tt;
IE7S
Pt
Y. foX-fl
7 DAY
The Hydrology and Mineralogy of
Deep Springs Lake Inyo County, California
GEOLOGICAL SURVEY PROFESSIONAL PAPER 502-AThe Hydrology and Mineralogy of Deep Springs Lake Inyo County, California
By BLAIR F. JONES
CLOSED-BASIN INVESTIGATIONS
GEOLOGICAL SURVEY PROFESSIONAL PAPER 502-A
A study of the relation between hydrologic factors and saline mineralogy for a small playa in the western Great Basin
UNITED STATES GOVERNMENT PRINTING OFFICE, WASHINGTON : 1965UNITED STATES DEPARTMENT OF THE INTERIOR STEWART L. UDALL, Secretary
GEOLOGICAL SURVEY Thomas B. Nolan, Director
For sale by the Superintendent of Documents, U.S. Government Printing Office
Washington, D.C. 20402CONTENTS
Page
Abstract_______________________________________________ A1
Introduction____________________________________________ 2
Acknowledgments____________________________________ 2
Geography__________________________________________ 3
Bedrock geology of the Deep Springs Valley_________ 6
Quaternary geology_________________________________ 8
Hydrology----------------------------------------------- 9
Precipitation______________________________________ 9
Evaporation_______________________________________ 12
Hydrography of the Deep Springs Valley____________ 12
Streams and drainage of the mountain areas__ 13
Springs north of Deep Springs Lake________________ 15
Springs east of Deep Springs Lake----------------- 15
Sag ponds_________________________________________ 17
Ground water______________________________________ 17
Comparison of discharge from streams and springs.. 20
Hydrography of the lake--------------------------- 20
Hydrochemistry_________________________________________ 25
Analytical methods________________________________ 25
Page
Hydrochemistry—Continued
Sampling____________________________________________ A25
pH and alkalinity____________________________________ 26
Major-element analysis_______________________________ 27
Sulfide.............................................. 27
Presentation and interpretation of data__________ 27
Inflow_______________________________________________ 32
Lake area____________________________________________ 40
Closed sag pond______________________________________ 42
Calculation of salinity from hydrographic data___ 43
Mineralogy________________________________________________ 43
Carbonate and saline mineralogy______________________ 43
Areal distribution in the lacustrine deposits________ 44
Saline crusts________________________________________ 45
Efflorescent crusts__________________________________ 48
Summary of carbonate and saline mineralogy________________ 48
Silicate mineralogy__________________________________ 49
Geochemistry______________________________________________ 49
References------------------------------------------------ 53
ILLUSTRATIONS
[Plates are in pocket]
Plate 1. Generalized geologic map of Deep Springs Valley. Page
2. Topography of the bed of Deep Springs Lake.
Figure 1. Index map of the Deep Springs Valley and vicinity, California_____________________________________________ A3
2. Index map of the Deep Springs Lake area__________________________________________________________________ 4
3. Aerial photograph of the Deep Springs Lake area__________________________________________________________ 5
4. View of fault scarps east of Deep Springs Lake___________________________________________________________ 6
5. Map showing stratigraphy of surficial deposits at Deep Springs Lake_____________________________________ 10
6. Graph showing precipitation in the Deep Springs Valley________________________________________________ 12
7. View across the bog-mound marsh_________________________________________________________________________ 15
8. Views of Buckhorn Springs 5 and 7_______________________________________________________________________ 16
9. Cemented fanglomerate on fault scarp above Buckhorn Springs 10 and 11_______________________________ 16
10. View of the closed sag pond at low stage________________________________________________________________ 17
11. Section of the alluvium in the Deep Springs Valley______________________________________________________ 18
12. High-altitude aerial photograph of the Deep Springs Valley______________________________________________ 19
13. Graph showing relation of water surface area to volume for Deep Springs Lake_____________________ 22
14. View of Deep Springs Lake at low stage__________________________________________________________________ 23
15. View of stagnant brine pool_____________________________________________________________________________ 23
16. Graph showing relation of lake volume to total dissolved salts at Deep Springs Lake_____________________ 24
17. View of windblown salt dust___________________________________________________________________________ 24
18. Graph showing relation of surface-water volume to dissolved-solids concentration for a receding phase of
Deep Springs Lake__________________________________________________________________________________ 25
19-22. Diagram showing-
19. Chemical composition of inflow to Deep Springs Lake____________________________________________ 31
20. Composition of well waters from the Deep Springs Valley________________________________________ 32
21. Major anions in waters from Deep Springs Lake__________________________________________________ 33
22. Trends in composition of dissolved solids in inflow to Deep Springs Lake_______________________ 33
mIV
CONTENTS
Figure 23-26. Diagram showing relation of ion concentration to discharge for— Pase
23. Wyman Creek_____________________________________________________________________________________ A36
24. Corral Spring 1__________________________________________________________________________________ 36
25. Corral Spring 2__________________________________________________________________________________ 36
26. Buckhorn Springs_________________________________________________________________________________ 37
27-30. Diagram showing relation of—
27. Ion concentration to total dissolved solids for Corral Spring inflow----------------------------- 38
28. Cation concentration to total dissolved solids for Buckhorn Spring inflow------------------------ 39
29. Anion concentration to total dissolved solids for Buckhorn Spring inflow------------------------- 39
30. pH to total dissolved solids in Buckhorn Spring inflow___________________________________________ 40
31. View of ephemeral runoff on interdunal flats north of Deep Springs Lake------------------------------- 41
32. View of gas vent on the north side of the Deep Springs Lake saltpan------------------------------------ 41
33. Diagram showing brines from Deep Springs Lake in terms of the reciprocal system Na2S04-K2S04-NaCl-
KC1................................................................................................ 42
34. Graph showing relation of lake stage to S04: Cl ratio__________________________________________________ 42
35. View of surface of salt crust at Deep Springs Lake_____________________________________________________ 45
36. Section of salt crust__________________________________________________________________________________ 47
37. Diagram showing composition of salt phases_____________________________________________________________ 50
38. Diagram showing mineral assemblages in equilibrium with saturated solutions---------------------------- 52
TABLES
Page
Table 1. Generalized column of sedimentary bedrock in the Deep Springs Valley drainage area________________________ A7
2. Climatic data for Deep Springs Valley and vicinity, California____________________________________________ 11
3. Estimated discharge of surface water units in Deep Springs Valley, 1959-61-------------------------------- 13
4. Data on water surface area, volume, and tonnage of dissolved solids at Deep Springs Lake__________________ 21
5. Hydrologic data from Deep Springs Lake as compared with some major closed lakes cited by Langbein--------- 22
6. Comparison of total C02, pH, and alkalinity determinations for samples of water from the Deep Springs Valley, 28
7. Analyses of waters from the Deep Springs Valley, 1959-61__________________________________________________ 29
8. Precipitate mineral species at Deep Springs Lake________________________________________________________ 44
9. Saline mineral assemblages in the system NaHCOs-Na^jCOj-NaaSO^LO at Deep Springs Lake_____________________ 51CLOSED-BASIN INVESTIGATIONS
THE HYDROLOGY AND MINERALOGY OF DEEP SPRINGS LAKE, INYO COUNTY, CALIFORNIA
By Blair F. Jones
ABSTRACT
Deep Springs Lake is a small intermittent saline lake at an altitude of about 5,000 feet within a completely closed basin in northern Inyo County, Calif. The basin is a relatively recently formed graben, in which the most active faulting occurred on the east, within or near the White-Inyo Mountains horst block.
The Deep Springs Valley was formerly occupied by a large lake which may have covered more than one-half of the valley floor as recently as 1,000 years B.P. Recent normal faulting has cut off outflow, tilted old shorelines, controlled the location of the playa in the southeast corner of the valley, and provided a locus for a major belt of springs which flow into the lake.
In the bedrock geology of the Deep Springs Valley the feature of greatest importance to the valley hydrology is the contact between granodioritic igneous rocks to the north and east and ancient sedimentary strata lying to the south and west. An intermediate belt of contact metamorphic rocks and associated mineralized zones is also very influential. The alluvium of the northern part of the valley is composed largely of fan materials which are derived from the Wyman-Crooked Creeks system, the major perennial drainage to the valley. The northern valley alluvium contains abundant zones of high permeability related to old channel deposits. Lacustrine materials can be recognized in the central part of the valley 5 miles from the present lake area and may contain impermeable confining layers composed of very compact clays or carbonate muds. The present playa area has three types of lacustrine deposits listed from the center of the playa outward or with increasing depth as (1) the saline crusts, (2) carbonate muds, and (3) fine-grained well-sorted silts and sands.
Annual precipitation in the Deep Springs Valley averages from about 5 inches on the valley floor to 15 inches on the basin divide. Distribution is intermediate between the summer minimum pattern of the California coast and the more uniform annual distribution of the Great Basin. The evaporation rate of the saline waters of Deep Springs Lake may be as low as 30 inches per year, but the mean is probably almost 52 inches per year.
The hydrography of the Deep Springs Valley is dominated by the perennial streams of the Wyman-Crooked Creeks and Birch-Antelope Creeks systems, and the three major spring groups around Deep Springs Lake—the Corral, Buckhorn, and bog-mound springs. Flow patterns suggest (a) that the Birch-Antelope Creeks system is connected with the western bog-mound springs, (b) and that the Wyman-Crooked Creeks
system is connected with the central and eastern bog-mound springs, and, to a lesser extent, (c) with the Corral and Buck-horn Springs along the prominent fault zone east of the lake. The Corral Springs discharge ground water from igneous rocks along the fault zone north of the lake; the Buckhorn Springs discharge largely ground water from sedimentary and contact metamorphic rocks intercepted by the fault zone south of the lake.
Discharge estimates indicate that inflow to Deep Springs Lake from the Wyman-Crooked Creeks, Corral Springs, or Buckhorn Springs systems is roughly equal. Seasonal fluctuation in the discharge of the springs with change in storm patterns suggests substantial contribution from sources other than the main valley ground-water reservoir.
Deep Springs Lake is generally pan shaped, and there is a roughly linear logarithmic relation between its area and volume. The lake demonstrates greater areal variation and shorter response time than anything considered in the closed lake compilation of Langbein (1961) and is thus most similar to the numerous small playas so common in the Great Basin. In spite of uncertainties in volume computations and the homogeneity of surface waters at Deep Springs Lake, a relation between volume and tonnage of dissolved salts was established. This relation follows, perhaps fortuitously in detail, the “ideal cycle” for a closed lake proposed by Langbein (1961). Continuity in the relations of water volume and total concentration is interrupted by the precipitation of sodium sulfate. Loss of dissolved salts by entrapment in marginal sediments during lake-level recession was quantitatively verified at Deep Springs Lake; lacustrine muds contain as much as 30 percent readily water-soluble material, and some contain over 80 percent of material of precipitate origin (that is, including alkaline earth carbonates).
Analytical studies at Deep Springs Lake have indicated that most presently available data on carbonate in concentrated natural waters can be considered only semiquantitative at best. Of all the constituents for which saline waters are normally analyzed, only sulfate and chloride are consistently reliable.
Chemical variation in more than 125 water samples from the Deep Springs Valley is most effectively shown by trilinear plotting. The chemical compositions of most waters considered inflow to the lake fall within fairly well-defined limits characteristic of the distinct hydrologic units. The general trend is toward increase of the percentage of alkali, sulfate, and chloride simultaneously with increase in total dissolved solids. The major factors affecting the differences in chemical com-
A1A2
CLOSED-BASIN INVESTIGATIONS
position of inflow units are gross environmental lithology and processes taking place within and between each unit. The chief processes involved are the precipitation of alkaline earth carbonate, sulfate reduction, and additional solution of alkali material. Cation exchange is considered a minor factor in affecting compositional change. The chemical data generally support the flow patterns deduced for the valley. Variations in water composition may also be related to seasonal fluctuations in discharge, temperature and biologic activity, but original lithologic control appears more important than any climatic factor. A shift of source areas due to the seasonal change of storm patterns may affect the composition of waters significantly, as indicated by the northern Buckhorn Springs. Substantial compositional changes take place between inflow springs and the lake itself; the major cause of these changes is evaporation associated with alkaline earth carbonate precipitation, and re-solution of capillary salts.
Ephemeral inflow to the lake very definitely reflects the immediate area from which it is derived, and tends to retain its identity within the playa area for some time.
Waters from Deep Springs Lake fall into relatively narrow ranges in percentage composition depending on the immediate environment. Surface waters generally have less than 10 percent total alkalinity, or about 5 percent below the average values for intercrustal brine. Sulfate and chloride vary about 20 equivalent percent, and potassium is stable at a relatively high level for natural waters. In surface waters, sulfate to chloride ratios fluctuate significantly with changes in stage and show the dominance of sulfate in the readily soluble salts of the saltpan.
The equation derived by Langbein (1961) for the calculation of salinity from hydrographic data for permanent closed lakes cannot be applied to a playa such as Deep Springs Lake because of the importance of solution kinetics, relative salt solubilities, thickness of exposed crust, and residual brines.
More than 15 different species of saline and carbonate minerals occur at Deep Springs Lake. There are three distinct patterns in their distribution: (1) areal zoning of precipitate minerals, (2) layer variation in saline crusts, and (3) local variation in capillary efflorescences. The areal sequence of mineral zones from playa margin to center is calcite and (or) aragonite, dolomite, gaylussite, thenardite, and burkeite. Dolomite, in part primary, is the most abundant mineral of the carbonate muds. Thenardite is the dominant mineral of the saline crusts. Burkeite is the most characteristic mineral of the central salt pan. The layer sequence of crustal minerals in ascending order is nahcolite, thenardite, burkeite, trona, and halite. Capillary efflorescent crusts are typical of the western part of the playa. These deposits are dominated by halite and thenardite, but variation is great and is indicative of the anion composition of ephemeral inflow. The silicate mineralogy is dominated by quartz and 10-A clay species, particularly 2Mi mica, but chlorite, talc, and poorly defined expandable lattice types were also identified. The major effects of the saline environment on the clays appear to be the “degradation” (probably oxidation and magnesium loss) in chlorite and possibly the reorganization of montmorillonite.
The data on the hydrochemistry and saline mineralogy indicate that Deep Springs Lake is basically a sodium sulfo-carbonate system. The most characteristic mineral of the central lake crusts is burkeite. In general, the formation of burkeite is favored by relatively high temperature, salinity, and low pCOa. Comparison of mineral assemblages at Deep
Springs Lake with experimental data on burkeite stability (Jones, 1962) with results of work by Teeple (1929) on the quinary NaAXL-NaHCCL-NaoSCh-NaCl-ILO suggests that the salines reflect their original sequence of precipitation but trend toward local equilibrium with associated brines.
INTRODUCTION
The intermontane basins of the Western States provide numerous examples of complete hydrologic units where the end product is a brackish or saline lake, a saltpan, or a playa. In such closed basins, both the derivation and disposition of solutes should be accountable in terms of processes taking place within a single confined area. Langbein (1961) has attempted to explain the variable salinity of closed lakes in terms of their geometric properties and hydrologic environment. A further step would be to relate the specific chemical nature of such waters and associated evapo-rite deposits to the hydrologic processes operative in each basin.
The waters of closed basins are often classified on the basis of the dominant anion, such as carbonate, sulfate, or chloride (Hutchinson, 1957). In the Great Basin, two main types are represented; the Bonneville basin, which is high in chloride, and the Lahontan basin, where alkali carbonates predominate. Further south in the Mojave Desert region, sulfate becomes a major component of the waters. These differences have been attributed to the source of solutes and conditions of evaporation, but outside of Great Salt Lake and Death Valley (Hunt, 1960) no detailed study has been made.
After a reconnaissance of the western Great Basin early in 1959, the Deep Springs Valley was chosen as a workable and reasonably representative area for the study of a closed basin of the sulfocarbonate type. Fieldwork was done during late spring and early fall from October 1959 to September 1961. The chemical variations in the waters could not be related either to hydrologic environment or source of solutes alone but to an intimate interrelation of both as evinced by the deposits which the waters leave behind.
ACKNOWLEDGMENTS
For much assistance in this study, I am especially grateful to Prof. Hans P. Eugster of the Johns Hopkins University, who participated in selection of the area for detailed study and provided detailed advice and constructive criticism of the manuscript. I am indebted to Prof. Charles B. Hunt also of the Johns Hopkins University for critical reading of the manuscript, and to Betty F. Jones for the manuscript outline and preliminary organization. Most of the analyticalHYDROLOGY AND MINERALOGY OF DEEP SPRINGS LAKE, CALIF.
A3
work was performed by Miss Shirley L. Rettig, Quality of Water Branch laboratory, Washington, D.C.
GEOGRAPHY
The Deep Springs Valley is a small intermontane basin (alt. about 5,000 ft) in northern Inyo County, Calif., which has a total drainage area of 200 square miles. To the west and northwest of this northeast-trending elongate basin are the White Mountains (altitudes up to 11,278 ft. in the valley drainage area). Most of the drainage to the valley is derived from precipitation in these mountains. The valley rim to the east and south rises gradually to a maximum altitude of 8,924 feet at the southernmost end of the valley, then merges with the White Mountains to form the Inyo Mountains. Beyond the mountains west of the Deep Springs Valley lies the Owens Valley and the high Sierra Nevada, which intercepts much precipitation that would otherwise fall within the Deep Springs Valley drainage area. To the east of the Deep Springs Valley lies the much larger and much lower Eureka Valley; to the north of both is the Fish Lake Valley. An index map of the Deep Springs Valley and vicinity is shown in figure 1.
The Deep Springs Valley floor is about 15 miles long, has a maximum width of 5 miles, and encloses an area of approximately 47 square miles. Deep
Figure 1.—Map of the Deep Springs Valley and vicinity, California, showing the major topographic and drainage features. The Owens Valley occupies the southwest corner of the map.
Springs Lake lies in the southeast corner of the valley. The saltpan and surrounding playa cover an area of about 5 square miles. The alluvial floor of the valley has an average gradient of about 50 feet per mile from the north end to the playa, where the average altitude is 4,920 feet; 4,917 feet is the lowest measured points in the lowest measured point. Altitudes of several points in the Deep Springs Lake area are given in figure 2.
The playa which constitutes Deep Springs Lake may be readily subdivided on the basis of distinctive surficial features (fig. 3). The east-central one-third of the playa is lowest and is marked by a thick porous varicolored salt crust broken into irregular polygonal units up to 100 feet across. Surrounding the large polygonal units is a narrow zone of thinner but coherent salt crust; this zone widens to the south and gradually passes westward into an area of mud flats, which widen markedly to the north. The western third of the playa area is 2 to 5 feet higher than the saltpan, is conspicuously marked by very shallow alluvial channels, and is normally covered by a thin efflorescent salt crust.
The most distinctive feature of the playa is a low levee up to 2 feet high enclosing the northeastern third of the area. This levee was built of material dredged from an adjacent trough and was part of an attempt in the early 1920’s to create a solar evaporation pond for the commercial production of potash salts (Palmer, 1922). The levee and trough seem to have little direct effect on the hydrography of the playa.
Southeast of Deep Springs Lake are several distinctive topographic features associated with recent faulting. This area has been described in detail by Miller (1928, p. 520-523). Here the mountain front stands behind the lake like a great wall, rising more than 2,000 feet in less than a mile. The base of the front is marked by distinct fault scarps for about 2 miles (fig. 4). The alluvial fans along the steepest part of the front are transected by a pronounced fault trough for a distance of about half a mile. This trough is terminated at both ends by sheer fault scarps up to 50 feet in height at the base of the mountain. The trough floor ranges from 100 to 300 feet in width and averages about 25 feet in depth. Lakeward from the main trough are some broader sags and ridges which trend roughly parallel, but somewhat arcuate, to the major fault zone, and decrease westward in mean altitude. The crest of the outermost ridge is slightly higher than the floor of the main trough. These sags and ridges are apparently the result of slumping in alluvial material associated with movement along the major fault.A4
CLOSED-BASIN INVESTIGATIONS
Thenardite
Calcite and (or) aragonite
than specific point
1000 0 1- I I I I I
1000
I
2000
i
3000 FEET
Gaylussite
CONTOUR INTERVAL 80 FEET DOTTED LINES REPRESENT GENERALIZED 20-F00T INTERVAL
Figure 2.—Map of the Deep Springs Lake area including altitudes of key points and sample localities. Dashed lines delineate the general saline mineral zoning in the surficial deposits of the playa.HYDROLOGY AND MINERALOGY OF DEEP SPRINGS LAKE, CALIF.
A5
Figure 3.—Aerial photograph of Deep Springs Lake and surrounding area at dry stage, July 1947. The faint border around the central saltpan is the levee mound. The straight part on the east side is about three-fourths of a mile long. The bog-mound springs are the dark spots north of the playa area. The Corral Springs and their inflow area are immediately east of the central saltpan, and the Buekhom Springs are to the south. Two sag ponds may be seen in the fault though between the two spring areas along the fault zone. Note the shoreline traces around the saltpan and the ancient lake shorelines at the south and west (left) extremes of the photograph. The sand dunes southwest of the playa are also conspicuous.
North of the playa is an elongate broad shallow depression that trends west-northwest and is separated in part from the playa by a low ridge. The floor of this depression slopes gently east and is marked by several circular or oval springs. On the ridge to the north and to the west and south of the playa, dune topography is common. Near the playa the dunal ridges are generally circular or oval, about 3 feet high, and are separated by broad smooth flats. The size and relief of the dunes increase with distance away from the playa to a maximum of about 500 feet long and 15 feet high about half a mile southwest. Most of the dunes are now stabilized by vegetation, but many show signs of recent activity.
Evidence indicates that at least the south half of the Deep Springs Valley was formerly occupied by a large lake. Although, as pointed out by Hubbs and Miller (1948, p. 89), shore features are not readily apparent in the valley, such features can easily be identified on aerial photographs of the region (fig. 3). Shoreline benches can be definitely delineated and examined at only a few places in the field, however. Such features can be distinguished with certainty up to 5,200 feet in altitude west of the playa or nearly 300 feet above the present lake level. However, faulting and a marked lowering of the southern part of the valley floor since the lake stood at the 5,100-foot level are suggested by the coincidence in altitude of
766-977A6
CLOSED-BASIN INVESTIGATIONS
Figure 4.—View looking southeast from the north side of Deep Springs Lake, showing the prominent fault scarps in alluvial
fans east of the playa.
fairly prominent benches on the southwest side of the valley and the top of prominent fault scarps on alluvial fans east of the lake. This hypothesis is further supported by the gradual transgression of strand lines by contours on the broad pediment formed by coalesced alluvial fans southwest of the playa.
As pointed out by Miller (1928, p. 525), the most likely outlet for the lake was through Soldiers Pass into the much lower Eureka Valley. Soldiers Pass at an altitude of about 5,500 feet is the lowest point on the east rim of the valley and consists of a deep canyon and a well-defined channel. Lustig (1962, p. 25) described remnants of a carbonate-cemented pebble conglomerate in the pass and noted that the beds dip to the west, which suggests local tilting. He also noted that C. A. Nelson (oral commun.) had mapped the large fault east of the lake as continuing through the pass and suggested a displacement of at least 200 feet on the basis of the difference in altitude from one side of the mountain mass to the other. Lustig (1962, p. 26) also described stream gravels which indicate uplifts of at least 200 feet in the northeast corner of the valley. Such tectonic movement makes it impossible to ascertain the maximum depth of the former lake. Miller’s (1928, p. 525) estimate of 400 to 500 feet is conjectural, and it is questionable whether or not the lake level was ever sufficiently high to reach the far north end of the valley.
Vegetation in the Deep Springs Valley is sparse. The playa itself supports no vegetation. Saltgrass and some pickleweed grow around the edge of the playa. Many species of grasses grow in the depression area north of the lake and around the spring channels east of the lake. Marsh grasses, reeds, and cattails grow around the ponds and spring orifices. The vegetation on the remaining valley floor is composed of shrubs
dominated by rabbitbush (Chrysothamnus sp.), saltbush (Atriplex canescens?), greasewood (Sarcobatus vermiculatus), and hopsage. Sagebrush becomes common only far outside and higher than the playa area. At altitudes above 7,000 feet the pinyon pine and, to a lesser extent, juniper are common, but on the valley floor the only trees outside the Deep Springs College and the State highway maintenance station are a small clump of cottonwoods at the north end of the spring zone east of the saltpan. A study of the distribution of xerophytes, phreatophytes, and plants indicative of saline conditions could provide additional information on the hydrology of Deep Springs Valley, but such a study was not undertaken here.
BEDROCK GEOLOGY OF THE DEEP SPRINGS VALLEY
The bedrock exposed within the drainage area of the Deep Springs Valley is divided about equally between intrusive rocks related to the Sierra Nevada bath-olith and a thick sequence of Cambrian and (or) Pre-cambrian sedimentary strata (Knopf and Kirk, 1918; Miller, 1928; Nelson, 1962). A metamorphic aureole is conspicuous along the contact between the igneous and sedimentary rocks. The contact trends north-northwest and roughly bisects the valley (pi. 1), as shown on the geologic map of the area by Nelson (1963) (table 1).
The oldest rocks in the area are thought to be Pre-cambrian. Kirk (in Knopf and Kirk, 1918, p. 9) attributed to the Precambrian three lithologic units with an aggregate thickness of 3,600 feet as follows: “a series of sandstones and thin-bedded impure dolomites at the bottom, the Reed Dolomite above these, and locally, the Deep Springs Formation at the top.” Nelson (1962) assigned Precambrian (?) age to the lowermost unit, the Wyman Formation, and indicated thatHYDROLOGY AND MINERALOGY OF DEEP SPRINGS LAKE, CALIF.
