lake: of annual stream discharge data (1949–2012) thereby decreasing tributary area with increas- from the nearest gauge maintained by the USGS ing lake surface area and sample altitude (Fig. AE×−()δδ18 O 18 O Q = LL e L. (9) on the South Fork of the Pit River, indicates that DR1 [see footnote 1]). The key parameters used w δδ18− 18 ()OOwL the modern runoff coeffi cient is 16.9% ± 2.6% to calculate changes in precipitation are: tem- σ (2 , n = 56 yr; http:// waterdata .usgs .gov /usa perature (T), lake hypsometry (AL), watershed δ18 δ18 δ18 Using the methods and assumptions outlined /nwis/rt; see Data Repository [footnote 1] for geometry (AT), OL (from Ocalcite), Oe, δ18 δ18 in Jones et al. (2007), we calculated Oe and additional details). Ow, relative humidity (Hr), and lake surface

evaporation rate (EL) (see Table 7 for all val- ues and related equations). We assumed that TABLE 7. ISOTOPE MODEL PARAMETER VALUES local temperatures were uniformly 7 °C lower noitpircseDsretemaraP eulaV during the LGM and 5 °C lower during the T Temperature (°C) Modern = 9.2, LGM = 2.2, post-LGM = 4.2, based on deglacial period, based on pollen assemblages Worona and Whitlock (1995) P Precipitation (L, lake) (mm/yr) Modern = 566 mm/yr (solved by model) at nearby Little Lake, Oregon (Fig. 1; Worona L and Whitlock, 1995). We assumed that tem- RT Runoff (surface + subsurface) Unknown (assumed 0.15 × PT , based on Menking et al., 2004) (mm/yr) perature changes were uniform among seasons H Relative humidity Unknown (modern = 0.57%) r δ18 (Jones et al., 2007). The OL was calculated EL Lake surface evaporation rate Linacre (1992) based on T, latitude (41.5°N), lake surface (mm/yr) altitude, and wind speed (modern = 1.9 m/s)* for each sample using the calcite-water frac- α tionation, w-c, of Kim and O’Neil (1997). The A Lake area (km2) Varies with lake surface level, based on modern topography, L δ18 calculated using fi fth-order polynomial fi t through results of Ow was assumed to be the average of mod- † δ18 − DEM analysis ern creek waters ( OL = 14.57‰ relative to 2 2 AB Basin area (km ) 3812 km based on modern topography (DEM analysis) Vienna Standard Mean Ocean Water [VSMOW]; A Tributary area (km2) A = A – A T T B L Table DR3 [see footnote 1]). Calculation of δ18O Lake water δ18O Calculated using α at given T from tufa sample L w-c δ18 measurements Oe assumed (1) a kinetic fractionation factor, δ18O Input fl uxes (precipitation + –14.57‰ (VSMOW, modern creek average; Table DR3 [see α α w kin, for wind speeds <6.8 m/s of kin = 0.994 runoff) δ18O text footnote 1]) δ18 δ18 (modern average wind speeds are 1.9 m/s), Oe Evaporation O Benson and White (1994) assuming fraction of atmospheric § (2) the fraction of atmospheric vapor in the water vapor in lake boundary layer, fad, is negligible (fad = 0)

Qw Total input fl ux Qp + Qr lake boundary layer equal to zero (Benson and Q Lake surface precipitation fl ux P × A p L L White, 1994; Jones et al., 2007), (3) no change Qr Runoff fl ux RT × AT in Hr from modern values, and (4) a tempera- Qe Evaporation fl ux EL × AL α # w-c Calcite-water fractionation Kim and O’Neil (1997) ture-dependent equilibrium fractionation factor, α Kinetic evaporation 0.994 for wind speeds <6.8 m/s (Majoube, 1971) kin α (Majoube, 1971). Lake surface evaporation fractionation eq α rate was calculated using temperature, latitude, eq Equilibrium evaporation Majoube (1971)** fractionation lake surface altitude (sample altitude), and Note: LGM—Last Glacial Maximum; DEM—digital elevation model; VSMOW—Vienna Standard Mean Ocean modern average wind speed. Jones et al. (2007) Water. noted that for similar midlatitude lakes during *E = [0.015 + 4 × 10–4T + 10–6z] × [480(T + 0.006z)/(84 − Lat) − 40 + 2.3u(T − T )], where T is mean annual L d the LGM, the combined temperature and ice- temperature, z is lake altitude (meters a.s.l.), u is wind speed, Lat is latitude, and Td is dew point temperature. Td 2 δ18 (°C) is calculated as: Td = 0.52 × Tmin + 0.6 × Tmax − (0.009 × [Tmax] ) − 2. Tmin and Tmax are assumed to uniformly volume effects on incoming precipitation O shift by the same magnitude as mean annual temperature (T) (Linacre, 1992; Jones et al., 2007). would only be between −0.5‰ and +1‰. Thus, †A = (–25.028) + (44.127 × h + ([–1.0357] × h 2) + (0.013333 × h 3) + ([–0.000088611] × h 4) + L L) L L L 18 18 5 6 the δ O (the δ O of runoff and precipitation) (0.00000029112 × hL ) + ([–0.00000000037455] × hL ), where hL = lake-surface altitude (altitude – 1355 m). W §δ18 α α α δ18 Oe = (Revap − 1) × 1000, where Revap = Rlake × ( kin/ eq) × (1/[1 − Hr + {Hr × kin}]) and Rlake = ( OL/1000) + 1 was held as the average of measured literature (Benson and White, 1994). # α 3 –1 values from Surprise Valley (Table 7; Table DR3 1000 × ln( calcite-water) = 18.03 × (10 × T ) − 32.42. T in Kelvin, assumed to be average summer temperature uniformly shifted by the same magnitude as mean annual temperature (T) (Kim and O’Neil, 1997). [see footnote 1]; Ingraham and Taylor, 1989; **α = exp(1137 × T–2 − 0.4156 × T–1 − 0.0020667). T in Kelvin (Majoube, 1971). eq Sladek et al., 2004).

20 Geological Society of America Bulletin, Month/Month 2014 Geological Society of America Bulletin, published online on 2 June 2014 as doi:10.1130/B31014.1

Lacustrine paleoclimate record from late Pleistocene Lake Surprise, California

Solving Equations 4 and 9 allows us to calcu- late the absolute and percent change in precipi- P Δ tation for each LGM and post-LGM tufa sample (%) δ18O measurement, given the associated lake surface altitude and lake surface area for each 619.7 10

sample (Table 8; Fig. 7B). Sensitivity analy- (mm/yr)

sis presented in Figure DR7 (see footnote 1) Precipitation demonstrates that the isotope model is most sensitive to the prescribed temperature anoma- w /yr) 3 lies and runoff coeffi cient. With an increased Q (m

temperature anomaly, the calculated precipita- † noitaluclacecnalabssamepotosI

tion amounts would decrease. Similarly, with (isotope model): an increased runoff coeffi cient, due to colder

temperatures and/or increased spring snowmelt, e O 18

the calculated precipitation amounts would also δ decrease (see Data Repository [footnote 1]). (‰, VSMOW) At the lake highstand, the HI calculation pre- dicts an 85.1% increase in precipitation rela- tive to modern, greater than the isotope mass Lake 575.9 LGM average precip balance model, which predicts a 75% increase (mm/yr) evaporation (Table 8). This HI-derived value for the Lake Surprise highstand is similar to those calcu- L

lated by Reheis (1999b) for other proximal O 18 lake systems, and Miffl in and Wheat’s (1979) δ

prediction of a 77.1% increase for south-cen- (‰, VSMOW) tral Oregon and northwestern Nevada (Table (isotope model): 2). At the LGM, the HI calculation predicts a calcite O

53% increase in precipitation compared to 10% 18 δ

increase predicted by the isotope mass balance (‰, VPDB) calculations (Table 8; Figs. 7B and 8). The

primary difference between the two models is *noit Δ P because the isotope mass balance calculations (%) a allow for sensitivity to changes in humidity luc l and temperature (see equations in Table 7). As ac ) I H

a result, the ratio of lake precipitation and run- (x (mm/yr) Precipitation off to net lake surface evaporation, Qw/Qe, is >1. e dn

Despite the potential uncertainties in the model i c i parameters in Table 7 and the sensitivity analy- go l ordy sis in Figure DR7 (see footnote 1), we regard index Hydrologic the isotope mass balance calculations as the H most accurate estimate of precipitation change at the LGM and post-LGM highstand because ) 2 the stable isotope values allow for the inclusion MASS BALANCE CALCULATIONS AND ISOTOPE 8. HYDROLOGIC INDEX (HI) TABLE area (km LGM average precip (HI): 867.6 53 LGM average lake evap

of processes that are dependent on temperature Tributary and evaporation rate (Table 7). These processes ) are the evaporation model (Linacre, 1992), and 2 area isotope fractionation equations (Majoube, 1971; (km Surface Benson and White, 1994; Kim and O’Neil, 1997) as assembled by Jones et al. (2007). (m) depth Our calculated changes in precipitation for Water Surprise Valley are on the lower end of the range of estimates for nearby Lake Lahontan (Fig. 8;

Miffl in and Wheat, 1979; Hostetler and Benson, altitude (m a.s.l.) 1990) and for the Basin and Range (Matsubara Lake level and Howard, 2009, their Table 1; Menking et al., 2004). This result suggests that decreases in sum- mer evaporation (36.4% annual decrease at the

LGM; Table 8; Fig. 8B) may have played a major 7 for isotope mass balance calculation assumptions and equations. Table See † : LGM—Last Glacial Maximum; VPDB—Vienna PeeDee Belemnite; VSMOW—Vienna Standard Mean Ocean Water. PeeDee Belemnite; VSMOW—Vienna Note : LGM—Last Glacial Maximum; VPDB—Vienna *Assumes no change in temperature or humidity relative to modern values. SVDI12-T4aSVDI12-T15 1439 84 1433.1 78.1 950 920 2861 2892 0.332 0.318 882.2 867.9 56 53 –3.827 –3.089 –4.152 –3.414 578.7 576.0 –21.33 –20.61 907069092 816317153 657.5 603.1 16 7 SVDI12-T4bSVDI12-T13SVDI12-T15 1439 1437.2 84 82.2 1433.1 941 78.1 950 920 2871 2861 2892 0.328 0.332 0.318 877.8 882.2 867.9 55 56 53 –3.343 –3.511 –3.129 –3.668 –3.836 –3.454 577.9 578.7 576.0 –20.86 –21.02 –20.65 857278731 880661324 819220785 625.1 638.3 605.3 10 13 7 SVDI12-T2c 1419.5 64.5 854 2957 0.289 835.8 48 –3.124 –3.449 569.7 –20.64 752560774 579.8 2 SVDI11-T14-1c 1478.4SVDI12-T10a 123.4 1163 1516.8SVDI12-T14 161.8 2648 1309 1530.7 175.7 2503 0.439 1354 2458 971.2 0.523 0.551 72 1029.3 1047.2 82 –2.851 85–3.844 –3.552 –2.737 –3.730 –3.438 779 805 798 –19.74 –20.72 1301799929 –20.43 1089307483 834.1 1593901125 990.9 47 946.2 75 67 SVDI12-T3a 1427.8 72.8 893 2918 0.306 855.2 51 –3.460 –3.785 573.6 –20.97 816361203 613.3 8 SVDI11-T3-1a 1437.7SVDI11-T4-1b 82.7 1430.6 944 75.6 2868 907 2905 0.329 0.312 879.1 55 861.9 52 –4.045 –3.225 –4.369 –3.550 578.1 574.9 –21.54 918508324 –20.74 668.6 813284920 18 605.7 7 SVDI12-T9SVDI12-T9SVDI12-T10b 1508.9SVDI12-T14 1508.9 153.9 1516.8 153.9 1284 161.8 1530.7 1284 1309 175.7 2528 1354 2528 2503 2458 0.508 0.508 0.523 1019.0 0.551 1019.0 1029.3 80 80 1047.2 82 –3.587 85 –3.465 –3.538 –3.473 –3.846 –3.352 –3.424 –3.733 794 794 798 805 –20.46 –20.34 –20.42 1560352962 –20.72 1543646580 1591921357 938.3 928.2 1089307483 945.0 66 64 991.1 67 75 SVDI12-T7 1472.5 117.5 1134 2678 0.424 959.6 70 –3.463 –3.350 776 –20.34 1332561746 867.7 53 SVDI12-T3b 1427.8 72.8 893 2918 0.306 855.2 51 –3.457 –3.782 573.6 –20.97 816126318 613.2 8 SVDI11-T3–2SVDI11-T4-1a 1437.7SVDI12-T1 1430.6SVDI12-T2a 82.7 75.6 944 1419.5 1419.5 907 64.5 64.5 2868 2905 854 854 2957 0.329 2957 0.312 879.1 0.289 861.9 0.289 55 52 835.8 835.8 –2.988 48 –3.238 48 –3.313 –3.696 –3.563 –3.491 578.1 –4.020 574.9 –3.816 –20.51 569.7 –20.75 569.7 833179570 814290165 –21.20 –21.00 606.5 606.5 792886253 777951961 7 7 610.8 599.3 8 6 SVDI12-T7 1472.5 117.5 1134 2678 0.424 959.6 70 –3.789 –3.675 776 –20.66 1371919912 893.3 58 SVDI12-T2b 1419.5 64.5 854 2957 0.289 835.8 48 –3.472 –3.796 569.7 –20.98 776575497 598.3 6 SVDI11-T14-1a 1478.4 123.4 1163 2648 0.439 971.2 72 –2.501 –2.387 779 –19.40 1264856983 810.5 43 SVDI11-T2-1SVDI11-T2–2 1453.5 1453.5 98.5 98.5 1031 1031 2781 2781 0.371 0.371 917.5 917.5 62 62 –3.123 –3.385 –3.448 –3.710 585.5 585.5 –20.64 –20.90 932804690 955057055 644.2 659.6 14 17 Sample Post-LGM samples role in driving and sustaining lake levels in the LGM samples northern Basin and Range during the late LGM.

