<<

REPRINTED FROM

MODELS AND PREHISTORY A SYMPOSIUM

EDITED BY DON D. FOWLER

DESERT RESEARCH INSTITUTE PUBLICATIONS IN THE SOCIAL SCIENCES NO, 12

EDITOR: DON D. FOWLER ASSOCIATE EDITOR: ALMA SMITH

1977 GREAT BASIN LATE QUATERNARY ENVIRONMENTS AND CHRONOLOGY'

PETER J. MEHRINGER, JR. WASHINGTON STATE UNIVERSITY

ABSTRACT

Evidence for the magnitude and chronology of Great Basin climatic change comes from plant macrofossils of dry caves, fluctuating lake levels, spring discharge and tree lines, dune activity, peat formation in salt marshes, and arroyo cutting and filling. About 12,000-10,000 B.P. pluvial lakes began to dry and by 7500 B.P. climatic conditions were much like the present. With minor fluctuations, relative aridity per- sisted until about 4000 B.P. and was followed by increased effective moisture. The dynamic nature of Great Basin environments is apparent whether measured by geological or biological criteria. There were several periods when regional climatic change was sufficient to warrant investigations of its possible influence on human populations; locally, volcanic or tectonic activity may have been even more important. How- ever, instability of the last 10,000 years is no more dramatic than the ecological variation encountered by Great Basin inhabitants within a single year. Variability itself may have been most important in shaping cultural or technological adaptations. The influence of climatic change on man is best considered in terms of evidence for its effect on local resources.

INTRODUCTION

Great Basin anthropologists have assumed an important link between man and environment, and interpretations of cultural change have been influenced by notions of the magnitude and chronology of climatic history. Conversely, varying views on climatic change have been influenced by interpretations of Great Basin cultural history. The relation between climate and culture has been expressed in concern with environmental influences on

113 population density and distribution, resource utilization, and technology (Ranere 1970). Environmental factors have undoubted- ly influenced the course of Great Basin technological and cultural adaptation, just as man's exploitation of natural resources continues to shape the course of civilizations. How- ever, as the smartest (if not always the wisest) and most adaptable of all creatures, man does not usually qualify as a sensitive biological indicator species.

Along with a discussion of the application of paleo- ecological models to Great Basin archeology (see Weide and Weide, this volume), it is important to reconsider the primary data from which inferences are drawn, and the geographical and chronological extent of their application to prehistory. Generalized models relating cultural change to synchronous and unidirectional climatic change are simple to construct and easy to use. For example, for more than 20 years the human response to a supposed extreme mid-postglacial drought pro- vided a focus for constructive discussion and unrestrained argument among Great Basin archeologists. However, the inter- pretation of paleoenvironmental data, the question of their correspondence over the entire Great Basin and their meaning to human history become increasingly less certain as informa- tion from specific localities reveals a chronological and geographical complexity few expected when the last 10,000 years could simply be divided into three geologic-climatic stages and the cultural sequences nearly always fit. We should, perhaps, retreat for a time to less encompassing models and explanations. Fortunately, the Great Basin pro- vides a laboratory of unusual variety and potential for studies of environmental change and human response. In this review I will attempt to describe some of what is known of the magnitude and chronology of Great Basin environmental changes and their possible influences on the prehistoric in- habitants.

The information reviewed herein is not fitted to an exist- ing climatic chronology or its presumed effect on man (D. Fowler 1972a); that has been done before (Baumhoff and Heizer 1965). Rather, the kinds of data, their interpretation, and possible relation to human occupation are presented both chronologically and regionally. Since man's tenure in North America remains uncertain, the last 40,000 years are included but the last 12,000 are emphasized. For convenience the re- view is artificially separated into geological and biological evidence for environmental change. Geological evidence in- cludes lake level fluctuations, glaciation, volcanic and tectonic activity, soils, sediments and erosion, and eolian activity. Biological evidence includes the records of plants and animals, dendroclimatology, and some biogeographic considerations. All of these may be indicators of past climates or directly affect important resources.

114 4

THE GEOLOGICAL EVIDENCE Pluvial Lakes Among the more important aspects of Great Basin geology is evidence for former moist-cool periods accompanying glaci- ation. While the concept of worldwide "pluvials" everywhere coeval with glaciation is questionable, the gross chrono- logical correspondence between Great Basin lakes and continen- tal glaciers clearly illustrates a related cause in Pleistocene atmospheric circulation over North America. Generally low lake levels from about 40,000-30,000 to 24,000 B.P. and deep lakes from 24,000 to 12,000 B.P. are of possible importance to Great Basin occupation. The evidence for pluvial lakes (Figs. 1,2,3) includes strandlines, algal tufas, deltas, bars, and subsurface stratigraphy and sediments. These obvious features of lake and river systems, common throughout the Great Basin, were studied during early geological and hydrological investi- gations (Gale 1914; Gilbert 1890; Meinzer 1922; Russell 1885). The use of "pluvial" to indicate increased precipitation is perhaps misleading, as the exact climatic parameters res- ponsible for any single lake level are unknown. The Spring Valley, , studies of Snyder and Langbein (1962) are the most thorough to date and illustrate the improbability of a single cause; both decreased evaporation and increased pre- cipitation were most probably required for maximum lake levels. Morrison's (1965a:267) use of "lacustral" avoids the connota- tions of pluvial. Galloway (1970:256) substitutes "minevaporal" for pluvial to indicate the importance of decreased temperatures. Estimates of climatic change required to account for high lake levels are given in Table 1 and reviewed by Morrison (1965a:281).

Table 1 Climatic Change Inferred from Estimates of Western Pluvial Lake Water Budgets, as Compared With Historic Mean Annual Climatic Data

Temperature Precipitation Reference (deg. F.) (inches) Meinzer 1922 -15 Leopold 1951 -11 to 12 +7 to 9 Antevs 1952 -5 +6 Broecker and Orr 1958 -9 +8 Snyder and Langbein 1962 -9* +8 Galloway 1970 -18 to 20 -0.6 to 1.7** *Estimated from a 30% (13 inch) decrease in evaporation given by Snyder and Langbein. **Calculated from Table 3 of Galloway.

115 LAKE BONNEVILLE

OREGON 1 NEVADA LAKE LAHONTAN

o 0 a er. '0 a 0 04 2

0 ; p UTAH ‘\ . 0 I ° 4d ,, b 4 \v:\

LAKE SEARLES., Ix\S

e \ , N MOHAVE

(7:9

0 50 100 150 200 250 MILES 0 100 200 300 400 KILOMETERS

Fig. 1. Great Basin pluvial lakes (after Snyder, Hardman and Zdenek 1964)

116 Fig. 2. The Great Bar at Stockton, Utah (Gilbert 1890 , plate IX), a gravel bar separating Tooele and Rush Valleys, is one of the more spectacular features of Lake Bonneville. Its wave cut surface is near- ly 1000 feet above the present Great Salt Lake (April 1970).

Fig. 3. Barren lake beds and strand lines of Lake Searles (Septem- ber 1967).

117 Four pluvial lakes (Bonneville, Lahontan, Searles, and Mohave) have been studied in considerable detail. The magni- tude and chronologies of their fluctuations illustrate present understanding of Great Basin pluvial lake history (Figs. 4,5). Interpretations of the radiocarbon chronologies vary; one im- portant reason is the differences in the reliability of materials dated. Thus, various interpretations depend on the researchers, the dates they accept, and the materials dated (wood, shell, whole core, core humates, strandline tufa, or subsurface car- bonates). For example, discrepancies are apparent in proposed chronologies for Lakes Lahontan and Bonneville (Fig, 4). Broecker and Kaufman's (1965:549, figs. 4,5,6) suggested very deep lake cycles after about 12,000 B.P. conflict with other interpretations (Born 1970:84; Eardley, Gvosdetsky and Marsell 1957:1169-1171; Morrison 1970; Morrison and Frye 1965:23). I interpret selected evidence (Bright 1966:27; Jennings 1957:97) as indicating that Lake Bonneville did not overflow after about 12,500 B.P., fell rapidly to below the level of Danger Cave (about 4300 feet and less than 100 feet above the highest his- toric stand of Great Salt Lake) by about 11,000 B.P. and did not rise above Danger Cave thereafter.

The Searles Lake chronology (Fig. 5), like that of Lahontan and Bonneville, reflects a late Wisconsin period of high water levels following smaller lakes correlated with the preceding interstadial (G. Smith 1967a, 1968, fig. 4; Smith, Friedman and Matsuo 1970). Lake history is recorded in the classic sub- surface stratigraphy with salts representing shallow lakes and muds representing deep lakes. Searles Lake received drainage from the Sierra Nevada via the Owens River, Owens Lake, and China Lake and overflowed into which in turn overflowed to , the ultimate pluvial-maximum sink for the Mohave and Amargosa rivers as well. , Because of its position in a long chain of lakes, the precise chronology of the last major Lake Searles overflow is critical to understand- ing the climatic and biographic history (Miller 1950) of the Death Valley system.

Hooke (1972:2093) assumes that lakes in Death Valley fluctuated in phase with Lake Searles and, further, that Pana- mint Valley most probably overflowed into Death Valley via Wingate Wash between 11,000 and 10,500 B.P. A shallow marsh deposit on the floor of North Panamint Valley is radiocarbon dated at 10,000-10,500 B.P. (Davis, Brott and Weide 1969:15; Mehringer 1967a:172, table 5). These dates place a minimum age on the last major overflow of Lakes Searles and Panamint.

Historically, Lake Mohave has been important in understand- ing man's possible utilization of lake resources during the waning phases of the last pluvial (Campbell and Campbell 1937). Recent geomorphic studies (Ore and Warren 1971, fig. 7) result- ed in the first chronological controls necessary to evaluate the contemporaneity of lake levels and human occupations. Based on dating of tufa or Anodonto carbonate, they suggest overflow

118 O . - A._X 5

VALDERAN

B.P. SUBSTAGE 10,000 — Formation TWOCREEKAN 1 I I I I 1 illiTT • T 7 Harmon School Bonneville Formation 5250 .dendriticSehoo Fm. mbr • Soil 15,000 — .. upper. „member : Bonneville shoreline WOODFORDIAN ` soil SUBSTAGE . . Bonneville Formation ..•.:Pmedale till of - 'lower-Sehoo member Formation 20,000 — TAGE ...white marl member.- S Alluvium, colluvium NAN FARMDALIAN 25,000 — NSI SUBSTAGE Ii I 1 iiIIMMI111} I I:11111111( M 1 Promontory Soil Churchill Soil O 1 { I 30,000 — WISC Alluvium, colluvium Wyemaho Formation

35,000 — ALTONIAN APPROXIMATE AGE IN RADIOCARBON YEARS SUBSTAGE 40,000 — Eetza Formation

45,000 —

Fig. 4. Correlation of Lakes Bonneville and Lahontan, continental and alpine glaciation, soils, and inferred climatic change (after Morrison and Frye 1965; also see Morrison 1965a,b). 960

LAKE MOJAVE WATER LEVELS 940 (Ore and Warren, 1971) ELEV. (FEET)

920

2500

SEARLES LAKE WATER LEVELS Beginning of salt deposition 2000 (Smith, 1970)

1500 ELEV. (FEET)

TULE SPRINGS GEOLOGIC EVENTS (Haynes, 1967)

TULE SPRINGS CLIMATIC EVENTS (Mehringer, 1967a)

YEARS B.P. stages prior to 14,500 B.P., 13,750-12,000 B.P. and 11,000- 9,000 B.P.; a smaller lake is dated at 8500-7500 B.P. (Fig. 5). While Great Basin pluvial lakes have shown significant water budget changes over the last 40,000 years and most, if not all, authors believe that water levels were as low or lower than historic levels by 7000 B.P., there is not necessar- ily agreement on the details of lake chronologies, magnitude of fluctuations, or ultimate climatic cause. In fact, there are good reasons to question climatic correlations over the entire Great Basin. Even if ultimately controlled by external causes, shifting atmospheric circulation patterns (Friedman and Smith 1972) might not be uniform through time over such a large area (Bryson, Baerreis, and Wendland 1970:55). Vari- ations might result from geographic relationship to climatic boundaries (Mitchell 1969, fig. 5.1) and changing latitudinal patterns of seasonal precipitation and temperature (Aschmann 1958). If climate were uniform, all four of the pluvial lake examples would still present certain problems in correlation. As part of large pluvial systems, their histories might be complicated by connecting basins of varying elevations, tribu- tary ice dams (Birkeland 1968), overflow and down cutting (Malde 1968), faulting and isostatic adjustments (Crittenden 1963; Scholz, Barazangi, and Sbar 1971:2980), and by vulcanism (Bright 1967; Mabey 1971). Ideally, a single closed basin that never overflowed, was not an important contributor to a regional ground water flow system (Mifflin 1968; Winograd 1971), and never completely dried at any time in the past 40,000 years would furnish the best regional climatic index for detailed analysis of past environments and their possible influences on man. Minor fluctuations of closed basin lakes over the last 7000 years should be excellent reflections of the climatically controlled balance between inflow and evaporation (Hardman and Venstrom 1941; Peck and Richardson 1966). Also, the avail- ability and aboriginal utilization of lake and marsh resources would have been closely linked to these fluctuations. Pre- liminary pollen and carbonate analyses of a core from Great Salt Lake, Utah, reveal a history of lake oscillations with higher than present levels about 6000, 3600 and 2800 B.P. Probable flooding of the Great Salt Desert, between 3500 and 2200 B.P., is recorded in the subsurface stratigraphy, and in fossil pollen and diatoms of a playa-edge spring-fed salt marsh west of the Terrace Range, Utah (Crescent Spring, Mun- dorff 1971:37). Morrison (1965a:281) suggests several rises of Great Salt Lake within the past 4000 years with the maximum at 4260 feet and equivalent to the Gilbert level (Eardley 1962: 18; Eardley, Gvosdetsky and Marsell 1957, pls. 1,2,3). The archeological chronology of the Great Salt Lake area furnishes additional evidence for former lake levels; Danger