A7
Table 1.—Generalized column of sedimentary bedrock in the Deep Springs Valley drainage area
[After Nelson (1962,1963)]
Age Stratigraphic unit and thickness (feet) Lithology
Middle Cambrian Gray shale, mudstone.
Lower Cambrian Mule Spring Limestone 1000 Massive blue-gray algal limestone.
Harkness Formation 2000 Gray-green shale, platy fine-grained sandstone, and siltstone.
Saline Valley Formation 850 Brown siltstone and quartzitic sandstone; locally metamorphosed to siliceous hornfels.
Poleta Formation 1200 Upper Gray-green shale, blue-gray limestone beds and quartzite.
Lower Massive gray-blue limestone, some dolomite.
Campito Formation 3500 Montenegro Member Gray shale interbedded with siltstone and sandstone.
Precambrian (?) Andrews Mountain Member Massive black quartzitic sandstone interbedded with gray siltstone and shale.
Deep Springs Formation 1500 Black quartzitic sandstone, dolomite, and limestone.
Reed Dolomite 2000 Massive gray to buff dolomite.
Wyman Formation 9000+ Thin bedded brown to dark-gray argillite, sandstone, siltstone, and interbedded lenticular gray-blue oolitic limestone or dolomite.
the overlying Reed Dolomite and Deep Springs Formation rest on regional unconformity and may possibly be Cambrian in age. These rocks have undergone some metamorphism and are commonly schistose. They underlie much of the upper parts of the major drainage systems which enter the Deep Springs Valley from the northwest.
Overlying the Deep Springs Formation is the Cam-pito Formation of Precambrian (?) and Early Cambrian age, about 3,500 feet thick. Lower Cambrian rocks conformably overlying the Campito Formation include the Poleta, Harkless, Saline Valley, and Mule Spring Formations, in ascending order (Nelson, 1962, p. 141-143). These formations have somewhat similar lithologies consisting of limestones (commonly dolomi-tic), quartzitic sandstones, and gray shales in varying proportions. Except where extensively metamorphosed by granitic intrusives, these rocks lie outside the drainage of any perennial streams and affect only ground-water and ephemeral surface flow in the southwestern part of the area.
Some Middle Cambrian rocks which immediately overlie the Mule Spring Limestone (Nelson, 1963) are included in the hornfels exposed near intrusive contact in the conspicuous canyon southeast of Deep Springs Lake.
The igneous rocks of the Deep Springs Valley have been described in general terms by Miller (1928, p. 512-514). He divided the intrusive rocks into two plu-tons. One is an older “quartz diorite varying to gab-
bro-diorite” and occurring “nowhere in large bodies,” but “in the form of very irregular inclusions ranging in size from less than an inch to half a mile long.” The other and much larger pluton, according to Miller, ranges in composition from granite to monzonite.
In the White Mountains north of the Deep Springs Valley, Anderson (1937) has delineated two principal units of plutonic rocks which he has called the Boundary Peak and Pellisier Granites. According to Anderson (1937, p. 8-9) the Boundary Peak Granite is a fairly uniform intrusive mass, whereas the Pellisier is decidedly variable in texture and composition. Anderson suggested that the Pellisier Granite was formed by the alteration and recrystallization of sediments under the influence of subjacent magma and is actually part of the Boundary Peak Granite contact aureole. In general, as Anderson described the two units they are quite similar except that the Pellisier Granite is foliated, somewhat more mafic, usually contains hornblende, and has been more strongly affected by secondary albitization. Although Anderson’s map shows only the Boundary Peak Granite extending southward toward the Deep Springs Valley, rocks which have petrographic affinities to both of Anderson’s units were recognized in the area. Anderson (1937, pi. 11) also mapped a body of older quartz diorite in the southeastern White Mountains; inclusions of such composition were also noted in Deep Springs Valley.
Emerson (1959) maintained that the Pellisier Granite as mapped by Anderson actually includes two dis-A8
CLOSED-BASIN INVESTIGATIONS
tinctive granodiorite plutons plus metamorphic rocks and that all the granitic bodies are in sharp intrusive contact.
The mountains immediately east and a little north of the lake area are composed of a monzonite mentioned by Knopf (1918, p. 60). Contrary to Miller’s (1928, p. 513) contention, the monzonite mass appears to be a separate pluton, rather than a more basic facies of the granitic rocks already described.
The plutonic rocks of the Deep Springs Valley are generally similar to rock units of the east-central Sierra Nevada batholith described by Bateman (1961). They underlie not only the lower segments of the perennial Wyman-Crooked Creeks system to the northwest, but also nearly all the ephemeral drainage area east of the valley. In addition to the plutonic rocks, basaltic lava occurs in patches at the north end of the Deep Springs Valley.
Contact metamorphic rocks are prevalent in three of the major drainageways leading into Deep Springs Valley. The South Fork and Lower Birch Creek Canyons are cut largely in such rocks. Contact metamorphic rocks are also exposed for more than a mile in Wyman Canyon, and all along the largest canyon immediately southeast of Deep Springs Lake (pi. 1). The rocks cover a wide range of calc-silicate and quartzofeldspathic hornfels and represent the contact metamorphism of nearly all the sedimentary rocks types exposed in the valley area.
Several ore zones, most of which are of minor proportions, are associated with areas of pronounced contact metamorphism, particularly in the drainage of Birch Creek, although ore zones are also present near the sources of Wyman and Crooked Creeks. Some study was made of compositional variations in the plutonic and metamorphic rocks and ore zones adjacent to the major stream courses. The author intended to correlate such variations with data on minor elements in solution in Deep Springs waters. Analytical difficulties have not been fully resolved, however, and further work is necessary.
QUATERNARY GEOLOGY
As with all the intermontane valleys of the Great Basin, the Deep Springs Valley floor is a great mass of alluvium1 derived from the erosion of the surrounding ranges. During drilling operations for the irrigation well at Deep Springs School, highly varied alluvium, including layers of clay and coarse gravel, was penetrated to a depth of more than 600 feet. Alluvial fans are conspicuous on all sides of the valley, espe-
1 As used here, the term “alluvium” denotes all detrital deposits resulting from fluvial agents and includes lacustrine detritus.
cially on the south and west. The largest single fan covers nearly all the north end of the valley. This deposit has been built by the perennial streams from the White Mountains. Delineation of individual fans at the base is difficult as they coalesce and overlap. The material in the fans ranges in size from fine silt to large boulders. As noted by Kesseli and Beatty (1959, p. 11), differences in size distribution of particles on the fans can be related to the bedrock of the catchment area as well as distance from source. Weathering along prominent joint planes in granitic rocks produces an abundance of large blocks which are generally absent form fans built with metamporphic and sedimentary debris. Though there is a very general decrease in particle size toward the center of the valley, the trend is highly irregular. Mudflows appear to be the dominant agents in transportation and deposition at present. Detailed information on the alluvial fan deposits in the Deep Springs Valley is given by Lustig (1962).
The alluvium on the valley floor decreases in particle size southward to the lake in a broad belt up to 2 miles wide between the alluvial fans at the valley margin. Aerial photographs of the region indicate that most of these central valley deposits are related to drainage from the far northwest end of the valley and that these sediments represent deposition by streamflow which reached part way or possibly all the way to the lake. Some of these deposits are well stratified and very fine grained. They may have been reworked by waters of the former lake.
Lacustrine deposits are not confined to the present area of Deep Springs Lake. Their extent, however, is difficult to ascertain, though they have been recognized to a depth of 30 feet and a distance of 5 miles north of the present playa. The lacustrine deposits of the playa may be divided into three facies: (1) the saline crusts, (2) carbonate muds, and (3) marginal silts and sands.
The saline crusts which occupy the central part of the playa are largely lacustrine rather than efflorescent in origin. They are fairly uniform in thickness, except where much efflorescent material is superposed upon them. From a maximum thickness of nearly 2 feet at the center of the salt pan (16 inches was the maximum measured thickness; material which appeared thicker was inaccessible to direct measurement) the crusts thin shoreward to a featheredge. The thicker crusts are composed of four or five distinct layers. Thinner crusts normally contain at least two layers. The average thickness of lacustrine crusts for the entire present area of Deep Springs Lake is about 2 inches.HYDROLOGY AND MINERALOGY OF DEEP SPRINGS LAKE, CALIF.
A9
A substantial part of the playa surface, especially in the south and west, is covered by a saline crust formed by capillary efflorescence. The saline crusts of lacustrine origin may be roughly distinguished from efflorescent crusts by their textures. Efflorescent crusts lack uniform layering and are highly variable in thickness. They form very bumpy dusty irregular surfaces marked by contorted ridges and mounds. The saline material usually includes substantial amounts of the underlying sediment, which is normally more granular than the sediment found underneath the lacustrine crusts. Where saline crusts are both lacustrine and efflorescent in origin, the distinction between the two types, of course, is difficult. The saline crusts are discussed in greater detail in the section entitled, “Mineralogy.”
Throughout most of the playa area, carbonate muds underlie the saline crusts. These muds are most typically very fine grained, smooth and even textured, plastic, and generally gray green. The top one-half inch or so is commonly light brown, apparently due to oxidation. In the central part of the saltpan, immediately beneath the thickest saline crusts, the mud is fluid, is dark green to black, and contains much organic matter. This black layer persists outward from the center of the lake at increasingly greater depths below the normal gray-green material, and finally lenses out. This distribution is illustrated by sections taken in shallow auger holes on both the north and south sides of the playa (fig. 5). Near its outer limits, the layer is faintly laminated in places.
The stratigraphy of the upper 26 feet of lacustrine carbonate mud in the playa area is given in the log of a shallow hole drilled with a jeep-mounted auger under the direction of Ward C. Smith of the U.S. Geological Survey (near loc. DL2, fig. 2; stratigraphic notes and material courtesy W. C. Smith and R. C. Erd). Alternating light and dark laminae mentioned below may be true varves (annular layers). These features are discussed in greater detail in the section entitled, “Mineralogy” (p. A43).
Feet
1.0- 2.0 Wet light green plastic mud.
2.0- 3.5 Laminated black and white mud, green near the
top. Laminae about 1 to 3 mm thick.
3.5- 6.0 Black plastic clay, smell of H2S, grading to unit
below.
6.0- 7.0 Gray-green mud which contains lenses of medium
to coarse sand.
7.0- 10.0 Laminated black and green mud in which green
dominates. Laminae are about 2 mm thick.
10.0-26.0 Crudely banded gray-green to yellow-green mud.
The lacustrine carbonate muds intertongue near the margins of the playa with the lenses of coarser grained
alluvial material, which ranges in particle size from fine silt to sand. A typical section is represented by a shallow auger hole in the southwest part of the playa (loc. 55, fig. 2):
Feet
0.2- 0.7 Moist granular textured silt and clay.
.7- 1.1 Dark brown silt to medium sand. Faintly bedded.
1.1- 1.6 Light brown silty mud.
1.6- 2.3 Dark brown fine silt to fine sand.
2.3- 3.0 Gray-green dense plastic mud.
All contacts are somewhat irregular and gradational.
Lacustrine deposits outside the playa have been extensively reworked by runoff and wind. These deposits are not readily distinguished from finer grained fluvial deposits presently being transported from the mountains, except where associated with ancient strand lines or where characterized by lacustrine faunal remains, chiefly gastropods of the family Planorbidae, probably Gyraulus (K. V. Slack, written commun., 1961).
Lustig (1962, p. 98) obtained a carbon-14 date of 1380±250 years B.P. for a fire hearth in deposits bordering the lower channel from Antelope Canyon (alt 5,080 ft). This site is close to some distinctive linear ridges presumably formed at a much higher stage of Deep Springs Lake. From consideration of the rates of erosion and deposition, Lustig (p. 99) dated this high stage at 700 to 1,000 years B.P.
HYDROLOGY
PRECIPITATION
Precipitation and temperature records have been obtained by the U.S. Weather Bureau at Deep Springs School (alt 5,225 ft) in the northern part of the valley since 1948 and at two research laboratories operated by the University of California on the crest of the White Mountains since 1955. The lower station (White Mountain No. 1, alt 10,150 ft) lies within the Deep Springs Valley drainage at the head of Crooked Creek; inasmuch as most of the inflow to the Deep Springs Valley is ultimately derived from the mountains to the northwest, this station is very advantageously located.
Climatic data for Deep Springs School and for White Mountain No. 1 for the period 1948-60 (table 2) indicate that maximum precipitation generally occurs in the winter or spring. Although no data are available for the specific area of Deep Springs Lake, the annual distribution of precipitation is presumably very similar to that at the school. Valley residents believe that the lake area receives more precipitation than does the school, and inasmuch as the White Mountain No. 1 station is near the highest altitudeA10
CLOSED-BASIN INVESTIGATIONS
give percent water-soluble material to nearest whole number
CONTOUR INTERVAL 80 FEET DOTTED LINES REPRESENT GENERALIZED 20-FOOT INTERVAL
Figure 5.—Map of the Deep Springs Lake area showing the location of sites outside the saltpan where cores were taken or pits dug to establish the stratigraphy of the surficial deposits. The general stratigraphy is shown in the small columns adjacent to each site. Numbers within the columns refer to the percent readily water soluble material within each stratigraphic unit.HYDROLOGY AND MINERALOGY OF DEEP SPRINGS LAKE, CALIF.
All
Table 2.—Climatic data for Deep Springs Valley and vicinity, California [M, missing; E, estimated]
Date Jan. Feb. Mar. Apr. May June July Aug. Sept. Oct. Nov. Dec. Annual
AIR TEMPERATURE (°C) Deep Springs School
1949... 12.4 22.2 37.0 52.5 56.5M 75.0M 70.2 68.0 51.0 47.5 30.9M
1950 26.2 38.9 43.1 51.6 59.5 65.3 71.2M 72.2M 62.3M 57.1M 44.8M
1951 29.3 42.0 54.6 59.8 67.5 76.1M 71.7M 67.8M 52.7M 40.7M 28.7M
1952 23.7 33.3 51.4 62.1M 67.4M 74.2M 65.9M 58. 6M
1953 38.3 37.9 44.2 50.4 52.2 66.4M 77.2 70.8 66.8 52. 0M 42. 7M 32.3 52. 6E
1954 33.9 42.3 41.0 56.3 63.5 70.5 78.5 71.2 65.4M 53.3 44.5 32.2 54.4
1955 21.6 33.3 42. 2M 47.4 57.5 74.2 55. 7M 41.2
1956 36.8M 33.8 44.2 50.4 59.8 69.6 72.4 69.6 66.6 52.5 40.8
1957 25.6 41.6 44. 4M 50. 5M 52.0 68.1 70.2M 70.2 49.6 38.3
1958 40. 4M 38. 7M 48.8M 61. 5M 69.5M 74.4M 74.9M 66.9M 59. 3M 42. 5M 40. 0M
1959 36.8 31.1 45. 5M 54.3M 57.4 71.2 77. 7M 73.4M 63.2 55.2 44.3 36.6 54.0
1960 28.4 36.3 48.6 53.6 60.9 74.7 75.1 73. 7M 68.3M 52.9 40.6 35.6 54.0
Average 28.5 35.6 42.8 51.8 58.6 69.0 74.7 71.8 66.1 54.2 42.5 33.8 53.8
White Mountain No. 1
[Alt 10,160 ft]
1955 38.5 30.0 23.0
1956 20.8M 16.5 23.9 25.3 34.2 46.1 49.3 47.6 46.3 33.7 28.6 24.5 33.1
1957 13.0 24.7 23.4 26.6 31.8 47.6 49.9 49.4 44.0 31. 2M 23.8 24.2 32.5
1958 20.2 20.7 12.7 24.3 38.6 44.0 50.3 52.6 46.4 39.5 27.7 30.3 33.9
1959 23.9 14.8 26.1 35.5 34.7 49.7 55.4 51.1 44.0 39.4 31.8 24.9 35.9
1960 15.6 19.9 29.9 32.2 37.9 49.9 53.0 52.4 47.2 35.4 25.3 19.2 34.8
Average 18.7 19.3 23.2 28.8 35.4 47.5 51.6 50.6 45.6 36.3 27.9 24.4 34.0
White Mountain No. 2
[Alt 12,470 ft]
1955 33.8 24.2
1956.... 14.5 8.8 18.4 18.3 26.5 38.7 43. 7M 42.7 41.9 26.5 26.1 22.0 27.4
1957 9.6 18.3 15.7 20.2 23.7 40.5 44.7 44.5 39.6 25.8 18.7 20.0 26.8
1958 17.2 21.5 14.2 22.3 31.5 33.1 41.3 44.9 37.8 32.0 20.2 22.6 28.2
1959.... 15.1 7.0 16.2 24. 4M 21.3 39.7 46.9 41.8 33.7 29.8 26.3
1960 5.3 7.9 18.4 20.9 27.1 41.6 43.0 44.8 39.5 29.2 19.5 16.3 26.1
Average 12.3 12.7 16.6 21.2 26.0 38.7 43.9 43.7 38.5 29.5 22.5 20.2 27.1
PRECIPITATION (INCHES) Deep Springs School
1949 0.61 0.40 0.15 0.03 1.20 0.13 0.00 Tr. 0.43 0.04
1950.... .04 .25 .65 .00 Tr. 0.00 0.26 . 11 1.46 0.29 . 55
1951.... .04 .39 1.23 .30 .25 .31 .25 .00 .04 .00 1.98
1952 2.07 .00 1. 55 2.58 .00 .30 .91 Tr. .08 .00 .25 . 83E 8. 57E
1953 .08 .15 .29 Tr. .30 .00 .50 Tr. .00 .06 .42 .02 1.82
1964 .64 .97 2.05 1.00 .20 .47 .52 .00 .03 .00 .42 .23 6.53
1955 . 94E .11 Tr. .22 1.36 . 00E Tr. . 28E . 20E .00 .01 1.75E 4.87E
1956 1. 26 .00 .00 3.80 .34 .00 .66 .00 .00 .42 .00 .00 6.48
1957 . 72E .31 .02 .66 1.46 .02 .06 .13 .20 1.42 .28 1. 61E 6.89E
1958 .98 1.66 1. 52 1.33 .52 Tr. .00 .19 .32 .03 .17 .00 6.72
1959. .04 1.61E .00 .04 .17 Tr. .15 .49 .17 Tr. .00 . 36E 3.03E
1960 . 23E . 83E . 07E .13 .00 .00 .05 Tr. .36 .03 3.46 .00 5.16E
Average 0.64 0.57 0.56 0.92 0.49 0.09 0.31 0.14 0.24 0.19 0.50 0.62 5.56
White Mountain No. 1
1955 0.00 0.36 6.03
1956. 2.79 0.16 0.02 4.15 1.79 Tr. 3.30 0.03 0.02 .96 Tr. .07 13.29
1957 2.03 1.25 .86 1.90 2. 74 Tr. .06 .00 1.50 1.54 1.22 1.22 14.32
1958.... 1.61 2. 05 3. 05 2.28 .94 Tr. Tr. 1.09 .30 .38 .84 .07 12.61
1959. .29 3.10 .04 .04 1.07 Tr. 2. 05 .75 2.25 .69 .00 .36 10.64
1960 .44 1.69 .14 .43 .09 Tr. .86 .57 .30 .72 4.29 .34 9.87
Average 1.43 1.65 0.82 1. 76 1.33 Tr. 1.25 0.50 0.87 0.72 1.12 1.35 12.15
White Mountain No. 2
1955 0.08 0.93
1956 2.30 0. 45 0.02 6.65 1. 67 6.18 2.93 0.26 0.00 1.95 .08 0.31 16.80
1957... 1.85 .92 2.41 1.59 4. 66 .26 .10 Tr. .49 3.33 1.56 .83 18.00
1958 1.41 2.30 2.39 1. 29 . 79 .48 32 4 30 .15 15.88
1959 1.07 3.83 .50 .62 3.11 Tr. .65 .40 1.58 2.26 .00
1960 1.37 1.21 .11 1.11 .38 .25 1.08 .75 .09 1.12 3.17 .19 10.83
Average 1.60 1.74 1.09 2.25 2.12 0.23 1.02 1.14 0.90 1.54 1.16 0.37 15.38A12
CLOSED-BASIN INVESTIGATIONS
within the Deep Springs Valley drainage, the two station records probably give a reasonable range for weather conditions in the entire basin.
According to Kesseli and Beatty (1959, p. 18), the White Mountains lie in a transition zone between two areas which have markedly different precipitation regimes. The Owens Valley, which has a typical California coastal regime, has little or no precipitation in summer, whereas the areas east of the main White Mountains, such as the Fish Lake Valley, have a fairly uniform distribution of precipitation throughout the year. Seasonal precipitation distribution within the Deep Springs basin appears to be intermediate between the two regimes (fig. 6).
Storm patterns in the Deep Springs area show marked seasonal variation. Winter precipitation is derived from cyclonic storms moving eastward across the Sierra Nevada where much of their moisture is lost. Thus, though precipitation may be of relatively long duration, it normally is only moderate to light in intensity. Strong northwest winds are common during the winter storms. During the summer, precipitation accompanies thunderstorms which frequently build up and move northward along the crests of the ranges on either side of the Deep Springs Valley. The summer thunderstorm period is often marked by strong southeasterly winds. Kesseli and Beatty (1959, p. 18) associated increased thunderstorm activity east
Figure 6.—Seasonal distribution of precipitation for the Deep Springs Valley.
of the White Mountains with a flow of warm moist air from the Gulf of Mexico, and they suggested that summer thunderstorms in this area are likely to produce at least as much precipitation as weakened winter cyclonic storms. As discussed in some detail by Kesseli and Beatty (1959, p. 22-23), thunderstorm precipitation is sporadic, both in time and place, and reliable accounts of intensities and total amounts are commonly not available in the established weather station records. However, the data for Deep Springs School and White Mountain No. 1 indicate that maximum precipitation in the Deep Springs drainage area usually occurs in April (fig. 6).
EVAPORATION
The Deep Springs Valley—because of size, higher altitude associated with lower mean annual temperature, and complete enclosure by steep mountain masses— does not have as high an evaporation rate as many nearby intermontane basins. Extrapolation from the evaporation maps prepared by the U.S. Weather Bureau (Kohler, Nordenson, and Baker, 1959) indicates that the average class-A-pan evaporation is about 65 inches per year, and lake evaporation is about 45 inches per year. Similarly, from the relation derived by Langbein (1961, p. 3), lake evaporation would be about 48 inches per year at a mean annual temperature of 54°F. (See table 2a.) At Deep Springs Lake, shallow relatively dilute surface waters probably evaporate at nearly the maximum pan rate, but the rate of inter-crustal and subsurface brine evaporation is probably close to or even lower than the rate for lakes where surface area and storage of radiation energy at depth are important factors. Because of the additional effect of the lowering of vapor pressure by high solute content, the evaporation rate for the perennial inter-crustal brines at Deep Springs Lake may be as low as 30 inches per year. An estimate of the net annual evaporation from lake waters indicates a rainfall equivalent to, or slightly higher than, that recorded at Deep Springs School, about 5 to 7 inches per year. (See table 2b.) On this basis the net evaporation rate for surface waters would range from 40 to 60 inches per year depending on salinity and depth. A reasonable estimate of the mean is about 4.25 feet per year.
HYDROGRAPHY OF THE DEEP SPRINGS VALLEY
Perennial surface waters are a major factor in the hydrology of Deep Springs Valley. In the mountains to the northwest of the valley, one stream is perennial and segments of three streams flow during most of the year. Most of the time, the total streamflow in the mountainous part of Deep Springs Valley drainageHYDROLOGY AND MINERALOGY
area is equaled or exceeded by discharge from springs near the mountain fronts and the lake. Several fairly large ephemeral drainage systems, which are normally dry for extensive periods, may carry considerable quantities of flow during times of storm runoff.
STREAMS AND DRAINAGE OF THE MOUNTAIN AREAS
The largest stream entering the Deep Springs Valley, Wyman Creek, has a drainage area of 29.7 square miles exclusive of its major tributary, Crooked Creek. Wyman Creek rises in a series of wet meadows called Roberts Ranch, about 1% miles above the contact zone between granitic rocks and Precambrian metasediments, and maintains perennial flow over a total length of about 7 miles. Throughout most of the distance of perennial flow, the Wyman Creek channel is narrow but well defined, and except in a few reaches closely bounded by bedrock, the channel is bordered and overgrown with brush. Two or three miles upstream from the head of its fan, the Wyman Creek channel is carved directly in granitic bedrock. Wyman Canyon 'is very narrow and steep sided through most of the distance underlain by igneous rocks. In contrast, above Roberts Ranch where Wyman Creek is largely ephemeral and underlain by sedimentary rocks, the valley is broader and is floored with much alluvium. Although distinct channels are present, no flow was observed in the Wyman Creek drainage above Roberts Ranch at any time except from Cedar Spring, a tributary channel about 3 miles above Roberts Ranch. Flow from this spring is light; less than 0.01 cfs (cubic feet per second) disappears into alluvium a very short distance from the source.