Geological Society of America Bulletin, Month/Month 2014 21 Geological Society of America Bulletin, published online on 2 June 2014 as doi:10.1130/B31014.1

Ibarra et al.

20 30 Previous investigations have proposed that Surprise Valley Meteorological Data the primary driver of maximum LGM and post- ABPMIP3 Ensemble 0 ka PMIP3 Ensemble 21 ka (LGM) LGM pluvial lake levels was increased precipi- 25 tation (Thompson et al., 1999; Menking et al., 15 2004; Lyle et al., 2012). Menking et al. (2004) 20 modeled Lake Estancia (34.6°N) and proposed that precipitation was doubled during Estancia’s 10 15 LGM highstands. During the Surprise Valley post-LGM highstand, we fi nd that precipitation 10 may have increased by up to 75% (Fig. 7B). Our 5 Annual) (% Runoff combined results bracket the “mean position” of Precipitation (% Annual) Precipitation (% 5 the polar and midlatitude storm tracks to the southern Basin and Range (~35°N) during the LGM and to the northern Basin and Range 0 0 Jan March May July Sept Nov Jan March May July Sept Nov (~41°N) during the late HS1, as proposed by Munroe and Laabs (2012). Figure 9. Seasonal precipitation and runoff predicted by the PMIP3 climate model ensem- ble. All climate models except COSMOS-ASO uniformly output both precipitation and run- Comparison of Precipitation Estimates to off in the PMIP3 database (Table DR6 [see text footnote 1]). (A) PMIP3 ensemble average Climate Model Predictions precipitation normalized to the percent of total annual precipitation for the 0 ka (preindus- Evaluation of PMIP3 climate models against trial) and 21 ka (Last Glacial Maximum [LGM]) experiments. Surprise Valley meteorologi- paleoclimate data provides an important oppor- cal data are the average of three weather stations (Fig. 2A). Climate models suggest that the tunity to test climate models at mean states very seasonality of precipitation during the LGM was unchanged. (B) PMIP3 ensemble average different than modern conditions. The direct runoff (surface and subsurface) normalized to the percent of total annual runoff for the 0 comparison between climate models and climate ka (preindustrial) and 21 ka (LGM) experiments. Climate models suggest that temperature records aids in understanding both the past and effects caused the decoupling of precipitation and runoff during the LGM, with peak runoff future regional sensitivity to changing climatic shifting to the late spring–early summer. See Figures DR8–DR12 and Table DR6 for indi- conditions (Braconnot et al., 2012; DiNezio and vidual climate model results (text footnote 1). Tierney, 2013; Hargreaves et al., 2013). The LGM precipitation anomaly varies widely among the PMIP3 ensemble members record (Worona and Whitlock, 1995). Addition- 75% increase in precipitation during late HS1. in both magnitude and direction of change ally, the assumption of a uniform temperature Given the range of uncertainty in some of the (Table DR6 [see footnote 1]). Eight of the anomaly is supported by the climate models, input parameters used in the isotope mass bal- nine climate models reproduce the modern which on average demonstrate a minimal range ance calculations (see Fig. DR7 for the sensitiv- seasonal precipitation patterns observed at the of variation in the monthly temperature anoma- ity analy sis of precipitation increase to RH, T, − − δ18 three meteorological stations in Surprise Val- lies of between 6.0 °C and 8.0 °C (Fig. DR10 OW, runoff coeffi cient, and wind speed [see ley, while most overpredict absolute precipita- [see footnote 1]). Corroborating evidence for footnote 1]), and the coarse scale of the climate tion amount (Figs. DR8 and DR11 [see foot- the importance of temperature on basin hydrol- models (~0.75° to 2° horizontal resolution; Figs. note 1]). All nine models suggest that the winter ogy is the seasonality of peak subsurface and DR11 and DR12 [see footnote 1]), further con- contribution to total annual rainfall decreased surface runoff predicted by the PMIP3 ensem- straints and/or modeling approaches are neces- during the LGM (Fig. 9; Figs. DR8 and DR11 ble. Runoff, which is coupled to precipitation in sary to fully assess model bias. [see footnote 1]). The results of the AOGCM modern times, is shifted from late winter–early analysis demonstrate a PMIP3 ensemble aver- spring in modern time to the late spring–early Regional Implications of the Latest age increase of 6.5% LGM precipitation rela- summer during the LGM (Fig. 9; Fig. DR9 [see Pleistocene Lake Surprise Lake Levels tive to modern, ranging from −14.5% to 50.3% footnote 1]). Finally, the PMIP3 ensemble sup- (Fig. 8A; Table DR6 [see footnote 1]). This is ports our assumption used in the isotope mass During the LGM, Lake Surprise stood at compara ble to the isotope mass balance calcula- balance calculations that relative humidity has moderate lake levels (65–99 m depth), similar tions, which estimate a 10% increase. The pre- not changed dramatically since the LGM (only to Lake Lahontan (Benson et al., 1995; Adams dicted total evaporation, a variable parameter- a 4.1% increase). et al., 2008), with an HI of 0.29–0.34 covering ized from the surface latent heat fl ux by climate The climate models and isotope mass balance 62%–72% of the ultimate highstand surface models, is predicted to decrease by an ensemble models presented here provide a framework for area (Table 2). During HS1, Lake Surprise rose average of 28.2% at the LGM (Fig. 8B; Fig. investigating whether wetter or cooler condi- to the highest latest Pleistocene lake level (176 DR12 [see footnote 1]). While not equivalent tions (or a balance of the two) drove pluvial lake m depth), peaking at 15.19 ± 0.18 ka cal. (radio- to lake surface evaporation rate (EL), these val- evolution during the LGM and deglaciation. carbon age). The HI increased by 39% to 0.56, ues agree with the 36% decrease in evaporation Based on our analysis, moderate LGM lake with Lake Surprise covering 36% of the termi- rate estimated for Lake Surprise at the LGM. levels at Lake Surprise were driven primarily nally draining watershed. The lack of substan- Among the PMIP3 ensembles, the mean annual by reduced temperatures and resultant reduced tial carbonate deposition and shoreline develop- temperature LGM anomaly is −7.0 °C (Table summer lake evaporation, with a minimal pre- ment (relative to LGM shorelines) suggests this DR6; Fig. DR10 [see footnote 1]), which is cipitation increase of 10% relative to modern. highstand was apparently brief. Additionally, the same value used in the isotope mass bal- Following moderate LGM lake levels, Lake Sr/Ca ratios of the post-LGM samples are very ance model and predicted by the nearby pollen Surprise lake levels responded rapidly to a similar to LGM samples, suggesting that dilute

22 Geological Society of America Bulletin, Month/Month 2014 Geological Society of America Bulletin, published online on 2 June 2014 as doi:10.1130/B31014.1

Lacustrine paleoclimate record from late Pleistocene Lake Surprise, California lake waters did not inhibit carbonate precipita- temperatures and attendant increases in precipi- instead been sustained by regional lake-atmo- tion. Recession was subsequently recorded in tation (Fig. 10H) during HS1. sphere effects (e.g., Hostetler et al., 1994; Lic- several stillstands between 15.2 and 12.7 ka. In Following the LGM, the temporal correspon- ciardi, 2001) brought about by the much larger the following synthesis, we link our new MIS dence among highstands at Lake Surprise and and longer-lasting Lahontan (highstand at 15.8 2 Lake Surprise hydrograph (Fig. 7A), isotope other small lake systems in the western United ka; Adams and Wesnousky, 1998) and Bonne- mass balance precipitation calculations (Figs. States refl ects a shift in Northern Hemisphere ville lake systems (recession from Provo shore- 7B and 8), and PMIP3 ensemble results (Figs. midlatitude atmospheric circulation during HS1 line between ca. 16 and 14.5 ka; Godsey et al., 8 and 9) to regional records (Fig. 10) from the (Fig. 10). Highstands and signifi cant stillstands 2011; Miller et al., 2013). In addition, the lack LGM through the deglaciation. of small lake systems during the latter parts of of a signifi cant latitudinal trend in highstands Regional lake and glacial records record the the LGM are attributed to a southward displace- and signifi cant stillstands (Fig. 10J) contrasts combined effects of temperature and precipita- ment of the mean position of the southern arm with the fi ndings of Lyle et al. (2012). Evidence tion during the LGM (Fig. 10). Coincident with of the split PJS (Polar Jet Stream) and dipping from Lake Elsinore, California (~34°N; Kirby a maximum winter insolation and a minimum westerlies, which would have brought more et al., 2013), and Cave of the Bells, Arizona summer insolation (Fig. 10C), Lake Franklin and precipitation to the region (Negrini, 2002; Mun- (~32°N; Wagner et al., 2010), provides addi- the much larger Lake Bonneville (Fig. 1) to the roe and Laabs, 2012; Kirby et al., 2013). Our tional support for the hypothesis that westerly, east rose gradually during the LGM (Fig. 10G), isotope mass balance results require that the North Pacifi c moisture sources were respon- although neither reached highstand levels until pluvial maximum of Lake Surprise was driven sible for post-LGM, mid- to late HS1, pluvial after the LGM (Oviatt et al., 1992; Munroe and by increases in precipitation (~75% increase maxima in the interior western United States. Laabs, 2013). Similarly, Lake Surprise records relative to modern), confi rming that displaced Cumulatively, this evidence suggests a latitudi- a gradual transgression from 22.13 to 19.22 midlatitude storm tracks likely drove increased nally broad and strengthened midlatitude west- ka cal. At the same time, glaciers were persis- lake levels during the latter part of HS1. Addi- erly storm track during HS1. tent: Glacial activity in the southeast Cascades tionally, reassessment of radiocarbon ages from Post-LGM glacial activity also supports gradually increased through the LGM (Fig. 10F; highstands and signifi cant stillstands spanning these conclusions. Although the Laurentide Rosenbaum et al., 2012), two glacial advances 31.8°N to 45.9°N (Fig. 10J; see Table DR7 ice sheet was receding (Fig. 10B) and North- are recorded in the Sierra Nevada (Tioga 1 and [see footnote 1]) indicates that the regional plu- ern Hemisphere temperature was increasing 2) at the beginning and end of the LGM (Phil- vial maximum lies between ca. 17 and 15 ka, (Figs. 10A and 10H), persistent glacial activity lips et al., 2009), a terminal moraine recorded an during the latter part of HS1. Prior to this and has been documented. Glacial activity in the age of 20.5 ka in the Ruby Mountains (Laabs, immediately following the LGM, during the southeastern Cascades was recorded until ca. et al., 2013), and a glacial maximum is observed early part of HS1 (ca. 19–17 ka), some authors 15 ka (Rosenbaum et al., 2012; Fig. 10F), two at 21.6 ka in the Wallowa Mountains (Fig. 10E; have suggested a regional “big dry” event based fi nal post-LGM advances are recognized in the Licciardi et al., 2004; radiocarbon ages recalcu- on desic cation in Lake Estancia, Sierra Nevada (Phillips et al., 2009; Tioga 3 lated in Rosenbaum et al., 2012). (Allen and Anderson, 2000), but this event is not and 4), four post-LGM recessional moraines Decreased insolation and reduced tem- apparent in the lakes in the northern Basin and are recorded from 17.2 to 14.8 ka in the Ruby peratures would have decreased summer lake Range (Broecker et al., 2009; Broecker and Put- Mountains (Laabs et al., 2013), and a glacial surface evaporation, decreased glacial melt/ nam, 2012; Munroe and Laabs, 2013). maximum has been observed (ca. 17.4 ka) in sublimation, and enhanced basin average run- Atmospheric forcing of precipitation in the the Wallowa Mountains (Fig. 10E; Licciardi et off. Summer insolation at 40°N was lower western United States is attributed to North al., 2004). Comparable to the small lake sys- than the long-term average between 27.1 and Atlantic cooling during HS1 (COHMAP, 1988; tems of similar latitude, these glacial records 14.5 ka (during HS1 and the late LGM), reach- Zic et al., 2002; Denniston et al., 2007; Wag- indicate increased winter precipitation during ing a minimum at 20.3 ka (Fig. 10I). During ner et al., 2010; Broecker and Putnam, 2012; HS1 despite increasing Northern Hemisphere the LGM, increases in regional moisture avail- Munroe and Laabs, 2012; Benson et al., 1998, temperatures. Finally, while the Laurentide ice ability refl ected by lake hydrographs and gla- 2003, 2013). The proposed mechanism is the sheet began to retreat at the end of the LGM, cial growth, the absence of abrupt shifts in lake winter enhancement of the Aleutian Low via the Cordilleran ice sheet advanced southward and/or glacial records, and decreasing summer reduced tropical Atlantic precipitation (Oku- into parts of Washington, Idaho, and Montana insolation all suggest that reduced temperatures mura et al., 2009) and the suppression of the until ca. 18–15 ka (Waitt, 1985; Clague and decreased lake surface evaporation. The results Atlantic Meridional Overturning Circulation James, 2002; Dyke, 2004) before fully melting from our isotope mass balance calculations and (McManus et al., 2004; Denton et al., 2010). during the Bølling-Allerød and Younger Dryas the PMIP3 climate model support the minimal The lake highstand compilation of Munroe from 14 to ca. 11.5 ka (Dyke, 2004). The per- increase in precipitation (2.5%–18.2% relative and Laabs (2012) demonstrated temporal syn- sistence of the Cordilleran ice sheet may have to modern; Figs. 7B and 8) and reduced lake chrony between some Basin and Range lake maintained the southward defl ection of the PJS surface evaporation (~36% decrease relative to highstands and the onset of iceberg discharge until the onset of the Bølling-Allerød inter- modern; Fig. 8). Such results are consistent with and reduced North Atlantic sea-surface tem- stadial (ca. 14.5 ka; COHMAP, 1988; Benson reduced temperatures due to lower atmospheric peratures during 1 at ca. 17 ka. et al., 1990; García and Stokes, 2006; Godsey