121 Cave was mentioned previously. Several sites along the Bear River and northeast arm of Great Salt Lake (Aikens 1967) lie at an elevation of about 4210 feet, a foot below the highest historic stand (4211 feet in 1873; Peck and Richardson 1966, fig. 4). The sites were flooded within historic time and must have been above the lake when occupied about 1000 B.P. Thus, about 1000 B.P. Great Salt Lake was below 4210 feet. Occupa- tion of Stansbury Island is thought to require a land connection with the mainland to the south. A 7000 B.P. date from Sandwich Shelter is interpreted as a minimum age for a water level lower than the connecting bar at 4206 feet (Marwitt, Fry and Adovasio 1971). Upwarping resulting from isostatic adjustment following initial draining and desiccation of Lake Bonneville (Crittendon 1963, fig. 3) complicates elevational and chronological com- parisons of strandlines. Doming and tilting from the center of the basin has controlled water depth, direction of flow, and threshold elevation between the Great Salt Lake and the Great Salt Desert. These factors in turn have contributed directly to the distribution, extent, and type of marsh and salt flat habi- tats available for human use through the last 11,000 years. Further, habitation sites and communication among aboriginal groups would have been influenced by the position and total area of salt water. In northwestern Utah, isostatic rebound may be archeologically more important than climatic change. Evidence for the recent history of former Lake Lahontan comes from Carson Sink and Pyramid Lake, Nevada. The level of Pyramid Lake is controlled primarily by discharge of the Truckee River originating at Lake Tahoe in the Sierra Nevada (Hyne et al. 1972). From 1840 to 1970 Pyramid Lake fluctuated about 100 feet in depth (Born 1970, fig. 36). However, prior to the 1905 diversion of the Truckee River, Pyramid Lake fluctuated only 20 feet between high water stands of about 3880 and 3860 feet. During the drought of the mid-1960's archeological sites at elevations below any prior historic stands were excavated; one of these at 3788 feet contained Artemisia wood radiocarbon dated at 24801120 B.P. (GaK-2386). Another site at 3878 feet may have been flooded by the highest historic stands; human bone collagen yielded a date of about 1800 B.P. (Tuohy and Stein 1969:100). Radiocarbon dating and stratigraphy of the Truckee River Delta at Pyramid Lake are interpreted as indicating two lake cycles in the past 4000 years (Born 1970, fig. 29). The lake is thought to have been low from 8000 to 3500 B.P. and again between about 2000 and 1000 B.P. The high stand of the most recent lake cycle dates from the past few hundred years. Five lake cycles are described from Carson Sink (Morrison and Frye 1965:19). They represent lake rises, with intervening desiccations or near desiccation, over the last 4000 years, following a long dry period (Morrison 1964:75). The last major lake maximum occurred within the past 100 years. The most

122 severe period of desiccation is thought to have occurred about 600 B.P. between the second and third lake cycles (Morrison 1965a:281). Lack of sufficient radiocarbon control and appar- ent differences in the ages and number of lake cycles recognized for Pyramid Lake and Carson Sink preclude correlation. Mor- rison (1964:105, fig. 39) relates lake history to archeological sites and speculates on the relationship between lake levels, climate, and aboriginal occupation.

In the Southern Great Basin recent lakes have been des- cribed from Searles and Death Valleys. The Searles 'Lake Over- burden Mud represents at least one shallow lake stand, in the last 6000 years or so, following a period of desiccation represented by the Upper Salt (Smith 1968). The radiocarbon dates, on a variety of material (G. Smith 1967b, fig. 12), do not permit fine chronological discrimination; however, one wood date of about 3500 B.P. serves as a maximum age for its burial in lake deposits. In Death Valley, a 30 foot deep lake is recorded in shore lines and salt deposition (Hunt et al. 1966:48; Hunt and Mabey 1966:82). A minimum age of about 2000 years is based on overlying dunes lacking artifacts older than those characteristic of Death Valley III occupation (Hunt 1960: 111).

Glaciation

The chronology and magnitude of alpine glaciation may be important to understanding prehistoric man in the Great Basin in the same ways as lake histories: 1. Glacial history may furnish clues to regional climatic change. 2. Former extent of glaciers clearly limits areas suitable for communication, exploitation, and occupation.

Possible correlations between pluvial lakes and glaciers In the Sierra Nevada or Wasatch Mountains (Gilbert 1890:318; Morrison 1965a:275, table 1; Richmond 1964, fig. 17; Russell 1889:369; Smith 1968, fig. 10) are based on a very few ex- posures of direct contact between glacial and lake deposits. Where they do occur, as at Little Cottonwood and Bells Canyon, Utah (Richmond 1964), they are probably older than 40,000 B.P. and thus beyond the scope of this review (Fig. 6). Glacial out- wash from the Sierra Nevada has been traced to and correlated with Lake Lahontan deposits (Birkeland 1968, table 2).

In the Sierra Nevada, three major Wisconsin glaciations are recognized (Tahoe, Tenaya, and Tioga). Within the last 10,500 years four glaciations are recognized: the Hilgard (10,500-9000 B.P.), Recess Peak (2600-2000 B.P.), Matthes (700 B.P.), and an unnamed advance of about 1100 B.P. (Curry 1971, table 1; 1969, fig. 10; Sharp 1972, table 1). In the Colorado Front Range, glacial advances of the last 10,000 years are dated at 10,000-7500, 4500-2600, 1850-950, and 350-100 B.P.

123 Fig. 6. View of the moraines and lake flats at the mouth of Little Cottonwood Canyon near Salt Lake City, Utah. Note the glacier-carved U-shaped canyon and the terminal moraine of Bells Canyon on the right (from Richmond 1964).

Fig. 7. View of Curelom Cirque, Raft River Mountains, Utah (August 1970). The bog, at 9300 feet, is formed behind the terminal moraine in the foreground and contains Mazama ash (Mehringer, Nash and Fuller 1971). A basal radiocarbon date of 12,000 B.P. gives a minimum age for ice withdrawal; in- itiation of organic deposition is dated at about 11,000 B.P.

124 (Benedict 1970, table 1). A radiocarbon date of about 12,000 B.P. from the Raft River Mountains, Utah (Fig. 7), establishes a minimum age for ice withdrawal (Mehringer, Nash and Fuller 1971).

Insufficient data preclude suggesting a Great Basin glacial chronology for the past 40,000 years. While radio- carbon dating, dendrochronology, lichenometry, tephrochronology, and degree of weathering have been applied in some areas, most of the isolated large mountains have little or no chronological information applicable to correlations with each other or archeological sequences. The distribution of glaciers and Pleistocene snowlines are given by Flint (1971, table 18-B, fig. 18-4). Reviews by Denton and Karl& (1973), Denton and Porter (1970), and Porter (1971) summarize the glacial chronol- ology for western North America.

Volcanic and Tectonic Activity

Volcanic ashes serve as important dated Pleistocene strati- graphic markers throughout the Great Basin (Izett, Wilcox, Powers, and Desborough 1970, fig. 1; Sheppard and Gude 1968). Extensive ash deposits of late Quaternary age are also wide- spread and distinguishable by their chemical composition, stratigraphic occurrence and age. The most important of these was derived from the eruption of Mount Mazama, the precursor of Crater Lake, Oregon, dated at about 7000 B.P. (Kittleman 1973: 2959; Randle, Gales, and Kittleman 1971:262). Its chrono- logical use is well established in archeological investigations in eastern Oregon (Bedwell 1970; Cressman and collaborators 1942). The reported distribution of Mazama ash also includes most of the northern Great Basin as far south , as the central Sierra Nevada, , and east to northwestern Utah (Adam 1967; Mehringer, Nash and Fuller 1971; Powers and Wilcox 1964). While not specifically identified, many of the ash occurrences from natural (Hawley and Wilson 1965:50; Morrison 1964:76-77) and archeological sites of northern Nevada •( Grosscup 1956; Heizer 1951) may be from Mount Mazama. It seems only a matter of time until many more archeological and paleoenvironmental localities from the northern Great Basin of Oregon, California, Nevada, Utah and Idaho will be precisely correlated, by the stratigraphic occurrence of ash marker beds, with time-parallel events throughout the northwestern United States and south- western Canada.

Other volcanic ejecta of probable chronological significance in the northwestern Basin and Range Province include those from Newberry Crater, Oregon; Mount St. Helens, Washington; and pyroclastic layers of uncertain affinities. The Newberry eruptions date from about 6400 to 1200 B.P. (Peterson and Groh 1969). Within the last 35,000 years over 40 pumice deposits have resulted from eruptions of Mount St. Helens (Borchardt, Harwood, and Schmitt 1971; Crandell and Mullineaux 1973; Mulli- neaux, Hyde, and Rubin 1972; Randle, Goles, and Kittleman 1971).

125 In eastern and northern California several active volcanos have produced chronologically significant regional deposits; the most important of these are from Mono-Inyo Craters (Chester- man 1971:141; Sheridan 1971, table 2). A dated stratigraphic sequence (Fig. 8), including fiv6 distinct layers of Mono Crater pumice, was recovered from 25 miles to the east at Black Lake, California (Batchelder 1970a). These five ash falls date from about 5000 to 1500 B.P. Three postglacial ashes have been re- covered from the Sierra Nevada, within or near Yosemite National Park; they may have been partly derived from Mono-Inyo craters (Wood 1972).

Although not well dated, recent volcanism in western Utah is indicated by the relationships of flows and pluvial lake strandlines (Condie and Barsky 1972, fig. 2), and in western Nevada by the association of volcanic ash and pluvial lake sediments (Morrison 1964:72). In Utah, activity continued after desiccation of Lake Bonneville and in Nevada through the waning phases of Lake Lahontan.

The Great Basin has been tectonically active throughout the late Cenozoic with recent activity concentrated in two north- south belts following the western and eastern boundaries. From 1961 to 1970, over 200 earthquakes were recorded in the Great Basin (Scholz, Barazangi, and Sbar 1971, fig. 2), and from 1852 to 1961 there were 1173 shocks recorded with Nevada epicenters alone (Slemmons, Jones, and Gimlett 1965, fig. 3). Faults con- trol the location and discharge of many Great Basin springs; thus, particularly in the more arid regions, their histories may be pertinent to aboriginal demography. Faulting has also alter- ed local patterns of erosion and sedimentation as well as the stratigraphic relationships of major geomorphic features (Clark 1972; Haynes 1967). The probable archeological importance of post-Lake Bonneville isostatic adjustments was mentioned pre- viously.

One can only speculate on the effect of a prehistoric earth- quake such as the Owens Valley disaster of 1872 (Hill 1972; Oake- shott, Greensfelder, and Kahle 1972), or the influence of an eruption such as that of Mount Mazama (Williams 1941:30) on, for example, resource availability (including new sources), settlement patterns or cosmology (Souther 1970). The results of historic volcanic eruptions and ash falls include not only the catastrophic loss of plant and animal life, but both immediate and long term changes in productivity. Along with detrimental effects, these changes may include mineral enrich- ment and increased soil moisture (Malde 1964; Wilcox 1959:462). The immediate and long term effects of a single major eruption and ash fall could be locally far more important to man than regional climatic change.

126 BLACK LAKE MONO CO., CALIFORNIA MAJOR POLLEN TYPES, CORE 2 c o P x I- Q7 Ijja. IL) Cf c- o Mono Craters Ash Carbon, Core 2 0 5 10 15 20 25% 0 40 80 0 30 0 i0 0 36 72 90% 0.00— Core 4 Core 2 I I I I

NWVS evVVV •

• 1.00- 2190 190

-‘.

S. • 2.00- S. • •

I 3890120 1 • • • CARBONATE • 3.00 - e' (X5) • • 5230±110 4580±00 • •

4.00- 4940*20

ORGANIC CARBON

• 5.00- •

11,350*350 •

6.00- 8550*210

Fig. 8. Mono Craters ashes in two cores from Black Lake Adobe Valley, California, and associated geochemical and pollen data. The high values of Cyperaceae (sedge) pollen and total carbon are thought to represent marsh environments of shallow lake phases (Batchelder 1970a,b; diagram provided by George L. Batchelder). Soils, Sediments and Erosion

Great Basin archeologists have depended on interpretations of geomorphic features and processes for both paleoclimatic and chronological data. Recognition of the importance of erosion, deposition and eolian activity to aboriginal subsistence (Bryan 1941; Hack 1942) and the effect of climate on geomorphic rates and processes, together with assumptions of synchronous and parallel regional climatic change (Antevs 1948, 1955:329), con- tinue to influence archeological interpretations. Repeated changes from deposition to erosion and increased eolian activity have usually been interpreted as indicative of reduced vegetation cover resulting from desiccation. Soil development in the Great Basin is generally thought to represent stable periods particularly characteristic of some climatic aspects of shallow lake phases. Such simplified explanations are improbable in some cases, and geologic-climatic assumptions and correlations require independent confirmation.