In the short reach of Wyman Canyon between Roberts Ranch and the first outcrops of intrusive rock, the canyon gradually narrows and the valley of Wyman Creek contains less fill as the metamorphic grade of underlying sedimentary sequences increases. The effect of decreased permeability in surrounding and underlying .bedrock is suggested by comparing flow estimates made on Wyman Creek just below the mouth of Mill Canyon and estimates made just above the confluence with Crooked Creek (table 3). Nearly two-thirds of the total flow of Wyman Creek was measured less than 1 mile from its source.
The main drainageway of Wyman Creek extends from the White Mountain ridge for a total linear distance of about 12 miles and enters the northwest corner of the Deep Springs Valley at the head of an extensive, low, and complex alluvial fan. The main channel curves very sharply at this point and for about 2 miles follows the outer limits of a small series of secondary fans before- reaching the central floor of the valley.
OF DEEP SPRINGS LAKE, CALIF. A13
Table 3.—Estimated discharge, in cubic feet per second, of surface water units in Deep Springs Valley, 1959-61
[Numbers in parentheses refer to sample localities shown in fig. 2; n.m., not measured]
Source Oct. 1959 June 1960 Sept. 1960 May 1961 Aug. 1961
Wyman Creek: 2.87
Above confluence with Crooked Creek (47) 3.71 4.12 1.92
Below confluence with Crooked Creek (56) 4.32 .62 2.05 No flow.
.65 .52 No flow.
Total, Wyman Creek plus
4.36 3.70 2.05 • 4.64 1.92
Corral Springs:
1.32 .57 n.m. n.m. .69 .31 1.44 .62 1.00 .30
Total
1.89 1.00 2.06 1.30
Outflow from the open sag pond (32) Buckhorn Springs: Nos. 1-5 (33)
.33 n.m. .60 1.67 .54
.38 .40 .21 .74 n.m. n.m. n.m. .44 .30 .21 .13 1.06 .43 .69 .32 1.59 .16 .49 .15 1.04
Nos. 6-7 (34)
Nos. 8-10(35)
Group 11 (36)
Total
1.73 1.70 3.03 1.84
Note.—Bog-mound springs discharge is very difficult to estimate but apparently near constant annually; northwest bog mound, about 0.01-0.02 cfs; western bog mound, at least 0.05 cfs; central bog mound, at least 0.20 cfs; and old artesian well (eastern bog mound) spring, about 0.03 cfs.
The valley channel is normally dry, as the waters of the creek are diverted into a canal which leads to the Deep Springs Junior College. This canal is headed by a sediment trap, which is periodically flushed; at such times water does occupy the former channel. However, through most of the year all the water of Wyman Creek is utilized by the school. Discharge from the school is released into a series of shallow channels directly south of the school lands, and what is not consumed or lost by evaporation eventually reenters Wyman Creek drainage. Inasmuch as creek water is stored in an open reservoir behind the school, evaporation losses may be high.
By far the largest tributary to Wyman Creek, Crooked Creek, enters the main stream about a mile above the sediment trap. Crooked Creek, about 14 miles long, drains 21.8 square miles. Although not perennial, Crooked Creek through most of the year maintains some flow from Dead Horse Meadow to its confluence with Wyman Creek, a distance of about 4 miles; the valley is especially narrow and the channel is carved in bedrock through most of this reach. This segment represents a former small tributary of Wyman Creek which by headward erosion and capture has diverted upper Crooked Creek drainage from the Fish Lake Valley. The valley of Crooked Creek above Dead Horse Meadow contains considerable fill, but the channel is commonly cut through almost to bedrock. Much of the Crooked Creek drainage area is underlain by granitic rocks and basalt. Discharge estimates
706-977 0—65-----3A14
CLOSED-BASEST INVESTIGATIONS
(table 3) indicate that Crooked Creek contributes up to 20 percent of the total flow of the Wyman-Crooked Creeks system.
Other major drainage systems tributary to the Deep Springs Valley on the west include Birch Creek, Antelope Creek, Beer Creek, and Payson Canyon. Birch Creek enters the valley about 6 miles southwest of Wyman Creek and about 4 miles due north of the lake. Though Birch Creek drains an area of 16.0 square miles, flow is perennial only in a short reach of bedrock channel which extends about iy2 miles from a meadow at the confluence of the north and south forks to the head of the relatively wide alluviated valley immediately north of its alluvial fan. The Birch Creek Canyon below the confluence of the north and south forks is cut through granitic and contact metamorphic rocks—chiefly calc-silicate hornfels, marbles, and schists. The actual source of Birch Creek surface water is a group of 16 springs in the wet meadow area at the confluence of the north and south forks. Discharge is relatively stable at the source, but streamflow decreases by more than two-thirds and fluctuates diurnally downstream. Maximum measured discharge is 0.5 cfs, but normal flow is about one-tenth as much. The active channel of the stream is tightly cemented by precipitated calcite throughout its length (Barnes, 1962).
A similar situation to the one at Birch Creek is found about 3 miles farther south. A little more than a mile above Antelope Spring, a very small surface stream originates from the base of alluvium at the head of a narrow gorge and flows with gradually decreasing discharge about y3 to % mile before disappearing. The exposed bedrock in the gorge consists of calcareous and argillaceous sediments which are not intensively metamorphosed. In contrast to Birch Creek, bed cementation with calcite is not as apparent in Antelope Canyon, which drains a total of 7.0 square miles and, in addition to its creek, includes a small spring near its mouth, where rapid downcutting along the main channel has cut under the alluvium filling a tributary valley.
The only surface waters outside the major drainage lines on the west side of the valley are south of Antelope Spring, where a small group of springs discharge from cemented fanglomerate about one-half mile south of the mouth of Antelope Canyon. These springs give rise to small vegetated channels which flow a maximum distance of about 100 yards before disappearing into valley alluvium. Though ostensibly derived from a small drainage upslope, much of the discharge seems to be derived from other systems by downvalley movement through local jointing along the mountain front;
this hypothesis is also supported by the water chemistry (p. A33).
Payson Canyon contains the largest fully ephemeral channel tributary to the Deep Springs Valley. Field observation and aerial photographs indicate that large quantities of surface flow may be carried from this drainage during periods of heavy runoff.
On the east side of the valley, the most conspicuous drainageway is a deep narrow canyon which lies very close to the contact of plutonic and sedimentary rocks southeast of the lake (pi. 1). The channel at the floor of the canyon is carved in bedrock for the most part, but a few small clumps of vegetation give some evidence of seepage.
There are many other ephemeral stream systems tributary to the Deep Springs Valley. All such drainage from the mountains discharges onto alluvial fans at the edge of the valley. Most of the fan surfaces are cut by several well-defined channels. The most marked channels, those of the Payson and Antelope Canyons systems, are on the largest fans of the southern valley area. Most channels reach a maximum depth where they cross remnant shorelines of the ancient lake which filled the valley. These systems rarely carry surface flow and are a result of intensive storms in their catchment areas. During the period of study, 1959-61, flash flooding was observed only in the Antelope Canyon system and in one of the drainage systems south of the playa; in each case flooding followed heavy thunderstorms. Only the flood from Antelope Canyon reached the lake. However, William A. Jenkins of the California Division of Highways has given the author accounts of flash flooding of major proportions in Payson Canyon in 1953 and 1958 and in Birch Creek Canyon in 1952. The Birch Creek flood carried a massive amount of material, and its effects are still evident. A road which formerly ran up the main stem and North Fork Birch Creek to Roberts Ridge was completely destroyed at this time. L. K. Lustig (oral com-mun.) reported minor flooding in Wyman Canyon in late August 1960; sediment-laden water moved down the old Wyman Creek channels to the center of the valley. Kesseli and Beatty (1959, p. 68-69) reported a flood of Antelope Canyon in August 1957.
In their study of the entire White Mountain range, Kesseli and Beatty (1959, p. 89) suggested that in the whole area there may be as many as 15 floods per decade of which 5 may be of major proportions. Flood frequency within individual drainages is more difficult to estimate, but major flooding has probably not occurred in many of the canyons in over 50 years, though the larger drainage systems probably undergo minor flooding in part every few years.HYDROLOGY AND MINERALOGY OF DEEP SPRINGS LAKE, CALIF.
A15
Most of the surface inflow to Deep Springs Lake is directly derived from springs near the lake area. These springs may be readily subdivided into groups based on location, setting, genesis, and chemical composition.
SPRINGS NORTH OF DEEP SPRINGS LAKE
The springs which occupy the broad shallow depression north of the playa have been given the field name bog-mound springs because they characteristically give rise to marshy muskeglike mounds which are roughly circular and which rise above the immediate surrounding area. Smaller mounds are occupied by marsh grass and reeds, commonly growing in clumps, and by standing water in the intervening spaces; no flow is apparent. The larger mounds have flowing springs. The large springs discharge from about 5 to 30 gpm (gallons per minute) into the surrounding reed-filled marsh which may contain discontinuous areas of open water. Flow can readily be detected as much as 150 feet from the spring orifice. Surface water does not usually extend farther than about 100 yards beyond the mound before disappearing. Much of this water, however, seems to reach the lake or nearby springs through a series of seepage lines trending southeast and marked by green vegetation. Material from holes put down in these areas has a much higher moisture content and shows more soil development than surrounding sediments.
The situation is somewhat analogous to the conditions described by Bunting (1961). Some seepage apparently escapes evapotranspiration and proceeds under the surface to enter Deep Springs Lake through a breech in the levee mound on the northeast side of the playa. In most of the bog mounds, the materials in which the springs arise are largely fine silts and clays admixed with a great deal of organic matter which is held together only through the binding action of roots. However, in the northwesternmost part of the bog-mount area, one of the springs arises in material which is largely of sand size, and the result is a treacherous quicksand pocket.
The vegetated mat around many bog mounds is spongy; trampling of this mat produces gas bubbles in abundance. Much of the gas is trapped air or carbon dioxide. However, around the orifice and immediately downstream from actual flowing springs, there is abundant white fungus similar to that found around hot springs. Traces of hydrogen sulfide may also be detected here.
The largest of the bog-mound springs is about a quarter of a mile due north of Deep Springs Lake. The mound and marsh area is about 200 feet long
Figure 7.—View looking southeast across the western bog-mound marsh.
and 75 to 90 feet wide. The mound and the springs are about 4 feet higher than the surrounding area.
Another large spring is about one-half mile west (fig. 7). Though the marsh area is large, the source of flow is difficult to detect, the mound is poorly formed, and there is much open water in a pond as much as 2 feet deep.
SPRINGS EAST OF DEEP SPRINGS LAKE
The major discharge of surface flow to Deep Springs Lake is derived from springs along the prominent fault zone to the south and east of the playa. These springs are separated into two groups by the fault trough and alluvial ridges at the steepest part of the mountain front. The northern group, referred to as the Corral Springs, lies approximately 11/2 miles north of the southern group, known as the Buckhorn Springs. The actual number of springs in the Corral Springs group is variable. The springs consist of rapid flows of clear water which emerge from beneath outcrops of cemented fanglomerate directly on the fault scarp and a series of marshy seeps which coalesce not far from their point of origin at the base of the scarp. These springs all appear fresh and clear, except for the southernmost of the group which is brackish, has an odor of hydrogen sulfide, and supports the growth of white fungus. About 100 feet from the point of origin, all outflow from the Corral Springs coalesces into a channel, as much as about 3 feet wide, which is partly filled with marsh vegetation. Although this channel leads directly westward toward the playa, much of the flow is diverted into a canal trending northward. Leakage from this canal toward the playa occurs at several places, and the canal is totally dry after carrying some flow a maximum distance of 500 feet. Most of the seepage from the canal rejoins flowA16
CLOSED-BASIN INVESTIGATIONS
Figure 8.—Buckhorn Springs 5 and 7, selected as representing extremes of channel types at the source. The broad vegetation-choked spring area is No. 5 (upper panel), whereas the trenched rocky channel is No. 7 (lower panel).
from the natural channel in a gently sloping grassy flat with little detectable flow. Eventually, the water is impounded east of the saltpan by the levee mound. A large pond forms here in the wet season, but nor-
mally the area is simply a damp mudflat. Apparently, the water eventually reaches the saltpan by seepage beneath the levee mound.
The Buckhorn Springs consists of 11 points of emergence where flow arises beneath the outermost coarse alluvial material; the line of springs parallels the approximate extension of the fault zone, which becomes less well defined southward. All the springs originate from depressions at the foot of an escarpment which results from faulting and may coincide with an old beach level. From south to north, the first five springs arise in rocky depressions and flow through channels choked with marsh vegetation for as much as 150 feet from the points of origin (fig. 8). The outflow channels are shallow, and many have little detectable flow which carries beyond the source. These channels coalesce in a large marshy area near the southeastern part of the playa and supply most of the surface water which accumulates on the lake muds south of the levee mound. Flow from springs 6 and 7 (from south to north) arises from sharp depressions as much as 4 feet deep at the foot of the escarpment and continues several feet through narrow channels (fig. 8). Springs 8 and 9 are small seeps which originate in wet grassy areas beneath coarse alluvial material. Springs 10 and 11 arise in the depression at the foot of a vertical fault scarp as much as 50 feet high and marked by outcrop of cemented gravel fanglomerate (fig. 9). Spring 11 results from the coalescence of flow from 17 seeps in the broad depression at the base of the scarp and provides the largest discharge in the Buck-horn group. The channel leading from spring 11 is narrow in its upper part and is as much as 5 feet deep. Fineness of bed material, low altitude, and the proximity of the steep scarp suggest that these springs originate directly from the fault zone.
Outflow from springs 6-11 discharges down a series of vegetated channels and coalesces in a large marsh
Figure 9.—Cemented fanglomerate on fault scarp above Buckhorn Springs 10 and 11. Buckhorn Spring 11 arises immediately below the center foreground.HYDROLOGY AND MINERALOGY OF DEEP SPRINGS LAKE, CALIF.
A17
which extends lakeward to the south end of the levee. Some of this flow is diverted from the main channel into the depressions between ridges in the alluvial mass which separates the Buckhorn from the Corral Springs. During the period of this study the south-westernmost depression contained a large pond which was not present when the aerial photographs were taken in 1947. Flow diverted into this pond returns to the marshy area south of the spring by seepage through the impounding ridge.
Attempts have been made to channel surface waters from several of the Buckhorn Springs into small canals lying transverse to the outflow channels. The purpose was to spread water over the grasslands lakeward from the springs. Most of these attempts have not greatly interfered with the general flow systems or patterns. All surface water from the springs which does not evaporate eventually appears to reach the central playa area through marshes and breeches in the southern part of the levee.
The temperature (about 20° C) of the waters from the Corral and Buckhorn Springs is consistently 5 to 10° C higher than that measured in the bog-mound springs or any of the mountain streams. This higher temperature suggests that the fault-zone springs issue water which has come from considerable depth.
SAG PONDS
In the fault trough between the Corral and Buckhorn Springs are two perennial sag ponds. The northern pond is about 500 feet long and 50 feet wide at its widest point. It is nearly surrounded by alluvial ridges as high as 50 feet, except at the northwest corner, where the confining ridge is only about 5 feet above the average pond level. At highest stage, outflow spills over this ridge, but normally the pond is closed. The pond is fed by seepage around the margin, which comes to the surface in small irregular channels on the north and east sides. A few reeds apparently supported by the seepage grow at the south end and along the east border (fig. 10). At all stages the waters of the pond are amber in color, and contain abundant algal material. In the spring high stage waters teem with brine shrimp. At lower stages, in late summer, the pond contains a salt crust as much as 2 inches thick which is partly submerged in the brine and irregularly eroded at the edges by seepage inflow from the margins. Efflorescent crust is abundant in the flat areas at the south and northeast ends of the pond at almost all stages.
A second sag pond occurs about 2,000 feet southwest of the northern pond. This one is as much as 750 feet long and about 75 feet wide at its widest point. It
Figure 10.—The closed sag pond at low stage; shown are the marginal seepage inflow to the pond and the growth of reeds in the seepage areas. The photograph is taken from the north end looking at the eastern shore. The white patches are efflorescent crusts.
is also surrounded by high ridges except on the northwest side. This pond, though also fed by marginal seepage, has a perennial outflow at the northwest and maintains fresh or slightly brackish water throughout the year. Marginal vegetation and aquatic life are abundant and green algal mats are often conspicuous. Overflow is through a relatively deep channel to the northwest. Some of this flow is diverted into a canal in an attempt to spread water over the grassland. A large breech in this old canal, however, allows much of the water to escape into a large interridge depression and locally gives rise to standing water. Ephemeral bodies of water also accumulate in other interridge areas lakeward from the sag ponds, especially during wet seasons. Marsh vegetation is common in those depressions which receive water most often. Water loss apparently takes place from all these depressions by seepage through the ridges toward the lake as well as by evaporation.
GROUND WATER
Information on ground water in the Deep Springs Valley is limited to what could be obtained from five wells in the valley area and from observations at the springs. The only fully utilized wells are in the north end of the valley at Deep Springs College and at the highway maintenance station. Although the college obtains most of its water supply from Wyman Creek, a large irrigation well has been installed on the north side of the school. This well was drilled more than 600 feet; unconsolidated alluvium was penetrated throughout the section. Depth to the water table averages about 185 feet. According to personnel at the college, extensive pumping has resulted in little draw-A18
CLOSED-BASIN INVESTIGATIONS
down in a short period. The well at the highway maintenance station has been drilled to approximately 280 feet, the depth to the water table being about 250 feet. This well supplies all water for operations of the maintenance station, but is pumped only for long enough periods to fill the reservoir tank. The water composition is very close to that of the school irrigation well and total solids average about 300 ppm (parts per million). There are two stock wells in the valley, one very close to the center of the valley and the other about one-half mile due west of the lake area. The well at the center of the valley is pumped infrequently and only in sufficient quantity to fill the stock tanks; depth to the water table is 45 feet. The well to the west of the lake is windmill operated and has a depth to water of 16 feet. There is very little increase in total dissolved solids in well water down-valley. The fifth valley well is a remnant of the old Inyo Chemical Co. workings, and was presumably used as a water supply for the pilot plant. This well was put down in the area of one of the bog-mound springs, and, even though the casing apparently is heavily corroded, it maintains at least 3 feet of artesian head.
Inasmuch as there is little subsurface data, it is very difficult to ascertain the extent of the ground-water reservoir in the Deep Springs Valley. The altitude and occurrence of springs, especially on the north side of Deep Springs Lake, suggest that the valley alluvium and thus the occurrence of ground water is subject to considerable variation both vertically and laterally. However, the available well data indicate that the piezometric surface slopes roughly parallel to the surface of the valley floor at a somewhat lower gradient from the mouth of Wyman Canyon to some-
where north of the playa (fig. 11). In the area north of the playa the ground water in the valley alluvium appears under artesian head, aided by gas pressure, as the bog-mound springs. The rough correspondence in maximum altitude of the mounds with the head measured in the old artesian well suggests that the mounds were built with material carried by upwelling ground waters.
Information on the areal pattern of water movement in the Deep Springs Valley can be obtained from field observations and inspection of aerial photographs (fig. 12). A complex pattern of distributary channels lead from Wyman Creek toward the lake and extend over a wide area in the central part of the valley. The most recent of these channels supports a substantial amount of vegetation as compared to the surrounding area. Prior to diversion, these channels carried the flow of Wyman Creek by the school. Their number and size suggest generally higher discharges than those now observed in the Wyman Creek system. The most recent channels are on the east side of the valley floor and two-thirds of the way downvalley they come very close to the mountain front. Although these channels are normally dry, they may carry some of the heaviest runoff from the Wyman Creek system. The multichannel patterns probably indicate the position of underlying gravel deposits related to former channels and mark the route of ground water contained therein. Channels related to the Wyman Creek system appear to cut off or deflect the lower parts of other drainages from the western mountain front farther south, such as the drainages from Birch Creek and Antelope Canyons. The Wyman Creek channels become less distinct near the lake and are more poorly defined below each old lake level. Most of the later channels may be traced
Area of recharge from the Wyman-Crooked Creeks system
Highway
station
HORIZONTAL SCALE
Figure 11.—North-south section of the alluvium in the Deep Springs Valley showing the estimated position of the piezometric
surface, 1959-61.HYDROLOGY AND MINERALOGY OF DEEP SPRINGS LAKE, CALIF.
A19
Figure 12.—High-altitude aerial photograph of the Deep Springs Valley north of Deep Springs Lake illustrating the Wyman Creek channel patterns and the mountain front fault zone on the east side of the valley. The dark patchwork at the north end of the photograph is the fields of the Deep Springs Junior College.
iffitfe.A20
CLOSED-BASIN INVESTIGATIONS
toward the fault zone north of the Corral Springs area. Older channels to the west may be traced directly into the bog-mound springs, but the foregoing observations suggest that recharge from the Wyman-Crooked Creeks system not only feeds some of the bog-mound springs but, to a lesser extent, the fault-zone springs as well. The slope of the piezometric surface in the valley (fig. 11) and the orientation of the major fault zone on the east side of the valley (pi. 1) strongly suggests that the fault-zone springs tap flow at depth under head in excess of the 30-foot differential between the highest spring altitudes and the saltpan. The discharge of the Corral Springs and especially the Buckhorn Springs may also consist wholly or in part of waters from drainage areas to the south and east from which waters are channeled through the major fault zone to the springs. Ground-water discharge in the area of Deep Springs Lake itself must be very slow because of the impermeability of the dense plastic carbonate muds which constitute a major part of the central lacustrine deposits.
The western bog-mound springs are probably recharged principally from the Birch and Antelope Creek systems. Small channels around the northwest-ernmost bog-mounds spring may actually be traced directly back into the Antelope Creek fan. Flash floods from Antelope Canyon enter the playa at the northwest corner. Similarly, the channels of the Payson Canyon system extend into this area. Inflow to the southwest corner of the playa is made exclusively by flash floods from the mountains to the waters and may be readily discerned from the very shallow alluvial channels in the southwest part of the playa.
The pattern of water movement to Deep Springs Lake is summarized in plate 1.
COMPARISON OP DISCHARGE FROM STREAMS AND SPRINGS
An effort was made to obtain some comparative data on the quantities of surface flow in the various water units of the Deep Springs Valley. Initial measurements of streams and spring outflow were made with a pygmy current meter. Such attempts were complicated by an abundance of vegetation, coarse alluvial material, and very shallow depths in many channels. Often, artificial improvements had to be made in the measuring section. The difficulties encountered rendered the resulting data rather poor. Subsequent to the initial measurements, estimates of discharge were made utilizing a folding pocket rule to obtain section parameters and velocities by measurement of differences in head on the face and edge. Such methods give semi-quantitative results at best, but they provide some
means of comparing flow in different water units. Data were obtained during each period of fieldwork in the course of the study and are presented in table 3. In addition to measurements on the major streams and springs, an estimate was obtained just northwest of the bog-mound-spring area on a minor flash flood from the lower reaches of Antelope Canyon. The measured discharge was half a cubic foot per second with a maximum velocity of 0.1 foot per second and was composed of 67 percent by weight of solid material.
The data presented in table 3 cannot be taken as truly representative of the range of discharges under so-called normal conditions, but the autumn measurements collected over the 3-year period 1959-61 might be considered as being close to minimum mean values for the total period. Though the valley flow patterns seem to substantiate the observation that water levels of the lake and associated springs respond principally to runoff from the White Mountains through Wyman Creek and its tributaries, the flow estimates lend support to the hypothesis that significant inflow to the lake area from the fault zone springs is derived from sources other than surface waters of the Wyman-Crooked Creeks system. The data of table 3 indicate that during each period of measurement, outflow from the Corral and Buckhorn Springs nearly equaled or exceeded the maximum possible surface water input from the Wyman-C rooked Creeks system. Furthermore, the substantial late spring increase (May, 1961) following a wet year suggests considerable contribution from sources in addition to the valley ground-water reservoir.
HYDROGRAPHY OF THE LAKE
Attempts were also made to obtain data on the hydrologic variations of Deep Springs Lake itself. Special attention was paid to variations in the salinity of the waters and resultant effects on the saline deposits.
Langebein (1961) showed that the salinity of closed lakes and, indirectly, of associated lacustrine salt deposits is strongly dependent on their hydrographic and hydrologic character. Among the most important parameters are the annual net evaporation, mean depth, the variability of lake area, and the geometric shape of the lake basin. Hydrographic data for Deep Springs Lake were difficult to obtain because of inaccessibility and frequent large fluctuations in level.
A staff gage was installed at the northwest corner of the leveed area close to the margin of the perennial brine body (fig. 2, loc. DL2). Staff levels were correlated with lake areas by plane-table leveling on nearby remnant shorelines which resulted from relativelyHYDROLOGY AND MINERALOGY OF DEEP SPRINGS LAKE, CALIF.
A21
stable intervals during an overall receding phase. Such shorelines are not everywhere continuous or readily apparent but may usually be distinguished by variation of moisture content, color, and texture retained in the surficial lacustrine salt crust. These shorelines are also conspicuous, though discontinuous, features on aerial photographs of the playa. Shorelines roughly corresponding to individual staff levels were sketched from aerial photograph enlargements, and their position compared with snapshots, tape measurements, field notes, and a plane-table map (scale, 1 inch=1,000 ft) of the lake area made during the dry stage of the early September 1960. Even so, it is not altogether certain that the selected shorelines were fully continuous or representative of a specific stage. The resultant approximation of the detailed topography of the lake as indicated by the shorelines is given in plate 2. The most questionable details are in the area around the southern part of the levee, where the faint shoreline traces are interrupted by alluvial wash, low mounds, and the Buckhorn Springs inflow patterns.