CO2 levels (Denton et al., 2010; Shakun et al., However, several highstands from smaller lakes et al., 2011). 2012), combined with lower summer insolation, near Bonneville and Lahontan postdate this Further evidence for differential seasonal as the key drivers for reduced evaporation and, period, as does the Lake Surprise highstand anomalies driving different paleorecords of the by extension, moderate LGM lake levels (Mun- radiocarbon age (15.19 ± 0.18 ka cal.; Table moisture balance in the interior western United roe and Laabs, 2013; Maher et al., 2014). These DR7 [see footnote 1]; Fig. 10J). Northern Basin States is demonstrated in the temporal asyn- conditions may have primed lake systems to and Range highstands that postdate the appar- chrony between the late-LGM and HS1 maxima respond rapidly to decreases in North Atlantic ent mid-HS1 maximum (ca. 17 ka) may have recorded in lakes and glaciers, with late MIS 3

Geological Society of America Bulletin, Month/Month 2014 23 Geological Society of America Bulletin, published online on 2 June 2014 as doi:10.1130/B31014.1

Ibarra et al.

YD B-A HS1 LGM MIS 1 MIS 2 MIS 1 MIS 2 MIS 3 SMOW 0 0 Cordilleran+Laurentide

Ice-Sheet Area (%) -35

-1 O(‰) 18 -40 -2 50 H Temp (°C) -3 -45 NGRIP δ North Hemisphere January Insol

-4 460 230 40°N (W/m ) AB100 2

440 220 ) 2 40°N (W/m 460 C 230 Winter Insol July Insol 2

I ) 40°N (W/m 420 Summer Insolation 210 440 220 Minimum 40 °N (W/m Summer Insol

2 North (45.9°N) )

1600 210 Lake D Surprise

1500 Highstand/Maxima Stillstand Lake Surprise 1400 Elevation (m a.s.l.)

RTRR E Glacial Records Western United States

WT TT Lake Highstands and Stillstands South (31.8°N) T4 T3 T2 T1 J 40 Glacial Flour Flux SE Cascades

(kgm Newark Valley, NV Diamond Valley, NV

-2 Fish Lake Valley, NV 20 yr Opal -1 Yucca Mountain, NV δ -15 Devil’s Hole ) K Barstow, CA 18 F P-ET Maxima pre-LGM Winter O(‰) 100 0 Precipitation/Infiltration -16

Maximum? VPDB -5 75 L -17 50 VPDB

O(‰) -10 (% Maximum) 25 18 Relative Lake Level G δ M 0 6 8 10 12 14 16 18 20 22 24 26 28 30 10 15 20 25 30 35 40 45 50 Age (ka) Age (ka) Figure 10. Paleoclimate archives and orbital parameters for comparison to the Lake Surprise hydrograph. (A) Proxy-derived Northern Hemisphere temperature stack (black line; dashed line is 1σ error bound) deviation from the early Holocene (11.5–6.5 ka) (compiled by Shakun et al., 2012). (B) North American ice-sheet percent area deglaciated, including both the Laurentide and Cordilleran ice sheets (compiled by Dyke, 2004); note scale is inverted (gray line). (C) Summer (red) and winter (blue) insolation for 40°N (Laskar et al., 2004). (D) The Lake Surprise hydrograph (this study); see the Figure 7 caption for explanation and radiocarbon ages are white circles. (E) Glacier advances (Tioga 1–4 [T1–T4]) in the Sierra Nevada, California, recorded by cosmogenic 36Cl ages (blue bars; Phillips et al., 2009), maximal glaciations (LGM maximum [TT] and post-LGM maximum [WT]) in the Wallowa Mountains, Oregon, documented by cosmogenic 10Be ages (gray bars; Licciardi et al., 2004), recalculated in Rosenbaum et al. (2012), and 10Be ages (Bayesian averages) of the terminal (RT) and range of recessional moraines (RR) from the Ruby Mountains, Nevada (Laabs et al., 2013). (F) Lacustrine record of glacial fl our fl ux recording glacial extent in the southeastern Cascades, calculated from the glacial fl our content in the Caledonia Marsh core from Upper Klamath Lake (Rosenbaum and Reynolds, 2004; Rosenbaum et al., 2012). (G) Lake-level curves plotted as the percent of Last Glacial Maxi- mum (LGM) maximum: Lake Bonneville, Utah (black dashed line; Oviatt et al., 1992; as compiled in McGee et al., 2012), Lake Lahontan, Nevada (black line; Benson et al., 1995; Adams et al., 2008), and Lake Franklin, Nevada (green line; Munroe and Laabs, 2013). (H) North Ice Core Project (NGRIP) Greenland ice-core δ18O on Greenland Ice Core Chronology 2005 (GICC05) time scale (gray line) smoothed with LOESS smoothing function (red line; Rasmussen et al., 2006). SMOW—Standard Mean Ocean Water. (I) Solar insolation as in C. (J) Lake highstands (black) and stillstands (white boxes) arranged from south to north, as compiled by Munroe and Laabs (2012), Lyle et al. (2012), and recalculated by this study in Table DR7 (see text footnote 1). (K) Great Basin and Mojave soil opal precipitation minus 234 238 evapotranspiration (P-ET) maxima (Maher et al., 2014) derived from ( U/ U)0 variations in soil opal (Maher et al., 2014) and vadose zone opal (Paces et al., 2010). (L) δ18O from Devils Hole, Nevada (Winograd et al., 2006), interpreted as decreased Pacifi c sea-surface tempera- tures during δ18O minima. VPDB—Vienna PeeDee Belemnite. (M) δ18O from speleothem records: Cave of the Bells, Arizona (Wagner et al., 2010), and Fort Stanton, New Mexico (Asmerom et al., 2010). Both records are interpreted to refl ect increased winter precipitation during δ18O minima. Both records are smoothed using a LOESS smoothing function (red lines). MIS—marine oxygen isotope stage; YD—Younger Dryas; B-A—Bølling-Allerød; HS1—Heinrich Stadial 1.

24 Geological Society of America Bulletin, Month/Month 2014 Geological Society of America Bulletin, published online on 2 June 2014 as doi:10.1130/B31014.1