Paleosols are routinely used in stratigraphy, in correla- tions (Birkeland 1967, table 1), and as indicators of past vege- tation and climate (Morrison 1967:54). Assumptions of contem- poraneity over great distances, total time involved in weathering episodes, and the climatic parameters responsible are discussed by Morrison (1967:43-48). He concludes that "geosols" are time- parallel and increased temperature is primarily responsible for intense weathering episodes.

One of the more commonly recognized and stratigraphically important soils is generally formed on an eroded surface in sediments deposited prior to 7000-6000 B.P., or on dunes (e.g. the Midvale and Toyeh soils, see Fig. 4); it.is overlain by deposits younger than 4500 B.P. At least one other distinctive weathering episode may be recognized in sediments younger than 3500 B.P., and has been related to archeological sequences in the Lake Lahontan and Death Valley regions. At Steamboat Hot Springs, Nevada, soil formation occurred between 3500-1400 B.P.; maximum age was determined by radiocarbon and minimum age by a distinctive artifact assemblage. This soil may be correlative with the L-Drain soil of the Carson Desert (Morrison 1964:91). The Steamboat Hot Springs sequence is further correlated with archeological sites, dune stratigraphy, and soils in the , Nevada (Davis and Elston 1972). In Ash Meadow, Nevada, two traceable weathered surfaces, radiocarbon dated between 3600-2000 and 1000-300 B.P., separate archeological sites in dunes.

The major recognized soil forming periods of the past 40,000 years, along with radiocarbon dating, serve to correlate Lakes Bonneville and Lahontan (Fig. 4) and appear chronologically similar to independently dated soil forming episodes in the Valley, Nevada (Haynes 1967). Correlations of late

128 Quaternary stratigraphy and soils of the western United States are given by Haynes (1967, fig. 7) and Birkeland, Crandell and Richmond (1971, chart B).

Cut-fill sequences, so well dated in parts of the Plains and Southwest (Haynes 1968, figs. 2,4), are poorly dated in the Great Basin except in the Las Vegas Valley. However, casual observa- tion reveals obvious evidence for several episodes of channel cutting and filling over the last 12,000 years in river valleys and tributary streams throughout the Great Basin.

The stage of channel erosion or aggradation along the Moapa and Virgin rivers and their tributaries serves as an example of possible importance of agriculture and settlement patterns. Severe erosion, destruction of the fertile flood plain, and lowering of water tables would be detrimental to flood plain farming. The valley of the Moapa River, choked with sediments, weeds, and mesquite thickets, and poorly drain- ed, seems an equally formidable deterrent to successful agricul- ture, but exceedingly productive for gatherers. While the alluvial history of the southeastern Great Basin is obviously germane to the question of agricultural expansions (Shutler 1961), the alluvial chronology is still too poorly understood to clearly demonstrate a relationship. The only radiocarbon dated evidence for the alluvial history of the region comes from Meadow Valley Wash, near Caliente, Nevada (Figs. 9,10), where initiation of aggradation following extensive erosion is dated at about 1700 and 450 B.P. (Madsen 1972).

While there are few examples from the Great Basin, sediment size analysis, degree of weathering, organic and trace element content, and mode of deposition of cave fill may contribute in- formation on past environments and cave occupation. Studies of Deer Creek Cave (Wilson 1963), Stuart Rock Shelter (Howell 1960; Sabels 1960) and Rampart Cave (Martin, Sabels and Shutler 1961) serve as examples. Varying percents of sand, silt and clay from Rock Creek, an open site, were used as an index of climatic history for correlation with cave sites (Green 1972:25, figs. 8,9). By contrast, similar studies of Weston Canyon Rock- shelter led Delisio (1970) to conclude that changing climate was not discernible from the sediments or stratigraphic record.

Wind Action

Dunes are conspicuous features found leeward of the rem- nants of virtually every pluvial lake and river. The position of presently active dunes results from availability of source material, topography, vegetation cover, fossil dune patterns, and wind direction, velocity and duration. All of these are illustrated in a comprehensive study of the Kelso Dunes, Mohave Desert, California (Sharp 1966).

129 Fig. 9. Alluvial fill of Meadow Valley Wash, about 5 miles south of Caliente, Nevada (August 1971). Fossil sedge seeds from the organic bands visible near the base of the fill provided a radio- carbon date of about 1700 B.F. as a minimum age for initiation of aggradation. Prior to this time the valley had been scoured to the depth of the present arroyo (Fig. 10).

Fig. 10. Modern arroyo and alluvial channel fill of Meadow Valley Wash, about 12 miles south of Caliente, Nevada (August 1971). Char- coal from the laminated sediments of this fill and others along Meadow Valley Wash provided a maximum age of about 450 years for initiation of aggradation. Thus, there was a major period of erosion sometime after deposition shown in Figure 10; it probably occurred shortly before 500 years ago. Repeated episodes of erosion and de- position may be related to climatic change, but more importantly, arroyo cutting lowers water tables and destroys fertile flood plains.

130 As paleoclimatic indicators, dunes are generally treated as evidence of aridity (H. Smith 1967:21). Increased eolian deposition during interlacustral phases (Morrison 1964) supports this assumption. However, examination of just one controlling factor, source material, reveals more complexity in relative dune action during the past 8,000 years. For example, accumu- lated eolian sediments of the Great Salt Desert, Utah, include gypsum being produced through evaporation at the surface of salt flat clays (Eardley 1962:16) and oolitic sands, formed in shallow water. Elsewhere the major source may be from alluvium, alluvial fans, beach sand, or playa surfaces (Denny and Drewes 1965:38; MacDonald 1970:5; Sharp 1966:1046; Wallace 1961). At least along the Mohave and Amargosa Rivers, the sources of sand for continued or renewed dune formation would be enhanced by increased rainfall and periodic flooding. The same might apply to shallow lakes where continued fluctuations produce annual in- crements of deflatable sand.

With few possible exceptions (Hunt and Mabey 1966:82), most presently active dune sand is derived from yet older dunes dating from the end of the last pluvial, a time characterized by a sub- sequently unequaled abundance of newly exposed and easily de- flatable sediments. Since established dune patterns may influ- ence source material, orientation and shape during reactivation(s) (Sharp 1966:1066), later sand movement is partly preconditioned by dune history. Thus, the relative effects of wind action may reflect dune history and initial post pluvial sediment avail- ability rather than degree of climatic change. Morrison (1964: 103) suggests that deflation of the last 4000 years was hampered by the stabilizing effect of soil formation.

Regardless of impressive recent wind action, there is con- siderable evidence for former periods of more' extensive activity, as well as greater stability (Allison 1966, table 5; Hawley and Wilson 1965:49; Morrison 1964:76, 84-85; Sharp 1966:1059; H. Smith 1967:6). In the Amargosa Desert evidence includes dune migration followed by deflation, stabilization, fine grained eolian deposition, weathering and stream dissection. Smith (1967:16) made similar observations and also reported rock detritus, on the modified surfaces of sand aprons, overlain by fresh sand'. He interprets this sequence as representing two distinct periods of wind action.

Recurrent eolian activity in the Amargosa Desert is radio- carbon dated from before 5300 to about 2500 B.P. (with probable brief reduction in activity about 4000 B.P.), from 2000 to 1000 B.P., and again within the past 400 years. A chronologically similar sequence occurs at Corn Creek Dunes, Las Vegas Valley, Nevada (Williams and Orlins 1963, fig. 1), where radiocarbon dated hearths place initial dune activity between 5000 and 4000 B.P. After a minor hiatus, eolian deposition resumes and is followed by stability and soil formation; the soil is partially deflated and buried by younger dunes (Haynes 1967:60-65, fig. 18). Absence of pre-Death Valley III artifacts (Hunt 1960:112;

131 Hunt and Mabey 1966:82) places a minimum age of about 2000 years on formation of some existing dune areas in Death Valley. Because of their ability to act as sponges, rapidly absorb- ing occasional rain but releasing it slowly (Sharp 1966:1047), dunes may have been important to aboriginal occupation. Mesquite (Prosopis juliflora), an especially important resource in the arid southern Great Basin, and rice grass (Oryzopsis hymenoides), a favored food throughout the Great Basin (Steward 1933:244, 1938: 18, 26, 28, 74, 96), may owe their existence in harvestable numbers to the presence of a semistable sand substrate. In the Death Valley region past dune activity has dammed spring-fed drainages to produce extensive marshes, thereby increasing local productivity and waterfowl and mammal resources. Examples of such dune-dammed marshes are found in Death Valley at Saratoga Springs (Wallace and Taylor 1959) and in Ash Meadows, Amargosa Desert, Nevada (Fig. 11). Radiocarbon dated hearths indicate sporadic use of dune campsites for at least 5000 years.

Fig. 11. View of Saratoga Springs and the , Death Valley, California (November 1966). The marsh is spring-fed and formed by a dune dam. Cores from the marsh reveal a complex history of spring discharge and vegetational response. Endemic desert pupfish have been isolated here since the springs last flowed to the Amargosa River.

132 THE BIOLOGICAL EVIDENCE

Plant Macrofossils

The history of vegetational fluctuations accompanying late Quaternary climatic change in the Great Basin is derived pri- marily from exceptionally well-preserved macrofossils from dry caves and rock shelters. Woodrat middens, sloth dung, and other debris of man and beast recovered from archeological cave exca- vations have provided most of the data. Ground sloth (Noth- rotheriops shastense) dung from Rampart and Muav caves in the Grand Canyon, Arizona, and Gypsum Cave near Las Vegas, Nevada (Laudermilk and Munz 1934, 1938; Long and Martin 1974), contains plants that still occur in the southern Great Basin. However, changes in the dominant plant types indicate significant vege- tational fluctuation 40,000-10,000 B.P. (Martin, Sabels and Shutler 1961:115). The dominant species before 12,000-10,000 years ago now grow at considerably higher elevations (Laudermilk and Munz 1934:34, pl. 11; Mehringer 1967b, figs. 4,5).

The single most important nonarcheological source of macro- is fossil woodrat (Neotoma) middens; they contain abun- dant and well-preserved plant remains, some of which have been dated to greater than 40,000 B.P. Because of the small home range of the woodrat, one can be reasonably certain that the plants came from the vicinity of the midden (Wells and Berger 1967). In southern Nevada, junipers (Juniverus osteosperma) descended from their present elevational limits by as much as 3000 feet (Fig. 12) to the Las Vegas Valley (Mehringer 1967a; 183). Throughout the Mohave Desert juniper and pinyon grew at lower elevations (Leskinen 1970; Mehringer 1967a:191; Van Deven- der and King 1971; Wells 1969), resulting in'connections between presently disjunct woodlands. The pluvial history of higher elevation montane forest is less certain but bristlecone (Pinus longaeva) and limber (P. flexilis) pines and white fir (Abies concolor) also expanded their ranges and grew at much lower elevations (D. Fowler 1972b; Madsen 1972, table 3; Mehringer and Ferguson 1969, fig. 6).

Harper and Alder's (1970, 1972) studies of plant remains from Hogup and- Danger caves provide evidence for long human utilization of playa-edge resources as well as for vegetational •history and climatic change. They conclude that the present climate is more arid than at any time during the past 10,000 years. They recognize a dry period about 10,000 B.P., followed by more moist conditions and then general drying 8000-3000 B.P. This long dry period is interrupted by a brief moist event about 6000 B.P. and followed by another significant moist interval between 1500 and 600 B.P. (Harper and Alder 1972, fig. 3).

While present distributions and ethnographic use and de- pendence on specific plants are essential considerations in interpretation of aboriginal subsistence, they are not necessar- ily a key to even the recent past. For example, the importance of pickleweed (Allenrolfea) for the past 10,000 years is well

133 documented (Harper and Alder 1972), but we do not know the post- pluvial geographic history or relative abundance of such import- ant resources as pinyon pine (Pinus monophylla) or mesquites (Prosopis juliflora and P. pubescens) (Steward 1933:241, 1928:20, 80). It may be convenient to assume that these trees, within their present ranges, have long been important to Great Basin inhabitants, but without direct evidence from archeological exca- tions or linguistic data (C. Fowler 1972) such assumptions lack verification. It seems likely that both trees have undergone considerable postpluvial range adjustment.

In the Death Valley region where mesquites were an import- ant ethnographic resource, the oldest known occurrence is dated at only 4500 B.P. (A-1269). It seems unlikely that pinyon pine occurred in southern Oregon between 10,000 and 7000 B.P., 200 miles beyond its present northern and western limits (Bedwell 1970:51-52, 83). But, at the same time, in the southern Basin pinyons were growing at lower elevations and were presumedly more abundant than now. Subsequent distributional changes of such an important resource could have had a significant effect upon prehistoric demography, and pinyon pine may be a relatively recent arrival at its present northern limits. It was not iden- tified from the early sediments of Danger (Levels I and II) nor Hogup (Levels 1 through 8) caves (Harper and Alder 1970, table 1; Jennings 1957:326).

7000 CLARK MT, CALIF A4m 6000 • P PRESENT LOWER LIMIT OF JUNIPER WOODLAND . W W w 5000 Nwl■ FUNERAL RANGE, CALIF. z milli 0 / • Is .0 ct 4000 lim

Ui 3000 ASH MEADOWS, NEV.

=M.=

2000 8000 10,000 12,000 14,000 RADIOCARBON YEARS BP Fig. 12. Elevations and radiocarbon dates of ancient woodrat middens from mountains of southern Nevada and adjacent California. Juniper (Juniperus osteosperma) macrofossils from these middens in- dicate downward displacement of woodland species and their persistence in present desert to 7800 B.P. The Clark Mountain midden also contains pinyon (Pious monophylla), limber pine (P. flexilis) and white fir (Shies concolor). Except as noted, the middens are from the Nevada Test Site) only one of these, at 5000 feet, contained pinyon (Wells and Berger 1967, fig. U.