Plate 2 indicates that Deep Springs Lake is generally pan shaped, having a flat bottom and very gently sloping sides. Level shots in the central leveed area indicated that the altitude of the saltpan surfaces was about 0.15 feet below the base of the staff (4,917.4 ± 0.2 ft, as controlled by given nearby point altitudes, U.S. Geol. Survey Blanco Mountain quad.). This level is thought to be virtually uniform on both sides of the central leveed area, with the exception of the thrusted edges on salt polygons and the delta around the Buckhorn Springs inflow. However, when seasonal desiccation begins, surface waters retreat to the southern part of the leveed area where inflow from the Buckhorn Springs stagnates. This is indicated by the successive lettered shorelines of plate 2, which repre-
sent nearly equal levels but successively smaller area.
For hydrographic calculations, water surface areas at Deep Springs Lake were obtained by planimeter from the plot of shoreline positions. Volumes were computed from water surface areas and corresponding staff-level measurements on the assumption of a trapezoidal cross section for the lake as a whole. Staff-level measurements could not be obtained for low stages because water levels fell below the base of the staff, whose position was determined by its stability and accessibility to an observer. Most of the staff-level data and samples for the relatively high lake stages were collected by William A. Jenkins of the California Division of Highways, inasmuch as water surface levels usually fell below the base of the staff during the author’s periods of observation. Volumes for lake stages below the staff base were obtained from crude depth measurements on the salt crust, and area estimates were made from aerial photographs.
Values for hydrographic parameters derived by use of methods just outlined must be viewed with caution. Lake area measurements may be in error by at least 5 percent, depending on the continuity and interpretation of shoreline positions. Such errors are compounded in the calculation of lake volumes, especially at low stage, by deviations from the assumed cross-sectional configuration. Nevertheless, estimates of area and volume were made to show, at least semiquantita-tively, the relations among hydrographic properties at Deep Springs Lake.
Derived values of area and volume for stages observed at Deep Springs Lake are given in table 4. The graphical relation between area and volume is shown in figure 13. This plot suggests a linear-logarithmic relation between area and volume for readily measurable stages of Deep Springs Lake. Such a rela-
Table 4.—Data on water surface area, volume, and tonnage of dissolved solids at Deep Springs Lake (Data on surface area are considered accurate to within 5 percent, volume and total tonnage to within 10 percent]
Sample 1 Date collected Staff reading (feet) Water surface area (acres) Volume (acre-ft) Concentration (ppm) Total tonnage dissolved solids2
DL13 10/3/59 1/30/60 2/4/60 2/20/60 2/29/60 3/19/60 3/31/60 5/5/60 6/5/60 1/14/61 260 13. 0 310, 000 67, 400 7, 840 43, 040 40, 850 35, 480 27, 950 20, 330 10, 460
2D 0. 97 518 450
2E . 75 459 350 8li 400 93, 300 108, 000
2F . 56 434 262
2G . 42 402 176
2H . 24 366 110 124, 000 122, 000 224, 000 295, 000 21, 900
21 . 04 314 57
2C ... Trace. 284 27 9^840 6, 750 16, 660
2B 265 13. 25
2K 1. 20 563 558
1 With the exception of DL13, all samples listed were collected at or near the staff gage (loc. DL2). The first sample was collected toward the center of the saltpan near locality DL13 (fig. 2) at very low stage.
* Calculated on the basis of the equation of Langbein (1961, p. 8): 0.00135Xvolume in acre-feetXconcentration in parts per million by weightXdensity.A22
CLOSED-BASIN INVESTIGATIONS
I ____________I I I I I I I I____________I I I I I I I
10 20 50 100 200 500 1000
CALCULATED VOLUME, IN ACRE-FEET
Figure 13.—Derived relation of water surface area to volume for Deep Springs Lake, October 1959 to Sepetember 1961.
tion was employed by Langbein (1961, p. 5) in calculating the variation in stage of much larger closed lakes. For comparison with closed-lake data computed by Langbein (1961, table 1), response time and area variation were also calculated for Deep Springs Lake.
Response time (coefficient K) is the measure of time, in years, that a closed lake takes to react to inflow fluctuation. Response time is defined by the equation
V" — V'
(Langbein, 1961, p. 6) E^A„ _ ^ , where V" (in
acre-ft) is the value for a high lake stage, V' is a low lake stage, A" and A' correspond to water surface areas (in acres), and E is net evaporation in feet per year (4.25 for Deep Springs Lake), which is assumed constant annually. Response time was computed utilizing volume and area data (table 5) for the highest level readily reached by the lake (based on textural and mineralogic criteria) when water covers about two-thirds of the unvegetated playa area and for the lowest level at which a coherent body of water still exists on the saltpan. Obviously, these limits are arbitrary.
Lake area variability may be characterized by the coefficient of variation, which is calculated from the equation (Langbein, 1961, p. 6)
TT 0.26ti IE(AtIAl)K D V (2+1/X)
where n is the exponent in the relation of lake area {Al) to volume (AL^ TV6 for Deep Springs Lake), D is mean depth, and AT is tributary area. Inasmuch as the mean depth for the period of study apparently remained just below the level of the staff base, the coefficient of area variation was computed for a mean depth of 0.1 foot. The results of these computations and the basic data from which they were made are compared with data of some of the more saline of Langbein’s examples given in table 5.
The calculated response time, K, is somewhat less than that observed for the well-documented recession of lake waters during early 1960 (to be discussed on p. A24) but is considerably more than that observed in October 1959 or September 1961. However, fluctuations resulting from protracted annual spring runoff might be expected to be of longer duration than fluctuations based on isolated, aperiodic runoff. The values of K and V given in table 5 should be close to a reasonable average for the two types of inflow events. As these values suggest, Deep Springs Lake more nearly resembles the typical small playa of the Great Basin than do any examples given by Langbein, who noted the lack of data on such lakes.
Actually, the period of study in the Deep Springs Valley was one of somewhat below-normal precipitation throughout the region. Local residents were able to recall times during the past when lake waters extended beyond the present playa margins. The highest
Table 5.—Hydrologic data from Deep Springs Lake as compared with some major closed lakes cited by Langbein 1
[Explanation of symbols: At, tributary drainage area in square miles; E, net evaporation in feet per year; U, coefficient of lake area variation; K, response time in years;
D, mean depth in feet; Al, lake area in square miles]
Lake At E U K Salinity D Al
Date Parts per million
Deep Springs, 200. 85 4. 25 7. 63 0. 32 Jan. 1961 21, 900 1. 05 0. 88
Oct. 1959 310, 000 . 1 . 41
Great Salt. . ....... 21, 000 2. 70 . 125 9. 0 1877 138, 000 18. 0 2, 200
1932 276, 000 13. 0 1, 300
Owens 2, 900 5. 00 . 10 10 1876 60, 000 24. 0 105
1905 213, 700 11. 0 76
Abert, Ore _ 900 2. 50 . 5 6 1902 76, 000 5. 0 50
1956-59 20, 000 10. 0 60
Lake Eyre, Australia . 550, 000 7. 0 2. 5 1. 5 1950 40, 000 8. 5 3, 100
1951 240, 000 2. 8 740
Tuz Golu, Turkey 4, 400 2. 4 . 5 1 1959 250, 000 2 650
■ Langbein (1961, p. 18, table 1).HYDROLOGY AND MINERALOGY OF DEEP SPRINGS LAKE, CALIF.
A23
Figure 14.—Deep Springs Lake at low stage prior to total desiccation, August 1961, as seen looking southeast from the staff
gage, locality DL2. Footprints give scale.
stages are normally recorded after periods of maximum winter precipitation, February through April, and are the result of snowmelt and associated runoff in the White Mountains. Marked short-term level rises result from intense thunderstorms in the surrounding mountains or directly over the lake in the late summer. The recession of lake waters usually takes place in a sporadic fashion as evaporation fluctuates with weather conditions; rapid late-summer level rises are usually followed by very rapid recession. The migration of receding shorelines is highly irregular locally, and on the flats just outside the leveed area, remnant puddles are left behind by receding lake waters in the most inconspicuous topographic depressions. Fractionation of lake waters into a group of isolated puddles becomes more conspicuous in the salt crust areas, where polygonal fracturing and thrusting of fractured crustal edges gives rise to sharp microrelief and promotes separation of lake waters into myraid shallow ponds prior to complete desiccation (fig. 14). At low stages, lake waters are commonly divided into three units separated by the physical features of the playa and density gradients due to variable salt concentrations. These units include: (1) inflow waters which may become somewhat stagnant and accumulate near the inflow channels of the bog-mound springs, Corral Springs, and particularly the Buckhorn Springs, (2) standing surface waters in the central or south-central part of the leveed area, and (3) a perennial brine body which occurs interstitially in the salt crusts and in a few stagnant ponds around large crustal fractures or in the trough just inside the levee. Mixing of these water units takes place more slowly as the stage rises, although circulation can be aided considerably by wind action. At lowest stage, surface
of Deep Springs Lake.
water at Deep Springs Lake is confined to a few small stagnant pools of dense amber-colored brine (fig. 15).
In spite of the problems of inhomogeneity in surface waters, as just outlined, an attempt was made to find what relations might exist between volume and the amounts of salts in solution in the lake waters. Samples were collected in the northwestern part of the leveed area, near the staff gage, where presumably at least the inhomogeneities resulting from inflow source would be held to a minimum.
Data on the relation between lake volume and tonnage of salts in solution at Deep Springs Lake, have been given in table 4, and are presented graphically in figure 16. It should be emphasized that whereas the points corresponding to the receding phase of the curve are derived from consecutive samplings, the other sampling points are not. For example, the sample of January 1961 was collected near the peak of a risingA24
CLOSED-BASIN INVESTIGATIONS
Figtjbe 16.—Relation of lake volume to total dissolved solids
at Deep Springs Lake for samples collected from October
1959 to January 1961. The insert shows the “ideal cycle”
of Langbein (1961, p. 9).
stage, but the area became inaccessible shortly thereafter. In spite of the lack' of samples during rising and initially receding stages, the dashed lines are thought to approximate the trends at Deep Springs Lake. The trend of the recession curve is rather regular, although it must be remembered that the salt tonnage in solution is derived in part from the volume data, and small errors in the volume values will have a large effect on the continuity of the volume-tonnage relation.
It should be noted in connection with figure 16 that salt precipitation apparently did not take place until just prior to March 19. Thus, it is apparent that nearly 50 percent of the total dissolved solids in the lake waters was lost from solution prior to solid precipitation.
The loss of dissolved salts during fluctuations in the volume of saline lakes has been discussed by Langbein (1961, p. 10). To explain such losses, he has proposed entrapment of salts in bordering sediments and removal of salines from dried parts of the lakebed by wind action. To test the first part of Langbein’s hypothesis, samples of the sediment from Deep Springs Lake were leached by repeated distilled-water washing. Data from the playa sediments outside the leveed area are incorporated with the stratigraphic information given in figure 5. The sediments were found to contain more than 30 percent material which was readily water soluble. Higher values were obtained for muds within the leveed area. Some samples contained more than 17 percent readily soluble material in the absence of a detectable saline phase, as determined by microscope
or X-ray techniques. In fact, some of the muds of Deep Springs Lake contain more than 80 percent material of precipitate origin (including alkaline earth carbonate as well as soluble salts).
The determination of the solubles content of muds of Deep Springs Lake indicates that dissolved salts may indeed be incorporated with bordering sediments as the surface waters recede. The stirring up of carbonate muds in the lake waters by wind action has been directly observed. The deposit of such sediment, one-tenth of a foot deep and containing 25 to 30 percent soluble salts on the -square mile area left exposed by recession of lake waters from January 30 to March 19, 1960, could readily account for the loss of 30,000 tons of dissolved salts (fig. 16) from the receding waters in that time.
The effect of wind action on the removal of salts from central Deep Springs Lake is considered to be relatively minor. Although the wind frequently stirs up clouds of salt-laden dust.from efflorescent crusts on the west side of the playa (fig. 17), the well-indurated crusts of the central lake area do not contribute much to aeolian transport. Observations throughout Deep Springs Valley suggest that most wind-transported material either remains within the drainage area or that the material lost is largely replaced by similar material carried into the basin.
Another mechanism for the removal of salts available for solution in Deep Springs Lake waters is the burial of saline muds and crusts by an influx of clastic sediments from one of the large ephemeral drainage systems. There is little evidence, however, that in recent years such an influx has carried as far as the saltpan.
Langbein (1961, p. 8) stated that the variation of salinity with lake volume is not directly a simple matter of concentration or dilution of a fixed mass of salts in a changing volume of lake water. The mass of salts in solution appears to decrease with contraction in lake volume,
Figure 17.—Salt-laden dust being blown off the west side of the Deep Springs Lake playa.HYDROLOGY AND MINERALOGY OF DEEP SPRINGS LAKE, CALIF.
A25
but although the mass in solution may increase with a recovery in lake volume, the total mass may be less than before the recession began. Completion of the cycle is accomplished during slow accumulation of salts carried in solution by the influent streamflow.
Langbein (1961, p. 9) also presented a schematic volume-salinity cycle and suggested that “actual volume-tonnage curves * * * may never describe the idealized cycle * * * but may consist of a series of zigzag curves of which each part is related to one of the limbs of the hypothetical cycle.”
The volume-tonnage relations for Deep Springs Lake illustrated in figure 16 appear to follow Lang-bein’s schematic cycle closely, but if one considers the potential errors in the volume data, the agreement in detail could be simply fortuitous. The only apparent disparity between the relations shown for Deep Springs Lake and Langbein’s ideal cycle is in the lower part of the recession curve. This disparity could be explained as an effect of direct salt precipitation from the lake waters.
In most closed lakes the proportions of constituents in solution remain virtually uniform throughout any decrease in volume and increase in concentration. This uniformity exists because of peripheral effects and lack of any significant precipitation, within the receding waters, of salts intermediate in solubility between alkaline earth carbonates and sodium chloride. As the concentration increases, each further reduction in volume is associated with a proportionately greater decrease of the total salts in solution. At Deep Springs Lake, however, the precipitation of sodium sulfate— the major salt in solution at high stage—apparently stabilizes the ratio of further volume-tonnage decrease. This would explain the linearity of the lower part of the curve representing the receding phase (fig. 16).
The effect of sulfate precipitation is more apparent in the relation of volume to concentration (fig. 18). During the receding phase, concentration increased regularly but slowly until precipitation began. The concentration remained constant until some sort of equilibrium with the precipitate was reached, and then increased rapidly, as the lake decreased further in volume, until complete desiccation was reached. This rapid rise was not noted by Langbein (1961, p. 10) in his volume-concentration relations hypothesized for a playa lake.
For the filling cycle of a playa lake. Langbein (1961, p. 10) hypothesized that fresh-water inflow will cause a rapid increase in dissolved-solids concentration owing to solution of crustal salts until the point of saturation is reached and then a gradual decrease in concentration as further increase in lake volume takes place after the crust is dissolved. These general trends
DISSOLVED SOLIDS, IN THOUSANDS OF PARTS PER MILLION
Figure 18.—Relation of surface-water volume to dissolved-solids concentration for a receding phase of Deep Springs Lake. The salt precipitated is mirabilite (NasSOulOHsO).
were noted at Deep Springs Lake, but the magnitude of an initial rise in concentration in the inflow waters was strongly dependent on the abundance of the different salts and their solubilities at the various points of inflow. The role of such factors is more evident in the section discussing the saline mineralogy of the lake area.
HYDROCHEMISTRY ANALYTICAL METHODS
The procedures employed in the collection and analysis of waters from the Deep Springs Valley were those in general use by the U.S. Geological Survey. Though some preliminary measurements were made in the field, most of the analytical work was done in the laboratory by Shirley L. Rettig, advised by H. R. Feltz. The reproducibility and limits of accuracy in the methods are generally the same as those set forth by Rainwater and Thatcher (1960) and Hem (1959), although special problems exist in the analysis of brines, and potential sources of error were greatly magnified when compared to determinations in more dilute waters. Some of the problems encountered and special techniques employed are discussed in greater detail below.
SAMPLING
A variety of containers were utilized in the sampling of waters. Initially, 1-liter Pyrex bottles with paraffin inner linings and Teflon-coated stoppers were used. Although contamination from glass was thus held to a minimum, these containers were too bulky
766-977 0—65
AA26
CLOSED-BASEST INVESTIGATIONS
and fragile for convenient field use. Further, the paraffin linings tended to crack and flake. These problems outweighed the possibility of gaseous leakage in polyethylene bottles, which were employed in subsequent work.
Where there was a distinct problem in retaining gaseous components present in solution at time of sampling, “citrate-of-magnesia” bottles were used initially. However, the pressure seals of these bottles were frequently defective, and greater success (plus less contamination during storage) was achieved by use of gas analysis tubes or fixatives.
Waters from the Deep Springs Valley were sampled during five periods of field observation from October 1959 to September 1961. As previously mentioned (p. A21), some additional samples taken from Deep Springs Lake at high stages during early spring were collected by W. A. Jenkins of the California Division of Highways. Bottles were simply dipped into waters of sufficient depth; but for very shallow waters, samples were taken by bulb suction, and care was taken not to pick up solids. During the period of study, up to 6 samples were collected from each major source of inflow to the lake, usually at the time that discharge estimates were made. For both inflow and the lake, samples were collected to establish point-to-point compositional variation as well as fluctuations with time.
Temperature measurements were made at the source of all samples, but some early measurements were in error because of a faulty thermometer.
pH AND ALKALINITY
It is well known that significant changes in the pH-alkalinity-carbon dioxide balance commonly take place between the time of water sample collection and analysis (Hem, 1959). These changes are brought about not only by inorganic loss or gain of carbon dioxide but also by subsequent variations in temperature, oxidation, precipitation, or the activity of microbiota. These factors are most pronounced where there are marked changes in conditions between the field and laboratory. Therefore, the pH and alkalinity for a majority of samples from the Deep Springs Valley were measured in the field.
The pH was measured electrometrically in the field by using several standard instruments in the course of the study. Measurements were normally made directly at the point of collection. The electrodes were kept as moist as possible and were standardized against sodium carbonate (pH'—10), borax (pH'—9), and phosphate (pH'—7), buffers. Suitable corrections
were made for temperature, which was brought to within 5° C of the sample wherever possible. Both electrode fatigue and instrumental drift were noted, and empirical corrections were applied by frequent buffer standardization. Nearly all measurements were reproducible and are considered accurate to at least ±0.15 pH unit, depending in many cases on temperature changes between collection and measurement on site. Sodium-ion and junction-potential effects expected in highly concentrated waters were found to be within the limits of error stated above by checking against saturated standard sodium carbonate solutions (Bedekar, 1955). These problems are discussed more fully by Bates (1954).
Alkalinity measurements in the field were made po-tentiometrically and were usually made immediately after the determination of pH. Quantitative distribution of C03~2 and HC03-1 were computed from the amount of standard sulfuric acid titrant necessary to reach the commonly assumed end points at pH of 8.2 and 4.5. These end points probably are sufficiently close to the true end points for waters having total alkalinity ranging from 100 to 450 ppm; therefore, no significant error is introduced. (For a detailed discussion of this problem, see Barnes, 1964.) However, for waters having higher alkalinity, errors resulting from variation of the end points from these standard values are to be expected. Because of the large amounts of titrant required and the expected interference of other weak acid radicals—such as borate, phosphate, and silica—alkalinity of the more concentrated waters from Deep Springs Lake was not measured in the field. Instead, an attempt was made to preserve the distribution of ionic species through use of gas analysis tubes. Use of this technique was only partially successful because of temperature variation and microbiotic activity. Comparison of gas sample tube analyses with field measurements in dilute waters suggests potential errors of more than 5 percent. Closer agreement was obtained where very small amounts of chloroform were added to the gas sample tubes to hinder biotic action, as suggested by K. V. Slack (oral commun.). A comparison of field data for pH and alkalinity with that determined in the laboratory on samples from bottles and gas tubes is presented in table 6. For waters having a total alkalinity of less than 500 ppm, the gas tube analyses reproduced field data to within 5 percent, but differences between field data and laboratory data on bottle samples exceeded 20 percent. For waters whose alkalinity is greater than 500 ppm, gas tube and standard laboratory data differed by even greater amounts. At high concentrations and pH’s,HYDROLOGY AND MINERALOGY OF DEEP SPRINGS LAKE, CALIF.
A27
there is a significant contribution to the total alkalinity by borate, phosphate, and silica, but inasmuch as there is no satisfactory data on the state of these elements in such solutions, no correction has been attempted.
In order to obtain additional data characterizing the carbonate-bicarbonate-carbon dioxide system in the Deep Springs waters, total carbon dioxide content of several samples was determined in the laboratory utilizing a manometric gas evolution technique modified from that described by Pro, Etienne, and Feeny (1959). These values, which were reproducible within 5 percent for highly concentrated waters, are also presented in table 6. Potential error in this method increases significantly at concentrations of less than about 800 ppm because of very small changes in manometer readings and difficulties in the calibration of the volume of the system. Bicarbonate concentrations calculated from the data on total carbon dioxide sometimes more nearly approximate field measurements than standard laboratory determinations, but the results are not consistent, and total range of variation in the three values may still exceed 20 percent.
Table 6 presents total carbon dioxide content data for many samples both as determined manometrically and by conversion of the titrated C03~2 and HCO3"1 values, including free carbon dioxide calculated as suggested by Rainwater and Thatcher (1961, p. 139) for waters containing less than 800 ppm in total alkalinity. The calculations of free carbon dioxide are based on the ionization constant for carbonic acid at 25° C and are uncorrected for ionic strength. Such errors are probably significant only for waters which have a temperature differential of 10° C and a pH of less than 7.5. Nevertheless, significant differences between the calculated and measured values for total carbon dioxide concentration are readily apparent. Calculated values of total carbon dioxide content are substantially lower than those measured in waters of low concentration and higher than those measured in waters of high concentration.
In view of the difficulties just outlined, data on the carbonate-bicarbonate-carbon dioxide system at Deep Springs Lake cannot be considered to be more than semiquantitative. From information currently at hand, it is not possible to establish the cause of all anomalies.
MAJOR-ELEMENT ANALYSIS
Calcium and total hardness were determined by EDTA titration, and magnesium was determined by difference. The values reported for calcium and magnesium in the highly concentrated lake brines are probably no more accurate than an order of magni-
tude, because of very small concentrations and the possibility of large interferences. The determination of calcium and magnesium in dilute waters from the valley may also be in error because of the precipitation of alkaline earth salts between time of collection and analysis. (For particular reference to Birch Creek, see Barnes, 1962.)
Alkalies in Deep Springs waters were determined by direct-reading flame photometry. Standards were run coincidentally with sample, and multiple readings were frequently made. The large dilutions necessary to analyze concentrated waters may have introduced small errors, but results were reproducible within a few percent. Gravimetric analysis for sodium in a few samples gave values within 10 percent of those determined by flame photometry.
Sulfate and chloride were determined gravimetri-cally as barium sulfate and silver chloride, respectively, in the concentrated lake waters. Sulfate was analyzed by the thorin method and chloride was determined volumetrically by Mohr titration in the more dilute inflow. These determinations are probably the most reliable of all data on Deep Springs Valley waters.
SUIiFUDE
Sulfide in some samples was preserved by collection in gas analysis tubes or by adding approximately 2 grams of zinc acetate to the sample at the time of collection. The iodometric method was utilized for both types of analysis.