Lacustrine paleoclimate record from late Pleistocene Lake Surprise, California and early LGM speleothem and vein calcite ACKNOWLEDGMENTS Benson, L.V., and White, J.W.C., 1994, Stable isotopes of δ18O minima recording wet and cold winter oxygen and hydrogen in the Truckee River–Pyramid We thank Benjamin Laabs (GSA Bulletin associ- Lake surface-water system. 3. Source of water vapor conditions (Winograd et al., 2006; Asmerom overlying Pyramid Lake: Limnology and Oceanogra- ate editor), Jeff Munroe, and an anonymous reviewer et al., 2010; Wagner et al., 2010; Figs. 10L and phy, v. 39, p. 1945–1958, doi: 10 .4319 /lo .1994 .39 .8 for thorough reviews and comments; Jorge Vazquez .1945 . 10M), and maximum (winter) infi ltration rates (second reader for Daniel Ibarra’s M.S. thesis) and Benson, L.V., Currey, D.R., Dorn, R.I., Lajoie, K.R., Oviatt, recorded in soil opal (Maher et al., 2014; Fig. Jeremy Caves for extensive comments and discus- C.G., Robinson, S.W., Smith, G.I., and Stine, S., 10K). Lakes, , glaciers, and speleothems sions on earlier versions of this manuscript; Marith 1990, Chronology of expansion and contraction of Reheis, Jessica Oster, David Miller, and Perach Nuriel four Great Basin lake systems during the past 35,000 are sensitive to different seasonal variations in for detailed discussions; Jonathan Glen and Sabina years: Palaeogeography, Palaeoclimatology, Palaeo- the hydrologic cycle. Late LGM lake levels are Kraushaar for fi eld work support; Guangchao Li for ecology, v. 78, p. 241–286, doi: 10 .1016 /0031 -0182 (90)90217 -U . particularly sensitive to summer insolation and cation analyses; Dave Mucciarone for stable isotope Benson, L.V., Kashgarian, M., and Rubin, M., 1995, Car- reduced temperatures, which in the midlati- analyses; David Medeiros and Claire Kouba from bonate deposition, Pyramid Lake subbasin, Nevada. tudes is more than twice winter insolation (Fig. the Stanford Geospatial Center for assistance on the 2. Lake levels and polar-jet stream positions recon- ArcGIS analysis; Kimberly Lau and Conni De Massi structed from radiocarbon ages and elevations of 10I). Given these observations, we propose that for help with laboratory work; Hari Mix for suggesting carbonates (tufas) deposited in the Lahontan Basin: changes in orbital conditions that infl uence sea- and discussing the Sr/Ca and stable isotope analyses; Palaeogeography, Palaeoclimatology, Palaeoecology, sonal insolation, particularly summer insola- and Matthew Winnick and Daniel Horton for assis- v. 117, p. 1–30, doi: 10 .1016 /0031 -0182 (94)00103 -F . Benson, L., White, L.D., and Rye, R., 1996, Carbonate depo- tion and its infl uence on lake evaporation (as tance with the PMIP3 climate model processing. Joseph Rosenbaum provided the raw data plotted for sition, Pyramid Lake Subbasin, Nevada: 4. Comparison well as evapotranspiration), is a key long-term of the stable isotope values of carbonate deposits (tufas) the glacial fl ux record from the Caledonia Marsh core and the Lahontan lake-level record: Palaeogeography, driver of hydrologic variability in the western from Upper Klamath Lake, Oregon, in Figure 10. Palaeoclimatology, Palaeoecology, v. 122, p. 45–76, United States. LiDAR data were collected by the National Center doi: 10 .1016 /0031 -0182 (95)00099 -2 . for Airborne Laser Mapping (NCALM) with funding Benson, L.V., Smoot, J.P., Kashgarian, M., SarnaWojcicki, CONCLUSION from the National Aeronautic and Space Administra- A., and Burdett, J.W., 1997, Radiocarbon ages and en- tion (NASA) through award 10-UAS10-0021 to Jona- vironments of deposition of the Wono and Trego Hot than Glen, Anne Egger, and Corey Ippolito. This work Springs tephra layers in the Pyramid Lake subbasin, Construction of the late Pleistocene lake was supported by the National Science Foundation Nevada: Quaternary Research, v. 47, p. 251–260, doi: 10 .1006 /qres .1997 .1897 . hydrograph for Lake Surprise, isotope mass (NSF) grant EAR-0921134 to Kate Maher and sup- Benson, L.V., Lund, S.P., Burdett, J.W., Kashgarian, M., balance calculations, and the regional synthesis port from Stanford University. Rose, T.P., Smoot, J.P., and Schwartz, M., 1998, Cor- herein reveal that previously overlooked con- relation of late-Pleistocene lake-level oscillations in REFERENCES CITED Mono Lake, California, with North Atlantic climate straints—namely, reduced temperature and the events: Quaternary Research, v. 49, p. 1–10, doi:10 effect of seasonal insolation variability on sum- Adams, K.D., and Wesnousky, S.G., 1998, Shoreline pro- .1006 /qres .1997 .1940 . mer lake evaporation—may have driven lake cesses and the age of the Lake Lahontan highstand in Benson, L.V., Lund, S., Negrini, R., Linsley, B., and Zic, M., the Jessup embayment, Nevada: Geological Society of 2003, Response of North American Great Basin lakes levels in the Basin and Range during MIS 2. America Bulletin, v. 110, p. 1318–1332, doi:10 .1130 to Dansgaard–Oeschger oscillations: Quaternary Sci- Based on isotope mass balance calculations, /0016 -7606 (1998)110 <1318: SPATAO>2 .3 .CO;2 . ence Reviews, v. 22, p. 2239–2251, doi: 10 .1016 /S0277 -3791 (03)00210 -5 . moderate LGM lake levels observed in Surprise Adams, K.D., Goebel, T., Graf, K., Smith, G.M., Camp, A.J., Briggs, R.W., and Rhode, D., 2008, Late Pleistocene Benson, L.V., Lund, S.P., Smoot, J.P., Rhode, D.E., Spen- Valley were driven by decreased evaporation and early Holocene lake-level fl uctuations in the La- cer, R.J., Verosub, K.L., Louderback, L.A., Johnson, (36% relative to modern), with minimal precipi- hontan Basin, Nevada: Implications for the distribution C.A., Rye, R.O., and Negrini, R.M., 2011, The rise and of archaeological sites: Geoarchaeology, v. 23, p. 608– fall of Lake Bonneville between 45 and 10.5 ka: Qua- tation increases (10% relative to modern). These 643, doi: 10 .1002 /gea .20237 . ternary International, v. 235, p. 57–69, doi: 10 .1016 /j results agree with climate model simulations, Allen, B.D., and Anderson, R.Y., 2000, A continuous, high- .quaint .2010 .12 .014 . which predict a 28% decrease in total evapora- resolution record of late Pleistocene climate variability Benson, L.V., Smoot, J.P., Lund, S.P., Mensing, S.A., Foit, from the Estancia basin, New Mexico: Geological So- F.F., and Rye, R.O., 2013, Insights from a synthesis tion and an average of 6.5% increase in precipi- ciety of America Bulletin, v. 112, p. 1444–1458, doi: of old and new climate-proxy data from the Pyramid tation relative to modern. Evidence presented 10 .1130 /0016 -7606 (2000)112 <1444: ACHRRO>2 .0 and Winnemucca Lake basins for the period 48 to here questions the long-standing paradigm of .CO;2 . 11.5calka: Quaternary International, v. 310, p. 62–82, Amelin, Y., and Back, M., 2006, Opal as a U-Pb geochro- doi: 10 .1016 /j .quaint .2012 .02 .040 . a “rainy LGM” in western North America and nometer: Search for a standard: Chemical Geology, Bischoff, J.L., and Fitzpatrick, J.A., 1991, U-series dating highlights the strong infl uence of temperature v. 232, p. 67–86, doi: 10 .1016 /j .chemgeo .2006 .02 .018 . of impure carbonates: An isochron technique using Anderson, R.F., Bacon, M.P., and Brewer, P.G., 1982, Ele- total-sample dissolution: Geochimica et Cosmochi- and seasonal insolation on water availability in vated concentrations of actinides in Mono Lake: Sci- mica Acta, v. 55, p. 543–554, doi: 10 .1016 /0016 -7037 the region. ence, v. 216, p. 514–516, doi:10 .1126 /science .216 (91)90011 -S . The moderate late LGM lake levels and sus- .4545 .514 . Bischoff, J.L., Fitts, J.P., and Fitzpatrick, J.A., 1997a, Re- Asmerom, Y., Polyak, V.J., and Burns, S.J., 2010, Variable sponses of sediment geochemistry to climate change tained glacial activity post-LGM, driven by winter moisture in the southwestern United States in Owens Lake sediment: An 800-ky record of sa- a minimum in summer insolation (20.3 ka), linked to rapid glacial climate shifts: Nature Geosci- line/fresh cycles in core OL-92, in Smith, G.I., and primed the Basin and Range lake systems to ence, v. 3, p. 114–117, doi: 10 .1038 /ngeo754 . Bischoff, J.L., eds., An 800,000-Year Paleoclimatic Bacon, M.P., and Anderson, R.F., 1982, Distribution of tho- Record from Core OL-92, Owens Lake, Southeast Cal- respond rapidly to increased rainfall during HS1 rium isotopes between dissolved and particulate forms ifornia: Geological Society of America Special Paper brought by latitudinally broad and strengthened in the deep-sea: Journal of Geophysical Research, 317, p. 37–47. v. 87, p. 2045–2056, doi: 10 .1029 /JC087iC03p02045 . Bischoff, J.L., Menking, K.M., Fitts, J.P., and Fitzpatrick, midlatitude westerly storm tracks. This rapid Bacon, S.N., Burke, R.M., Pezzopane, S.K., and Jayko, A.S., J.A., 1997b, Climatic oscillations 10,000–155,000 yr response was recorded in highstands across the 2006, Last Glacial Maximum and Holocene lake levels B.P. at Owens Lake, California, refl ected in glacial Basin and Range (32°N–42°N) via atmosphere- of Owens Lake, eastern California, USA: Quaternary rock fl our abundance and lake salinity in core OL-92: Science Reviews, v. 25, p. 1264–1282, doi: 10 .1016 /j Quaternary Research, v. 48, p. 313–325, doi: 10 .1006 ocean teleconnection to reduced North Atlan- .quascirev .2005 .10 .014 . /qres .1997 .1933 . tic sea-surface temperatures (ca. 17.5–16.5 ka Benson, L.V., and Paillet, F.L., 1989, The use of total lake- Blard, P.H., Sylvestre, F., Tripati, A.K., Claude, C., Causse, until ca. 14.7 ka). The rapid response of lake surface area as an indicator of climatic change: Exam- C., Coudrain, A., Condom, T., Seidel, J.-L., Vimeux, F., ples from the Lahontan Basin: Quaternary Research, Moreau, C., Dumoulin, J.P., and Lave, J., 2011, Lake systems to substantial increases in precipitation v. 32, p. 262–275, doi: 10 .1016 /0033 -5894 (89)90093 -8 . highstands on the (Tropical Andes) contem- emphasizes the role of atmosphere-ocean tele- Benson, L., and Paillet, F., 2002, HIBAL: A hydrologic-iso- poraneous with Heinrich 1 and the Younger Dryas: topic-balance model for application to paleolake sys- New insights from 14C, U-Th dating and δ18O of car- connections in water availability in the western tems: Quaternary science reviews, v. 21, p. 1521–1539, bonates: Quaternary Science Reviews, v. 30, p. 3973– United States. doi: 10 .1016 /S0277 -3791 (01)00094 -4 . 3989, doi: 10 .1016 /j .quascirev .2011 .11 .001 .

Geological Society of America Bulletin, Month/Month 2014 25 Geological Society of America Bulletin, published online on 2 June 2014 as doi:10.1130/B31014.1

Ibarra et al.