Fossil Pollen

The pioneering efforts of Henry P. Hansen (1947a,b) estab- lished the presence of significant postpluvial changes in the pollen deposition of southeastern Oregon. He suggested exten- sive pine forests during the waning phases of the last pluvial and composite pollen indicated dryer and warmer conditions 8000 to 4000 years ago. While his studies were completed before radiocarbon dating, they illustrate the usefulness of pollen analysis for correlation and for deciphering vegetational history in western North America.

134 Pollen records from pluvial lake deposits have confirmed Hansen's findings on conifer expansion prior to 12,000-10,000 years ago. At Searles Lake (Leopold 1967, 1970; Roosma 1958) and Tule Springs (Mehringer 1967a) abundant pine and juniper pollen indicate considerable expansion of woodland near the sites and of forests at higher elevations. Sagebrush (Arte- misia) pollen is also more abundant than during the past 10,000 years. To the north, in the Pluvial Lake Bonneville Basin, spruce accompanies pine as an important indicator of pluvial forest expansion (Martin and Mehringer 1965, fig. 2). Artemisia pollen is dominant during the preceding shallow lake phase and more abundant than present throughout the last pluvial (Fig. 13). While there was major pluvial forest and woodland expansion, the details of vegetational history are still poorly understood. For example, Major and Bamberg (1967) consider some Sierra Nevada-Rocky Mountain alpine disjuncts and the differences in forest species on both sides of the northern Great Basin as evidence for very cold-dry conditions. They suggest a nearly treeless full glacial landscape and widespread permafrost (Fosberg 1965; Malde 1961). On the other hand, Billings (1950) believes that conifer stands in the Virginia Range, Nevada, re- sulted from pluvial migrations of higher elevation Sierra Nevada forest species. Loope (1969:259) supports Billings in suggest- ing that the distribution of limber pine (Pinus flexilis) is the result of pluvial migrations (Critchfield and Allenbaugh 1969, fig. 1). While paleoclimatic arguments based on present plant distributions are academically interesting, definitive answers must await firm evidence in the form of fossil records. Pollen records covering the last 10,000-12,000 years are few and indicate either individual problems of interpretation, chronological control, ecological site sensitivity, and/or regional differences in post-pluvial climatic and vegetational history. In the Las Vegas Valley, Nevada (Fig. 14), the major trend toward warmer and dryer conditions is marked by a change from juniper-sagebrush to sagebrush-shadscale starting about 12,000 B.P. By 7000 B.P. the vegetation was probably similar to the present lower elevation Mohave Desert. Short-term re- versals of this drying trend are dated at about 10,500-10,000 B.P. and 8500-8000 B.P. The Tule Springs record is incomplete after 7000 B.P. The spring-fed salt marshes of the Great Basin (Fig. 11) offer the potential for establishing a history of spring dis- charge and marsh resources. Changes are reflected by the sediments as well as fossil pollen, molluscs and diatoms that furnish data on past salinity and water depth. Salt marsh pollen records of the Death Valley area contain evidence for nearly total desiccation and deflation, as well as higher than present spring discharge and preserved peat deposits. In both Panamint Valley and Ash Meadows, initiation of a major period

135

PLUVIAL LA.KE BONNEVILLE CRESCENT SPRING, UTAH

u) Summary Diagram CJzd 1— op7 u) z . e,,,, z \C\ 0 P 4CP 'cc w ,c), A.\(\ \C) 1--. w C Os\ g7V \ 07' CÝ CNV ww w ? oM u) 06 0 20 40 40 540 20 360 12 0 18 -2.66 2.70 -

2.80 - non- calcareous 2.90 - dark gray to green gray 3.00 - clayey silt

3.10 -

3.20

3.30

calcareous 3. 40* green gray to yellow gray 3.50- clayey silt

3.60-

3.70 -

3.80 - banded calcareous 3.90 - dark gray 26,700±900 to 1-4409 yellow gray 4.00 - sand,silt,& oolite I

4.10 -

4.20 - 4.26 GREAT SALT LAKE, 0-7000 B.P I 1

Fig. 13. A summary pollen diagram of Lake Bonneville sediments and average pollen COUNT FOR the last 7000 years of a core from the Great Salt Lake. The diagram shows THE transition from shallow water and marshes to an increase in water depth and CONIFER pollen importance, including spruce (Picea). This transition marks the beginning OF the last pluvial about 24,000 B.P. The higher values of sagebrush (Artemisia) and lower values of juniper distinguish the Holocene from the last WISCONSIN interpluvial. Presumably the latter period was considerably colder.

136 Fig. 14. Shallow lake and marsh deposits, Tule Springs site, Las Vegas Valley, Nevada (June 1965). The sediments, plant macro- fossils and pollen, vertebrate remains, and molluscs from these deposits have provided clues to the pluvial climates of the Mohave Desert. of peat formation and accompanying dominance of marsh pollen and macrofossils is dated at about 3600 B.P. These deposits are eroded and weathered, probably as a result of significant desiccation (after 2200 B.P.?). Another important episode of marsh growth started about 400 years ago. Yet older evidence for equally significant post-pluvial variation in spring dis- charge is undated. Madsen (1972, fig. 3) interprets a 7000 radiocarbon year pollen record from O'Malley Shelter as indicative of a change from juniper-sagebrush to grass-sagebrush about 5200 B.P., followed by a return to juniper-sagebrush beginning 3900 B.P. Thereafter there was a gradual increase in arboreal pollen, first juniper, then oak and pine. By 900 B.P. vegetation apparently similar to the present juniper-pinyon woodland pre- vailed. The pollen records are partly substantiated by macro- fossils from two woodrat middens. One shows arboreal dominance

137 (Juniperus osteosperma) some time within the past few hundred years, while juniper is rare in another dating from about 4400 B.P. (Madsen 1972, table 2).

Pollen profiles are available from two localities in northeastern California--Osgood Swamp, near the south end of Lake Tahoe, and Adobe Valley, 15 miles southeast of Mono Lake. Osgood Swamp was formed within a terminal moraine of Tioga age. Radiocarbon dated and Mazama ash further serve to estab- lish the depositional chronology. The pollen diagram (Fig. 15) is interpreted by Adam (1967) as indicating colder and dryer conditions in the basal inorganic zone dominated by,Artemisia and juniper. The most obvious change occurs at the transition to organic sediment and conifer pollen dominance about 10,000 B.P. The diagram is quite uniform for the remainder of post- glacial time. Adam, primarily utilizing aquatic and bog-edge types as indicators of water depth, recognizes a two-part post- glacial climatic optimum with the latter temperature maximum terminating about 2900 B.P. He correlates the uppermost in- crease in fir (Abies) and Ericaceae with the return of cooler conditions.

Paleoenvironmental studies at Black Lake include plant microfossils, fossil molluscs and the geochemistry of sedi- ments (Fig. 8), postdating drying of the larger pluvial lake that occupied Adobe Valley and overflowed southward to the Owens River (Fig. 16). Batchelder (1970a, b), relying pri- marily on plant indicators of water depth and lake size, recognizes several important climatic changes. Organic sedi- ments dating from about 11,500 B.P. are thought to indicate conditions considerably dryer and somewhat colder than present. This was followed by a trend toward a deeper lake culminating in an important regional increase in effective moisture about 8500-8000 B.P. Dryer conditions prevailed from 8000 to 4500 B.P. The most xeric interval, about 7000-6000 B.P., is follow- ed by a minor increase in water depth before 5300 B.P. After 2000 B.P. temperature and moisture approached their modern values. The xeric mid-postglacial interval recognized by Batchelder at Black Lake corresponds chronologically to Hansen's xerothermic period of southern Oregon. Several pollen records are available from the northeastern Great Basin; the most important of these is from Swan Lake (Bright 1966). Swan Lake occupies a shallow depression about 4.5 miles south of Red Rock Pass, the overflow outlet of Plu- vial Lake Bonneville to the Snake River. Basal deposition, postdating the last overflow of Lake Bonneville to the Snake River, is radiocarbon dated at 12,000 B.P. (Fig. 17). At this time coniferous forests were dominant. With minor fluctuations the forests gradually retreated to higher elevations as the climate warmed. By 10,000 B.P. they were replaced by Artemisia steppe. The lowest values of arboreal pollen indicating a warm semiarid climate similar to the present, occurred between 8000 and 3000 B.P. Increased pine pollen values from about 3000 to 1700 B.P. probably resulted from a return to more effective moisture and minor forest expansion.

138

OSGOOD SWAMP, CALIFORNIA

ARBOREAL HERBS and SHRUBS AQUATIC

rPiP A 4. Ae A 0° ,A 1 ,c,i .#.• 1 50" ,o- cc0- t . ‘e vier i.O' AI vco A co. 0 os C 4 .S.S“ .S110,1 0 J —I DATES

• 2 ACONITU11

• • CAPRIFOLIACEAE I 2°31.:200 • ACON1TUM

24 • EPHEORA

2b 4444 4 /A , VOLCANIC ASH 1•NAZAMA ,1

• ONAGRACEAE •

2e • BERGER1S

• EPHEPRA

3

• • EPHEERA •

• GERANIACEAE • GERANIACEAE •

SUM. TOTAL POLLEN - r bly_3• 100 SUM • TOTAL POLLEN- piNus 00

Fig. 15. Pollen diagram from the Sierra Nevada, California (from Adam 1967). The volcanic ash is radio- carbon dated at about 6900 B.P. Fig. 16. The outlet canyon of Pluvial Adobe Lake, California, look- ing east toward Boundary Peak in the White Mountains (October 1966). The outlet gorge was cut when the pluvial lake overflowed to the Owens River. Cores from this valley have provided information on volcanic ash chronology and lake history of the last 11,000 years (Fig. 8).

A preliminary pollen diagram from the Raft River Mountains, Utah (Fig. 7) is quite similar to the Swan Lake diagram. From about 12,000 to 10,500 B.P. the vegetation is represented by high values of Artemisia pollen, apparently representing cold, nearly treeless conditions. During the latter part of this period there is a gradual upward migration of coniferous trees, first spruce followed by pine, fir and Douglas fir. By about 10,000 B.P. coniferous pollen reaches its highest relative value; these values are maintained with a few minor fluctua- tions until shortly before the fall of Mazama ash. About 7500 B.P. Artemisia replaces conifer pollen, presumably as a re- sult of climatic conditions more nearly like those of the present than the preceding 5,000 years. Following deposition of Mazama ash there is a sharp relative increase of coniferous pollen which might correlate with a similar event near the middle of Zone S3 at Swan Lake. By extrapolation of sedimen- tation rates this event occurs about 6000 B.P. As at Swan Lake, there are significant fluctuations recording a return to increased effective moisture, indicated by a return to coni- fer dominance, in the last 3500 years or so.

140

0' SWAN LAKE ,i k. ,N I .. - BANNOCK COUNTY, I DAHO t (■ 0- to- R.C. Bright, 4 965 . t Q cr s .\0 t.■ '' 0 o 4) '■ C, o ci c, ..X* C5 , 4' 6' '0 • c, 6 • c, •\;) A (Q ,4 Pollenen k• ' IQ. Zones \ AY4\ 1 , I P(3 1 (‘'.. F;

s1 C14 DATES (years B.P)

S2 W — 4340 • 1,8501200

S3

S4 W — 4 339 4 0,4 90:1250

S5

S6 ......

S7 W-4338 • 42,09W:30011 r 1 I I h r 10 20 30 40 50 60 70 %

Fig. 17. A summary pollen diagram from Swan Lake, Idaho, shows vegetational variation from forest to steppe and postdates the last overflow of Lake Bonneville ffrom Wright 1971). Unpublished analyses of cores from the Great Salt Lake and Crescent Spring, near Hogup Cave, as well as pollen studies at Hogup Cave sediments, clearly indicate significant changes in regional vegetation during the past few thousand years. Most important are two distinct trends toward increased Artemisia and conifer pollen relative to that of Chenopodiaceae, between about 3600 and 2200 radiocarbon years ago. They may correlate with Zone S2 at Swan Lake. The increase in grass pollen from Hogup Cave (Level 12) starting about 1500 B.P. (Kelson 1970, fig. 2) is indicative of an important change in the local vege- tation producing pollen spectra unlike those of the preceding 7000 radiocarbon years. The low relative frequency of Cheno- podiaceae pollen about 8000 B.P. (levels 1 and 2) may indicate more effective moisture and/or less area of halophytic cheno- pod .

Analysis of cave fill from the Toquima Range, Nevada, produced inconclusive results. Kautz and Thomas (1973) suggest the possibility of a change to more mesic conditions, indicat- ed by relative increase in pine and juniper pollen, about 3400 B.P. Interpretation of significant changes in pollen spectra from Fishbone and Guano Caves, Nevada (Sears and Roosma 1961) is hampered by inadequate dating and sampling, and question- able identifications. I believe the reported spruce pollen to be misidentified because there is a lack of substantiating evidence for a southeastward pluvial expansion of spruce from northern California.

Dendroclimatology

Studies of past tree growth and distribution provides the most precise evidence for the chronology and magnitude of climatic change, as well as dating of glacial features (Currey 1965) and geomorphic processes (LaMarche 1968). These studies are unique, in their time depth, to the Great Basin where living bristlecone pines (Bailey 1970) may be older than 4000 years and their wood survives many thousands of years past death of the trees (LaMarche 1969). Using both living trees and wood remnants, Ferguson (1968, 1969, 1970) established a continuous 7484 year chronology and reported yet older rem- nants. The long bristlecone pine chronology is of prime im- portance in understanding variation in radiocarbon production (Damon, et al. 1974).