PRESENTATION AND INTERPRETATION OF DATA
The analytical results for 125 samples of water taken from the Deep Springs Valley during the 2^ -year period of study are presented in table 7. Sample numbers correspond to the points of collection shown on the location maps (figs. 2 and pi. 1). The analyses are grouped roughly with decreasing distance from the perennial brine body at Deep Springs Lake. Thus, they are listed as: (1) streamflow including springs immediately within major drainage units, (2) wells, except for the old artesian well adjacent to the lake area, (3) springs of the lake area including the artesian well spring, (4) ponds, (5) surface waters of the lake area, and (6) interstitial brine. Except for the closed sag pond, all water outside the immediate playa area has been considered as inflow to Deep Springs Lake. Because of the analytical problems discussed previously, a strictly quantitative evaluation of the data is impossible, but significant qualitative relations are apparent.A28
CLOSED-BASIN INVESTIGATIONS
Table 6.— Total CO2, pH, and alkalinity for samples of water from the Deep Springs Valley
[Sample numbers refer to localities shown in fig. 2 and pi. 1. Only samples for which more than the normal laboratory measurement of alkalinity was made are included. Concentrations are in parts per million. Symbols: Asterisk (*) indicates data from gas analyses tube; n.a., indicates data not analyzed]
Sample Total CO2 pH Alkalinity
Measured Calculated from— Field Laboratory HCOs CO,
Laboratory data Field or gas tube* data Field Laboratory Field Laboratory
DSH1 184 193 7.50 7.8 250 245
DSS1 - 202 7.45 7.7 268
DL1 21, 500 24,400 9.15 12,600 20,900
2 19i 500 23',250 9.4 12; 400 19i 500
2B S, 730 8' 710 9.3 5; 330 Q, 640
2C 6, 890 6' 800 9.4 3; 770 5'570
2D — 2,750 2,740 9.6 b 680 2, 080
2E 2,760 2,980 9.5 2,000 2,100
2F 2'800 2^870 9.5 2,150 1,800
2G_ 3' 260 3, 410 9.3 3; 160 1, 540
21 3; 450 3| 630 9.5 2,390 2 , 600
2J IS, 400 * 18,600 9.58 9.4 i; 690 23i 400
* 9.60 * 7,860 * 17,600
2L 9.5 22,100
2P 17, 500 9.35 1,210 22j 700
3A 19,000 24' 700 9.30 9.35 15,300 IS, 400
3B IS, 700 21,900 11; 100 IS, 900
13 19' 900 2l’, 300 10.20 9.3 8,720 20', 500
14 5' 220 2,970 10.00 9.35 371 3,680
21 12, 600 13, 600 9.65 9.2 5,320 14,700
22 22j 700 24,700 9.70 9.1 16,000 IS, 200
24 17; 500 17', 100 9.70 9.15 S, 730 17^400
27C ' 135 7.23 7.3 169 .0
29A 10,300 10,700 9.75 8.8 11,200 3, 630
30A S, 190 7', 190 9.40 8.85 7,000 2,920
31 ' 618 '578 8.85 8.1 788
32 218 251 211 8.35 7.7 290 334 9.8 .6
32C 206 8.79 7.8 304
33 190 142 8.45 7.2 195 232 .6
33C 148 7.78 7.5 192 .0
34 133 152 157 7.75 8.4 210 209 .0 .0
34C 158 8. 01 7.7 210 .0
35 256 274 172 7.55 7.4 225 350 .0 .0
35C 165 8.13 7.8 220 .0
36 280 349 185 7.65 7.7 245 464
36D 176 * 180 * 8.6 7.6 * 216 232 * 20 .0
36E 178 177 7.97 7.65 240 236 .0 .0
* 175 *7.75 * 233
37 258 317 241 7.70 7.80 320 426
37 C 215 8.11 7.8 291
38A 1,080 1,310 771 8.10 7.80 1,050 1,750 .6 .0
38 B 1,070 676 8.35 7.20 930 1,300 .0 .0
39A 50, 600 73,100 10.20 9.7 46,800 53,600
39B '560 '658 397 7.90 8.1 535 896 .0 .0
39C 29, 500 30,900 9.6 18,700 23,800
39D-E 55; 400 * 48,800 10.18 10.1 12; 500 63' 300
* 10.10 * 10,400 * 56,300
39 G 22, 200 9.7 11,000 19, 500
40A — 730 '748 792 9.05 8.9 970 919 123 113
40B 2,340 10.15 9.2 2,220 1,020
41A 190 '217 215 7.70 7.8 285 291 .0 .0
41F 212 228 8.08 7.6 310 278 .0 .0
*212 *7.9 *289 *.0
42 344 260 233 7.10 7.1 275 308 .0 .0
43 . 111 161 7.65 7.9 218 .0
44 110 143 8.80 7.8 192 .0
44B 191 7.93 7.5 248 .0
45 263 260 8. 20 8.4 350 354 4.9 8
CC47 87 132 8.10 7.9 179 .0
WC47 186 212 8.10 7.7 282 .0
47 E 195 201 8.60 7.9 265 263 12.8 .0
DL48A3 143 *135 *8.9 7.7 *154 190 *33 .0
48B3 353 *372 *8.7 7.3 *476 441 *38
48A4 147 8.36 7.7 196 .0
48B4 279 8.61 7.75 371 .0
48C 186 184 8.29 8.0 253 252 .0 .0
51A 22,900 *20, 500 9.58 9.7 *2,430 9, 360 *25, 500 22,100
BC70 73 104 *110 8.52 8.0 *144 142 *.0 .0
*7.6
AS71A 160 *158 7.22 8.1 *210 218 *.0 .0
*7.7
71C 151 153 7.89 7.8 207 203 .0 .0
DL73A 132 152 *160 8.01 7.6 *218 200 *.0 .0
*8.1
73C — 152 157 156 8.26 7.65 213 208 .0 .0
*158 *8.0 *215 *.0
73D 325 388 7.95 7.8 525 436 .0 .0
78 145 137 8.84 7.8 188 194 17.2 .0
AS79 143 *145 8.69 8.0 *200 197 *.0 .0
*8.0 HYDROLOGY AND MINERALOGY OF DEEP SPRINGS LAKE, CALIF.
A29
Table 7.—Analyses of waters from the Deep Springs Valley, 1959-61
l Sample numbers refer to localities shown in fig. 2 and pi. 1. Locations and samples are listed in approximate geographical order from valley margins to the playa area.
Concentrations are in parts per million]
Location Sample Date of Silica Cal- cium Mag- ne- Sodium Potas- sium Bicar- bonate Car- bonate Sulfate Chloride Flu- oride Boron Total dis-
collection (SiO.) (Ca) sium (Mg) (Na) (K) (HCOj) (CO3) (SO,) (Cl) (F> (B) solved solids
Stream flow
Crooked Creek:
Just above confluence CC47B Oct. 20,1959 20 46 6.3 20 7.0 179 0 21 4.5 0.2 0. 06 213 7.9
with Wyman Creek. 47 June 7,1960 28 42 5.8 9.8 2.3 150 0 18 2.5 .2 .01 183 7.8
47D May 20,1961 24 46 5.1 11 1.8 170 0 20 4.0 .2 .07 172 7.6
Wyman Creek:
Just above confluence WC47 Oct. 21,1969 16 69 17 12 6.0 282 0 35 2.0 .2 .03 296 7.7
with Crooked Creek. 47B June 7,1960 20 58 14 9.4 3.0 238 0 41 2.0 .2 .01 265 7.9
47C Sept. 1960 27 85 12 12 8.5 286 0 43 1.0 .3 .20 330 7.6
47E May 20,1961 22 69 16 9.5 2.7 265 0 40 2.8 .1 .10 291 7.9
At gage just above 56A June 7,1960 21 59 13 9.5 2.9 226 0 37 2.5 .3 .02 256 8.0
sediment trap. 56B Sept. 13,1960 23 65 18 10 3.5 264 0 40 3.0 .2 .03 293 8.2
Spring beneath inclu- 57 June 7,1960 16 55 11 7.3 2.3 198 0 36 2.0 .1 .00 228 7.6
sion riddled outcrop, Yz mile above confluence with Crooked Creek.
mile below Mill Canyon. 58 19 56 11 7.3 2.3 198 0 37 1.0 .2 .02 232 7.8
Source near Roberts 59 A Sept. 27,1959 20 53 16 12 6.0 234 0 17 4.5 .1 .08 244 7.5
Ranch. WS59B June 7,1960 37 112 19 18 3.0 366 0 60 5.0 .3 .04 435 7.5
Birch Creek:
Near point of disap- BC70 Sept. 11,1960 22 42 23 13 3.3 142 0 104 4.0 .7 .02 282 8.0
pearance (other points on fig. 23 from Barnes, 1962). Antelope Creek:
AS79 Aug. 16,1961 Sept. 11,1960 Aug. 16,1961 19 44 27 10 3.2 197 0 70 5.5 . 1 .00 302 7.9
Antelope Springs: 71 20 56 26 13 4.0 218 0 93 6.0 .4 .01 325 8.1
71C 20 56 24 12 3.6 203 0 90 8.5 .3 .00 314 7.6
Northern spring 2 72 Sept. 11,1960 21 55 25 12 4.3 202 0 100 5.0 .7 .03 323 7.9
Wells
DSHl 4.1959 9.1960 16 69 16 14 7.0 245 0 45 5.0 0.1 0.03 293 7.8
2 June 21 64 15 22 10 230 0 88 6.0 .2 .07 339 8.0
3 May 20,1961 24 67 16 10 2.5 252 0 39 6.5 .1 .10 317 7.3
Deep Springs School DSS1 Oct. 8,1959 18 76 17 15 7.0 268 0 57 4.1 .1 .05 326 7.7
irrigation well. Stock well, Central Val- DL43 Oct. 20,1959 35 42 10 88 15 218 0 121 10 2.5 .22 431 7.9
ley. 43B Oct. 10,1961 39 35 12 55 3.3 186 0 100 5. 1 2.5 344 8.1
Stock well, west of Deep 44 Oct. 20,1959 26 32 31 45 12 192 0 133 12 1.4 .11 388 7.8
Springs Lake. 44B Aug. 16,1961 73 49 32 26 10 248 0 109 9.5 1.1 .10 431 7.4
Springs of the Lake Area
Bog-mound springs: Northwest bog-mound DL27A Oct. 7,1959 39 40 22 37 10 230 0 105 3.0 1.8 0.14
spring. 27AS Feb. 1960 86 46 26 39 9.0 270 0 98 9.0 1.8
27 C Aug. 14,1961 48 38 22 28 7.0 169 0 100 9.0 1.8 .05
Western bog-mound pond. Central bog-mound 78 42 28 16 39 12 194 0 59 8.5 1.4 .03
73 A Sept. 12,1960 48 37 13 18 7.6 200 0 30 4.0 .4 .06
spring. 73B May 17,1961 48 39 14 20 7.7 200 0 30 7.0 .3 .14
73C Aug'. 16,1961 46 36 17 15 7.2 208 0 30 4.0 .2 .41
Central bog-mound 73D 56 78 31 26 11 436 0 27 4.0 .8 .34
marsh. Eastern bog mounds: Artesian well spring.. 41A Oct. 19,1959 41 14 14 40 45 291 0 1.0 4.0 1.2 .33
41E May 28,1960 46 16 13 35 48 288 0 6.6 .0 .43
41F Aug. 12,1961 46 14 14 36 46 278 0 8.0 3.8 1.1 .34
Northeast bog 42 Oct. 19,1959 42 19 18 40 46 308 0 17 5.0 1.2 .36
mound. Corral Springs: 48 A Mar. 28, 1959 46 17 14 44 16 186 4.8 36 3.6 .4 .10
48 A1 Oct. 28, 1959 41 18 10 48 20 202 0 40 6.0 .4 .05
48A2 Sept. 9,1960 42 17 9.1 46 16 188 0 37 6.5 .4 .09
48A3 May 22, 1961 39 18 10 45 15 190 0 38 7.0 .3 .13
48A4 Aug. IS, 1961 43 16 11 44 15 196 0 41 6.8 .3 .03
48B Mar. 28, 1959 61 10 7.5 234 48 402 78 107 7.2 . 9 1. 0
48B1 Oct. ' 1959 6.4 2.9 230 55 420 14 193 57 .9 .87
48B2 Sept. 9,1960 29 8.0 2.4 280 52 496 0 143 87 1.0 1.0
48B3 May 22, 1961 26 6.0 2.4 205 50 441 0 82 56 .8 .68
Spring intermediate 48B4 Aug. 15, 1961 32 8.0 5.6 222 34 371 0 161 77 .8 .06
between Nos. 1 and 2. 48Int___ May 22, 1961 39 14 8.6 56 15 192 0 44 7.0 .4 .11
Outflow from Nos. 1 48 C Aug. 15, 1961 39 13 9.2 112 23 252 0 82 30 .5 . 15
and 2 combined. Corral Spring inflow 50 Mar. 28, 1959 43 16 18 769 118 661 113 593 385 1.0 2.6
to the lake. Buckhorn Springs: 33 Oct. 13, 1959 41 40 20 60 18 232 0 121 1.5 .5 .09
33B May 22', 1961 45 39 21 48 10 192 0 119 10 .5 . 14
33C Aug. 12, 1961 44 38 22 44 9.0 192 0 120 8.0 .5 .07
34... Oct. 13, 1959 37 34 21 75 22 209 4 138 11 .6 . 14
34B May 22, 1961 42 39 19 68 13 216 0 134 10 .5 .15
34C Aug. 13, 1961 41 34 22 65 14 210 0 135 9.5 .5 .07
35.. Oct. 13, 1959 38 35 20 88 18 350 o 127 2. 0 .6 . 12
35C Aug. 13, 1961 41 32 23 68 12 220 0 140 12 .6 .07
7166-977 O—65- 5
371
447
337
303
257 267
258 484
304
307
447
341
273
284
267
264
273
754
774
848
698
770
271
278
2,390
416
387
380
446
431
424
501
436
7.5 8.1 7.0 7.7
7.6 7.4 7. 55
7.6
7.8
8.2
7.4
7.1
8.3
7.8
7.9
7.7
7.8 9.1
8.4
7.9 7.3 7. 65
7.9
8.0
9.0
7.2
7.4 7.35
8.4
7.4 7.6
7.4 7. 65A30
CLOSED-BASIN INVESTIGATIONS
Table 7.—Analyses of waters from the Deep Springs Valley, 1959-61—Continued
[Sample numbers refer to localities shown in fig. 2 and pi. 1. Locations and samples are listed in approximate geographical order from valley margins to the play area.
Concentrations are in parts per million]
Location Sample Date of Silica Cal- cium Mag- ne- Sodium Potas- sium Bicar- bonate Car- bonate Sulfate Chloride Flu- oride Boron Total dis-
collection (SiO,) (Ca) sium (Mg) (Na) (K) (HCOs) (COs) (SO,) (Cl) (F) (B) solved solids
Springs of the Lake Area—Continued
Buckhorn Springs—Con.
DL36 Oct. 13, 1959 34 35 19 100 22 464 0 62 2.0 0.6 0.15 504 7.7
36C Sept. 1960 39 36 27 89 16 233 0 178 16 .8 .15 517 7.8
36D May 22, 1961 41 38 24 82 15 232 0 188 14 .8 .20 517 7.6
36E Aug. 13, 1961 39 35 25 88 13 236 0 181 16 .6 .08 513 7.5
Buckhorn Springs in- 38 A Oct. 14, 1960 77 32 177 450 100 1,750 0 325 125 2.7 .78 2,150 7.8
38B 68 18 130 605 110 1,300 0 691 250 2.8 1.6 2,520 7.2
38D Aug. 13, 1961 52 62 42 154 28 '518 0 227 28 1.8 .36 850 7.7
38E 26 42 434 94 598 15 556 155 2.3 1.3 1,650 8.3
40 A Oct. 14, 1959 25 80 1,200 230 919 113 1,370 695 2.8 4.2 4,190 8.9
40B 11 52 11,700 3,910 2,220 1, 020 10,500 10,400 3.6 24 38, 700 9.2
40C Aug. 13, 1961 19^ 500 2,430 2,180 2,140 17,300 14,100 29 44 56,800 9.6
54A 26 60 '800 105 '735 0 1,080 390 1.7 2.4 2,940 8.1
54B do 22 37 1,620 188 1,570 45 1,960 648 2.4 4.6 5,370 8.5
Ponds
Ephemeral: 8.
North end of slump ridges east of lake. DL31 Oct. 9, 1959 51 134 597 84 78 0 1, 440 246 2.27 2,974
8.
Marsh fed by outflow from open sag pond. 45 Oct. 20, 1959 42 31 250 40 354 8 333 98 1.0 .43 1,010
8.
61 Sept. 1960 12 3.4 1,200 552 828 90 638 1,360 4.8 22 4,320
runoff northwest corner of playa. 9.
Accumulated runoff, 28 June 6, 1960 38 12 4.4 240 92 540 74 59 30 2.7 10 828
carbonate flats area, north of playa. 7.
32 Oct. 9, 1959 May 22, 1961 36 31 26 190 20 334 0 281 62 .8 .34 812
32B 32 23 24 209 23 293 0 279 61 .8 .36 824 7.
32C Aug. 15, 1961 34 24 27 186 23 304 0 276 63 .9 .07 845 7.
37 Oct. 14, 1959 28 46 24 125 25 426 0 199 1.0 .8 .01 692 7 A
37B May 22, 1961 18 51 32 118 18 331 0 235 21 .9 .20 674 7.
37 C Aug. 15, 1961 37 44 32 101 21 291 0 226 20 .8 .35 654 7.
Closed sag pond: 49,000 50 303 .68 279,200 1, 210
Low stage 39 A 39B Oct. 14,1959 Oct. 19,1959 13 7.7 92,200 400 9,150 50 46,800 896 53,600 0 50,100 200 1.1 9. 8.
39C June 6,1960 18 8.8 48, 700 2,710 18, 700 23,800 35, 700 11,900 41 .00 132, 000 9.
High stage Low stage 39D&E. 39G Sept. 9,1960 May 22,1961 12 3.0 101,000 36,600 15, 700 2,400 12, 500 11,000 63,300 19,500 40, 500 27, 200 58,100 8,460 171 11 532 36 286,000 97,300 10. 9.
Deep Springs Lake Playa Area
Surface waters:
DL29A Oct. 8,1959 54, 000 14, 000 11,200 3,630 8,180 79,000 94 165,000 8. 8i
north side. 30 A Oct. 9', 1959 37', 900 10', 100 7,000 2,920 11,400 53,800 75 120,000 8. 8c
30 E May 20j 1961 50 158 3,650 1,510 2,040 0 365 5,930 2.4 3.1 13,000 7.4
2B 18 15 95, 500 3,760 5,330 6, 640 152,000 31,700 20 .00 293, 000 9.3
2C May 5,1960 13 6.1 73,000 4,150 3,770 5, 570 109,000 29,900 22 .00 224,000 9.4
2D Jan. 30,1960 10 23 21,400 3,040 1,680 2,080 18, 500 20,300 2.1 .00 66,300 9.6
2E Feb. 4,1960 11 24 26,500 3,670 2,000 2,100 21,100 26,200 2.3 43 80,800 9.5
2F Feb. 20,1960 12 20 29,800 2,790 2,150 1,800 40, 500 16,800 1.6 .00 90, 500 9.5
20 Feb. 29,1960 13 11 37,900 2,670 3,160 1,540 51,100 16,400 2.0 .00 111,000 9.3
2H Mar. 19,1960 11 52 39,900 3,820 2,270 2,790 51,700 23,200 1.8 22 123,000 9.5
21 Mar. 31,1960 15 46 38,800 3,740 2,390 2,600 50,600 18, 500 1.8 33 116,000 9.5
2K Jan. 14,1961 25 2.3 7,370 926 1,400 89 7,640 5,280 1.4 11 21,900 8.6
Intercrustal Brines
Trough, north side of central lake area.
Trough at staff gage..
Southeast central lake area.
Southern thin crust, near Buckhorn inflow.
Southwest central area...
Western area, near gas vent.
Northwest trough.......
South central lake area near toe of Buckhorn.
Inflow delta___________
DL1 . Sept. 29,1959 99,000 19,000 12,600 20,900 38,800 108, 000 407 291,000
3A Sept. 30,1959 103,000 22; 600 15,300 18,400 27, 700 127,000 364 314,000
3B Oct. 2,1959 102,000 19,400 11,100 18,900 37,100 116,000 432 306, 000
2 May 31' I960 6.1 3.6 110^000 24, 500 12,400 19,500 57,800 120,000 41 230 340,000
2J Sept. 1960 1.5 3.0 106,000 21,200 1,690 23,400 62,600 113, 000 40 335 329,000
2L May 17,1961 3.1 1.3 116,000 24,800 22,100 56,900 118, 000 290 334,000
2M 110', 000 18,600 24,000 52,200 119,000 358 330,000
2N May 20,1961 1.5 1.1 108,000 24'000 22,500 55,200 120,000 255 333,000
2P Aug. 11,1961 106,000 19,900 1,210 22,700 63,100 98,500 149 305 316,000
13 Oct. 3,1959 102,000 21,000 8, 720 20,500 43,000 111,000 13.0 544 310, 000
14 Oct. 5,1959 37,000 4,610 371 3,680 37,000 28,400 17.4 104 110,000
21 Oct. 6,1959 97, 700 12,600 5,320 14,700 54,100 101,000 17.6 310 288,000
22 Oct. 7,1959 88,800 16,000 16,000 18, 200 62, 700 75,000 13.0 344 277, 000
24 102, 000 20,400 8,730 17,400 41,000 119,000 12.0 392 311,000
68 Sept. 9,1960 6.2 3.6 105; 000 18; 800 12,800 20,400 74,100 94,600 99 935 321, 000
53 June 1,1960 12 15 108,000 30,900 12, 700 19,200 53, 700 120,000 29 53 341,000
53B do 6.3 29 110,000 25,200 0 26,600 54,100 117,000 111 752 336,000
9.15
9.30
9.35
9.4
9.4
9.4 9.45
9.5
9.35
9.30
9.35
9. 20 9.10
9.15
9.6
9.5
9.3
Interstitial Brines
In muds DL29C Oct. 8,1959 3.9 1.1 88,300
51 Sept. 12,1960 3.1 1.2 111,000
15,600 8,640 15,900 37,600 103,000 62 166
19,500 9,360 22,000 57,100 119,000 92 739
265.000
335.000
9.4
9.7HYDROLOGY AND MINERALOGY OF DEEP SPRINGS LAKE, CALIF.
A31
Along with the analyses, the equivalent percentages of the major cations and anions have been computed, compiled in table 7, and plotted on trilinear diagrams (figs. 19, 20 and 21) such as proposed by Piper (1944). This system brings out the chemical relations among waters more clearly than a simple numerical tabulation. However, these plots are chiefly illustrative, and processes suggested by the relations shown must be verified by further evidence, as pointed out by Hem (1959, p. 183). For example, a change in water composition of Deep Springs waters resulting from the precipitation of calcite cannot be resolved from changes brought about by combination of base exchange and sulfate reduction. In spite of limitations, trilinear plotting is very useful in showing systematic
differences among waters, as also indicated by Hem (1959, p. 184). Probably the chief shortcoming of the method is that the plot does not indicate total concentration. In the diagram for Deep Springs Valley waters, total concentration generally increases to the lower right of the diagram; that is, toward the alkali and chloride corners. This is not true of streamflow, however, and this factor renders the Piper diagram less satisfactory for illustrating trends in streamflow units than for illustrating the valley waters as a whole. Also, the criteria for mixtures described by Piper (1944, p. 920) cannot be fully utilized because of solid precipitation and reactions taking place in many waters.
A summary of the compositional trends in Deep
EXPLANATION
Buckhorn Springs
Wyman Creek ■
Birch Creek
v
Antelope Springs ★
Crooked Creek ▲
Central bog-mound spring
A
Northwest bog-mound spring
X
Eastern bog-mound spring
+
Corral Spring 1
□
Corral Spring 2 <>
Open sag pond Ephemeral runoff
Y
Buckhorn inflow channels 80
Corral channel
Figure 19.—Chemical composition of waters from the Deep Springs Valley which can be considered inflow to Deep Springs Lake.A32
CLOSED-BASIN INVESTIGATIONS
EXPLANATION
• DSH State Highway maintenance station
Deep Springs School irrigation well
• 43
Central valley stock well
• 44
Stock well west of Deep Springs Lake
Ca
CATIONS
EQUIVALENT PERCENT
Cl
ANIONS
Figure 20.—Composition of well waters from the Deep Springs Valley. The numbers refer to localities shown on plate 1
and the corresponding analyses of table 7.
Springs Valley waters suggested by the trilinear plot is given in figure 22.
INFLOW
As illustrated in the central part of the Piper diagram (the diamond field, fig. 19), the chemical compositions of most waters which constitute inflow to Deep Springs Lake fall within distinctive limits; these limits are based on the ratio of alkaline earth to total alkalies and on the ratio of alkalinity to sulfate plus chloride. Towards the lake area, the concentration of
alkalies, sulfate, and chloride increases simultaneously with an increase in total dissolved-solid concentration. To a lesser extent these trends may be seen on the individual cation and anion plots as well. The diamond field is the most useful for demonstrating trends in the inflow waters, fig. 19; a separate anion diagram (fig. 21) better illustrates differences in lake and ponded waters. It is interesting to compare the anion diagrams (figs. 19 and 21) with the diagram given by Hutchinson (1957, p. 566) for the general anionic composition of inflow and closed lakes in the westernHYDROLOGY AND MINERALOGY OF DEEP SPRINGS LAKE, CALIF.
A33
so4
Figure 21.—Proportions of major anions, in equivalent percent, in waters from Deep Springs Lake. The points correspond to analyses given in table 7.
Great Basin. The waters of the Deep Springs Valley follow a trend similar to that in Hutchinson’s diagram, but in the course of their evolution they show a pronounced offset toward sulfate.
Gross differences in chemical composition of inflow waters are readily apparent in the diagram (fig. 19). The waters of Birch Creek and Antelope Springs are very similar in ionic proportions, and they differ from the waters of the Wyman Creek and Crooked Creek drainages by containing higher concentrations of sulfate and magnesium. In the same way, Buckhorn Springs water can usually be separated from Corral Springs outflow. The easternmost bog-mound springs, on the other hand, are characterized by very low proportions of sulfate. Low concentrations of alkaline earth which result in higher alkali percentages serve to distinguish Crooked Creek from Wyman Creek waters. The open sag pond differs from the Buckhorn Springs by containing higher alkali and chloride percentages.
The major factors influencing gross differences in the chemical composition of inflow units are the original mineral source of the dissolved constituents and processes taking place within each hydrologic unit. Thus, the higher proportions of sulfate in Birch Creek as compared with Wyman Creek reflect the greater number of mineralized zones found within the Birch Creek drainage. Large segments of the channels of Birch Creek and its tributaries parallel contact aureoles formed in the Precambrian and Cambrian rocks by the intrusion of the granitic rock suite. (See pi. 1.) Hydrothermal veins and sulfide mineralization are closely associated with this metamorphism, especially in the drainage of the North Fork Birch Creek. Wyman Creek, on the other hand, crosses this zone at right angles, and a much smaller part of its drainage
comes in contact with such rocks. This chemical composition of Antelope Springs may reflect either an extension of mineralization beyond the immediate contact zone or, more likely, a connection with Birch Creek drainage through fractures or joints related to faulting along the west margin of the valley.