Braconnot, P., Harrison, S.P., Kageyama, M., Bartlein, P.J., Denton, G.H., Anderson, R.F., Toggweiler, J.R., Edwards, Hall, B.L., and Henderson, G.M., 2001, Use of uranium- Masson-Delmotte, V., Abe-Ouchi, A., Otto-Bliesner, R.L., Schaefer, J.M., and Putnam, A.E., 2010, The last thorium dating to determine past 14C reservoir effects B., and Zhao, Y., 2012, Evaluation of climate models glacial termination: Science, v. 328, p. 1652–1656, doi: in lakes: Examples from : Earth and Plan- using palaeoclimatic data: Nature Climate Change, 10 .1126 /science .1184119 . etary Science Letters, v. 193, p. 565–577, doi: 10 .1016 v. 2, p. 417–424, doi: 10 .1038 /nclimate1456 . DiNezio, P.N., and Tierney, J.E., 2013, The effect of sea /S0012 -821X (01)00524 -6 . Bradbury, J.P., Colman, S.M., and Dean, W.E., 2004, Lim- level on glacial Indo-Pacifi c climate: Nature Geosci- Hargreaves, J.C., Annan, J.D., Ohgaito, R., Paul, A., and nological and climatic environments at Upper Klamath ence, v. 6, p. 485–491, doi: 10 .1038 /ngeo1823 . Abe-Ouchi, A., 2013, Skill and reliability of climate Lake, Oregon, during the past 45 000 years: Journal Doebbert, A.C., Carroll, A.R., Mulch, A., Chetel, L.M., model ensembles at the Last Glacial Maximum and of Paleolimnology, v. 31, p. 167–188, doi: 10 .1023 /B: and Chamberlain, C.P., 2010, Geomorphic controls mid-Holocene: Climate of the Past, v. 9, p. 811–823, JOPL .0000019232 .74649 .02 . on lacustrine isotopic compositions: Evidence from doi: 10 .5194 /cp -9 -811 -2013 . Brady, E.C., Otto-Bliesner, B.L., Kay, J.E., and Rosen- the Laney Member, Green River Formation, Wyo- Hedel, C.W., 1980, Late Quaternary Faulting in Western Sur- bloom, N., 2013, Sensitivity to glacial forcing in the ming: Geological Society of America Bulletin, v. 122, prise Valley, Modoc County, California [M.S. thesis]: CCSM4: Journal of Climate, v. 26, p. 1901–1925, doi: p. 236–252, doi: 10 .1130 /B26522 .1 . San Jose, California, San Jose State University, 284 p. 10 .1175 /JCLI -D -11-00416 .1 . Dyke, A.S., 2004, An outline of North American deglaciation Hedel, C.W., 1984, Maps Showing Geomorphic and Geo- Broecker, W.S., and Putnam, A.E., 2012, How did the hydro- with emphasis on central and northern Canada, in Ehlers, logic Evidence for Late Quaternary Displacement logic cycle respond to the two-phase mystery interval?: J., and Gibbard, P.L., eds., Developments in Quaternary along the Surprise Valley and Associated Faults, Quaternary Science Reviews, v. 57, p. 17–25, doi: 10 Sciences, Quaternary Glaciations–Extent and Chronol- Modoc County, California: U.S. Geological Survey .1016 /j .quascirev .2012 .09 .024 . ogy, Part II: Amsterdam, Elsevier, p. 373–424. Miscellaneous Field Studies Map MF-1429, scale Broecker, W.S., McGee, D., Adams, K.D., Cheng, H., Egger, A.E., and Miller, E.L., 2011, Evolution of the north- 1:62,500, 2 sheets. Edwards, R.L., Oviatt, C.G., and Quade, J., 2009, A western margin of the Basin and Range: The geology Hendriks, A.J., Schipper, A.M., Caduff, M., and Huijbregts, Great Basin–wide dry episode during the fi rst half of and extensional history of the Warner Range and envi- M.A.J., 2012, Size relationships of water infl ow into the mystery interval?: Quaternary Science Reviews, rons, northeastern California: Geosphere, v. 7, p. 756– lakes: Empirical regressions suggest geometric scal- v. 28, p. 2557–2563, doi:10 .1016 /j .quascirev .2009 .07 773, doi: 10 .1130 /GES00620 .1 . ing: Journal of Hydrology (Amsterdam), v. 414–415, .007. Egger, A.E., Glen, J.M.G., and McPhee, D.K., 2014, Struc- p. 482–490, doi: 10 .1016 /j .jhydrol .2011 .11 .025 . Carmichael, I.S., Lange, R.A., Hall, C.M., and Renne, P.R., tural controls on geothermal circulation in Surprise Hernández-Mendiola, E., Bernal, J.P., Lounejeva, E., 2006, Faulted and tilted Pliocene olivine-tholeiite lavas Valley, California: A re-evaluation of the Lake City Mortimer, G.E., and McCulloch, M.T., 2011, U- near Alturas, NE California, and their bearing on the fault zone: Geological Society of America Bulletin, series dating of carbonates using inductively coupled uplift of the Warner Range: Geological Society of v. 126, p. 523–531, doi: 10 .1130 /B30785 .1 . plasma–quadrupole mass spectrometry: Quaternary America Bulletin, v. 118, p. 1196–1211, doi:10 .1130 Ehlers, J., Gibbard, P.L., and Hughes, P.D., 2011, Quater- Geochronology, v. 6, p. 564–573, doi: 10 .1016 /j .quageo /B25918.1 . nary Glaciations: Extent and Chronology: A Closer .2011 .09 .001 . Cassata, W.S., Singer, B.S., Liddicoat, J.C., and Coe, R.S., Look: Amsterdam, Elsevier, 1108 p. Hostetler, S.W., and Benson, L.V., 1990, Paleoclimatic im- 2010, Reconciling discrepant chronologies for the Eugster, H.P., and Kelts, K., 1983, Lacustrine chemical sedi- plications of the high stand of Lake Lahontan derived geomagnetic excursion in the Mono Basin, California: ments, in Goudie, A.S., and Pye, K., eds., Chemical from models of evaporation and lake level: Climate Insights from new 40Ar/39Ar dating experiments and a Sediments and Geomorphology: Precipitates and Re- Dynamics, v. 4, p. 207–217, doi: 10 .1007 /BF00209522 . revised relative paleointensity correlation: Quaternary sidua in the Near-Surface Environment: San Francisco, Hostetler, S.W., and Benson, L.V., 1994, Stable isotopes of Geochronology, v. 5, p. 533–543, doi: 10 .1016 /j .quageo California, Academic Press, p. 321–368. oxygen and hydrogen in the Truckee River–Pyramid .2010 .02 .001 . Felton, A., Jewell, P.W., Chan, M., and Currey, D., 2006, Lake surface-water system. 2. A predictive model of Chamberlain, C., Wan, X., Graham, S.A., Carroll, A.R., Controls of tufa development in Pleistocene Lake δ18O and δ2H in Pyramid Lake: Limnology and Ocean- Doebbert, A.C., Sageman, B.B., Blisniuk, P., Kent- Bonneville, Utah: The Journal of Geology, v. 114, ography, v. 39, p. 356–364, doi: 10 .4319 /lo .1994 .39 .2 Corson, M.L., Wang, Z., and Chengshan, W., 2013, p. 377–389, doi: 10 .1086 /501218 . .0356 . Stable isotopic evidence for climate and basin evolu- Fleischer, R.L., 1982, Alpha-recoil damage and solution Hostetler, S.W., Giorgi, F., Bates, G., and Bartlein, P., 1994, tion of the Late Cretaceous Songliao Basin, China: effects in minerals: Uranium isotopic disequilibrium Lake-atmosphere feedbacks associated with paleolakes Palaeogeography, Palaeoclimatology, Palaeoecology, and radon release: Geochimica et Cosmochimica Bonneville and Lahontan: Science, v. 263, p. 665–668, v. 385, p. 106–124, doi: 10 .1016 /j .palaeo .2012 .03 .020 . Acta, v. 46, p. 2191–2201, doi: 10 .1016 /0016 -7037 doi: 10 .1126 /science.263 .5147 .665 . Cheng, H., Edwards, R., Hoff, J., Gallup, C., Richards, D., (82)90194 -6 . Hubbs, C.L., and Miller, R.R., 1948, The zoological evi- and Asmerom, Y., 2000, The half-lives of uranium-234 Fletcher, K.E., Sharp, W.D., Kendrick, K.J., Behr, W.M., dence: Correlation between fi sh distribution and hydro- and thorium-230: Chemical Geology, v. 169, p. 17–33. Hudnut, K.W., and Hanks, T.C., 2010, 230Th/U dating graphic history in the desert basins of western United Cheng, H., Lawrence Edwards, R., Shen, C.-C., Polyak, of a late Pleistocene alluvial fan along the southern San States, in The Great Basin with emphasis on glacial V.J., Asmerom, Y., Woodhead, J., Hellstrom, J., Wang, Andreas fault: Geological Society of America Bulletin, and postglacial times: Bulletin of the University of Y., Kong, X., and Spötl, C., 2013, Improvements in v. 122, p. 1347–1359, doi: 10 .1130 /B30018 .1 . Utah, v. 38, no. 20, p. 17–166. 230Th dating, 230Th and 234U half-life values, and U-Th Gabitov, R.I., and Watson, E.B., 2006, Partitioning of Ingraham, N.L., and Taylor, B.E., 1989, The effect of snow- isotopic measurements by multi-collector inductively strontium between calcite and fl uid: Geochemistry melt on the hydrogen isotope ratios of creek discharge coupled plasma mass spectrometry: Earth and Plan- Geophysics Geosystems, v. 7, Q11004, doi:10 .1029 in Surprise Valley, California: Journal of Hydrology etary Science Letters, v. 371, p. 82–91, doi: 10 .1016 /j /2005GC001216 . (Amsterdam), v. 106, p. 233–244, doi: 10 .1016 /0022 .epsl .2013 .04 .006 . García, A.F., and Stokes, M., 2006, Late Pleistocene high- -1694 (89)90074 -7 . Clague, J.J., and James, T.S., 2002, History and isostatic ef- stand and recession of a small, high-altitude pluvial Irwin, R.P., and Zimbelman, J.R., 2012, Morphometry of fects of the last ice sheet in southern British Columbia: lake, Jakes Valley, central Great Basin, USA: Quater- Great Basin pluvial shore landforms: Implications for Quaternary Science Reviews, v. 21, p. 71–87, doi: 10 nary Research, v. 65, p. 179–186, doi: 10 .1016 /j .yqres paleolake basins on Mars: Journal of Geophysical Re- .1016 /S0277 -3791 (01)00070 -1 . .2005 .08 .025 . search–Planets (1991–2012), v. 117, E07004, doi:10 Clawson, R.F., Gibson, L.R., and , S.L., 1986, Sur- Garnett, E.R., Gilmour, M.A., Rowe, P.J., Andrews, J.E., .1029 /2012JE004046 . prise Valley Ground Water Basin Water Quality Study: and Preece, R.C., 2004, 230Th/234U dating of Holocene Israelson, C., Björck, S., Hawkesworth, C.J., and Possnert, Sacramento, California Department of Water Re- tufas: Possibilities and problems: Quaternary Science G., 1997, Direct U-Th dating of organic- and carbon- sources–Northern District, 180 p. Reviews, v. 23, p. 947–958, doi:10 .1016 /j .quascirev ate-rich lake sediments from southern Scandinavia: COHMAP, 1988, Climatic changes of the last 18,000 years: .2003 .06 .018 . Earth and Planetary Science Letters, v. 153, p. 251– Observations and model simulations: Science, v. 241, Gent, P.R., Danabasoglu, G., , L.J., Holland, M.M., 263, doi: 10 .1016 /S0012 -821X (97)00184 -2 . p. 1043–1052, doi: 10 .1126 /science.241 .4869 .1043 . Hunke, E.C., Jayne, S.R., Lawrence, D.M., Neale, Jaffey, A.H., Flynn, K.F., Glendenin, L.E., Bentley, W.C., Currey, D.R., 1990, Quaternary palaeolakes in the evolution R.B., Rasch, P.J., Vertenstein, M., Worley, P.H., Yang, and Essling, A.M., 1971, Precision measurement of of semidesert basins, with special emphasis on Lake Z.-L., and Zhang, M., 2011, The Community Climate half-lives and specifi c activities of U-235 and U-238: Bonneville and the Great Basin, U.S.: Palaeogeogra- System Model Version 4: Journal of Climate, v. 24, Physical Review C, v. 4, p. 1889–1906, doi:10 .1103 phy, Palaeoclimatology, Palaeoecology, v. 76, p. 189– p. 4973–4991, doi: 10 .1175 /2011JCLI4083 .1 . /PhysRevC .4 .1889 . 214, doi:10 .1016 /0031 -0182 (90)90113 -L . Godsey, H.S., Oviatt, C.G., Miller, D.M., and Chan, M.A., James, S.R., 1983, Surprise Valley settlement and subsis- Davis, S.J., Mix, H.T., Wiegand, B.A., Carroll, A.R., and 2011, Stratigraphy and chronology of offshore to near- tence: A critical review of the faunal evidence: Jour- Chamberlain, C., 2009, Synorogenic evolution of shore deposits associated with the Provo shoreline, nal of California and Great Basin Anthropology, v. 5, large-scale drainage patterns: Isotope paleohydrology Pleistocene Lake Bonneville, Utah: Palaeogeography, p. 156–175. of sequential Laramide basins: American Journal of Palaeoclimatology, Palaeoecology, v. 310, p. 442–450, Jewell, P.W., 2007, Morphology and paleoclimatic signifi - Science, v. 309, p. 549–602, doi: 10 .2475 /07 .2009 .02 . doi: 10 .1016 /j .palaeo .2011 .08 .005 . cance of Pleistocene Lake Bonneville spits: Quaternary Denniston, R.F., Asmerom, Y., Polyak, V., Dorale, J.A., Car- Haase-Schramm, A., Goldstein, S.L., and Stein, M., 2004, Research, v. 68, p. 421–430, doi: 10 .1016 /j .yqres .2007 penter, S.J., Trodick, C., Hoye, B., and González, L.A., U-Th dating of Lake Lisan (late Pleistocene Dead Sea) .07 .004 . 2007, Synchronous millennial-scale climatic changes aragonite and implications for glacial East Mediter- Jones, M.D., and Imbers, J., 2010, Modeling Mediterranean in the Great Basin and the North Atlantic during the ranean climate change: Geochimica et Cosmochimica lake isotope variability: Global and Planetary Change, last interglacial: Geology, v. 35, p. 619, doi:10 .1130 Acta, v. 68, p. 985–1005, doi: 10 .1016 /j .gca .2003 .07 v. 71, p. 193–200, doi: 10 .1016 /j .gloplacha .2009 .10 /G23445A .1 . .016 . .001 .

26 Geological Society of America Bulletin, Month/Month 2014 Geological Society of America Bulletin, published online on 2 June 2014 as doi:10.1130/B31014.1