Analysis of climatic factors responsible for formation of wide or narrow rings, together with the use of statistical models of tree response to environmental change, will un- doubtedly lead to the reconstruction of Great Basin climatic circulation patterns of the past hundreds if not thousands of years (Fritts 1965, 1966, 1971; Fritts, Smith and Stoken 1965; Fritts, Blasing, Hayden and Kutzbach 1971; LaMarche and Fritts 1971). Figure 18 illustrates the use of tree growth in re- constructing 10 year climatic anomalies for the western United States and the regional importance of such anomalies.

142 . :;_ " 1 4 7 2 lik, 1 \VAILI ' N . i.ii, w _ • 1 - 1 / 01111111111FVE 4 72 I fr 4 • . til ins i .1 ), 4 . I,„MV,II A Z ' i difkr,„441 ;,,i.i.oli : ' „ 4,..4reia,„ ligl- 1 - 1101r. . l'e77AT k ,,..• . .4 ,,,, 14 1 re f 1 1 .. ,'i 13 2 11 1 ' 7 1 1.2 . r:4 , ,.., io ' 4Ztill' 14111 441111 11xil - ' ----40- Idift 4Him' 1 i il 0 4„ illi■ . 11 I i 3 3 0 / "--, , . i 11 • • . . I11411 4844 1576-1585 IL I 1616-1625 . . ..1/ 1 1521-1530 N /4 .1551-1560 I — MILLS

. 1 ,.111 911111i - 2 .,,„0.„ k .41, IL TO l inti1 t 0 H NIP NI . i t, \ _ _ ,!..; .„___..,. IT, ill C RISEN %MI ( ( 03 ' • 44 2 . .111 11 -14 ' %14 ‘1071 11S V 1 E p I. *111 . , 1 .1,1,10,r IA* 11 It 10 1 AO; Ail r / 1 3,, '1111!,:;111 II 144 111)1. I 11111' X 411111116. 1111116 ,1 416110 . , ito 1 40 i wr . , It . . . 0 ,,.. , 4 I , 1626-1635 *11! 1731-1740 i1741-1750I mr6-1765 4 4 t ... V ". - 1 61 I mom - \ '

Fig. 18. A map of regional variation in climate based upon ten year relative departures in tree-ring indices from western North America. The symbols H and L designate areas of significant high and low tree growth (from Fritts 1971). High growth (H) indicates moist-cool climatic anomalies, while low growth (L) indicates dry-warm climatic anomalies.

143 In some cases most of the Basin is affected by significant climatic variability, while at other times only a small por- tion of the total area is affected. The influence of a short-term climatic change on resources and their users might be either similar or quite different across the Great Basin. For example, compare the geographic patterns of tree growth for A.D. 1776-1785 with A.D. 1521-1530. Tree line fluctuations of the last 6000 years are record- ed by the remains of dead bristlecone pines (Fig. 19) up to 450 feet above present tree line (LaMarche 1973; LaMarche and Mooney 1967) and by dated changes in growth form (LaMarche and Mooney 1972, fig. 10). Figure 20 illustrates the history of tree line in the White Mountains, California. Summer tempera- tures are important in determining tree line elevation. Thus, former higher tree lines are probably indicative of higher temperatures. During most of the past 6000 years, tree line in the White Mountains has been higher than during the last few hundred.

Fig. 19. Dead bristlecone pines above present tree line in the White Mountains, California (September 1967). These trees probably became established during a time of higher temperatures.

144 RADIOCARBON YEARS B.P. 6000 5500 5000 4500 4000 3500 3000 2500 2000 1500 1000

150

— METERS

100

CAMPa0 MTN TREEUNE,

- MODERN

L _ 50

SHEEP MTN. ALTITUDE

5500 5000 4500 4 000 3500 300 0 2 500 2000 I 500 100 0 500 B.C. A.D. 500 I 000 1 500 -

CALENDAR YEARS

Fig. 20. Envelope curves showing minimum past altitudes of the upper tree line at two localities in the White Mountains, California (from LaMarche 1973). Fossil Animals

No biological event of the late Pleistocene is more im- portant than the sudden and massive extinction of the North American megafauna by 10,000 years ago (Martin 1967, 1973). The present deer, mountain sheep and antelope are but a small remnant of the mammals that once existed and remain now only as fossils, attesting to the long, and for a time successful, experiment in evolution and coexistence of autochthonous and immigrant species from Eurasia and South America. From Fossil Lake, Oregon (Allison 1966), to the Las Vegas Valley, Nevada (Mawby 1967), and from caves and lake beds betWeen comes evidence for a fauna whose demise must have had con- siderable influence on man, the surviving fauna, and flora as well. All comparisons of present with pre-extinction species associations seem limited without consideration of the mega- faunal influences upon the landscape. For this inquiry we must turn.to range management studies and, for the best com- parisons, to Africa where a varied megafauna still exists. Such efforts are beyond the scope of this review. However, the reader should be warned that any statements comparing pluvial with present biotic communities should include the disclaimer--except for the uncertain influence of horse, camel, sloth, bison, mammoth, musk ox and associated predators and scavengers. In this regard, present species associations and our understanding of their ecology seem an uncertain measure of even the recent past.

Hubbs and Miller's (1948) study of fish distribution and hydrographic history remains the most ambitious and comprehen- sive investigation of Great Basin biogeography. Their analy- ses illustrate former connections of Pleistocene lakes, river courses, estimates of times of isolation and most importantly the magnitude of climatic and biological change over a geo- logically short time. Other studies of fish distributions and fossils provide information on specific paleoenvironments (Smith, Stokes and Horn 1968) and past distributions (Hubbs and Miller 1970; Miller 1965). Data on the paleoecology of aquatic habits also come from studies of the ubiquitous molluscan remains (Roscoe 1963; Taylor 1967, 1970).

During the last pluvial small mammals extended their ranges accompanying changes in (Fig. 21). Subsequent range contraction, continuing into postpluvial times, is in- dicated by the presence of pika (Ochotona) prior to about 8000 B.P. at Connley Cave (Bedwell 1970:85, 256) and pika and marten (Martes) in the lower levels of Deer Creek Cave (Ziegler 1963:16). In both cases these mammals are inter- preted as indicating cooler climates during early postpluvial times. The occurrence of the pallid bat (Antrozoas pallidus) at Hogup Cave (Levels 4,5,7,8,9) is the single instance of a possible significant northward range extension indicative of higher temperatures (Durrant 1970:242).

146 .. , ...... ow&m...... maglAWANCEARAWANOWnvei.CiimMaImmiimmomimaillig, 747..riumusemiumraumw.r.,Empagrompaummoreumemraama IIIA101111.111WW1111••11/4/40.U.I &I .12011111111•••••••raurami/IONSMONIUMMIMILWAMONIM1/ 40111/Un Ww41■1•1•11■110/4r4c a Pa UMMIII••••••••112FAMIPAPUR1111W4101/411=1101r4/A&WWW.A&NIMO micdr.■••■•■". AN .NOMMEMINOINIOUROmParrE/Me%WIRIITA MPAIMPAIWO irimPar.,•21511■WW/IPMINIIII■1•11111UM.r. UM ZUVIIMIIIIMMISMOVKI/11, AIIIIMMV:IPAPN&MIEN'Il&r.//r 111/MPUKIN$ 01 •741.74111111111••••••11OMPAW S: !PAROMMII•••••1111%I.IPME NEVADA "ArArlmw"R""2.".MIJOUN1141/2.$2011/111,4Ed mWA% UTAH IIIumgrawliMIPUMft U? 0Pa U11•11=11111•MINMWI&P.M21011M&RAUWW,di&IMAI&IMMOMEM2 FAVAMMIINNEMINIVIAPACV KIP.:1102$;10111111111111•11M/M1AMIUM/MEWIUMOUCIA/ANIPUROW&EAUMMINIUMI/MAPIII■•=1•11/2Prero,$ pi raM0IWIIIIIIIIIIIIIIIIIME&NOUICIPW&INAUROMIARIMIUMN12/4511■114MEMUIEN ■■411/411■1•11•40.1■Fo 4: 4■10401■11••••••••■11WOMPIPMUSOMIWIIMUMMIUMUKOMPAIWOM0 UPAP:MMENIIIIMINIIIIINISFAM.. 14, •,,y"... agliThil MIIMIIINVAAPK• W. 4 21ROWARAIAPWAY IMI,M2/40M wArArA■Nesk...•10 r gunwamodmiamermarmau%ownparAwarArraw•miararagrAmr s %.• 12 1111 .$1■■riN re 0•KOAQU$211MME WAVAIIONUMW$U0•767~41rOUROU111,MAIINPIVE 2/4/41/ 41■A,4u OiSiraMeTa%011•1 r spesompaP2P-12PAMilr.aiirAnaratindpurAplin/P2 r.ilr .v4W t w.AP:11:4P:VAGEM FA/ 2hranirallerarigirorannininfradirIndOWNP2it Ear . •0-, p. er.m.ownsaPAPIIN Am ripAniplIrroNW4rirArdr4r.a.%141MTNOR• MI AIP 40 4, 4 $.4.`OPIMICVAINNUM I • Alre.4° , 4■41.:1ft."mmum. FP OrOmeAnwanadymmionimegmriumlimm AP •,w1P ,4%moar.pa2110■A .....4 I .•. • ..•* 4 4■ 01 ....■coNK.EANICOMPAW AmeugardorM 010 4P CR MM I 7 wIr 74R0 7. 7 44N wOallowiolrAPAPEARPi■HSmOaMlPnInCi Iii ■A $ 4 40IPi&■ M p lIOU 8 111 . 0 :44 )4W poom.Rwmwra 00, MI 1•111011 71 a 0 .p4meamewans 4 E MEW4 g PP004kr11 1 e O...... i• •1 2100 111=111111111!4&%0:& SO, e -- 7 1111111111MMONVP40 0 . 0016tWON 21MIKVA■ 40 • hi • • • ,ownirai..).v 4,4 .4 =11•1/4■021TiOW/ A 4P. ,44 sweiriwt,■.., L, NO:401■474 $ •1■ 4■.04.:• - - 4 .4 9 ..wmmo t 04004 $1WW44% o■ w 4■ t4 1"4. wivanLy, u ' .040 e , ■'04 4, "ON' 4 *■ 0■4.4 $ dio-Wa6gb.7 • V4A010• * ry .04 1r04: r4 0AMO vivo "%ow 1, 0 pwoolo■+4. 0 *4 00000 $10■00, 0 rvt4. 4 Mercury Ridge 04■A ...004M AIM 0■ Tde Spring ce 0 440 // MIL= 4401 MEW 01111, 4 4 010 IMP 41 Rampart Cave ...... ' 441% *440 -0+ AiAlesa'll Cove CALIFORNIA ARIZONA

Ochotona princeps EgMarmota flaviventris Citellus lateralis

Sylvilagus idahoensis MILES o 50 100 Microtus californicus 0 75 '150 KILOMETERS

Fig. 21. Pluvial occurrence (circles) and modern distribution of pika, marmot, golden-mantled ground squirrel, sagebrush rabbit and meadow mouse in the southern Great Basin (from Mehringer and Ferguson 1969).

147 Mammal records from other archeological sites are consider- ably more equivocal. However, Harper and Alder (1970:231-235, fig. 3; 1972:18-19) use mammal remains from Hogup and Danger Caves to support their previously mentioned interpretations of postpluvial climatic change. Further they agree with Butler (1972) that a climatically induced decrease in rodents from archeological sites 8000-7000 radiocarbon years ago is represen- tative of a decline in human carrying capacity. The remarkable report of bog lemmings (Synaptomys) from O'Malley Shelter within the last 1000 years (Fowler, Madsen and Hattori 1973:51) and the apparent sudden disappearance of turkey from southern Oregon about 7000 B.P. (Bedwell 1970:87, 1971) remain to be explained. Both occurrences seem unlikely.*

SUMMARY

The dynamic nature of Great Basin environments is apparent whether measured by geological or biological criteria. However, instability of the last 10,000 years is no more dramatic than the ecological variation encountered by Great Basin inhabitants within a single year. Variability itself may have been most important in shaping cultural or technological adaptations. There were several periods when regional climatic change was sufficient to warrant investigations of its possible influence on human populations; locally, volcanic or tectonic activity may have been even more important. During the last pluvial many basins, now dry and salt encrust- ed, were fed to overflowing by cool waters and joined by great rivers. As woodlands descended to the treeless , glaciers carved beds in the snow capped mountains. Herds of camels, horses and mammoths grazed the steppes and fertile marshes. And then, within 2000 years (12,000-10,000 B.P.), lakes shrank, rivers ceased to flow, and springs began to dry. Plants and animals began the long retreat northward and to higher elevations and man witnessed demise of the Pleistocene megafauna. By com- parison, all subsequent environmental changes have been minor. A trend toward aridity prevailed for the next few thousand years. As lakes grew even smaller and spring discharge decreased, with the dwindling supply of pluvial age ground water, both plants and animals continued to adjust their ranges. Short term reversals of this trend probably occurred shortly before 10,000

Ed.'s note. The specimens identified as Synaptomys from the O'Malley Shelter site were subsequently re-examined by T.R. Van Devender, Department of Geosciences, University of Arizona, and his colleagues. Their determina- tion was that the specimens in question are Microtus sp. (T.R. Van Devender, personal communication, 1 July, 1974). A recent re-examination of the avi- fauna material from southern Oregon indicates that the material in question is sage grouse (Donald K. Grayson, 1977, A Review of the Evidence for Early Holocene Turkeys in the Northern Great Basin. American Antiquity 42(1):110-12)

148 38. and 8000 B.P. By 7500 radiocarbon years ago conditions were much like the present. Some researchers have suggested the per- sistence of extreme arid climates, hotter and drier than the present, for the next 3000 years. Perhaps this view is over- simplified as there is some evidence for a brief increase in effective moisture sometime between 6500-5500 radiocarbon years ago. This includes plant remains from cave deposits and pollen from lakes and bogs of northwestern Utah. The pollen records from O'Malley shelter and Osgood swamp also show changes at this time, and it is a time of peat growth in the spring-fed marshes of the Amargosa Desert. Considerably more data are' required to establish the detailed chronology and magnitude of Great Basin climatic change from 7500-4000 B.P. As for the rest of post- pluvial time, the influence of climatic change on man is best considered in terms of evidence for its effect on local resources.