There is adequate source for the calcium, magnesium, and carbonate in the calcareous sedimentary and metamorphic rocks in both the Wyman Creek and Birch Creek basins. The somewhat higher calcium content and the still higher magnesium content of Birch
Figure 22.—Generalized trends in the composition of dissolved solids in inflow to Deep Springs Lake, based on the plot of water compositions shown in figure 19. Large arrows indicate the two major compositional trends in inflow to Deep Springs Lake. Small arrows indicate the compositional effect of organic respiration, decay, and sulfate reduction on each major trend.A34
CLOSED-BASIN INVESTIGATIONS
Creek reflect the fact that a larger area is underlain by the Eeed dolomite and its metamorphic equivalents in the Birch Creek drainage. The alkalies in both basins are primarily derived through leaching of the feldspathic and micaceous components of granitic rocks. This is shown by the low alkaline earth concentration and resulting high alkali concentration in the waters of Crooked Creek, where the drainage is principally underlain by granitic rocks.
Precipitation of calcite on the streambed of Birch Creek has resulted in a decrease in the concentration of calcium and alkalinity in the waters downstream and a concurrent decrease in total dissolved-solids concentration. Wyman Creek waters decrease slightly downstream in total concentration, but despite some calcite cementation of the bed, there is little change in percentage composition of the streamflow.
The effects of lithologic control and mineralogic source of solutes may also be noted in the compositional variation of the Buckhorn Springs water. From north to south along the spring belt, individual springs receive more drainage from areas underlain by Precambrian rocks which are at increasingly greater distance from the contact with intrusives. (See index map, fig. 1; table 7; and fig. 19; locality numbering from south to north is from DL33 to DL36.) The contact zone, a continuation of that found along Birch Creek, is exposed along the prominent canyon immediately southeast of the lake, and the effects of thermal metamorphism in the sedimentary rocks decrease steadily to the southwest. Correlatively, the proportion of alkalies in the Buckhorn Springs decreases from north to south. Total dissolved-solids concentration and sulfate concentration also decrease in the springs to the south.
The distinctive compositional differences of the bog-mound springs and Corral Springs are enhanced by secondary effects superposed on the lithologic control. The chief processes involved are the precipitation of alkaline earth carbonate, sulfate reduction, and additional solution of material as water moves through the valley deposits.
The patterns of water movement deduced from the hydrography of Deep Springs Valley suggest that waters from the Wyman-Crooked Creeks system move down the hydraulic gradient in the valley alluvium and reappear in the central and eastern bog-mound springs and Corral Springs. (See p. 81 and pi. 1.) No information is available on the flow rates of individual hydrologic units, but comparisons may be made on the assumption that the composition of Wyman Creek waters are not greatly altered between the time they enter the valley ground-water reservoir and the
time they issue from the springs. The chief difference in chemical composition between Wyman Creek streamflow and the Corral Springs water is the higher percentage of alkalies and increased total dissolved-solids concentration in the springs; anionic proportions are strikingly similar. (See fig. 19 and table 7.) Alkali concentrations range from 20 to 25 percent of the cations in the central bog-mound spring (loc. DL73), from 55 to 60 percent in the eastern bog-mound group (DL41-42) and Corral Springs 1 (DL48A), and from 90 to 95 percent in Corral Spring 2 (DL48B). Such change in composition would at first appear to result primarily from cation exchange. Montmorillonite clays which might be capable of releasing alkali and fixing calcium have been identified in the surficial valley alluvium but they probably form a minor part of the total valley fill and even in the clay fraction are subordinate to calcium-saturated types. Although ion exchange may be a significant mechanism for alkali enrichment in waters permeating fine-grained lacustrine deposits such as in the bog-mound-spring area, it is more likely that calcite precipitation balanced by additional alkali solution from granitic alluvium accounts for the compositional difference between the streams and spring outlets. Calcite cement has been recognized in several places in the valley alluvium and is most conspicuous in the fanglomerate around the north end of the Corral Springs area. The fact that total dissolved-solids concentration does not decrease in the waters associated with the calcite deposits suggests that precipitation loss is balanced by additional alkali and bicarbonate picked up as the waters percolate through the granitic alluvium and old lacustrine deposits. The obvious and abundant calcite deposits around the orifice of the Corral Springs may be associated with the loss of gaseous carbon dioxide as ground waters come to the surface.
Biotically induced sulfate reduction is another process causing differences in chemical composition between streams and springs. Good evidence of sulfate reduction is the hydrogen sulfide detected in all the bog-mound springs and also at Corral Spring 2. White fungus growth is common too. In the easternmost bog-mound springs 0.1 ppm S'2 as hydrogen sulfide was measured, and at Corral Spring 2 up to 11 ppm was determined. The easternmost bog-mound springs contain the lowest sulfate concentrations and, in addition, are notably high in K+1. In fact, were it not for the high potassium content the water of these springs would be cationically intermediate between the large central bog-mound spring and Corral Spring 1. The K+1 is probably released by hydrogen-ion reaction with micaceous clays; such a process is conceivableHYDROLOGY AND MINERALOGY OF DEEP SPRINGS LAKE, CALIF.
A35
in a strong reducing environment resulting from organic decomposition. The higher dissolved-solids concentration and alkali carbonate enrichment in water of Corral Spring 2 (fig. 2, loc. DL48B) and of the inflow to the closed sag pond (loc. DL39B) as compared with water of Corral Spring 1 (loc. DL48A) are also most likely the result of biotic activity. Where the discharge of springs is slow, marsh vegetation has taken hold and flourished. As a result, evapotranspi-ration increases total concentration, and respiration and decay keep carbon dioxide concentration high. To a lesser extent, the same processes coupled with evaporation from open water can account for an increased dissovled-solids concentration and alkali proportion in the open sag pond (fig. 2, loc. DL32) and related waters (loc. DL45) as compared with the outflow from the northern Buckhorn Springs (loc. DL36). The effects of biota and evaporative processes on water composition in the immediate spring areas is shown by comparison of samples from the central bog-mound spring taken consecutively at the orifice (loc. DL73C) and at the edge of the surrounding marsh (loc. DL73D). From the spring orifice to the marsh edge total dissolved-solids concentration of the waters nearly doubles as alkalies increase 5 percent and alkalinity increases 10 percent. (See fig. 19 and table 7.)
The lower concentration of alkali in the central bog-mound spring waters as compared with that in the eastern bog-mound group and Corral Springs (fig. 19) may be indicative of variations in the lithologic source of the waters and valley alluvium. The Corral Springs most likely receive all contributions from the small drainages on the east side of the valley which are underlain exclusively by granitic rocks. Furthermore, what little return flow there is from irrigation at Deep Springs School enters old channels on the west side of the valley floor. The central bog-mound spring, on the other hand, is probably fed through alluvial deposits which contain a small amount of material derived from sedimentary and metasedimentary strata.
Waters which by their composition can be traced to Birch Creek-Antelope Springs drainage occur only in the two western bog-mound springs. Even here, some mixture with waters derived from the Wyman-Crooked Creeks system is suggested by the composition of waters from the larger of the two springs (loc. DL78). The chemical composition of its water plots in the diamond field of the Piper diagram (fig. 19) intermediately between the most northwestern bog-mound spring (loc. DL27) and Corral Spring 1 (loc. DL48A). The difference in samples from the most northwestern bog-mound spring (table 7) to some ex-
tent reflects different points of collection; sample DL27C was obtained closer to the spring orifice. The trend from the Birch-Antelope Creeks drainage system to the western bog-mound springs is similar to the trend from Wyman Creek to the Corral Springs (fig. 22); no doubt these similar trends from similar processes.
The single sample (AS79) obtained from surface waters in Antelope Canyon shows significant cationic differences from the waters of Antelope Springs; magnesium and sodium are both about 5 percent higher (table 7). This sample most closely resembles samples from the well due west of the lake, and is probably representative of waters in contact with relatively unmetamorphosed Cambrian strata or alluvium derived therefrom. Relatively high Mg*2 content might be expected to occur in waters in contact with dolomitic argillites rather than waters in contact with the metamorphic equivalents, in which Mg42 is locked in insoluble silicates.
The composition of most well waters varies from place to place in accordance with the flow patterns suggested for the valley (pi. 1). In the northern part of the valley, well samples (fig. 23) either reflect the composition of the Wyman-Crooked Creeks system very closely (table 7, samples DSH-1 and DSS-1) or show the effect of calcite precipitation plus alkali and sulfate enrichment (sample DSH-2). Well-water samples from the central valley (samples DL43A and DL43B) unfortunately could only be obtained from the stock tank. Thus, the percentage composition reflects a rise in temperature, evaporation, loss of carbon dioxide, and the precipitation of alkaline earth carbonate as scum. The effect of these factors is also shown in a comparison of two samples from the well west of the lake area (table 7, samples DL44 and DL44B); the sample taken directly from the outflow pipe had higher calcium and bicarbonate concentrations than the sample collected from the stock tank when the well was not operating.
Variations in the chemical composition of inflow are also caused by seasonal fluctuations in discharge, source areas, temperature, and biologic activity.
Figure 23 shows the relations of constituent concentration and percentage composition to discharge for Wyman Creek. Except where there is some contribution from overland flow, total dissolved-solids concentration decreases with an increase in discharge which decreases the concentration of alkaline earths and bicarbonate; however, percentage composition remains relatively uniform. The addition of overland flow raises the amounts of calcium and bicarbonate significantly because calcareous rocks are plentiful in the upperCLOSED-BASIN INVESTIGATIONS
A36
SAMPLE NO: 47E 56B 47 47B 56A 47C
DATE COLLECTED: 8/61 9/60 10/59 5/61 6/60 8/60
DISSOLVED SOLIDS: 291 293 296 265 256 330
ESTIMATED DISCHARGE, IN CUBIC FEET PER SECOND
<
K
O
Figure 23.—Relation of major ion concentration and percentage composition to estimated discharge for Wyman Creek. Dissolved solids are given in parts per million.
parts of the drainage area where most such flow originates. Magnesium is the only constituent which shows a consistent trend; its concentration decreases as discharge increases.
No clear-cut relation of either constituent concentration or percentage composition to total dissolved solids concentration or discharge is apparent from the scanty data available for the Corral and Buckhorn Springs (figs. 24, 25 and 26), though the composition of the Buckhorn Springs water faintly suggests that calcium increases as the discharge increases. Whatever relation might exist is apparently masked by prior seasonal fluctuations, particularly in the waters feeding the springs. This masking is especially evident in the northern Buckhorn Springs. Samples collected after an early winter-type northwest storm in October 1959, in which most of the precipitation fell in the north half of the valley, showed the composition of the northern Buckhorn Springs water to be very close to that of the Corral Springs—higher bicarbonate, lower sulfate, and slightly higher alkali than usual. (See fig. 19 and table 7, sample DL36. ) This effect was not as pronounced in the Buckhorn springs to the south. These conditions were probably the result of increased flow southward along the major fault zone. The southern Buckhorn Springs were not as strongly affected because of distance and a reversed hydraulic gradient to the south (the Buckhorn Springs farthest north, loc. DL36, are the lowest points in the system).
SAMPLE NO: 48A2 48A4 48A1 48A3 48A DATE COLLECTED: 9/60 9/61 10/59 5/61 3/59 DISSOLVED SOLIDS: 267 273 284 264 273
ESTIMATED DISCHARGE, IN CUBIC FEET PER SECOND
Figure 24.—Relation of major ion concentration to estimated discharge for Corral Spring 1. Dissolved solids are in parts per million. Dashed lines suggest trend associated with higher, but unknown, discharges, such as that in March 1959.
SAMPLE NO: 48B3 48B4 48B 48B1 48B2
DATE COLLECTED:5/61 8/61 3/59 10/59 6/60
TOTAL DISSOLVED SOLIDS, IN PARTS PER MILLION
Figure 25.—Relation of major ion concentration to total dissolved solids for Corral Spring 2.HYDROLOGY AND MINERALOGY OF DEEP SPRINGS LAKE, CALIF.
A37
SAMPLE NO: 35C 35
DATE COLLECTED: 8/61 10/59 DISSOLVED SOLIDS: 436 501
33C 33B 34 34B 36 36E 36C 36D
8/61 5/61 10/59 8/61 5/61 10/59 8/61 9/60 5/61
380 387 446 424 431 504 513 517 517
EXPLANATION
• A o X
Springs 1-5 Springs6-7 Springs 8-10 Springs 11
Figure 26.—Relation of major ion concentration to estimated discharged for the Buckhom Springs. Dissolved solids are in
parts per million.
Seasonal variation in the composition of water may also be noted in the Corral Springs (figs. 24 and 25), where the composition of water affected by winter storms bears little resemblance to compositional trends of waters during the summer. Such variation is also related to seasonal differences in temperature and biotic activity.
Substantial changes in total concentration and percentage composition of waters contributed to Deep Springs Lake take place between the orifices of inflow springs and the central lake area. The course of these
changes can be traced definitely only when the waters which reach the lake remain on the land surface, as in the main inflow channels of the Corral and Buck-horn Springs. Lack of prior knowledge prevented the selection of the best sites to represent these trends in adequate detail, but general relations are given in figures 27, 28 and 29, in which the change of percentage composition with increasing total concentration is shown. In these plots, lines are drawn in a down-channel sequence and connect values for each major constituent at each site, but because of the small numb-A38
CLOSED-BASIN INVESTIGATIONS
er of sites not much significance can be attached to point-to-point variation. Futrhermore, the sampling at the sites was not time sequential.
The major cause of the compositional changes downchannel is evaporation coupled with alkaline earth carbonate precipitation and re-solution of capillary salts. Biologic, particularly floral, activity probably plays a significant catalytic role in these processes which are all accelerated as the inflow channels become less distinct and, near the lake, open onto marshy flats having very low gradients.
Compositional change in the Corral Springs inflow (fig. 27) appears to follow a regular pattern, if one assumes that the small seeps intermediate in pattern between springs 1 and 2 and Corral Spring 2 itself are in downgradient sequence as the result of a progressive increase in evaporation at each site. Precipitation of calcite, probably impure, is indicated by near
< O 00
00 CO 00 o
^ ^ LO
Figure 27.—Relation of major ion concentration to total dissolved solids for Corral Springs waters along the course of inflow to Deep Springs Lake. Sampling locations at top of graph: 48A, Corral Spring 1; 48C, Corral Spring Canal; 48B, Corral Spring 2; and 50, lower channel.
constancy of calcium and to a lesser extent magnesium. An increase of alkalinity due to evaporation is offset by carbonate precipitation and loss of carbon dioxide to the atmosphere; so, alkalinity declines as sulfate and chloride contents increase.
The fact that compositional trends shown by the main Buckhorn Springs inflow are generally the same as those in the Corral Springs sequence (compare fig. 27 with 28 and 29) suggests that the same processes are operative. In detail, however, there are much greater fluctuations, particularly in the anionic composition. The trend of each constituent during different time and seasonal intervals is roughly parallel but displaced relative to total dissolved-solids concentration. The same is true of the downchannel rise in pH (fig. 30). This displacement occurs because the higher temperatures and increased floral activity in summer bring about a loss of carbon dioxide, rise in pH, and precipitation of alkaline earth carbonate at lower levels of total dissolved-solids concentration. The cationic trends suggest that seasonal factors are especially important in the precipitation of a magnesium-bearing carbonate phase (dolomite). Such seasonal effects are somewhat similar to those described by Alderman and Skinner (1957) in their explanation of dolomite precipitation in the lagoons of southeastern Australia.
In summary, the chemical composition of most waters from the Deep Springs Valley falls into distinct units according to the ratios of alkaline earths to alkalies and of alkalinity to sulfate plus chloride—the basis of the Piper diagram (fig. 19). These units include the Wyman-Crooked Creeks system, Birch Creek plus Antelope Springs, the eastern, central, and western bog-mound springs, the Corral Springs 1 and 2, the Buckhorn Springs, and the open sag pond. In the course of movement from the mountains to the lake, the water composition is initially controlled by bedrock lithology and is subsequently influenced by other factors. If cations or anions are considered individually, compositional change follows a single trend from the drainage divides to the lake area; the percentages of alkalies and sulfate plus chloride generally increase as the total dissolved solids concentration increases. However, the combination of cations and anions suggests two major compositional trends in the inflow to Deep Springs Lake (fig. 22). One group of waters is predominantly associated with sedimentary lithologies and is characterized by higher concentrations of sulfate. This group includes Birch Creek, Antelope Springs, the northwest bog-mound springs, and Buckhorn Springs, and the open sag pond. The other group is predominantly associated with igneous rocks and includes the Wyman-Crooked Creeks system, the eastHYDROLOGY AND MINERALOGY OF DEEP SPRINGS LAKE, CALIF.
A39
Figure 28.—Relation of major cation concentration to total dissolved solids for Buckhom Springs waters along the course of inflow to Deep Springs Lake. Though collected in June 1960, samples DL54A and DL54B are included with samples collected in August 1961 because hydrologic conditions were very similar.
SAMPLE NO:
o uj O Q LU < CO < < CO
ro ro 00 CO 00 00 ■vf o
coco coco coro 00 00 00 00 LO
00
o
<3-
o
o
<3-
DISSOLVED SOLIDS, IN PARTS PER MILLION
Figure 29.—Relation of major anion concentration to total dissolved solids for Buckhom Springs waters along the course of inflow to Deep Springs Lake. Though collected in June 1960, samples DL54A and DL54B are included with samples collected in August 1961 because hydrologic conditions were very similar.A40
CLOSED-BASIN INVESTIGATIONS
o
SAMPLE NO: fo
*1
uj O Cl
in ld rv oo
ro co co co co
"1 UJ < CD < < *1 CO
00 co 00 CO 00 CO in o m
CO
o
"1
CJ
o
Figure 30.—Relation of pH to total dissolved solids in Buckhorn Springs waters along the course of inflow to Deep Springs Lake. Though collected in June 1960, samples DL54A and DL54B are included with the August 1961 sequence because hydrologic conditions were nearly identical.
and central bog-mound springs, and Corral Springs. The two groups have compositional trends which are distinct but roughly parallel up to the area adjacent to the play a, where they begin to converge (fig. 22). The compositional trends further corroborate the flow patterns worked out for the valley (p. A20). Compositional change in inflow units is brought about by precipitation of alkaline earth carbonate, sulfate reduction, additional alkali solution, and evaporation.
Some variation in stream composition is associated with changes in the amount of discharge, but spring composition is more affected by prior seasonal variation in waters feeding the springs. Compositional changes in all waters may be correlated with seasonal changes in temperature and biologic activity.
LAKE AREA
The chemical composition of waters from Deep Springs Lake varies through wide extremes dependent on stage and specific influences at the point of collection. Lake waters become more uniform in composition as the stage rises but at low stages the waters become physically separated, as previously described (p. A23), and their composition is more dependent on their immediate environment. Surface waters entering on the playa, especially at low stage, often become stagnant before mixing and retain chemical characteristics
related to a particular inflow source. Thus, stagnant waters around the immediate inflow area of the Corral Springs may be distinguished from ponded waters near the mouth of the Buckhorn channels by higher ratios of alkalinity to sulfate. (See fig. 19; compare samples DL50 and DL54, table 7.) The most extreme examples of such differences are in puddles and ponds resulting from heavy local precipitation and runoff. (See fig. 31.) These waters reflect the most abundant and soluble surficial materials in the immediate area. Runoff from the carbonate-rich interdunal flats on the north side of the lake area was found to carry more than 80 equivalent percent bicarbonate (fig. 19; table 7, DL28), whereas ponded water from the northwest corner of the playa where there is much halite in the efflorescent crusts contained nearly 70 equivalent percent chloride (table 7, sample DL61). Interstitial solutions and surface pools derived from ground-water inflow are also closely related to their source (table 7, sample DL3A). Brines around the north margin of the central lake area are influenced by the low sulfate content of the eastern bog-mound springs.
The most distinctive waters associated with specific playa features are found around gas vents (fig. 32). These vents are roughly conical in shape and extend into the underlying mud. They are most common inHYDROLOGY AND MINERALOGY OF DEEP SPRINGS LAKE, CALIF.
A41
the northern part of the central lake area and are associated with small pools even at lowest lake stage. Gaseous activity is seldom apparent in most of the vents, though one or two vents near the north limits of the leveed area may show vigorous bubbling. The gases are mostly air and carbon dioxide, but hydrogen sulfide and methane were also detected. The waters associated with these vents (fig. 21; table 7, samples DL29A, 30A, and 30E) are characteristically high in bicarbonate. They may also be high in potassium and chloride, though concentrations are % to %o that of the perennial brine. These waters were probably once similar to those of the eastern bog-mound springs, but they have since picked up chloride by seepage in the northern lake area, which is characteristically high in chloride and in which are located the halite “beaches” described on p. A46. Samples of intercrustal brine from the northern part of the trough are also high in Cl'1 (table 7, sample DL3B).
Despite considerable variation in absolute quantities, equivalent percentages of the major anions in waters from the central lake area fall within narrow ranges
Figure 31.—Puddle of ephemeral runoff left behind on in-terdunal flats just north of Deep Springs Lake. Suspended material is almost entirely alkaline earth carbonate. Knife gives scale.
Figure 32.—Typical small gas vent on the north side of Deep Springs Lake saltpan. Gas bubbles were not present at the time the photograph was taken but appeared later. The surrounding crust was less than 2 inches thick. Hunting knife gives scale.
(fig. 21). All samples of surface water taken from the main lake area, regardless of sampling point, contained less than 14 equivalent percent carbonate species (less than 10 equivalent percent if associated with a lake stage in excess of 0.1 foot). In contrast, all samples of intercrustal brine had higher bicarbonate contents and ranged from 14 to 20 equivalent percent in total alkalinity (such values would be changed only 2 or 3 equivalent percent with errors up to 20 percent in the alkalinity determination). Ratios of S04:C1 generally are slightly higher in brines from the southern part of the lake area, but variations are less than 25 equivalent percent of either constituent in intercrustal brines regardless of the sampling point. Two samples of interstitial brines from shallow muds immediately outside the leveed area (table 7, samples DL29C, and DL51) did not differ significantly from the central area brines.
In the absence of almost any detectable alkaline earth, cation variation of the brines can be expressed primarily on the basis of the relative amounts of sodium and potassium. There appears to be little areal variation of Na :K ratios in the intercrustal brines of the central lake area, though the ratios (3.5-7.7) are consistently very low for natural solutions (even compared to such a brine as that from Owens Lake, 13 +). When considered in the simplified reciprocal system Na-K-S04-Cl (fig. 33), the intercrustal brines show three times as much variation in the S04 :C1 as in the Na:K ratios. This indicates that potassium does not precipitate solely as the chloride but also as a sulfateA42
CLOSED-BASIN INVESTIGATIONS
FiotTBE 33.—Composition of intercrustal brines from Deep Springs Lake in terms of the reciprocal system Na2S04-KaSOi-NaCl-KCl. Points represent analyses given in table 7.
salt (identified as aphthitalite, K3Na(S04)2)- Slightly higher K+1 contents may be associated with brines from the trough, especially on the north side.
Significant variations in composition of the surface waters at Deep Springs Lake are associated with changes in stage. (See table 7, and fig. 34.) Rapid rises in stage resulting from rainfall directly on the lake area are characterized by S04:C1 ratios in the waters which are nearly 10 times as great as those in intercrustal brines. This change in the ratios reflects the quantitative dominance of sulfate over chloride in the saline crusts. The dominance of sulfate is also shown in changes in the SO* :C1 ratios during recession of seasonal high stages. After a rise in stage, the S04 :C1 ratios are initially somewhat similar to intercrustal brine, but continued re-solution of sulfate after most of the chloride has been redissolved brings about an increase in the S04: Cl ratios until saturation with
sodium sulfate is reached. Precipitation of the sulfate temporarily reduces the ratio again, but continued desiccation eventually leads to even higher values (fig. 34). The relatively low S04 :C1 ratios of intercrustal brine reflect higher chloride solubility, the formation of additional sulfate salts prior to chloride precipitation, and the bacterial reduction of sulfate in the brines. Substantial sulfide has been detected in all intercrustal brines analyzed.
In summary, waters from the Deep Springs Lake playa itself may be categorized as marginal inflow (ephemeral or perennial), surface waters of the main lake, or intercrustal brines. Marginal inflow is diverse in composition, depending on its source, but surface waters and brines show consistent compositional traits; intercrustal brines are higher in alkalinity and have a lower S04 :C1 ratio than any surface waters of measurable stage.
CLOSED SAG POND
Variation in the percentage composition of waters (table 7, samples DL39A and DL39C-G) from the closed sag pond is limited almost exclusively to the S04 :C1 ratio. Alkalinity varies less than 3 equivalent percent in all samples analyzed and is about 30 percent higher than the waters of the main lake. Available data indicate that the S04:C1 ratio remains at the level found in the inflow seepage (table 7, sample DL39B) until saturation with sodium sulfate is reached, and then the ratios shifts to lower values.
STAGE, IN FEET
Figure 34.—Relation of lake stage to SOi: 01 ratio for surface waters at Deep Springs Lake.HYDROLOGY AND MINERALOGY OF DEEP SPRINGS LAKE, CALIF.