Lacustrine paleoclimate record from late Pleistocene Lake Surprise, California

Jones, M.D., Roberts, C.N., and Leng, M.J., 2007, Quantify- Astronomy and Astrophysics, v. 428, p. 261–285, doi: Majoube, F., 1971, Fractionnement en oxygène-18 et un ing climatic change through the last glacial–intergla- 10 .1051 /0004 -6361: 20041335 . deutérium entre l’eau et sa vapeur: The Journal of cial transition based on lake isotope palaeohydrology Li, H.C., and Ku, T.L., 1997, δ13C–δ18C covariance as a Chemical Physics, v. 187, p. 1423–1436. from central Turkey: Quaternary Research, v. 67, paleohydrological indicator for closed-basin lakes: Matsubara, Y., and Howard, A.D., 2009, A spatially explicit p. 463–473, doi: 10 .1016 /j .yqres .2007 .01 .004 . Palaeogeography, Palaeoclimatology, Palaeoecol- model of runoff, evaporation, and lake extent: Applica- Kageyama, M., Braconnot, P., Bopp, L., Caubel, A., Foujols, ogy, v. 133, p. 69–80, doi:10 .1016 /S0031 -0182 tion to modern and late Pleistocene lakes in the Great M.A., Guilyardi, E., Khodri, M., Lloyd, J., Lombard, (96)00153 -8 . Basin region, western United States: Water Resources F., Mariotti, V., Marti, O., Roy, T., and Woillez, M.-N., Licciardi, J.M., 2001, Chronology of latest Pleistocene lake- Research, v. 45, p. 1–33, doi: 10 .1029 /2007WR005953 . 2013a, Mid-Holocene and Last Glacial Maximum cli- level fl uctuations in the pluvial Lake Chewaucan basin, McGee, D., Quade, J., Edwards, R.L., Broecker, W.S., mate simulations with the IPSL model-part I: Compar- Oregon, USA: Journal of Quaternary Science, v. 16, Cheng, H., Reiners, P.W., and Evenson, N., 2012, ing IPSL_CM5A to IPSL_CM4: Climate Dynamics, p. 545–553, doi: 10 .1002 /jqs .619 . Lacus trine cave carbonates: Novel archives of paleo- v. 40, p. 2447–2468, doi: 10 .1007 /s00382 -012 -1488 -8 . Licciardi, J.M., , P.U., Brook, E.J., Elmore, D., and hydrologic change in the Bonneville Basin (Utah, Kageyama, M., Braconnot, P., Bopp, L., Mariotti, V., Sharma, P., 2004, Variable responses of western U.S. USA): Earth and Planetary Science Letters, v. 351, Roy, T., Woillez, M.N., Caubel, A., Foujols, M.-A., glaciers during the last deglaciation: Geology, v. 32, p. 182–194, doi: 10 .1016 /j .epsl .2012 .07 .019 . Guilyardi , E., Khodri, M., Lloyd, J., Lombard, F., p. 81, doi: 10 .1130 /G19868 .1 . McManus, J.F., Francois, R., Gherardi, J.M., Keigwin, L.D., and Marti, O., 2013b, Mid-Holocene and Last Glacial Lillquist, K.D., 1994, Late Quaternary Lake Franklin: and Brown-Leger, S., 2004, Collapse and rapid re- Maximum climate simulations with the IPSL model: Lacus trine Chronology, Coastal Geomorphology, and sumption of Atlantic meridional circulation linked to Part II: Model-data comparisons: Climate Dynamics, Hydro-Isostatic Defl ection in Ruby Valley and North- deglacial climate changes: Nature, v. 428, p. 834–837, v. 40, p. 2469–2495, doi: 10 .1007 /s00382 -012 -1499 -5 . ern Butte Valley, Nevada [Ph.D. Thesis]: Salt Lake doi: 10 .1038 /nature02494 . Kaufman, A., 1971, U-series dating of Dead Sea Basin car- City, Utah, University of Utah, 370 p. Menking, K.M., Anderson, R.Y., Shafi ke, N.G., Syed, K.H., bonates: Geochimica et Cosmochimica Acta, v. 35, Lin, J., Broecker, W., Anderson, R., Hemming, S., Ruben- and Allen, B.D., 2004, Wetter or colder during the Last p. 1269, doi: 10 .1016 /0016 -7037 (71)90115 -3 . stone, J., and Bonani, G., 1996, New 230Th/U and 14C Glacial Maximum? Revisiting the pluvial lake question Kaufman, A., 1993, An evaluation of several methods for ages from Lake Lahontan carbonates, Nevada, USA, in southwestern North America: Quaternary Research, determining 230Th-U ages in impure carbonates: Geo- and a discussion of the origin of initial thorium: Geo- v. 62, p. 280–288, doi: 10 .1016 /j .yqres .2004 .07 .005 . chimica et Cosmochimica Acta, v. 57, p. 2303–2317, chimica et Cosmochimica Acta, v. 60, p. 2817–2832, Miffl in, M.D., and Wheat, M.M., 1979, Pluvial Lakes and doi: 10 .1016 /0016 -7037 (93)90571 -D . doi: 10 .1016 /0016 -7037 (96)00136 -6 . Estimated Pluvial Climates of Nevada: Nevada Bureau Kaufman, A., and Broecker, W.S., 1965, Comparison Lin, J.C., Broecker, W.S., Hemming, S.R., Hajdas, I., An- of Mines and Geology Bulletin 94, 57 p. of Th230 and C14 ages for carbonate materials from derson, R.F., Smith, G.I., Kelley, M., and Bonani, G., Miller, D.M., Oviatt, C.G., and McGeehin, J.P., 2013, Stra- Lakes Lahontan and Bonneville: Journal of Geo- 1998, A reassessment of U-Th and 14C ages for late- tigraphy and chronology of Provo shoreline deposits physical Research, v. 70, p. 4039–4054, doi:10 .1029 glacial high-frequency hydrological events at Searles and lake-level implications, late Pleistocene Lake /JZ070i016p04039 . Lake, California: Quaternary Research, v. 49, p. 11– Bonneville, eastern Great Basin, USA: Boreas, v. 42, Kigoshi, K., 1971, Alpha-recoil thorium-234: Dissolution 23, doi: 10 .1006 /qres .1997 .1949 . p. 342–361, doi: 10 .1111 /j .1502 -3885 .2012 .00297 .x . into water and the uranium-234/uranium-238 disequi- Linacre, E., 1992, Climate Data and Resources: A Reference Moeyersons, J., Nyssen, J., Poesen, J., Deckers, J., and Haile, librium in nature: Science, v. 173, p. 47–48, doi:10 and Guide: London, Routledge, 366 p. M., 2006, Age and backfi ll/overfi ll stratigraphy of two .1126 /science.173 .3991 .47 . Ludwig, K.R., 2012, User’s Manual for Isoplot 3.75: A tufa dams, Tigray Highlands, Ethiopia: Evidence for Kim, S.J., Crowley, T.J., Erickson, D.J., Govindasamy, B., Geochronologic Toolkit for Microsoft Excel: Berkeley late Pleistocene and Holocene wet conditions: Palaeo- Duffy, P.B., and Lee, B.Y., 2008, High-resolution cli- Geochronology Center Special Publication 5, 75 p. geography, Palaeoclimatology, Palaeoecology, v. 230, mate simulation of the Last Glacial Maximum: Cli- Ludwig, K.R., and Paces, J.B., 2002, Uranium-series dating p. 165–181, doi: 10 .1016 /j .palaeo .2005 .07 .013 . mate Dynamics, v. 31, p. 1–16, doi: 10 .1007 /s00382 of pedogenic silica and carbonate, Crater Flat, Nevada: Müller, G., Irion, G., and Forstner, U., 1972, Formation -007 -0332 -z . Geochimica et Cosmochimica Acta, v. 66, p. 487–506, and diagenesis of inorganic Ca-Mg carbonates in the Kim, S.T., and O’Neil, J.R., 1997, Equilibrium and nonequi- doi: 10 .1016 /S0016 -7037 (01)00786 -4 . lacustrine environment: Naturwissenchaften, v. 59, librium oxygen isotope effects in synthetic carbonates: Ludwig, K., and Titterington, D., 1994, Calculation of iso- p. 158–164, doi: 10 .1007 /BF00637354 . Geochimica et Cosmochimica Acta, v. 61, p. 3461– chrons, ages, and errors: Geochimica et Cosmochimica Munroe, J.S., and Laabs, B.J.C., 2011, New investigations of 3475, doi:10 .1016 /S0016 -7037 (97)00169 -5 . Acta, v. 58, p. 5031–5042, doi: 10 .1016 /0016 -7037 Pleistocene glacial and pluvial records in northeastern Kirby, M.E., Feakins, S.J., Bonuso, N., Fantozzi, J.M., and (94)90229 -1 . Nevada, in Lee, J., and , J.P., eds., Geologic Field Hiner, C.A., 2013, Latest Pleistocene to Holocene Luo, S., and Ku, T.L., 1991, U-series isochron dating: A gen- Trips to the Basin and Range, Rocky Mountains, Snake hydroclimates from Lake Elsinore, California: Qua- eralized method employing total-sample dissolution: River Plain, and Terranes of the U.S. Cordillera: Geo- ternary Science Reviews, v. 76, p. 1–15, doi: 10 .1016 /j Geochimica et Cosmochimica Acta, v. 55, p. 555–564, logical Society of America Field Guide 21, p. 1–25. .quascirev .2013 .05 .023 . doi: 10 .1016 /0016 -7037 (91)90012 -T . Munroe, J.S., and Laabs, B.J., 2012, Temporal correspon- Ku, T.L., 2000, Uranium-series methods, in Noller, J.S., Luo, X.Z., Rehkamper, M., Lee, D.C., and Halliday, A.N., dence between pluvial lake highstands in the southwest- Sowers, J.M., and Lettis, W.R., eds., Quaternary Geo- 1997, High precision 230Th/232Th and 234U/238U mea- ern US and Heinrich Event 1: Journal of Quaternary chronology: Methods and Applications, AGU Refer- surements using energy-fi ltered ICP magnetic sector Science, v. 28, no. 1, p. 49–58, doi: 10 .1002 /jqs .2586 . ence Shelf: Washington, D.C., American Geophysical multiple collector mass spectrometry: International Munroe, J.S., and Laabs, B.J.C., 2013, Latest Pleistocene Union, p. 101–114. Journal of Mass Spectrometry, v. 171, p. 105–117, doi: history of pluvial Lake Franklin, northeastern Nevada, Ku, T.L., and Liang, Z.-C., 1984, The dating of impure 10 .1016 /S0168 -1176 (97)00136 -5 . USA: Geological Society of America Bulletin, v. 125, carbonates with decay-series isotopes: Nuclear In- Lyle, M., Heusser, L., Ravelo, C., Yamamoto, M., Barron, p. 322–342, doi: 10 .1130 /B30696 .1 . struments and Methods in Physics Research, v. 223, J., Diffenbaugh, N.S., Herbert, T., and Andreasen, D., Negrini, R.M., 2002, Pluvial lake sizes in the northwest- p. 563–571, doi: 10 .1016 /0167 -5087 (84)90710 -5 . 2012, Out of the Tropics: The Pacifi c, Great Basin ern Great Basin throughout the Quaternary Period, in Ku, T.L., Luo, S., Lowenstein, T.K., Li, J., and Spencer, R.J., lakes, and late Pleistocene water cycle in the western Hershel, R., ed., Great Basin Aquatic Systems His- 1998, U-series chronology of lacustrine deposits in United States: Science, v. 337, p. 1629–1633, doi:10 tory: Washington, D.C., Smithsonian Institute Press, Death Valley, California: Quaternary Research, v. 50, .1126 /science .1218390 . p. 11–52. p. 261–275, doi: 10 .1006 /qres .1998 .1995 . Maher, K., DePaolo, D.J., Conrad, M.E., and Serne, R.J., Neymark, L.A., 2011, Potential effects of alpha-recoil on Kurth, G., Phillips, F.M., Reheis, M.C., Redwine, J.L., and 2003, Vadose zone infi ltration rate at Hanford, Wash- uranium-series dating of calcrete: Chemical Geology, Paces, J.B., 2011, Cosmogenic nuclide and uranium- ington, inferred from Sr isotope measurements: Water v. 282, p. 98–112, doi: 10 .1016 /j .chemgeo .2011 .01 series dating of old, high shorelines in the western Resources Research, v. 39, 1204, doi:10 .1029 .013 . Great Basin, USA: Geological Society of America /2002WR001742 . Neymark, L.A., and Paces, J.B., 2000, Consequences of Bulletin, v. 123, p. 744–768, doi: 10 .1130 /B30010 .1 . Maher, K., DePaolo, D.J., and Christensen, J.N., 2006, U-Sr slow growth for 230Th/U dating of Quaternary opals, Laabs, B.J.C., Munroe, J.S., Best, L.C., and Caffee, M.W., isotopic speedometer: Fluid fl ow and chemical weath- Yucca Mountain, NV, USA: Chemical Geology, v. 164, 2013, Timing of the last glaciation and subsequent de- ering rates in aquifers: Geochimica et Cosmochimica p. 143–160, doi: 10 .1016 /S0009 -2541 (99)00142 -4 . glaciation in the Ruby Mountains, Great Basin, USA: Acta, v. 70, p. 4417–4435, doi: 10 .1016 /j .gca .2006 .06 O’Connell, J.F., and Inoway, C.M., 1994, Surprise Valley Earth and Planetary Science Letters, v. 361, p. 16–25, .1559 . projectile points and their chronological implications: doi: 10 .1016 /j .epsl .2012 .11 .018 . Maher, K., Wooden, J.L., Paces, J.B., and Miller, D.M., Journal of California and Great Basin Anthropology, Laîné, A., Kageyama, M., Salas-Mélia, D., Voldoire, A., 2007, 230Th-U dating of surfi cial deposits using the ion v. 16, no. 2, p. 162–198. Rivière, G., Ramstein, G., Planton, S., Tyteca, S., and microprobe (SHRIMP-RG): A microstratigraphic per- Okumura, Y.M., Deser, C., Hu, A., Timmermann, A., and Peterschmitt, J.Y., 2009, Northern Hemisphere storm spective: Quaternary International, v. 166, p. 15–28, Xie, S.P., 2009, North Pacifi c climate response to fresh- tracks during the Last Glacial Maximum in the PMIP2 doi: 10 .1016 /j .quaint .2007 .01 .003 . water forcing in the subarctic North Atlantic: Oceanic ocean-atmosphere coupled models: Energetic study, Maher, K, Ibarra, D.E., Oster, J.L., Miller, D.M., Redwine, and atmospheric pathways: Journal of Climate, v. 22, seasonal cycle, precipitation: Climate Dynamics, v. 32, J.L., Reheis, M.C., and Harden, J.W., 2014, Ura- p. 1424–1445, doi: 10 .1175 /2008JCLI2511 .1 . p. 593–614, doi: 10 .1007 /s00382 -008 -0391 -9 . nium isotopes in soils as a proxy for past infi ltration Orme, A.R., 2008, Pleistocene pluvial lakes of the American Laskar, J., Robutel, P., Joutel, F., Gastineau, M., Correia, and precipitation across the western United States: West: A short history of research: Geological Society, A.C.M., and Levrard, B., 2004, A long-term numeri- American Journal of Science, v. 314, p. 821–857, doi: London, Special Publications, v. 301, p. 51–78, doi:10 cal solution for the insolation quantities of the Earth: 10.2475 /04 .2014 .01 . .1144 /SP301 .4 .