Many lines of evidence are suggestive of a significant change to more effective moisture starting about 4000-3000 B.P. and ending before 2000 B.P. At this time minor changes to deeper lakes are recorded from Great Salt Lake, Pyramid Lake, Searles Lake and Death Valley. In the north, pollen records reveal an increase in arboreal types, and in the White Mountains there is a lowering of the bristlecone pine tree line. These events are accompanied by changes in eolian activity and from erosion to deposition and soil development.

It becomes more difficult to make broad regional generaliza- tions as the data become more abundant. Fortunately, tree-ring studies probably will establish regional climatic patterns for the last 2000 years in a more precise manner than is otherwise possible (LaMarche 1974). During the past 2000 years, as during the preceding 10,000, there was geological and biological in- stability of sufficient magnitude to affect the abundance of local resources. The evidence includes plant macrofossils from caves of northwestern Utah, suggestive of a significant moist interval 1500-600 B.P. Fluctuations in lake levels, lowering of tree line, renewed dune activity and stabilization, peat formation in desert salt marshes, arroyo cutting and filling and significant tectonic and volcanic activity all continued through the past 1000 years.

NOTES

1. This paper was initially prepared for the forthcoming Handbook of North American Indians, and submitted and distributed for review in 1973. However, because of delays in publication of the Great Basin Handbook volume, it is included here to pro- vide a summary of environmental information essential to con- siderations of paleoecological models in Great Basin prehistory.

149 BIBLIOGRAPHY '/ Adam, David P. 1967 Late-Pleistocene and Recent Palynology in the Central Sierra Nevada, California. In E.J. Cushing and H.E. Wright, Jr. (eds.), Quaternary Paleoecology. New Haven: Yale University Press. Pp. 275-301. Aikens, C. Melvin 1967 Excavations at Snake Rock Village and the Bear River No. 2 Site. University of Utah Anthropological Papers, No. 87. Allison, Ira S. 1966 Fossil Lake, Oregon: Its Geology and Fossil Faunas. Oregon State Monographs, Studies in Geology, No. 9. /Antevs, Ernst 1948 The Great Basin, with Emphasis on Glacial and Post- glacial Times: Climatic Changes and Pre-White Man. Bulletin of the University of Utah 38(20):168-191. 1952 Cenozoic Climates of the Great Basin. Geologische Rundschau 40:94-108. / 1955 Geologic-Climatic Dating in the West. American Antiquity 20(4, pt. 1):317-335. Aschmann, Homer 1958 Great Basin Climates in Relation to Human Occupance. Reports of the University of California Archaeological Survey, No. 47:23-40. Berkeley. Bailey, D.K. 1970 Phytogeography and Taxonomy of Pinus Subsection Bal- fourianae. Annals of the Missouri BotanicalGarden 57:210-249. Batchelder, George L. 1970a Post-Glacial Fluctuations of Lake Level in Adobe Valley, Mono County, California. Abstracts of the First Meeting of the American Quaternary Association, p. 7. v1970b Post-Glacial Ecology at Black Lake, Mono County, California. Unpublished Ph.D. dissertation, Arizona State University. iiaumhoff, Martin A. and Robert F. Heizer 1965 Postglacial climate and Archaeology in the Desert West. In H.E. Wright, Jr. and David G. Frey (eds.), The Quaternary of the United States. Princeton: Princeton University Press. Pp. 697-707.

150 Bedwell, Stephen F. 1971 New Evidence for the Presence of Turkey in the Early Postglacial Period of the Northern Great Basin. The Great Basin Naturalist 31(2):48-49. 1970 Prehistory and Environment of the Pluvial Fort Rock Lake Area of South Central Oregon. Unpublished Ph.D. dissertation, University of Oregon. Benedict, James B. 1970 Downslope Soil Movement in a Colorado Alpine Region: Rates, Processes, and Climatic Significance. and Alpine Research 2(3):165-226. Billings, W.D. 1950 Vegetation and Plant Growth as Affected by Checmically Altered Rock in the Western Great Basin. Ecology 31(1):62-74. Birkeland, P.W. 1968 Correlation of Quaternary Stratigraphy of the Sierra Nevada with that of the Lake Lahontan Area. In Roger B. Morrison and Herbert E. Wright, Jr. (eds.), Means of Correlation of Quaternary Successions. Salt Lake City: University of Utah Press. Pp. 469-500. 1967 Correlation of Soils and Stratigraphic Importance in Western Nevada and California, and Their Relative Rates of Profile Development. In R. B. Morrison and H.E. Wright, Jr. (eds.), Quaternary Soils. Proceed- ings of the VII Congress of the International Association for Quaternary Research, Vol. 9. Reno: Desert Research Institute. Pp. 71-91. //Birkeland, Peter W., Dwight R. Crandell, and Gerald M. Richmond 1971 Status of Correlation of Quaternary Stratigraphic Units in the Western Coterminous United States. Quaternary Research 1(2):208-227. t--Borchardt, G.A., M.E. Harward, and R.A. Schmitt 1971 Correlation of Volcanic Ash Deposits by Activation Analysis of Glass Separates. Quaternary Research 1(2):247-260. Born, Stephen M. 1970 Deltaic Sedimentation at Pyramid Lake, Nevada. Unpublished Ph.D. dissertation, University of Wis- consin. oliright, Robert C. 1966 Pollen and Seed Stratigraphy of Swan Lake, South- eastern Idaho: Its Relation to Regional Vegetational History and to Lake Bonneville History. Tebiwa: The Journal of the Idaho State University Museum -§77):1-47.

151 1967 Late-Pleistocene Stratigraphy in Thatcher Basin, Southeastern Idaho. Tebiwa: The Journal of the Idaho State University Museum 10(1):1-5.

P4raecker, Wallace S. and Aaron Kaufman 1965 Radiocarbon chronology of Lake Lahontan and Lake Bonneville II, Great Basin. Geological Society of America Bulletin 76(5):537-566. vBroecker, W.S. and P.C. Orr 1958 Radiocarbon Chronology of Lake Lahontan and Lake Bonneville, Geological Society of America Bulletin 69(8):1009-1032. a/Bryan, Kirk 1941 Pre-Columbian Agriculture in the Southwest, as Con- ditioned by Periods of Alluviation. Annals of the Association of American Geographers 31(4):219-242.

'Bryson, Reid A., David A. Baerreis, and Wayne M. Wendland 1970 The Character of Late-Glacial and Post-Glacial Climatic Changes. In Wakefield Dort and J. Knox Jones, Jr. (eds.), Pleistocene and Recent Environ- ments of the Central Great Plains. Lawrence: Univer-gity Press of Kansas. Pp. 53-74.

Butler, B. Robert 1972 The Holocene or Postglacial Ecological Crisis on the Eastern Snake River Plain. Tebiwa: The Journal of the Idaho State University Museum 15(1):49-63.

/Campbell, Elizabeth W. Crozer and William H. Campbell 1937 The Archeology of Pleistocene Lake Mohave: The Lake Mohave Site. Southwest Museum Papers, No. 11.

Chesterman, Charles W. 1971 Volcanism in California. California Division of Mines and Geology, California Geology 24(8):139-147.

(..-Clark, Malcolm M. 1972 Range-Front Faulting: Cause of Anomalous Relations among Moraines of the Eastern Slope of the Sierra Nevada, California. Geological Society of America, Cordilleran Section, 68th Annual Meeting Abstracts with Programs 4(3). p. 137.

"tondie, Kent C. and C.K. Barsky 1972 Origin of Quaternary from the Black Rock Desert Region, Utah. Geological Society of America Bulletin 83(2):333-352.

Crandell, Dwight R. and Donald R. Mullineaux 1973 Pine Creek Volcanic Assemblage at Mount St. Helens, Washington. U.S. Geological Survey Bulletin 1383-A. Washington: Government Printing Office. Pp. 1-23.

152 Cressman, Luther S. (and collaborators) 1942 Archaeological Researches in the Northern Great Basin. Carnegie Institution of Washington, Publication 538. Pp. 1-158. Critchfield, William B. and Gordon L. Allenbaugh 1969 The Distribution of Pinaceae in and Near Northern Nevada. Madrono 20(1):12-26.

"Crittenden, Max D., Jr. 1963 New Data on the Isostatic Defortmation of Lake Bonne- ville. U.S. Geological Survey Professional Paper 454-E. Washington: Government Printing Office. Pp. 1-IL

Currey, Donald R. 1965 An Ancient Bristlecone Stand in Eastern Nevada. Ecology 46(4):564-566.

(Currey, Robert R. 1969 Holocene Climatic and Glacial History of the Central Sierra Nevada, California. In S.A. Schumm and W.C. Bradley (eds.), United States Contributions to Quaternary Research. Geological Society of America Special Paper 123. Pp. 1-47. 1971 Glacial and Pleistocene History of the Mammoth Lakes Sierra, California--A Geologic Guidebook. University of Montana Department of Geology, Geological Series, Publication No. 11.

Damon, P. E., C.W. Ferguson, A. Long, and E. I. Wallick 1974 Dendrochronologic Calibration of the Radiocarbon Time Scale. American Antiquity 39(3, pt. 1):350-366.

Davis, E .L., C.W. Brott, and D.L. Weide 1969 The Western Lithic Co-Tradition. San Diego Museum Papers, No. 6.

Davis, Jonathan 0. and Robert Elston 1972 New Stratigraphic Evidence of Late Quaternary Climatic Change in Northwestern Nevada. In Don D. Fowler (ed.), Great Basin Cultural Ecology: A Symposium. Desert Research Publications in the Social Sciences, No 8:43-56.

Delisio, Mario P. 1971 Preliminary Report on the Weston Canyon Rockshelter, Southeastern Idaho: A Big Game Hunting Site in the Northern Great Basin. In C.M. Aikens (ed.), Great Basin Anthropological Conference, 1970, Selected Papers. University of Oregon Anthropological Papers No. 1. Pp. 43-58.

Denny, Charles S. and Harold Drewes 1965 Geology of the Ash Meadows Quadrange Nevada-California. U.S. Geological Survey Bulletin 1181-L. Washington: Government Printing Office. Pp. 1-56.

153 Denton, George H. and Wibjorn Karlen 1973 Holocene Climatic Variations--Their Pattern and Possible Cause. Quaternary Research 3(2):155-205.

Denton, George H. and Stephen C. Porter 1970 Neoglaciation. Scientific American 222(6):101-110.

Durrant, Stephen D. 1970 Appendix II: Faunal Remains as Indicators of Neo- thermal Climates at Hogup Cave. In C. Melvin Aikens, Hogup Cave. University of Utah Anthropological Papers, No. 93. Pp. 241-245.

Eardley, Armand J. 1962 Gypsum Dunes and Evaporite History of the . University of Utah, College of Mines and Mineral Industries, Special Studies, No. 2.

./Eardley, A.J., Vasyl Gvosdetsky, and R.E. Marsell 1957 Hydrology of Lake Bonneville and Sediments and Soils of Its Basin. Geological Society of America Bulletin 68(9):1141-1201.

Ferguson, C.W. 1968 Bristlecone Pine: Science and Esthetics. Science 159(3817):839-846. 1969 A 7104-Year Annual Tree-Ring Chronology for Bristle- cone Pine, Pinus-aristata, from the White Mountains, California. Laboratory of Tree-Ring Research, Tree- Ring Bulletin 29(3-4):3-29. 1970 Dendrochronology of Bristlecone Pine, Pinus aristata: Establishment of a 7484-year Chronology in the White Mountains of East-central California, U.S.A. In I.U. Olsson (ed.), Radiocarbon Variations and Ab- solute Chronology. New York: John Wiley. Pp. 237-59.

Flint, Richard Foster j 1971 Glacial and Quaternary Geology. New York: John Wiley. Fosberg, Maynard A. 1965 Characteristics and Genesis of Patterned Ground in Wisconsin Time in a Chestnut Soil Zone of Southern Idaho. Soil Science 99(1):30-37.

Fowler, Catherine S. 1972 Some Ecological Clues to Proto-Numic Homelands. In Don D. Fowler (ed.), Great Basin Cultural Ecology-F- A Symposium. Desert Research Institute Publications in the Social Sciences, No. 8. Pp. 105-121.

Fowler, Don D. 1972a Editor's Introduction. In Don D. Fowler (ed.), Great Basin Cultural Ecology: A Symposium. Desert Research Institute Publications in the Social Sciences, No. 8. Pp. 1-3.

154 1972b Current Research: Great Basin and California. Ameri- can Antiquity 37(4):577-579.