A43
CALCULATION OF SALINITY FROM HYDROGRAPHIC
DATA
From data compiled from 24 closed lakes in all parts of the world, Langbein (1961, p. 1&—14) has derived an empirical equation for salinity based solely on hydrographic factors. Computations of solution concentrations for various stages of Deep Springs Lake utilizing this relation departed from actual measurements irregularly and by large amounts. However, only two or three of Langbein’s examples could be considered playa lakes having response time of about 1 year or less. The rapid fluctuations, shallow depths, and high concentrations of water bodies such as Deep Springs Lake increase the importance of nonhydrographic factors, such as solution kinetics, relative salt solubilities, thickness of exposed salt crust, and residual brines. At Deep Springs Lake the thickness of salt crust exceeds the mean depth, and the major constituent of that crust—thenardite (sodium sulfate)—may not dissolve in appreciable quantities until the stage has exceeded a quarter of the highest annual level rise. The initial surge of fresh inflow is expended in the dilution of residual brine and the solution of surficial salts. A minimum estimate of the total quantity of salts available for solution can be derived by considering the salt crust as a spherical segment having a maximum thickness of 1 foot and a minimum diameter of 1 mile. If a minimum specific gravity of 2.0 is assumed, the crustal salts at Deep Springs Lake would exceed 106 tons. This quantity is more than could have been dissolved at the highest lake stage observed during the entire course of this study.
MINERALOGY
The deposits of Deep Springs Lake contain a rather wide variety of mineral constituents. Inasmuch as most of the material is very fine grained, it was seldom possible to precisely identify even the major mineral species in the field. Thus, samples were collected for detailed laboratory examination from more than 100 points in the playa area at or close to the localities shown in figure 2. Sample points were generally selected with reference to the major surficial features of the playa; location was in large measure controlled by obvious changes in texture or color. One great limitation on sampling was the inaccessibility of substantial parts of the playa which, especially within the leveed area, could not be reached with safety. At most points, particular attention was given to the deposits on or near the surface which might be related to present hydrologic conditions. Several replicate samples were collected in areas subject to seasonal hydrologic fluctuation.
Surficial materials were collected by cutting out small sections with a hunting knife or trowel. Where direct observation of material at depth was desirable and possible, pits were manually dug. Otherwise, samples from depths as much as 5 feet were obtained by a plunger-type hand auger having a 1-inch-diameter core barrel. Some samples were also examined from the 28-foot auger hole drilled through the levee on the north side of the lake near the level staff. Samples collected in the field were packed in two to three layers of heavy polyethylene to prevent drying and then placed in 1-pint cartons.
CARBONATE AND SALINE MINERALOGY
The saline minerals in the samples from Deep Springs Lake were identified principally by X-ray powder diffraction, utilizing a Norelco recording geiger-counter diffractometer. The instrumental techniques employed were virtually those of Zen (1957, p. 890). Small portions of all samples were first ground fine and analyzed in bulk. Both aluminum holders and glass slides were used for mounting. When possible, a part of the material was X-rayed in its original wet state and compared with another portion of the same sample ground in acetone. Where distinction was doubtful between reflections belonging to precipitate or detrital minerals, the samples were treated with distilled water and (or) dilute acid and reexamined.
Several attempts were made to separate the individual saline mineral species occurring in any single sample. Spot tests using several compounds that characteristically react with various salts were unsuccessful. Centrifuging in liquids of known specific gravity was also largely unsuccessful. Some alkali carbonate, however, was separated from saline crusts by centrifuging at high speed. In many samples, specific layers, spots, or individual mineral grains were picked out under the binocular microscope and X-rayed separately. In addition, most samples were examined under the petrographic microscope to provide textural information and some check on the X-ray results; however, examination was severely limited by the extremely fine-grained nature of most of the material. In the coarsely crystalline samples, some minerals were separated by disaggregation and sieving into selected size fractions. A few thin sections cut from material impregnated with wax by the method of Tourtelot (1961) were examined.
The carbonate and saline minerals and the general type of deposit in which they occur are given in approximate order of abundance in table 8. Identification is based on the comparison of bulk-sample X-ray pat-A44
CLOSED-BASIN INVESTIGATIONS
terns with standards and on an optical check of most of the samples.
The detailed mineralogy of the Deep Springs Lake deposits is strongly dependent on the hydrography of the area. The distribution and character of the carbonate and saline minerals are most readily discussed in terms of three distinct patterns found in the saline deposits: (1) relatively regular areal zoning of precipitate minerals in the surficial (top 2+ ft) lacustrine deposits of the entire playa, (2) layer variation in mineralogy of lake-deposited saline crusts, and (3) highly irregular and local mineralogical variation in capillary efflorescences.
Table 8.—Precipitate mineral species at Deep Springs Lake [Minerals are listed in approximate order of abundance]
Mineral Occurrence
Dolomite, CaMg(COj)2................. Mud.
Calcite, aragonite, CaCOs............ Do.
Thenardite, NajSCL___________________Mud, saline crust, and efflorescence.
Halite, NaCl......................... Do.
Gaylussite, NajCa(CC>3)i-5HjO_.......Mud.
Burkeite, Na«(S04)jC03............... Mud, saline crust, and efflorescence.
Trona, Na3H(C03)3-2Hj0...............Saline crust and efflorescence.
Aphthitalite, KsNaCSOO*-------------- Do.
Pirssonite, Na2Ca(C03)2-2H30......... Efflorescence.
Nahcolite, NaHC03--.................. Mud.
Thermonatrite, Na2C03-H20..........Efflorescence.
Glauberite, Na2Ca(S04)2.............. Do.
Sylvite, KC1.......................... Saline crust and efflorescence.
(?)Analcite, NaAlSi206-H20...........Saline crust.
(?)Bloedite, Na2Mg(S04)s............. Do.
AREAL DISTRIBUTION IN THE LACUSTRINE DEPOSITS
Hunt (1960) has delineated mappable zones in the saltpan at Death Valley based on dominant anion composition and related to the sequence of precipitates formed on evaporation of an average brine. Similarly, the zoning of precipitate minerals in the deposits at Deep Springs Lake can be utilized in areal mapping of the playa (fig. 2).
Except for calcite and dolomite, zonal boundaries are drawn along the outer limit of key precipitate minerals in the surficial playa deposits. Calcite-arag-onite and dolomite zones are based on the relative amounts of these minerals. The sequence of key minerals from playa margin to center is calcite and (or) aragonite, dolomite, gaylussite, thenardite, and burkeite. Except for aragonite and gaylussite, all the key minerals persist from the outer limits of their occurrence to the center of the saltpan. No aragonite could positively be identified in deposits containing thenardite. Gaylussite crystals persist only into the outer reaches of the thenardite zone.
No zone based on the relative amount of halite can be established, inasmuch as halite crystallizes from evaporating waters over a wide area under a variety of conditions. Other saline minerals occur irregularly
or in response to conditions other than fluctuating lake levels.
Calcite and (or) aragonite strongly dominate over dolomite only on the playa’s far west side. Typically, the surficial sediments of these areas contain a high percentage of detrital material. Most of the carbonate is present as an extremely fine-grained aggregate. A few relatively large anhedral grains suggest that at least some of the calcite is detrital in origin, but such occurrences are confined to the marginal channel-scarred parts of the playa which most frequently receive surface runoff from the surrounding area. In the borderlands immediately outside the playa, the in-terdunal flats are characteristically covered by white to gray sediment which has a shiny smooth surface. This surficial sediment remains in milky suspension for a considerable time in the standing puddles which accumulate. This sediment is composed chiefly of calcite but includes some aragonite, dolomite, and minor amounts of silicate clay. Such material has a physical appearance very similar to the sediments, described by Alderman and Skinner (1957), from the shallow saline interdunal basins of southeast Australia (fig. 34).
Dolomite is the dominant precipitate mineral constituent in the gray-green and black muds of the lake area. Though dolomite may make up more than 70 percent of some near-surface muds, the amount of dolomite relative to calcite and clay generally decreased with depth. In most samples, the dolomite is so fine grained that individual grains can not be resolved with very high magnification, but a few tiny euhedral rhombs were observed. The dolomite-clay aggregate is commonly coated with hydrous iron oxide.
It is difficult to establish a precise boundary that separates areas containing mostly dolomite from those containing mostly calcium carbonate. This boundary is closely dependent on the microrelief of the playa surface and apparently lies near the highest shoreline reached by the ephemeral lake during an annual climatic cycle. (See highest contour in fig. 12.) Precipitation of dolomite directly from solution is suggested by the occurrence of euhedral grains and the distribution of the dolomite; muds within areas of relatively frequent flooding by lake waters contain the highest percentage of dolomite. Bulk-sample X-ray data on the dolomite muds were obscured by the presence of substantial clay impurity, but variation in spacings, broad lines, and weak order reflections in many samples suggested the presence of primary calcic protodolomite (Goldsmith and Graf, 1958), though detrital dolomite was identified in sediment carried from Antelope Canyon to the lake by flash flood. More detailed X-ray andHYDROLOGY AND MINERALOGY OF DEEP SPRINGS LAKE, CALIF.
A45
isotopic studies by Peterson, Bien, and Berner (1963) have since shown that much of the dolomite is definitely of recent primary origin.
Near the west-central part of the playa in the area of broad mud flats, gaylussite occurs as clear euhedral flattened wedge-shaped crystals disseminated in a carbonate-clay matrix that is mostly dolomite but that also contains calcite and white lamina of aragonite. The gaylussite crystals are fairly uniform in size and average 1 to 2 mm in maximum dimension; they are present only in permanently wet mud. They are oriented with long axes approximately parallel to traces of stratification and appear to interrupt such traces. Furthermore, their absence within aragonite laminae and their association with calcite suggest that the gaylussite formed by reaction of calcite with sodium-rich solutions. This reaction may have occurred at the bottom of the desiccating lake or diagenetically after burial. The distribution of the crystals within the mud varies considerably; in places the crystals make up nearly 35 percent of the total sediment, whereas a short distance away they may be entirely absent.
Examination of the two or three deeper core sections (Iocs. DL2, 62, and 65) in the central playa suggests that “varved” strata are characteristic of the older lacustrine carbonate muds of Deep Springs Lake. The “varves” are formed by the alternation of white aragonite lamina with darker colored layers composed primarily of dolomite and calcite in varying proportions, plus a little clastic quartz and 10-A clay. Megascopic gaylussite crystals are found only in the darker layers. The “varved” strata are very similar in appearance to some of the mud layers in the deposits of Searles Lake as described by Smith and Haines (1964) and may be the result of a true annual cycle of lake-level rise and all. Aragonite may be precipitated from relatively fresh spring inflow to the lake (probably stratified at high stages) whereas calcite is formed in more saline waters which result from late summer desiccation; this idea is supported by the absence of aragonite and the presence of calcite in association with thenardite at Deep Springs Lake. In the Searles Lake occurrences, the darker layers often contain considerable microscopic gaylussite, dolomite, and very little or no calcite. The dominance of calcite over gaylussite in the more recent dolomitic deposits of Deep Springs Lake suggests that the formation of gaylussite is continuous but very slow.
The playa immediately west and south of the levee consists of thin nearly monomineralic thenardite crust overlying dolomitic mud. Enclosed in the mud are individual crystals of thenardite. The proportion of crystals decreases abruptly with depth. Textures sug-
Figure 35.—View directly down on salt crust at Deep Springs
Lake. Intricate pattern is formed by network of acicular
mirabilite crystals now altered to powdery thenardite.
Knife gives scale.
gest that the very fine grained dolomite-clay aggregate has been forced aside by the growth of the thenardite crystals.
Thenardite is also the major constituent of the saline crusts which cover nearly the whole eastside of the playa. The upper surface of many crusts is covered with a network of skeletal powdery mirabilite (Na2SO4-10H2O) crystals altered to thenardite (fig. 35). The mirabilite forms with the solution and reprecipitation of sulfate as a result of fluctuation in lake levels. Subsequent alteration to thenardite takes place as the temperature rises or as evaporation and concentration of lake waters progress.
Inside the area enclosed by the levee, the layered crust of massive thenardite overlying brown and green dolomitic muds thickens abruptly and becomes brilliant white. As the crust becomes thicker, the mud underneath becomes darker and more fluid. In the central saltpan the average total thickness of saline layers exceeds 1 foot, and the crust is tan, pink, or red. Where the saline crusts exceed 2 or 3 inches in thickness, burkeite becomes a persistent component of the upper crustal layers. Thus, burkeite may be considered the characteristic mineral of the saltpan.
SALINE CRUSTS
Distinctive crustal layering may be detected in the saline crusts at Deep Springs Lake wherever the thickness exceeds about one-half inch. Although individual mineral species can seldom be identified megascopical-ly, differences in mineral assemblage can be roughly discerned by variations in the texture of crustal layers. Layering becomes more distinctive as thickness increases.
The layering of the average saline crust is based on variations in texture and mineralogy and usually falls into a three- to five-fold sequence. Local mineral-A46
CLOSED-BASEST INVESTIGATIONS
ogic differences are common, but in a typical specimen of thick crust, the topmost layer is very thin, white or clear, highly discontinuous, and confined to the center of a crustal polygon surface. This layer is composed chiefly of halite, commonly in skeletal hoppershaped crystals. A similar delicately thin layer at places coats vestigal pools of brine on crustal surfaces. In contrast, on the north and east sides of the leveed area, cuspate beachlike deposits of halite crystal mush accumulate on top of layered crust. These deposits are usually brilliant white when dry, and commonly tan when wet. They are formed through redistribution of the uppermost crustal layers by shallow, wind-driven, surface waters saturated with sodium chloride. As a result, a distinct surface layer of halite is commonly missing from the thick crusts in the southern part of the saltpan, where normally there is some surface water derived from the Buckhom Springs.
Other minerals of the uppermost layer representing the final stages of precipitation from the lake waters include sylvite, bloedite(?), and analcite(?). The syl-vite is present mostly as minute cubic inclusions in halite. Analcite has been identified interstitially in tiny icositetrahedrons, but its precipitate origin has not been definitely established.
Rarely, distinct, but very thin patchy lenses of trona occur at the base of the uppermost layer. More commonly, myriad trona needles occur as mats included in halite—as if the trona precipitated just prior to crystallization of halite.
Beneath the surface layer of the thicker crusts are white to pink flaky highly porous gently undulatory laminae usually dominated by burkeite. These laminae are at some places very thin or absent, especially in the crusts outside the central saltpan. Burkeite is commonly present only in traces in other layers. The burkeite is usually massive, anhedral, or tabular, but it may be faintly prismatic or spherulitic (probably due to twinning). (See Foshag, 1935, p. 52-53.) The material is clouded by myriad inclusions, both solid and liquid. Rather than distinct layers, burkeite may also form little clusters surrounding thenardite crystals. Optically, burkeite is most readily distinguished from thenardite by higher birefringence; both minerals are commonly tabular parallel to (110) or (111). The refractive indices of burkeite from Deep Springs are virtually the same as those given by Foshag (1935, p. 54) for material from Searles Lake, but the 2V is slightly lower (approximately 30° as compared with 34°).
Underlying the flaky laminae are massive relatively coarse-grained layers of thenardite which compose more than 75 percent of the total crustal thickness.
For the most part, these layers are strikingly uniform in thickness, but in the area of the central saltpan they may be somewhat irregular and possess cavities a few inches across. Single layers have a maximum thickness of about 5 inches. Dry layers are usually dull white. Wet layers may be tan or, more commonly, pink, red, or deep purple. These colors result from algal or bacterial colonies which thrive in the interstitial brine. Surfaces around cavities are commonly darkest colored and colloform or “warty” in texture. The thickest layers are nearly 100 percent thenardite, but aphthitalite is a common minor accessory, and burkeite may be found in traces toward the top. Thenardite habit may differ between layers; most of the thenardite is massive anhedral, but some is dypyramidal or, rarely, even prismatic in form.
The bottom layer of saline crust commonly exhibits great irregularity, especially in the central saltpan area, where the layer appears as a group of crystal clusters projecting downward into the underlying mud. Thickness of the layer may thus range from 2 to 10 inches in a single sample. The crystals composing the clusters are generally dipyramidal; thenardite is usually the sole mineral species, although traces of aphthitalite are also present in the thicker masses. Saline crusts less than 2 inches thick have irregular bottoms, though the irregularity is not great.
The individual disseminated thenardite crystals previously noted in gray-green dolomitic muds near the levee are not found in the fetid black ooze immediately underlying the thick crusts in the central area of the lake. The only salt mineral identified in this material was nahcolite at one or two localities.
Thus, the layer sequence of saline minerals upward from the carbonate muds in the central area of the lake is nahcolite, thenardite, burkeite, trona, halite, and sylvite(?). Small quantities of aphthitalite are persistent throughout the thicker crustal layers.
A definite mineralogical sequence can also be found in the crust which forms on the small closed sag pond of the fault zone at low stage. Although the layer sequence of minerals is the same in the sag pond as it is in the main lake crust, relative quantities of the minerals are markedly different. In the sag pond, the thin surface layer dominated by halite is underlain by delicate porous somewhat granular orange laminae consisting largely of trona; burkeite occurs in the lower layers. These laminae may account for more than 50 percent of the crustal thickness. Distinct thenardite layers are present, but large crystal clusters are not found on the bottom of these layers. Nahcolite, associated with some gaylussite, is common in the under-HYDROLOGY AND MINERALOGY OF DEEP SPRINGS LAKE, CALIF.
A47
lying mud. Aphthitalite is confined to the upper parts of the thenardite layer.
The structure of the saline crusts in the central saltpan appears dependent largely on crystallization and growth in the thenardite layers. As is apparent in aerial photographs (fig. 3), the crusts are broken into rough polygonal units, usually crudely hexagonal, which range in size from 5 feet across where first discernible outside the levee the more than 100 feet across in the center of the lake. These polygons all have distinct raised edges which are commonly thrust over one another. Crustal fragments show little or no growth on their outer margins that is not clearly secondary and the result of capillary draw along the already raised edge. Indeed, the outer margins of thrusted fragments on adjacent polygonal units may commonly be fitted almost exactly. This suggests than the main force which causes the raising and thrusting of the polygon edges comes from crystallization and resulting expansion near the center of the polygonal unit and beneath the thrusted layer. This relation may be seen in a section of thick crust (fig. 36). The thenardite layers generally dip toward the center of the crustal ploygon, and all layers thicken toward the center of the polygonal unit. One of the upper thenardite layers usually extends the farthest laterally to form the ploygonal edge. Though crustal porosity is extremely variable, the upper thenardite layer is usually rather massive as a result of frequent re-solution and recrystallization. Thus, additional crystallization in intralayer pore space would probably force crustal growth and expansion to take place.
The width of the raised edge seems to be roughly proportional to the overall size of crustal polygons.
Polygons about 5 feet across have an edge 1 to 2 inches wide, whereas 15-foot polygons may have edges 4 to 5 inches wide. Larger polygons were inaccessible to detailed measurement. The relation is probably not regular, however, because of variation in thickness and in number of layers plus the effects of frequent resolution and redistribution by surface inflow.
A secondary crystal growth is common beneath the raised edge of a crustal polygon, as well as on the outer margin of the edge fragments. This material is dominated by halite but includes minor burkeite, sylvite, and aphthitalite. These minerals form through capillary draw and complete evaporation of underlying brine in the crustal cracks.
Two specific kinds of features of the central saltpan area are worth special attention.
1. The man-made trough adjacent to the levee is covered by a layered saline crust nearly identical in composition and structure with the thick crusts near the center of the playa. This crust averages about 3 to 5 inches thick and is underlain by fluid black to dark-green dolomitic mud. However, about 8 inches below the trough surface is another saline layer about 2 inches thick composed of coarsely crystalline thenardite and minor amounts of halite. A third layer of the same composition about 3 inches thick occurs about 12 inches below the trough surface and is separated from the layer above by the same highly fluid black mud. These subsurface saline layers have probably formed through alternate partial filling and evaporation of waters in the trough until the present playa level was reached. Interlayer muds were probably washed in or precipitated during high stages.
Figure 36.—Section of representative salt crust polygon from the saltpan at Deep Springs Lake. The most complete mineral
sequence within the crustal layers is shown.A48
CLOSED-BASIN INVESTIGATIONS
2. Other features of mineralogical interest are associated with the circular gas vents (fig. 32) in the areas marginal to the central saltpan. Only the smaller vents were accessible for detailed observation ; these are as much as several inches in diameter and as previously mentioned (p. A41) are roughly conical. These vents are lined with black to dark-green dolomitic mud and are surrounded by salt crust of variable thickness. Gas bubbles vigorously from the smallest vents, which occur north of the central saltpan, but in the thick crusts, the vents appear to show little, if any, gaseous activity. Even when the lake is near complete desiccation, the vents are usually covered by pools of highly colored brine and seem to be surface outlets for subcrustal brine. Saline crusts adjacent to the vents (fig. 32) are usually orange and yellow. Mineralogically, the sequence is the same as elsewhere, except that trona is more abundant in the upper layers, and traces of nahcolite may be present.
In spite of the almost monomineralic composition of the thin thenardite crusts outside the central saltpan, these deposits are complex with regard to structure, texture, and color. Much of this complexity is due to the rapid fluctuations of water levels in this area. Frequent level changes leave their mark on the surface of the saline crust. When submerged beneath a thin sheet of highly mineralized water, the crustal surface is smooth. As surface waters retreat, recrystallization forms an irregular surface of new material. Mean temperature during a particular cycle determines whether the crust will consist of a layer of thenardite or a network of long prismatic mirabilite crystals (fig. 35). Further desiccation promotes the fracturing of the crust into large polygonal patterns. The raising and thrusting of the polygon edges isolates any remaining surface water into pockets for final evaporation. As desiccation and crystallization proceed, smaller polygons form within the larger features. As additional cracks form in the crustal surface, brine trapped underneath is drawn to the surface by capillary action, and ridges of halite form along the fractures. As desiccation proceeds to the underlying mud, further polygonal subdivision occurs and the halite ridges grow large, irregular, and very hard. The entire sequence, as just outlined, may be readily observed in the flats immediately south of the leveed area. Notable differences in color commonly reflect the amount of interstitial brine in wet crusts or underlying mud.
EFFLORESCENT CRUSTS
Capillary efflorescence is primarily responsible for salt crusts found on the west third of the playa and
around inflow channels outside the playa area proper. The mineral assemblages in such deposits changed as seasonal variations in temperature and hydrology took place. Any zoning is subject to local variation.
Outside the playa, to the west and south, the saline efflorescences consist largely of halite mixed with small quantities of alkaline earth carbonates in thin white coatings on clays and silts. In the spring areas to the north and east of the lake, surfaces efflorescences are commonly dark in color (so-called black alkali) and commonly contain small amounts of pirssonite. Pirs-sonite also occurs in the efflorescences on the playa not far from the west edge. Nearer the saltpan, thenardite becomes a common constitutent of the efflorescent crust. In the playa area marked by alluvial channels, the surface efflorescences found on the channel bottoms contrast with those on the divides. The surficial materials on channel bottoms are soft, powdery, and mixed with underlying mud; the divides are covered with very hard, silty, and irregularly ridged efflorescence. Thenardite dominates in the channel bottoms; halite, on the divides. In the admixed mud, dolomite is prevalent in the channels, and calcite, on the divides. Where the channels coalesce and on the mud flats outside the leveed area, the efflorescences are chiefly thenardite and commonly contain small amounts of glauberite and trona.
Efflorescences dominated by halite or thenardite can be distinguished roughly throughout the lake area on the basis of texture. Thenardite crusts are usually soft and powdery. Halite forms hard ridges or large “popcorn balls” in a random pattern. Where the two minerals are present in nearly equivalent amounts, ridges high in halite are separated by smooth platy material containing mostly thenardite.
Efflorescences near springs or the outflow channels from springs reflect the nearby water composition. Seepage channels from the bog-mound springs on the north side of the lake commonly support soft crusts composed of trona, burkeite, halite, aphthitalite, and thermonatrite. The salt forming near channels and around roots near the Buckhorn Springs is mostly thenardite but also includes some halite and traces of aphthitalite. Capillary salts around seeps feeding the closed sag pond include trona, thenardite, aphthitalite, and thermonatrite.
SUMMARY OF CARBONATE AND SALINE MINERALOGY
There are three distinct patterns in the carbonate and saline mineralogy of Deep Springs Lake which include: (1) The areal zoning of precipitate minerals, (2) the layer variation in saline crusts, and (3) theHYDROLOGY AND MINERALOGY OF DEEP SPRINGS LAKE, CALIF.
A49
local variation in capillary efflorescences. The minerals, their composition, and the type of deposit in which they occur are given in table 8. The areal sequence of mineral zones from playa margin to center is calcite and (or) aragonite, dolomite, gaylussite, thenardite and burkeite.
Calcite, aragonite, and dolomite are the major components of Deep Springs Lake muds which contain up to 80 percent material of precipitate origin. The dolomite is, at least in part, primary. Some of the carbonate muds which lie 3 feet or so below the top appear “varved” because of light-colored aragonite laminae alternating with darker layers containing calcite and dolomite.