Geological Society of America Bulletin, Month/Month 2014 27 Geological Society of America Bulletin, published online on 2 June 2014 as doi:10.1130/B31014.1

Ibarra et al.

Osmond, J.K., May, J.P., and Tanner, W.F., 1970, Age of the Analytical Chemistry, v. 361, p. 400–403, doi: 10 .1007 Sharp, W.D., Ludwig, K.R., , O.A., Amundson, Cape Kennedy barrier-and-lagoon complex: Journal /s002160050914 . R., and Glaser, L.L., 2003, Dating fl uvial terraces by of Geophysical Research–Planets (1991–2012), v. 75, Rasmussen, S.O., Andersen, K.K., Svensson, A.M., Stef- 230Th/U on pedogenic carbonate, Wind River Basin, p. 469–479. fensen, J.P., Vinther, B.M., Clausen, H.B., Sig- Wyoming: Quaternary Research, v. 59, p. 139–150, Oster, J.L., Montañez, I.P., Sharp, W.D., and , K.M., gaard-Andersen, M.L., Johnsen, S.J., Larsen, L.B., doi: 10 .1016 /S0033 -5894 (03)00003 -6 . 2009, Late Pleistocene California droughts during de- Dahl-Jensen, D., Bigler, M., Rothlisberger, R., , Shen, C.C., Lawrence Edwards, R., Cheng, H., Dorale, J.A., glaciation and Arctic warming: Earth and Planetary H., Goto-Azuma, K., Hansson, M.E., and Ruth, U., Thomas, R.B., Bradley Moran, S., Weinstein, S.E., and Science Letters, v. 288, p. 434–443, doi: 10 .1016 /j .epsl 2006, A new Greenland ice core chronology for the Edmonds, H.N., 2002, Uranium and thorium isotopic .2009 .10 .003 . last glacial termination: Journal of Geophysical Re- and concentration measurements by magnetic sector Oster, J.L., Ibarra, D., Harris, C.R., and Maher, K., 2012, search, ser. D, Atmospheres, v. 111, D06102, doi:10 inductively coupled plasma mass spectrometry: Chem- Infl uence of eolian deposition and rainfall amounts .1029 /2005JD006079 . ical Geology, v. 185, p. 165–178, doi: 10 .1016 /S0009 on the U-isotopic composition of soil water and soil Redwine, J.L., 2003, The Quaternary Pluvial History and Pa- -2541 (01)00404 -1 . minerals: Geochimica et Cosmochimica Acta, v. 88, leoclimate Implications of Newark Valley, East-Central Shen, C.C., Wu, C.C., Cheng, H., Edwards, R.L., Hsieh, Y.-T., p. 146–166, doi: 10 .1016 /j .gca .2012 .04 .004 . Nevada; Derived from Mapping and Interpretation of Gallet, S., Chang, C.C., Li, T.Y., Lam, D.D., Kano, A., Oviatt, C.G., Currey, D.R., and Sack, D., 1992, Radiocarbon Surfi cial Units and Geomorphic Features [M.S. thesis]: Hori, M., and Spoetl, C., 2012, High-precision and high- chronology of Lake Bonneville, eastern Great Basin, Arcata, California, Humboldt State University, 385 p. resolution carbonate 230Th dating by MC-ICP-MS with USA: Palaeogeography, Palaeoclimatology, Palaeo- Reeder, R.J., Nugent, M., Tait, C.D., Morris, D.E., Heald, SEM protocols: Geochimica et Cosmochimica Acta, ecology, v. 99, p. 225–241, doi: 10 .1016 /0031 -0182 S.M., Beck, K.M., Hess, W.P., and Lanzirotti, A., 2001, v. 99, p. 71–86, doi: 10 .1016 /j .gca .2012 .09 .018 . (92)90017 -Y . Coprecipitation of uranium (VI) with calcite: XAFS, Sims, K.W., Gill, J.B., Dosseto, A., Hoffmann, D.L., Lund- Paces, J.B., Neymark, L.A., Wooden, J.L., and Persing, micro-XAS, and luminescence characterization: Geo- strom, C.C., Williams, R.W., Ball, L., Tollstrup, D., H.M., 2004, Improved spatial resolution for U-series chimica et Cosmochimica Acta, v. 65, p. 3491–3503, Turner, S., and Prytulak, J., 2008, An inter-laboratory dating of opal at Yucca Mountain, Nevada, USA, using doi: 10 .1016 /S0016 -7037 (01)00647 -0 . assessment of the thorium isotopic composition of syn- ion-microprobe and microdigestion methods: Geo- Reheis, M.C., 1999a, Extent of Pleistocene Lakes in the West- thetic and rock reference materials: Geostandards and chimica et Cosmochimica Acta, v. 68, p. 1591–1606, ern Great Basin: U.S. Geological Survey Miscellaneous Geoanalytical Research, v. 32, p. 65–91, doi: 10 .1111 /j doi: 10 .1016 /j .gca .2003 .08 .022 . Field Studies Map MF-2323, scale 1:800,000, 1 sheet. .1751 -908X .2008 .00870 .x . Paces, J.B., Neymark, L.A., Whelan, J.F., Wooden, J.L., Reheis, M., 1999b, Highest pluvial-lake shorelines and Sladek, C., Arehard, G.B., and Benoit, W.R., 2004, Geochem- Lund, S.P., and Marshall, B.D., 2010, Limited hydro- Pleistocene climate of the western Great Basin: Qua- istry of the Lake City Geothermal System, California, logic response to Pleistocene climate change in deep ternary Research, v. 52, p. 196–205, doi:10 .1006 /qres USA: Geothermal Resources Council Transactions, vadose zones—Yucca Mountain, Nevada: Earth and .1999 .2064 . v. 28, p. 363–368. Planetary Science Letters, v. 300, p. 287–298, doi:10 Reheis, M.C., Stine, S., and Sarna-Wojcicki, A.M., 2002, Smith, G.I., 1984, Paleohydrologic regimes in the south- .1016 /j .epsl .2010 .10 .006 . Drainage reversals in Mono Basin during the late Plio- western Great Basin, 0–3.2 my ago, compared with Personius, S.F., Crone, A.J., Machette, M.N., Mahan, S.A., cene and Pleistocene: Geological Society of America other long records of “global” climate: Quaternary and Lidke, D.J., 2009, Moderate rates of late Quater- Bulletin, v. 114, p. 991–1006, doi:10 .1130 /0016 -7606 Research, v. 22, p. 1–17, doi: 10 .1016 /0033 -5894 nary slip along the northwestern margin of the Basin (2002)114 <0991: DRIMBD>2.0 .CO;2 . (84)90002 -4 . and Range Province, Surprise Valley fault, northeast- Reheis, M.C., Bright, J., Lund, S.P., Miller, D.M., Skipp, G., Soligo, M., Tuccimei, P., Barberi, R., Delitala, M.C., Mic- ern California: Journal of Geophysical Research– and Fleck, R.J., 2012, A half-million-year record of cadei, E., and Taddeucci, A., 2002, U/Th dating of Planets (1991–2012), v. 114, B09405, doi:10 .1029 paleoclimate from the core, Mojave Des- freshwater travertine from Middle Velino Valley (cen- /2008JB006164 . ert, California: Palaeogeography, Palaeoclimatology, tral Italy): Paleoclimatic and geological implications: Phillips, F.M., , A.R., Smith, G.I., and Bischoff, Palaeoecology, v. 365, p. 11–37, doi:10 .1016 /j .palaeo Palaeogeography, Palaeoclimatology, Palaeoecol- J.L., 1994, Interstadial climatic cycles—A link be- .2012 .09 .002 . ogy, v. 184, p. 147–161, doi: 10 .1016 /S0031 -0182 tween western North America and Greenland: Geol- Reimer, P.J., and 29 others, 2013, IntCal13 and Marine13 (02)00253 -5 . ogy, v. 22, p. 1115–1118, doi: 10 .1130 /0091 -7613 radiocarbon age calibration curves 0–50,000 years cal Southon, J., Noronha, A.L., Cheng, H., Edwards, R.L., and (1994)022 <1115: ICCALB>2.3 .CO;2 . B.P.: Radiocarbon, v. 55, p. 1869–1887, doi:10 .2458 Wang, Y., 2012, A high-resolution record of atmo- Phillips, F.M., Zreda, M., Plummer, M.A., Elmore, D., and /azu_js_rc.55 .16947 . spheric 14C based on Hulu Cave speleothem H82: Qua- Clark, D.H., 2009, Glacial geology and chronology of Rosenbaum, J.G., and Reynolds, R.L., 2004, Record of Late ternary Science Reviews, v. 33, p. 32–41, doi: 10 .1016 Bishop Creek and vicinity, eastern Sierra Nevada, Cali- Pleistocene Glaciation and Deglaciation in the South- /j .quascirev .2011 .11 .022 . fornia: Geological Society of America Bulletin, v. 121, ern Cascade Range. II. Flux of Glacial Flour in a Sedi- Steinman, B.A., Abbott, M.B., Nelson, D.B., Stansell, N.D., p. 1013–1033, doi: 10 .1130 /B26271 .1 . ment Core from Upper Klamath Lake, Oregon: Journal Finney, B.P., Bain, D.J., and Rosenmeier, M.F., 2013, Placzek, C., Patchett, P.J., Quade, J., and Wagner, J., 2006a, of Paleolimnology, v. 31, p. 235–252, doi: 10 .1023 /B: Isotopic and hydrologic responses of small, closed Strategies for successful U-Th dating of paleolake JOPL .0000019229 .75336 .7a . lakes to climate variability: Comparison of measured carbonates: An example from the Bolivian Altiplano: Rosenbaum, J.G., Reynolds, R.L., and Colman, S.M., 2012, and modeled lake level and sediment core oxygen Geochemistry Geophysics Geosystems, v. 7, Q05024, Fingerprinting of glacial silt in lake sediments yields isotope records: Geochimica et Cosmochimica Acta, doi: 10 .1029 /2005GC001157 . continuous records of alpine glaciation (35–15ka), v. 105, p. 455–471, doi: 10 .1016 /j .gca .2012 .11 .026 . Placzek, C., Quade, J., and Patchett, P.J., 2006b, Geochro- western USA: Quaternary Research, v. 78, p. 333–340, Stirling, C.H., Andersen, M.B., Potter, E.K., and Halliday, nology and stratigraphy of late Pleistocene lake cycles doi: 10 .1016 /j .yqres .2012 .06 .004 . A.N., 2007, Low-temperature isotopic fractionation of on the southern Bolivian Altiplano: Implications for Rosholt, J.N., 1976, 230Th/234U dating of travertine and ca- uranium: Earth and Planetary Science Letters, v. 264, causes of tropical climate change: Geological Society liche rinds: Geological Society of America Abstracts p. 208–225, doi: 10 .1016 /j .epsl .2007 .09 .019 . of America Bulletin, v. 118, p. 515–532, doi:10 .1130 with Programs, v. 8, p. 1076. Stuiver, M., and Reimer, P.J., 1993, Extended 14C database /B25770.1 . Rouchy, J.M., Servant, M., Fournier, M., and Causse, C., and revised Calib 3.0 14C age calibration program: Placzek, C.J., Quade, J., and Patchett, P.J., 2011, Isotopic 1996, Extensive carbonate algal bioherms in upper Radio carbon, v. 35, p. 215–230. tracers of paleohydrologic change in large lakes of Pleistocene saline lakes of the central Altiplano of Bo- Sueyoshi, T., Ohgaito, R., Yamamoto, A., Chikamoto, M.O., the Bolivian Altiplano: Quaternary Research, v. 75, livia: Sedimentology, v. 43, p. 973–993, doi: 10 .1111 /j Hajima, T., Okajima, H., Yoshimori, M., Abe, M., p. 231–244, doi: 10 .1016 /j .yqres .2010 .08 .004 . .1365 -3091 .1996 .tb01514 .x . O’ishi, R., Saito, F., Watanabe, S., Kawamiya, M., and Polyak, V.J., and Asmeron, Y., 2001, Late Holocene climate Russell, I.C., 1884, A Geological Reconnaissance in South- Abe-Ouchi, A., 2013, Set-up of the PMIP3 paleocli- and cultural changes in the southwestern United States: ern Oregon: Annual Report of the Director of the mate experiments conducted using an Earth system Science, v. 294, p. 148, doi: 10 .1126 /science .1062771 . United States Geological Survey, v. 4, p. 431–464. model, MIROC-ESM: Geoscientifi c Model Develop- Polyak, V.J., Asmerom, Y., Burns, S.J., and Lachniet, M.S., Russell, R.J., 1927, The Land Forms of Surprise Valley, ment, v. 6, p. 819–836, doi: 10 .5194 /gmd -6 -819 -2013 . 2012, Climatic backdrop to the terminal Pleistocene Northwestern Great Basin: University of California Sylvestre, F., Servant, M., Servant-Vildary, S., Causse, C., extinction of North American mammals: Geology, Publications in Geography, v. 2, p. 323–358. Fournier, M., and Ybert, J.P., 1999, Lake-level chro- v. 40, p. 1023–1026, doi: 10 .1130 /G33226 .1 . Sack, D., 2002, Fluvial linkages in Lake Bonneville sub- nology on the southern Bolivian Altiplano (18°–23°S) Potter, E.K., Stirling, C.H., Andersen, M.B., and Halliday, basin integration: Great Basin aquatic systems history, during late-glacial time and the early Holocene: Qua- A.N., 2005, High precision collector MC- in Hersler, R., ed., Great Basin Aquatic Systems His- ternary Research, v. 51, p. 54–66, doi:10 .1006 /qres ICPMS thorium isotope ratio determination: Interna- tory: Washington, D.C., Smithsonian Institution Press, .1998 .2017 . tional Journal of Mass Spectrometry, v. 247, p. 10–17, p. 129–144. Tackman, G., 1993, Middle and Late Pleistocene Hydrologic doi: 10 .1016 /j .ijms .2005 .08 .017 . Schofi eld, I., Jewell, P., Chan, M., Currey, D., and Gregory, History of Diamond Valley, Eureka and Elko Counties, Przyblowicz, W., Schwarcz, H.P., and Latham, A.G., 1991, M., 2004, Shoreline development, longshore transport Nevada, with Climatic and Isostatic Implications [M.S. Dirty calcites: 2. Uranium-series dating of artifi cial and surface wave dynamics, Pleistocene Lake Bonne- thesis]: Salt Lake City, Utah, University of Utah, 384 p. calcite detritus mixtures: Chemical Geology, v. 86, ville, Utah: Earth Surface Processes and Landforms, Talbot, M.R., 1990, A review of the palaeohydrological p. 161–178. v. 29, p. 1675–1690, doi: 10 .1002 /esp .1121 . interpretation of carbon and oxygen isotopic ratios in Raptis, K., Mayer, K., Hendrickx, F., and De Bièvre, P., Schwarcz, H.P., and Latham, A.G., 1989, Dirty calcites: primary lacustrine carbonates: Chemical Geology– 1998, Preparation and certifi cation of new thorium 1. Uranium-series dating of contaminated calcite using Isotope Geoscience Section, v. 80, p. 261–279, doi:10 isotopic reference materials: Fresenius’ Journal of leachates alone: Chemical Geology, v. 80, p. 35–43. .1016 /0168 -9622 (90)90009 -2 .