Fowler, Don D., David B. Madsen, and Eugene M. Hattori 1973 Prehistory of Southeastern Nevada. Desert Research Institute Publications in the Social Sciences, No. 6. Friedman, Irving and George I. Smith 1972 Deuterium Content of Snow as an Index to Winter Climate in the Sierra Nevada Area. Science 176(4036): 790-793.

Fritts, Harold C. 1965 Tree-Ring Evidence for Climatic Changes in Western America. Monthly Weather Review 93(7):421-443. 1966 Growth-Rings of Trees: Their Correlationwith Climate. Science 154:973-979. 1971 Dendroclimatology and Dendroecology. Quaternary Research 1(4):419-449.

Fritts, Harold C., Terence J. Blasing, Bruce P. Hayden and John E. Kutzbach 1971 Multivariate Techniques for Specifying Tree-Growth and Climate Relationships and for Reconstructing Anomalies in Paleoclimate. Journal of Applied Meteorology 10:845-864.

Fritts, Harold C., David G. Smith, and Marvin A. Stokes 1965 The Biological Model for Paleoclimatic Interpretation of Mesa Verde Tree-Ring Series. American Antiquity 31(2):101-121.

Gale, H. S. 1914 Salines in the Owens, Searles, ana Panamint Basins, Southeastern California. U. S. Geological Survey Bulletin 580-L. Washington: Government Printing Office. Pp. 251-323.

-e/' Galloway, Robert W. 1970 The Full-Glacial Climate in the Southwestern United States. Annals of the Association of American Geographers 60(2):245-256.

4.-/' Gilbert, Grove Karl 1890 Lake Bonneville. Monographs of the U. S. Geological Survey, Vol. I. Washington: Government Printing Office.

Green, James Patten 1972 Archaeology of the Rock Creek Site 10 CA 33 Sawtooth National Forest, Cassia County, Idaho. Unpublished M.A. thesis, Idaho State University.

155 Grosscup, G. R. 1956 The Archeology of the Carson Sink Area. Reports of the University of California Archeological Survey, No. 33, Paper 43. Berkeley. Pp. 58-64.

Hack, John T. 1942 The Changing Physical Environment of the Hopi Indians of Arizona. Papers of the Peabody Museum of American Archaeology and Ethnology, Vol. 35, No. 1.

Hansen, Henry P. 1947a Postglacial Forest Succession, Climate, and Chron- ology in the Pacific Northwest. Transactions of the American Philosophical Society Vol. 37 (n.s0:1-130. 1947b Postglacial Vegetation of the Northern Great Basin. American Journal of Botany 34(3):164-171.

Hardman, George and Cruz Venstrom 1941 A 100-Year Record of Truckee River Runoff Estimated from Changes in Levels and Volumes of Pyramid and Winnemucca Lakes. American Geophysical Union Transactions 22:71-90.

Harper, K. T. and G. M. Alder 1970 Appendix I: The Macroscopic Plant Remains of the Deposits of Hogup Cave, Utah, and Their Paleo- climatic Implications. In C. Melvin Aikens , Hogup Cave. University BY Utah Anthropological Papers, No. 93. Pp. 215-240. 1972 Paleoclimatic Inferences Concerning the Last 10,000 Years from a Resampling of Danger Cave, Utah. In Don D. Fowler (ed.), Great Basin Cultural Ecology: A Symposium. Desert Research Institute Publications - in the Social TET&IEes, No. 8. Pp. 13-23. Ha*ley, John W. and William E. Wilson III j1965 Quaternary Geology of the Winnemucca Area, Nevada. Desert Research Institute Technical Report, No. 5.

Haynes, C. Vance, Jr. 01967 Quaternary Geology of the Tule Springs Area, Clark County, Nevada. In H. M. Wormington and Dorothy Ellis (eds.), PleTitocene Studies in Southern Nevada. Nevada State Museum Anthropological Papers, No. 13. Pp. 15-104. 1968 Geochronology of Late-Quaternary Alluvium. In Roger B. Morrison and Herbert E. Wright, Jr.-Teds.), Means of Correlation of Quaternary Successions. University of Utah Press. Pp. 591-631.

Heizer, Robert F. 1951 Preliminary Report on the Leonard Rockshelter Site, Pershing County, Nevada. American Antiquity 17(2):89-97.

156 Hill, Mary R. 1972 The Great Owens Valley Earthquake of 1872. California Division of Mines and Geology, California Geology 25(3):51-54. / Hooke, Roger LeB. V 1972 Geomorphic Evidence for Late-Wisconsin and Holocene Tectonic Deformation, Death Valley, California. Geological Society of America Bulletin 83(7):2073- 2098.

Howell, Paul W. 1960 Interpretation of the Past Climate Represented at Stuart Rockshelter from the Soluble Salt Content of Its Strata. In Dick Shutler, Jr., Mary Elizabeth Shutler, and James S. Griffith (eds.), Stuart Rock- shelter: A Stratified Site in Southern Nevada. Nevada State Museum Anthropological Papers, No. 3. Pp. 23-25. k, Hubbs, Carl L. and Robert R. Miller 1948 The Great Basin, with Emphasis on Glacial and Post- glacial Times: The Zoological Evidence. Bulletin of the University of Utah 38(20):17-166. 1970 Paleohydrography and Remant Fishes of the North- Central Great Basin. Abstracts of the First Meeting of the American Quaternary Associaion. P. 68.

Hunt, Alice 1960 Archeology of the Death Valley Salt Pan. University of Utah Anthropological Papers No. 47.

Hunt, Charles B. and Don R. Mabey 3.966 Stratigraphy and Structure, Death Valley, California. U.S. Geological Survey Professional Paper 494-A. Washington: Government Printing Office. Pp. 1-162.

Hunt, Charles B., T. W. Robinson, Walter A. Bowles, and A. L. Washburn 1966 Hydrologic Basin, Death Valley, California. U.S. Geological Survey Professional Paper 494-B. Washington: Government Printing Office. Pp. 1-138.

LAyne, Norman J., Paul Chelminski, James Court, Donn Gorsline, and Charles Goldman 1972 Quaternary History of Lake Tahoe, California-Nevada. Geological Society of America Bulletin 83(5):1435- 1448. Izett, G. A., Ray E. Wilcox, H. A. Powers, and G. A. Desborough 1970 The Bishop Ash Bed, a Pleistocene Marker Bed in the Western United States. Quaternary Research 1(1): 121-132.

157 Jennings, Jesse D. 1957 Danger Cave. University of Utah Anthropological Papers, No. 27.

Kautz, Robert R. and David Hurst Thomas 1973 Palynological Investigations of Two Prehistoric Cave Middens in Central Nevada. Tebiwa: The Journal of the Idaho State University Museum 15(2):43-54.

Kelso, Gerald 1970 Appendix IV: Hogup Cave, Utah: Comparative Pollen Analysis of Human Coprolites and Cave Fill. In C. Melvin Aikens, Hogup Cave. University of Utah Anthropological Papers, No. 93. Pp. 251-262.

Kittleman, Laurence R. 1973 Mineralology, Correlation, and Grain-Size Distribu- tions of Mazama Tephra and Other Postglacial Pyroclastic Layers, Pacific Northwest. Geological Society of America Bulletin 84(9):2957-2980.

LaMarche, Valmore C., Jr. 1968 Rates of Slope Degradation as Determined from Botanical Evidence: White Mountains, California. U.S. Geological Survey Professional paper 352-1. Washington: Government Pringing Office. Pp. 341- 377. 1969 Environment in Relation to Age of Bristlecone Pines. Ecology 50(1):53-59. 1973 Holocene Climatic Variations Inferred from Treeline Fluctuations in the White Mountains, California. Quaternary Research 3(4):632-660. 1974 Paleoclimati-C-YEETences from Long tree-Ring Records. Science 183(4129):1043-1048.

LaMarche, Valmore C., Jr. and H. C. Fritts 1971 Anomaly Patterns of Climate over the Western United States, 1700-1930, Derived from Principal Component Analysis of Tree-Ring Data. Monthly Weather Review 99(2):138-142.

LaMarche, Valmore C., Jr. and Harold A. Mooney 1967 Altithermal Timberline Advance in Western United States. Nature 213(5080):980-982. 1972 Recent ClIMYETE Change and Development of the Bristlecone Pine (P. longaeva Bailey), Krummholz Zone, Mt. Washington, Nevada. Arctic and Alpine Research 4(1):61-72.

Laudermilk, J. D. and P. A. Munz 1934 Plants in the Dung of Nothrotherium from Gypsum Cave, Nevada. Carnegie Institution of Washington, Publication 453. Pp. 29-37.

158

1938 Plants in the Dung of Nothrotherium from Rampart and Muav Caves, Arizona. Carnegie Institution of Washington, Publication 487. Pp. 271-281.

Leopold, Estella B. 1-7 1951 Pleistocene in New Mexico. American Journal of Science 249:152-168. 17'1967 Summary of Palynological Data from Searles Lake. In Guidebook for Friends of the Pleistocene, Pacific Coast Section. [Meno Park?]: U.S. Geological Sur- vey. Pp. 52-66. 1970 Late Wisconsin and Modern Pollen Rain at Searles Lake, California. Supplement to the Abstracts of the First Meeting of the American Quaternary Association. Pp. 11-12.

Leskinen, Paul H. 1970 Late Pleistocene Vegetation Change in the Christmas Tree Pass Area, Newberry Mountains, Nevada. Unpublished M.A. Thesis, University of Arizona.

Long, Austin and Paul S. Martin 1974 Life and Death of the Shasta Ground Sloth. Science 186(4164):638-640.

Loope, Lloyd Lee 1969 Subalpine and Alpine Vegetation of Northeastern Nevada. Unpublished Ph.D. Dissertation, Duke University.

Mabey, D. R. 1971 Geophysical Data Relating to a Possible Pleistocene Overflow of Lake Bonneville at Gem Valley, South- eastern Idaho. U.S. Geological Survey Professional Paper 750-B. Washington: Government Printing Office. Pp. 122-127.

MacDonald, Angus A. 1970 The Northern 's Little . California Division of Mines and Geology, Mineral Information Service 23(1):3-6.

Madsen, David B. 1972 Paleoecological Investigations in Meadow Valley Wash, Nevada. In Don D. Fowler (ed.), Great Basin Cultural Ecology: A Symposium. Desert Research Institute Publications in the Social Sciences, No. 8. Pp. 57- 65.

Major, Jack and Samuel A. Bamberg (i 1967 Some Cordilleran Plants Disjunct in the Sierra Nevada of California, and Their Bearing on Pleisto- cene Ecological Conditions. In H. E. Wright, Jr. and W. H. Osburn (eds.), Arctic and Alpine Environments. Bloomington and London: India University Press. Pp. 171-188.

159 Malde, Harold E. 1961 Patterned Ground in the Western Snake. River Plain, Idaho, and Its Possible Cold-Climate Origin. Geological Society of America Bulletin 75(3):191-208. vf964 The Ecologic Significance of Some Unfamiliar Geo- logic Processes. In The Reconstruction of Past Environments. Fort Burgwin Research Center, Publica- tion No. 3. Pp. 7-13. ,2968 The Catastrophic Late Pleistocene Bonneville Flood in the Sanke River Plain, Idaho. U.S. Geological Survey Professional Paper 596. Washingtoh: Government Printing Office. Pp. 1-52.

Martin, Paul S. 1967 Prehistoric Overkill. In P. S. Martin and H. E. Wright, Jr. (eds.), Pleistocene Extinctions: The Search for a Cause. New Haven and London: Yale University Press. Pp. 75-120. 1973 The Discovery of America. Science l79(4077):969- 974.

Mavtin, Paul S. and Peter J. Mehringer, Jr. d1965 Pleistocene Pollen Analysis and Biogeography of the Southwest. In H. E. Wright, Jr. and David G. Frey (eds.), The Quaternary of the United States. Princeton: Princeton University Press. Pp. 433-451.

Martin, Paul S., Bruno E. Sabels and Dick Shutler, Jr. 1961 Rampart Cave Coprolite and Ecology of the Shasta Ground Sloth. American Journal of Science 259:102- 127.

Marwitt, J. P., G. F. Fry, and J. M. Adovasio 1971 Sandwich Shelter. In C. M. Aikens (ed.), Great Basin Anthropological Conference, 1970, Selected Papers. University of Oregon Anthropological Papers, No. 1. Pp. 27-36.

Mawby, John E. 1967 Fossil Vertebrates of the Tule Springs Site, Nevada. In H. M. Wormington and Dorothy Ellis (eds.), Pleistocene Studies in Southern Nevada. Nevada State Museum Anthropological Papers, No. 13. Pp. 105-128.

Mehringer, Peter J., Jr. 967a Pollen Analysis of the Tule Springs Area, Nevada. I/ In H. M. Wormington and Dorothy Ellis (eds.), Nevada State Museum Anthropological Papers, No. 13. Pp. 129-200. 1/ ,967b The Environment of Extinction of the Late-Pleistocene 1 Megafauna in the Arid Southwestern United States. In P. S. Martin and H. E. Wright, Jr. (eds.), Pleisto- cene Extinctions: The Search for a Cause. New Haven and London: Yale University Press. Pp. 247-266.

160 Mehringer, Peter J., Jr. and Charles W. Ferguson 1969 Pluvial Occurrence of Bristlecone Pine (Pinus aristata) in a Mohave Desert Mountain' Range. Journal of the Arizona Academy of Science 5(4):284-292.