28 Geological Society of America Bulletin, Month/Month 2014 Geological Society of America Bulletin, published online on 2 June 2014 as doi:10.1130/B31014.1

Lacustrine paleoclimate record from late Pleistocene Lake Surprise, California

Tessier, A., Campbell, P.G., and Bisson, M., 1979, Se- Vazquez, J.A., and Lidzbarski, M.I., 2012, High-resolution Bulletin, v. 107, p. 867–876, doi:10 .1130 /0016 -7606 quential extraction procedure for the speciation of tephrochronology of the Wilson Creek Formation (1995)107 <0867: LQVACH>2 .3 .CO;2 . particulate trace metals: Analytical Chemistry, v. 51, (Mono Lake, California) and Laschamp event using York, D., 1968, Least squares fi tting of a straight line with p. 844–851, doi: 10 .1021 /ac50043a017 . 238U-230Th SIMS dating of accessory mineral rims: correlated errors: Earth and Planetary Science Letters, Thomas, L.E., Hawkesworth, C.J., Van Calsteren, P., Turner, Earth and Planetary Science Letters, v. 357, p. 54–67, v. 5, p. 320–324, doi: 10 .1016 /S0012 -821X (68)80059 -7 . S.P., and Rogers, N.W., 1999, Melt generation beneath doi: 10 .1016 /j .epsl .2012 .09 .013 . Yukimoto, S., Adachi, Y., and Hosaka, M., 2012, A new ocean islands: A U-Th-Ra isotope study from Lan- Voldoire, A., Sanchez-Gomez, E., Salas y Melia, D., global climate model of the Meteorological Research zarote in the Canary Islands: Geochimica et Cosmo- Decharme, B., Cassou, C., Senesi, S., Valcke, S., Beau, Institute: MRI-CGCM3: Model description and basic chimica Acta, v. 63, p. 4081–4099, doi: 10 .1016 /S0016 I., Alias, A., , M., Deque, M., Deshayes, performance (Special Issue on Recent Development -7037 (99)00310 -5 . J., Douville, H., Fernandez, E., Madec, G., Maison- on Climate Models and Future Climate Projections): Thompson, R.S., Anderson, K.H., and Bartlein, P.J., 1999, nave, E., Moine, M.-P., Planton, S., Saint-Martin, D., Journal of the Meteorological Society of Japan, v. 90, Quantitative Paleoclimatic Reconstructions from Late Szopa, S., Tyteca, S., Alkama, R., Belamari, S., Braun, p. 23–64, doi: 10 .2151 /jmsj .2012 -A02 . Pleistocene Plant Macrofossils of the Yucca Mountain A., Coquart, L., and Chauvin, F., 2013, The CNRM- Zic, M., Negrini, R.M., and Wigand, P.E., 2002, Evidence Region: U.S. Geological Survey Open-File Report CM5.1 global climate model: Description and basic of synchronous climate change across the Northern 99–338, 38 p. evaluation: Climate Dynamics, v. 40, p. 2091–2121, Hemisphere between the North Atlantic and the north- Thurber, D.L., Broecker, W.S., Blanchard, R.L., and Potratz, doi: 10 .1007 /s00382 -011 -1259 -y . western Great Basin, United States: Geology, v. 30, H.A., 1965, Uranium-series ages of Pacifi c atoll coral: Wagner, J., Cole, J., Beck, J.W., Patchett, P., Henderson, G., p. 635–638, doi: 10 .1130 /0091 -7613 (2002)030 <0635: Science, v. 149, p. 55–58, doi:10 .1126 /science .149 and Barnett, H.R., 2010, Moisture variability in the EOSCCA>2.0 .CO;2 . .3679 .55 . southwestern United States linked to abrupt glacial cli- Zimbelman, J., Garry, W., and Irwin, R.P., III, 2009, Preci- Torfstein, A., Goldstein, S.L., Kagan, E.J., and Stein, M., mate change: Nature Geoscience, v. 3, p. 110–113, doi: sion Topography of Pluvial Features in Western Nevada 2013, Integrated multi-site U-Th chronology of the last 10 .1038 /ngeo707 . as Analogs for Possible Pluvial Landforms on Mars, in glacial Lake Lisan: Geochimica et Cosmochimica Acta, Waitt, R.B., 1985, Case for periodic, colossal jokulhlaups 40th Lunar and Planetary Science Conference, Houston, v. 104, p. 210–231, doi:10 .1016 /j .gca .2012 .11 .003 . from Pleistocene Glacial : Geologi- Texas, Lunar and Planetary Institute, p. 1360. Trumbore, S.E., 2000, Radiocarbon geochronology, in cal Society of America Bulletin, v. 96, p. 1271–1286, Zimmerman, S.R.H., Hemming, S.R., Hemming, N.G., Noller , J.S., Sowers, J.M., and Lettis, W.R., eds., Qua- doi: 10 .1130 /0016 -7606 (1985)96 <1271: CFPCJF>2 .0 Tomascak, P.B., and Pearl, C., 2011, High-resolution ternary Geochronology: Methods and Applications, .CO;2 . chemostratigraphic record of late Pleistocene lake- AGU Reference Shelf: Washington, D.C., American Wells, S.G., Brown, W.J., Enzel, Y., Anderson, R.Y., and level variability, Mono Lake, California: Geological Geophysical Union, p. 41–60. McFadden, L.D., 2003, Late Quaternary geology and Society of America Bulletin, v. 123, p. 2320–2334, doi: Turner, S., Hawkesworth, C., Rogers, N., and King, P., 1997, paleohydrology of pluvial , southern 10 .1130 /B30377 .1 . U-Th isotope disequilibria and ocean island basalt California, in Enzel, Y., Wells, S.G., and Lancaster, Zimmerman, S.R.H., Steponaitis, E., Hemming, S.R., and generation in the Azores: Chemical Geology, v. 139, N., eds., Paleoenvironments and Paleohydrology of the Zermeno, P., 2012, Potential for accurate and precise p. 145–164, doi: 10 .1016 /S0009 -2541 (97)00031 -4 . Mojave and Southern Great Basin Deserts: Geological radiocarbon ages in deglacial-age lacustrine carbon- Turner, S., Van Calsteren, P., Vigier, N., and Thomas, L., 2001, Society of America Special Paper 368, p. 79–114, doi: ates: Quaternary Geochronology, v. 13, p. 81–91, doi: Determination of thorium and uranium isotope ratios in 10 .1130 /0 -8137 -2368 -X .79 . 10 .1016 /j .quageo .2012 .07 .003 . low-concentration geological materials using a fi xed Winograd, I.J., Landwehr, J.M., Coplen, T.B., Sharp, W.D., multi-collector-ICP-MS: Journal of Analytical Atomic Riggs, A.C., Ludwig, K.R., and Kolesar, P.T., 2006, SCIENCE EDITOR: A. HOPE JAHREN Spectrometry, v. 16, p. 612–615, doi: 10 .1039 /b008351h . Devils Hole, Nevada, delta O-18 record extended to ASSOCIATE EDITOR: BENJAMIN J.C. LAABS Vacco, D., Clark, P., Mix, A., Cheng, H., and Edward, R., the mid-Holocene: Quaternary Research, v. 66, p. 202– 2005, A speleothem record of Younger Dryas cool- 212, doi: 10 .1016 /j .yqres .2006 .06 .003 . MANUSCRIPT RECEIVED 16 OCTOBER 2013 ing, Klamath Mountains, Oregon, USA: Quaternary Worona, M.A., and Whitlock, C., 1995, Late Quaternary REVISED MANUSCRIPT RECEIVED 16 APRIL 2014 Research, v. 64, p. 249–256, doi: 10 .1016 /j .yqres .2005 vegetation and climate history near Little Lake, central MANUSCRIPT ACCEPTED 6 MAY 2014 .06 .008 . Coast Range, Oregon: Geological Society of America Printed in the USA

Geological Society of America Bulletin, Month/Month 2014 29 Geological Society of America Bulletin, published online on 2 June 2014 as doi:10.1130/B31014.1

Geological Society of America Bulletin

Rise and fall of late Pleistocene pluvial lakes in response to reduced evaporation and precipitation: Evidence from Lake Surprise, California

Daniel E. Ibarra, Anne E. Egger, Karrie L. Weaver, Caroline R. Harris and Kate Maher

Geological Society of America Bulletin published online 2 June 2014; doi: 10.1130/B31014.1

Email alerting services click www.gsapubs.org/cgi/alerts to receive free e-mail alerts when new articles cite this article Subscribe click www.gsapubs.org/subscriptions/ to subscribe to Geological Society of America Bulletin Permission request click http://www.geosociety.org/pubs/copyrt.htm#gsa to contact GSA Copyright not claimed on content prepared wholly by U.S. government employees within scope of their employment. Individual scientists are hereby granted permission, without fees or further requests to GSA, to use a single figure, a single table, and/or a brief paragraph of text in subsequent works and to make unlimited copies of items in GSA's journals for noncommercial use in classrooms to further education and science. This file may not be posted to any Web site, but authors may post the abstracts only of their articles on their own or their organization's Web site providing the posting includes a reference to the article's full citation. GSA provides this and other forums for the presentation of diverse opinions and positions by scientists worldwide, regardless of their race, citizenship, gender, religion, or political viewpoint. Opinions presented in this publication do not reflect official positions of the Society.

Notes

Advance online articles have been peer reviewed and accepted for publication but have not yet appeared in the paper journal (edited, typeset versions may be posted when available prior to final publication). Advance online articles are citable and establish publication priority; they are indexed by GeoRef from initial publication. Citations to Advance online articles must include the digital object identifier (DOIs) and date of initial publication.

Copyright © 2014 Geological Society of America