Mehringer, Peter J., Jr., William P. Nash, and Richard H. Fuller 1971 A Holocene Volcanic Ash from Northwestern Utah. Proceedings of Utah Academy of Sciences, Arts, and Letters 48(1):46-51.

Meinzer, 0. E. 1922 Map of the Pleistocene Lakes of the Basin and Range Province and Its Significance. Geological Society of America Bulletin 33(3):541-552.

Mifflin, M. D. 1968 Delineation of Ground-Water Flow Systems in Nevada. V Desert Research Institute Technical Report Series H-W. Hydrology and Water Resources Publication, No. 4.

LAiiller, Robert Rush 1950 Speciation in Fishes of the Genera Cyprinodon and Empetrichthys, Inhabiting the Death Valley Region. Evolution IV(2):155-163. IV 1965 Quaternary Freshwater Fishes of North America. In H. E. Wright, Jr. and David G. Frey (eds.), The -- Quaternary of the United States. Princeton: Princeton University Press. Pp. 569-596.

Mitchell, Val Leonard 1969 The Regionalization of Climate in Montane Areas. Unpublished Ph.D. Dissertation, "University of Wisconsin.

Morrison, Roger B. 1964 Lake Lahontan: Geology of Southern Carson Desert, Nevada. U.S. Geological Survey Professional Paper 401. Washington: Government Printing Office. Pp. 1-156. r 7-1965a Quaternary Geology of the Great Basin. In H. E. Wright, Jr. and David G. Frey (eds.), The Quaternary of the United States. Princeton: Princeton University Press. Pp. 265-285. 1965b New Evidence on Lake Bonneville Stratigraphy and History from Southern Promontory Point, Utah. U.S. Geological Survey Professional Paper 525-C. Washington: Government Printing Office. Pp. 110-

V 1967 Principles of Quaternary Soil Stratigraphy. In R.B. Morrison and H. E. Wright, Jr.(eds.), Quaternary Soils. Proceedinvs of the VII Congress of the Inter- national Association--YOr Quaternary Research-T-Vol. 9. Reno: Desert Research Institute. Pp. 1-69.

161 Plorrison, Roger B. 04 1970 Conflicting Pluvial-Lake Evidence on Climatic Changes between 14 and 9 Millenia Ago, with Particu- lar Reference to Lakes Lahontan, Bonneville, and Searles. Abstracts of the First Meeting of ---the American Quaternary Association. Pp. 97-7L Morrison, R. B. and John C. Frye v/ 1965 Correlation of the Middle and Late Quaternary Suc- cessions of the Lake Lahontan, Lake Bonneville, Rocky Mountain (), Southern Great'Plains, and Eastern Midwest Areas. Nevada Bureau of Mines Report 9. Reno: University of Nevada.

Mullineaux, Donald R., Jack H. Hyde, and Meyer Rubin 1972 Preliminary Assessment of Upper Pleistocene and Holocene Pumiceous Tephra from Mount St. Helens, Southern Washington. Geological Society of America, Cordilleran Section, 68th Annual Meeting Abstracts with Programs 4(3). P. 204.

Mundorff, J. C. 1971 Nonthermal Springs of Utah. Utah Geological and Mineralogical Survey, Water-Resources Bulletin 16.

Oakeshott, Gordon B., Roger W. Greensfelder, and James E. Kahle 1972 One Hundred Years Later. California Division of Mines and Geology, California Geology 25(3):55-61.

\)401,ie, H. Thomas and Claude N. Warren 1971 Late Pleistocene-Early Holocene Geomorphic History of Lake Mojave, California. Geological Society of America Bulletin 82(9):2553-2562.

Peck, Eugene L. and E. Arlo Richardson V1966 Hydrology and Climatology of Great Salt Lake. In William Lee Stokes (ed.), Guidebook to the Geology of Utah, No. 20. Salt Lake City: Utah Geological Society. Pp. 121-134.

Pe4Arson, N. V. and E. A. Groh C- 1969 The Ages of Some Holocene Volcanic Eruptions in the Newberry Volcano Area, Oregon. The Ore Bin 31(4): 73-87.

Irter, Stephen C. Pl' 1971 Fluctuations of Late Pleistocene Alpine Glaciers in Western North America. In Karl K. Turekian (ed.), Late Cenozoic Glacial Ages. New Haven and London: Yale TEITia7gIty Press. Pp. 307-329.

Powers, Howard A. and Ray E. Wilcox 1964 Volcanic Ash from Mount Mazama (Crater Lake) and from Glacier Peak. Science 144(3624):1334-1336.

162 Randle, Keith, Gordon G. Goles, and Laurence R. Kittleman 1971 Geochemical and Petrological Characterization of Ash Samples from Cascade Range Volcanoes. Quaternary Research 1(2):261-282.

Ranere, Anthony J. 1970 Prehistoric Environments and Cultural Continuity in the Western Great Basin. Tebiwa: The Journal of the Idaho State University Museum 1777):52-72.

Richmond, Gerald M. 1964 Glaciation of Little Cottonwood and Bells Canyons, Wasatch Mountains, Utah. U.S. Geological Survey Professional Paper 454-D. Washington: Government Printing Office. Pp. 1-41.

jitioosma, Aino 1958 A Climatic Record from Searles Lake, California. Science 128:176.

,,,Roscoe, Ernest 1963 Stratigraphic Summary of Quaternary Bonneville Basin Mollusca. Sterkiana 34:1-23.

, Israel Cook t”11181 Geological History of Lake Lahontan: A Quaternary Lake of Northwestern Nevada. Monographs of the U.S. Geological Survey, Vol. XI. Washington: Government Printing Office. 1889 Quaternary History of Mono Valley, California. Eighth Annual Report of the U.S. Geological Survey, 1887. TiWiTIEgton: Government Printing Office. Pp. 261-394.

Sabels, Bruno E. 1960 Trace Element Studies on Cave Strata with Comments on Stuart Rockshelter. In Dick Shutler, Jr., Mary Elizabeth Shutler, and James S. Griffith (eds.), A Stratified Site in Southern Nevada. Nevada State Museum Anthropological Papers, No. 3. Pp. 17-23.

Scholz, Christopher H., Muawia Barazangi, and Marc L. Sbar 1971 Late Cenozoic Evolution of the Great Basin, Western United States, as an Ensialic Interarc Basin. Geological Society of America Bulletin 82(11):2979- 2990.

Sear, Paul B. and Aino Roosma 1961 A Climatic Sequence from Two Nevada Caves. American Journal of Science 259:669-678.

Sharp, Robert P. 1966 Kelso Dunes, Mojave Desert, California. Geological Society of America Bulletin 77(10):1045-1074.

163 Sharp, Robert P. 1972 Pleistocene Glaciation, Bridgeport Basin, California. Geological Society of America Bulletin 83(8):2233- 2260.

Sheppard, Richard A. and Arthur J. Gude, III pi 1968 Distribution and Genesis of Authigenic Silicate Minerals in Tuffs of Pleistocene , Inyo County, California. U.S. Geological Survey Profes- sional Paper 597. Washington: Government PTIRTIrig Office. Pp. 1-38.

Shctidan, Michael F. 3/1971 Guidebook to the Quaternary Geology of the East- Central Sierra Nevada. Phoenix: Lebeau Printing Company.

Shutler, Richard, Jr. 1961 Lost City: Pueblo Grande de Nevada. Nevada State Museum Anthropological Papers, No. 5.

Slemmons, David B., Austin E. Jones, and James I. Gimlett 1965 Catalogue of Nevada Earthquakes, 1852-1960. Seismological Society of America Bulletin 55:519- 565.

Smith, George I. 1967a Geological Survey Research 1967: Chapter A. U.S. Geological Survey Professional Paper 575-A. Washington: Government Printing Office. P. 89. 1967b Field Guide to Examples of Later Quaternary Geology, Searles Valley, California. In Guidebook for Friends of the Pleistocene, Pacific Coast Section. Menlo Park?]: U.S. Geological Survey. Pp. 1-51. 17 Late-Quaternary Geologic and Climatic History of Searles Lake, Southeastern California. In Roger B. Morrison and Herbert E. Wright, Jr. (eds77, Means of Correlation of Quaternary Successions. Salt Lake n-ty: University of Utah Press. Pp. 293-310. 1970 Late Wisconsin Lake Fluctuations in Searles Valley, California. Abstracts of the First Meeting of the American Quaternary Associaion. P. 124.

Smith, George I., Irving Friedman, and Sadao Matsuo 1970 Salt Crystallization Temperatures in Searles Lake, California. Mineralogical Society of America Special Publication, No. 3. Washington. Pp. 257- 259.

4Tith, G. R., W. L. Stokes, and K. F. Horn V 1968 Some Late Pleistocene Fishes of Lake Bonneville. Copeia 4:807-816.

164 Smith, H. T. U. 1967 Past versus Present Wind Action in the Mohave Desert Region, California. AFCRL-67-0683, Scientific Report, No. 1. Bedford: U.S. Air Force Cambridge Research Laboratories. Snyder, C. T., George Hardman, and F.F. Zdenek / 1964 Pleistocene Lakes in the Great Basin. U.S. Geolog- ical Survey, Map 1-416. Washington: Government Printing Office. Snyder, Charles T. and Walter B. Langbein 1962 The Pleistocene Lake in Spring Valley, Nevada, and Its Climatic Implications. Journal of Geophysical Research 67(6):2385-2394. Souther, J. G. 1970 Recent Volcanism and Its Influence on Early Native Cultures of Northwestern British Columbia. In R. A. Smith and J. W. Smith (eds.), Early Man and Environ- ments in Northwest North America. Calgary: University of Calgary Archaeological Association. Pp. 53-64. Steward, Julian H. 1933 Ethnography of the Owens Valley Paiute. University of California Publications in American Archaeology and Ethnology, Vol. 33, No. 3. Pp. 233-350. 1938 Basin-Plateau Aboriginal Sociopolitical Groups. Bulletin of the Bureau of American Ethnology, No. 120. Washington: Government Printing Office. Taylor, Dwight W. 1967 Late Pleistocene Molluscan Shell8 from the Tule Springs Area. In H. M. Wormington and Dorothy Ellis (eds.), Pleistocene Studies in Southern Nevada. Nevada State Museum Anthropological Papers, No. 13. Pp. 395-399. 1970 West American Freshwater Mollusca, 1: Bibliography of Pleistocene and Recent Species. San Diego Society of Natural History Memoir 4. Tuohy, Donald R. and Mercedes C. Stein 1969 A Late Lovelock Shaman and His Grave Goods. Nevada State Museum Anthropological Papers, No. 14, Miscel- laneous Paper, No. 4. Pp. 96-130. Van Devender, Thomas R. and James E. King 1971 Late Pleistocene Vegetational Records in Western Arizona. Journal of the Arizona Academy of Science 6(4):240-244.

165 Wallace, Robert E. 1961 Deflation in Buena Vista Valley, Pershing County, Nevada. U.S. Geological Survey Professional Paper 424-D. Washington: Government Printing Office. Pp. 242-244.

Wallace, William J. and Edith S. Taylor 1959 A Preceramic Site at Saratoga Springs, Death Valley National Monument. Contributions to California Archaeology, No. 3, Pt. II. Los Angeles: Archae- ological Research Associates. Pp. 1-13. '

Wells, Philip V. 1969 Preuves Paleontologiques dire Vegetation Tardi- Pleistocene (Datee par le C) dans les Regions Aujourd'hui Desertiques d'Amerique du Nord. Translated by C. Vergnaud-Grazzini. Revue de Geographie Physique et de Geologie Dyn'irligue 11(2):335-340.

'Wells, Philip V. and Rainer Berger 1967 Late Pleistocene History of Coniferous Woodland in the Mohave Desert. Science 155(3770):1640-1647.

Wilcox, Ray E. 1959 Some Effects of Recent Volcanic Ash Falls with Especial Reference to Alaska. U.S. Geological Survey Bulletin 1028-N. Washington: Government Printing Office. Pp. 409-476.

Williams, Howel 1941 Crater Lake: The Story of Its Origin. Berkeley: University of California Press.

Williams, Pete A. and Robert I. Orlins 1963 The Corn Creek Dunes Site: A Dated Surface Site in Southern Nevada. Nevada State Museum Anthropolog- ical Papers, No. 10.

Wilson, George M. 1963 Geology of Deer Creek Cave and the Jarbidge Area. In Mary Elizabeth Shutler and Richard Shutler, Jr. (eds.), Deer Creek Cave, Elko County, Nevada. Nevada State Museum Anthropological Papers, No. 11. Pp. 9-13.

Winograd, Isaac Judah 1971 Origin of Major Springs in the Amargosa Desert of Nevada and Death Valley, California. Unpublished Ph.D. Dissertation, University of Arizona.

166 Nfrood, Spencer H. 1972 Preliminary Holocene Volcanic Ash Chronology for the Yosemite National Park Region of the Sierra Nevada, California. Geological Society of America, Cordilleran Section, 68th Annual Meeting, Abstracts with Programs 4(3). P. 263.

/Wright, H. E., Jr. 1971 Late Quaternary Vegetational History of North America. In Karl K. Turekian (ed.), Late Cenozoic Glacial Ages. New Haven and London: Yale University Press. Pp. 425-464.

Ziegler, Alan C. 1963 Unmodified Mammal and Bird Remains from Deer Creek Cave, Elko County, Nevada. In Mary Elizabeth Shutler and Richard Shutler, Jr. (eds.), Deer Creek Cave, Elko County, Nevada. Nevada State Museum Anthropological Papers, No. 11. Pp. 15-22.

167

.