The saline crusts of the Deep Springs Lake saltpan are composed primarily of thenardite (sodium sulfate), but other saline minerals are contained in specific crustal layers. The layer sequence of crustal minerals is, in ascending order, nahcolite, thenardite, burkeite, trona, and halite. Burkeite is a characteristic and abundant mineral in the thick crusts of the central saltpan. Aphthitalite also persists in small amounts in the thick layers. The crusts formed on the closed sag pond show a layer sequence similar to the crusts in the Deep Springs Lake saltpan, but trona (sodium sesquicarbonate) dominates rather than thenardite. Crustal expansion appears to take place predominantly by growth in intralayer pore space. A complete sequence in surface forms is associated with the desiccation saline crust. In the central saltpan, somewhat more sodium carbonate is common in the crusts near gas vents, and additional thenardite layers are present at depth in the trough. Capillary efflorescences are characteristic of the surficial deposits of the western part of the playa. These deposits are dominated by halite and thenardite; the most abundant of these salts is locally indicated by texture. Pirssonite, glauberite, and thermonatrite are restricted to such deposits. Efflorescences around specific inflow areas reflect the anion distribution in the associated inflow waters.
SILICATE MINERALOGY
Silicates compose from nearly 100 percent of the lacustrine deposits well outside the margins of the present playa area to less than 20 percent of the muds underlying the crusts of the central saltpan. The minerals are primarily quartz, feldspar, and clays and traces of detrital amphibole. All but the amphibole can be found even in submicron-size fractions.
Although feldspar was identified in most lacustrine materials, little attempt was made to subdivide the species inasmuch as very fine grain size and interference of other phases commonly obscured characteristic
X -ray reflections. However, much of the feldspar is apparently albite or sodic plagioclase and is probably detrital.
Generally, quartz and 10-A clay minerals are the dominant silicates in the lacustrine deposits. The assemblage quartz, dioctahedral and trioctahedral mica plus minor chlorite, talc, and expandable lattice clays is common throughout the Deep Springs Lake area. The muds of the central saltpan contain a very well ordered mica-type clay close to 2Mi muscovite (Yoder and Eugster, 1955) and minor amounts of a poorly defined high-iron “sedimentary chloride” (Warshaw and Boy, 1960, p. 1492), similar to that described by Nelson (1960). Evidence of montmorillonite is sparse, consisting only of very weak reflections about 12 A. Outside the central saltpan area, normal magnesium chlorite, talc, and expandable-lattice clays are more abundant, especially in the parts of the playa which receive detritus from sedimentary and metamorphic rock drainage. To the southwest and northeast of the actual saltpan, chlorite is more abundant, better crystallized, and apparently contains a high proportion of magnesium.
Subordinant amounts of montmorillonite in muds underlying the central saltpan, despite its abundance in marginal lacustrine deposits and its availability from the weathering of either sedimentary or igneous rocks in the basin, suggests diagenetic alteration in the saline environment. Such alteration is probably the result of the high potassium content of the brine at Deep Springs Lake. The chlorite alteration suggested by poor crystallinity and low magnesium content could be the result of leaching and iron oxidation.
GEOCHEMISTRY
The data on the hydrochemistry and saline mineralogy indicate that the Deep Springs Lake saltpan is basically an alkali sulfocarbonate system. Thus, the general geochemistry may be illustrated by phase equilibria in the system Na20-S03-C02-H20. The compositional relations of the major phases in the aqueous salt subsystem Na2C03-NaHC03-Na2S04-H20 are shown in figure 37.
The interpretation of environment and brine composition in a natural salt system depends on the stability relations of the coexistent minerals. Milton and Eugster (1959) have considered the stability relations of the sodium carbonate minerals as functions of temperature and pC02 with reference to the Green Kiver Formation. The assemblages at Deep Springs Lake illustrate primarily the addition of sulfate to the sodium carbonate system. The only mineral compound in the system involving both sodium sulfate and car-A50
CLOSED-BASIN INVESTIGATIONS
h2o
so3
Figure 37.—Major phases in the aqueous salt system NaHCOs-N a2C0r-Na2S04-H20. The larger quaternary Na20-SOs^C02-H20 in which the salt subsystem is outlined, is shown in the smaller diagram to right.
bonate is burkeite, 2Na2S04'Na2C03. The stability of burkeite has been studied experimentally (Jones, 1962) in solutions of nahcolite-trona and trona-natron (Na2CO3-10H2O) buffer assemblages which approximate the total pC02 range in nature. The lower stability limits of burkeite were found to be 47° and 24° C, respectively. Addition of saturated sodium chloride solution decreased the lower limit to 33° and 14° C.
The experimental results indicate that the formation of burkeite is favored by relatively high temperature and salinity. Also, above the lower stability limit, burke-ite-bearing assemblages are favored over other sulfo-carbonate mineral combinations by relatively low partial pressures of carbon dioxide. These results explain why at Deep Springs Lake burkeite is a persistent phase only in the thick central crusts of the main lake, where the concentration of the intercrustal brine and the average annual temperature are relatively high. In the marginal crustal areas, trona plus thenardite or mirabilite—in saturated sodium carbonate solution the mirabilite-thenardite transition is 28 °C according to Makarov and Blidden (1928), and in saturated sodium chloride solution without carbonate the mirabilite-thenardite transition is 18°C according to D’ans (1933)—is a more common assemblage because of more dilute inflow and lower temperature. Similarily, trona plus thenardite (or mirabilite in cold weather) predominates in surficial crustal deposits because surface lake waters are relatively dilute.
The saline mineral assemblages at Deep Springs Lake also correlate with experimental conditions rela-
tive to pC02. Thus, the sodium sulfocarbonate mineral assemblage in the fetid muds beneath the crusts is nahcolite plus thenardite (high pC02 assemblage), except where thin crust may allow the formation of burkeite because of higher temperature. Where there is sufficient carbonate, trona plus thenardite occur in the upper parts of crusts where the temperatures or salinities are not high enough for burkeite formation. Ther-monatrite, indicative of high temperature and (or) salinity or low pC02 occurs only in efflorescent crusts, usually accompanied by burkeite. Most thermonatrite was probably derived from the dehydration of natron. Natron plus mirabilite assemblages have never been observed, but might precipitate from relatively dilute surface waters in very cold weather and are not preserved. Much of the thenardite in surficial crustal deposits appears to have been derived from original mirabilite, but in most of the crustal layers, salinity was sufficiently high that thenardite crystallization was probably primary. A detailed itemization of the sodium sulfocarbonate assemblages identified at Deep Springs Lake, including location and nature of the deposit, is given in table 9. A specific example of the interpretative application of experimental data is at locality 22. Here, the coexistence of trona, nahcolite, thenardite, and burkeite fixes the temperature at 33°C to 47°C depending on the total salinity of the intercrustal brine. The total dissolved-solids content for brine sample DL22 (277,000 ppm) suggests that temperature in excess of 35°C may not have been necessary for burkeite to form.HYDROLOGY AND MINERALOGY OF DEEP SPRINGS LAKE, CALIF.
A51
Table 9.—Saline mineral assemblages in the system NaHC03“Na2C(>3 — Na2S(>4 — H20 at Deep Springs Lake [In some assemblages, thenardite or thermonatrite have derived from their equivalent decahydrates, mirabilite or natron, prior to analysis]
Assemblage
Burkeite-f thenardite...........
Burke ite+trona.................
Burkeite+trona+thermonatrite...
Burkeite+nahcolite..............
Burkeite-f nahcolite+trona......
Burkeite-f thenardite-ftrona....
Thenardite-j-trona..............
Thenardite+nahcolite____________
Thenardite-ftrona+thermonatrite
Locality No. (fig. 2) Type of deposit
9,11,24,63
21,22,68
39..
51....
39
22
49 Efflorescent crust around “seep channel" to
8,11,16 lake from bog mound springs.
22
22
1,2,3,8,10,13,39
11,12,24,29
37 Efflorescence around Buckhorn pond seepage. _.
Relative condition
Low pCOa.
Hi{g temperature and (or) salinity.
Do!
Average pCOa.
High temperature and (or) salinity.
High temperature and average pCOa. High salinity.
High temperature, very high pCOa.
High temperature and (or) salinity and high pCOa.
High temperature and (or) salinity, average pCOj.
Low temperature and (or) salinity, average pCOa.
Low temperature and (or) salinity, high pCOj.
Low temperature and salinity, low pCOa.
To a considerable extent, the sodium sulfocarbonate mineral sequences indicate original sequences of precipitation as well as gradients in temperature or present solution composition. Thus the lower and outermost lacustrine crusts consist of sulfate, alkali carbonate occurring areally inward or vertically above. Where burkeite is present, it is between and (or) intermixed with the sulfate and carbonate phases. In the seasonally formed crusts of the sag pond where carbonate predominates, nahcolite and thenardite are succeeded upward by burkeite and thence by trona alone. This mineral layering reflects both upward decrease in pC02 and the original precipitation sequence. Generally, it appears that the saline minerals and associated brines at Deep Springs Lake have reached local equilibrium in the sense of Thompson (1959, p. 430).
Equilibrium between brines and sulfocarbonate mineral assemblages may be tested in part for conditions at 20°C, the only temperature for which data on the system Na2C03-NaHC03-Na2S04-NaCl-II2C) are available (Teeple, 1929). The data are shown graphically in a triangular projection (fig. 38). This plot represents the composition of saturated solution surfaces with NaHC03 is projected to the Na2C03 corner. The letters refer to the experimentally determined points; the connecting curves are reduced to straight lines because of lack of data. The divariant fields are labeled according to the solids which coexist with saturated solution; fields which have a single solid phase are undersaturated with carbonate. Teeple’s (1929, p. 163) actual data are also listed in fig. 38. If additional data were available each divariant field might be contoured according to pC02 and total concentration, but the burkeite field very likely contains the points of maximum total con-
centration and minimum pC02 of solution in equilibrium with a carbonate phase.
Figure 38 also illustrates the relative composition of a few brines from Deep Springs Lake which had temperatures near 20°C at time of collection. The mineral assemblages found coexistent with these brines were as follows:
1 -------------------------- Trona+thenardite.
2 __________________________ Thenardite.
3A __________________________ Trona+halite.
3B __________________________ Trona + thenardite.
13 __________________________ Do.
24___________________________ Thenardite.
290 _________________________ Thenardite+nahcolite.
If one realizes that much thenardite may be derived from the evaporative dehydration of mirabilite during sampling (as observed at No. 13) or analysis, then the overall agreement is good. The few disparities can readily be explained by the effect of additional components in the system. The additional components in Deep Springs Lake brines would tend to lower the activity of H20 and carbon dioxide in the system and expand the fields representative of brines which are in equilibrium with solids of lower hydration and car-bonation state. An increase in temperature would have the same effect. The burkeite and thenardite fields would be especially favored by such trends and would expand primarily at the expense of mirabilite. Thus, it is not surprising that brines coexistent with these phases at higher temperatures would plot within the mirabilite field on the 20°C diagram. (Compare fig. 38 and fig. 21.) The relations emphasize the need for additional data on the quinary system; data for 30° to 35 °C should be particularly useful to the interpretation of saline occurrences at Deep Springs Lake.A52
CLOSED-BASIN INVESTIGATIONS
Figure 38.—Mineral assemblages in equilibrium with saturated solution in the system NaHC03-Na2C0rNa2S04-NaCl-H20 at 20° C and 1 atmosphere total pressure based on the data of Teeple (1929, p. 163). Isobaric invariant assemblages consist of points where there are four solid phases plus saturated solution. Field boundaries in the diagram represent univariant curves with three solids plus saturated solution and are shown straight for lack of further data. Numbered points refer to brine samples which were near 20° C at time of collection. Nah, nahcolite; So, natron; Tr, trona; Th, thenardite; M, mirabi-lite; Bu, burkeite; H, halite.
Invariant point on fig. 38 Teeple’s number Percent of equivalents per million
HCOj+COa S04 Cl
A 253 Nah + So + M + Tr Nah + M + Th + Tr Nah + M + Tr Nah + Th + Tr Nah + Th + H + Tr So + M + Bu + Tr So H 4- Bu 4- Tr 56.0 24.3 19.8
B 256 11.7 16.2 72.1
C 36.0 23.2 40.8
D 9.8 12.7 77.5
E 258 9.1 11.4 79.5
F 254 43.8 21.3 34.8
G 2 55 32.3 5.9 61.8
H 257 M H - Th + Bu + Tr - Th + Tr - Th + Bu + Tr - Th + Nah 17.4 16.7 65.9
I H H 9.0 11.2 79.8
J 259 H M 14. 2 11.1 74.6
K 250 M H 4.2 15.6 80.2
L 251 Th + Nah + H 3.2 12.0 84.8
34.8 65.2
Th + H Th + M Nah + Tr + So Tr + So + H Nah + Tr + H So + Nah + M 11.7 88.3
243 247 248 252 245 15.6 84.4
83.8 30.9 9.5 70.0 16.2
69.1
90.5
30.0
HYDROLOGY AND MINERALOGY REFERENCES
Alderman, A. R., and Skinner, H. C. W., 1957, Dolomite sedimentation in the southeast of South Australia: Am. Jour. Sci., v. 255, p. 561-67.
Anderson, G. H., 1937, Granitization, albitization, and related phenomena in the northern Inyo Range of California— Nevada: Geol. Soc. America Bull., v. 48, p. 1-74.
Barnes, I. K., 1962, Geochemistry of Birch Creek, Inyo County, California [abs.] : Jour. Geophys. Research, v. 67, no. 9, p. 3540.
------1964, Field measurement of alkalinity and pH: U.S.
Geol. Survey Water-Supply Paper 1535-H. p. 1-17.
Bateman, P. M., 1961, Granitic formations in the east-central Sierra Nevada near Bishop, California: Geol. Soc. America Bull., v. 72, p. 1521-1538.
Bates, R. G., 1954, Electrometric pH measurements: New York, John Wiley and Sons, 331p.
Bedekar, S. G., 1955, Properties of sodium carbonate-sodium bicarbonate solutions: Jour. Appl. Chemistry, v. 5, p. 72.
Bunting, B. T., 1961, The role of seepage moisture in soil formation, slope development and stream initiation: Am. Jour. Sci., v. 259, p. 503-518.
D’ans, J., 1933, Die Losungsgleichgewichte der systeme der salze oceanisher Salzablagerungen: Berlin.
Emerson, D. O., 1959, “Degranitization” of the northern Inyo Range, California-Nevada [abs.] : Geol. Soc. America Bull., v. 70, p. 1717.
Foshag, W. F., 1935, Burkeite, a new mineral species from Searles Lake, California: Am. Mineralogist, v. 20, p. 50-56.
Garrels, R. M., 1960, Mineral equilibria: New York, Harper Bros., 254 p.
Goldsmith, J. R., and Graf, D. L., 1958, Structural and compositional variations in some natural dolomites: Jour. Geology, v. 66, p. 678-693.
Grim, R. E., 1960, Clay mineralogy: New York, McGraw-Hill Book Co., 304 p.
Hem, J. D., 1959, Study and interpretation of water analyses: U.S. Geol. Survey Water-Supply Paper 1473, 269 p.
Hubbs, C. L., and Miller, R. R., 1948, The Great Basin, part 2: The zoological evidence: Univ. Utah Bull., v. 38, p. 103-113.
Hunt, C. B., 1960, The Death Valley salt pan, a study of evaporite in Short papers in the geological sciences: U.S. Geol. Survey Prof. Paper 400-B, p. B456-B457.
Hutchinson, G. E., 1957, A treatise on limnology, vol. 1, geography, physics and chemistry: New York, John Wiley and Sons, 1015 p.
Jones, B. F., 1962, Stability of burkeite and its significance in lacustrine evaporites [abs.] : Jour. Geophys. Research, v. 67, no. 9, p. 3569.
Kesseli, J., and Beatty, R., 1959, Desert floods in the White Mountains of California and Nevada: Quartermaster Research and Engineering Center, Tech. Rept. EP-108, 108 p., 1 pi.
ICnopf, Adolph, 1918, A geologic reconnaissance of the Inyo Range and the eastern slope of the southern Sierra Nevada, California, with a section on the stratigraphy of the Inyo Range, by Edwin Kirk: U.S. Geol. Survey Prof. Paper 110, 130 p.
Kohler, M. A., Nordenson, T. J., and Baker, D. R., 1959, Evaporation maps for the United States: U.S. Weather Bureau, Tech. Paper 37, 13 p.
OF DEEP SPRINGS LAKE, CALIF. A53
Langbein, W. B., 1961, Salinity and hydrology of closed lakes: U.S. Geol. Survey Prof. Paper 412, 20 p.
Lustig, L. K., 1962, Clastic sedimentation in a bolson environment : Unpub. Ph.D. dissertation, Harvard Univ., 102 p.
Makarov, S. Z., and Blidden, B. P., 1938, Polytherms of the quaternary system Na2C03-Na2SC>4-NaCl-H20 and solid solutions of the burkeite type: Izvest. Akad. Nauk. SSSR, v. 10, no. 4, p. 866-892.
Miller, W. J., 1928, Geology of the Deep Springs Valley, California : Jour. Geology, v. 36, p. 510-525.
Milton, C., and Eugster, H. P., 1959, Mineral assemblages of the Green River formation in P. H. Abelson, ed., Researches in Geochemistry: New York, John Wiley and Sons, p. 118-150.
Nelson, B. W., 1960, Clay minerals of the bottom sediments of the Rappahannock River, Virginia, in Clays and clay minerals, Proceedings of the 7th National Conference, NAS-NRC, Monograph no. 5: New York, Pergamon Press, p. 135-147.
Nelson, C. A., 1962, Lower Cambrian-Precambrian succession, White-Inyo Mountains, California: Geol. Soc. America Bull., v. 73, p. 139-144.
------ 1963, Geology of the Blanco Mountain Quadrangle, California : U.S. Geol. Survey Mineral Inv. Map, MF256.
Palmer, L. A., 1922, Evaporation of potash brines at Deep Spring Lake, California: Chem. and Metall. Eng., v. 26, p. 1034-1037.
Peterson, M. N. A., Bien, G. S. and Berner, R. A., 1963, Radiocarbon studies of recent dolomite from Deep Spring Lake, California: Jour. Geophys. Research, v. 68, no. 24, p. 6493-6505.
Piper, A. M., 1944, A graphic procedure in the geochemical interpretation of water analyses: Am. Geophys. Union Trans., v. 25, p. 914-923.
Pro, M. J., Etienne, A., and Feeny, F., 1959, Determination of carbon dioxide in wines using a vacuum system: Jour. Assoc, of Official Agr. Chemists, v. 42, p. 679-683.
Rainwater, F. H., and Thatcher, L. L., 1960, Methods for collection and analysis of water samples: U.S. Geol. Survey Water-Supply Paper 1454, 301 p.
Smith, G. I., and Haines, D. V., 1964, Character and distribution of nonclastic components in the Searles Lake evaporite deposits: U.S. Geol. Survey Bull. 1181-P, 58 p.
Teeple, J. E., 1929, Industrial development of Searles Lake brines: Am. Chem. Soc. Mono. no. 49, 182 p.
Thompson, J. B., 1959, Equilibrium in metasomatic processes in P. H. Abelson, ed., Researches in Geochemistry: New York, John Wiley and Sons, p. 427-457.
Tourtelot, H. E., 1961, Thin sections of clay and shale: Jour. Sed. Petrology, v. 31, p. 131-132.
Warshaw, C. M., and Roy, R., 1960, Classification and a scheme for the identification of layer silicates: Geol. Soc. America Bull., v. 72, p. 1455-1492.
Yoder, H. S., and Eugster, H. P., 1955, Synthetic and natural muscovites: Geochem. et Cosmochim. Acta., v. 8, p. 225-280.
Zen, E-an, 1957, Preliminary report on the mineralogy and petrology of some marine bottom samples off the coast of Peru and Chile: Am. Mineralogist, v. 42, p. 889-903.INDEX
A Page
Acknowledgments............................... 2
Albite....................................... 49
Albitization.................................. 7
Alkaline earth, carbonates................. 2,34
concentration...........................33,34
precipitation............................. 38
Alkalinity................................. 26
Alluvial fans............................... 3,8
Alluvium...........................8,13,17,18,34
Amphibole.................................... 49
Analcite..................................... 46
Analysis, gravimetric........................ 27
major-element............................. 27
methods................................... 25
waters..................................28,30
X-ray..................................... 44
Anderson, G. H., cited....................... 7
Antelope Canyon.........................14,20,35
Antelope Creek............................ 14,20
Antelope Spring...................... 14,33,35,38
Aphthitalite......................... 42,46,47,49
Aragonite.............................. 44,45,49
Auger hole................................. 9,43
B
Barnes, I. K., cited........................... 14
Basalt......................................... 13
Bateman, P. M., cited........................... 8
Batholiths...................................... 8
Beatty, R., and Kesseli, J., cited.......... 8,12
Beer Creek................................ 14
Bibliography................................... 53
Birch Creek....................... 8,14,33,34,35,38
Blanco Mountain quadrangle..................... 21
Blidden, B. P., and Makarov, S. Z., cited___ 50
Bloedite.................................... 46
Bog-mound springs. 15,18,23,33,34,35,38,40
Bonneville basin............................ 2
Boundary Peak Granite........................... 7
Brine, intercrustal.......................... 30
interstitial............................... 30
shrimp in.................................. 17
Buckhorn Springs.............................. 15,
16,20,21,23,33,34,35,36,37,38,46,48 Burkeite....................... 44,45,46,49,50
C
Calcite.............................. 44,45,48,49
deposits.................................. 34
precipitation........................ 34,35,38
Calcium determination......................... 27
Campito Formation.............................. 7
Carbon-14 dating............................... 9
Carbon dioxide..................... 27,35,41,49,50
Carbonate.............................. 44,50,51
mineralogy................................. 43
muds___________________________________ 8,9
precipitation.............................. 38
Cedar Spring.................................. 13
Chloride................................ 2,49
precipitation............................. 42
See also Sulfate-chloride ratios.
Chlorite...................................... 49
Class-A-pan evaporation....................... 12
Clastic sediments........................... 24
Clays................................... 45,49
See also Montmorillonite.
I Italic page numbers indicate major references]
Page
9
Contact metamorphism 8,14
Corral Springs........ 15,20,23,33,34,35,36,37,38,40
Crooked Creek............................8,13,33,34
D
Dead Horse Meadow___________________________ 13
Death Valley.................................. 2
Deep Springs Formation..................... 6,7
Deep Springs Junior College.................. 13
Deep Springs School.......................8,9,12
Desiccation_____________________________21,25,48
Discharge, from streams and springs........13,20
Dissolved solids, factors affecting concentration__________________________________________ 25
loss..................................... 24
tonnage................................ 21
Dolomite.......................... 6,44,45,48,49
precipitation............................. 38
Drainage.............................. 15,33,34
Dunes....................................... 5
Page
Hornblende.............................- 7
Homfels................................. 7
calc-silicate__________________________ 14
Hutchinson, G. E., cited............... 2,32
Hydrochemistry-------------------------- 25,49
Hydrogen sulfide______________________15,34,41
Hydrography............................ 12,20
Hydrology-................-............ 9
Hydrothermal veins________________________ 33
I
Igneous rocks-----------
Inflow...................
Introduction...........
Inyo Chemical Co.........
Inyo Mountains...........
Iodometric method________
Ion exchange.............
7,13,38 32 2 18 3 27 34
J
Junction-potential effects.
26
E
Efflorescent crusts..................9,17,48,50
Equations................................... 22
Escarpment................................ 16
Eugster, H. P., and Milton, C., cited....... 49
Eureka Valley............................... 3
Evaporation...............................12,02
F
Fanglomerate...........................14,16,34
Faulting............................ 3,5,15,16
Feldspar.................................... 49
Fish Lake Valley........................3,12,13
Flame photometry.......................... 27
Flash flooding____________________________14,20
Flood frequency............................ 14
Foshag, W. F., cited........................ 46
G
Gas, analysis tubes........................ 26
evolution, technique..................... 27
in springs............................ 15
vents................................ 40,48
Gastropods.................................. 9
Gaylussite........................... 44,46,49
Geochemistry............................... 49
Geography................................... 3
Geology, bedrock............................ 6
Quaternary—.............................. 8
Glauberite................................ 48
Granite...................................... 7
Granodiorite................................ 8
Gravimetric analysis________________________ 27
Great Basin............................... 2,8
Great Salt Lake______________________________ 2
Green River Formation_______________________ 49
Ground water_____________________________ 17
H
Halite__________________________________ 44,46
beaches................................ 41
Hardness................................. 27
Harkness Formation___________________________ 7
Hem, J. D., cited________________________ 26,31
Highway, maintenance station..._____________ 17
K
Kesseli, J., and Beatty, R., cited------- 8,12
Kirk, Edwin, quoted---------------------- 6
Knopf, Adolph, cited..................... 8
L
Lacustrine deposits....................8,9,20,44
Lahontan basin______________________________ 2
Lake, area variability equation.............. 22
composition of waters.................... 40
evaporation............................... 12
level__________________________________ 5, 20
volume, relation to salt tonnage in
solution.......................... 23
Langbein, W. B., cited............... 2,20,22,43
quoted_________________________________24,25
Leaching..................................... 34
Levee..............................3,16,21,45,47
Limestone............................-.... 7
Lustig, L. K., cited...................... 6, 9
M
Magnesium.............................. 35,36,49
determination............................ 27
Makarov, S. Z., and Blidden, B. P., cited- 50
Marble...................................... 14
Marsh..................................... 15,35
Metamorphic, aureole.......................... 6
rocks_________________________________ 8,33
Metamorphism............................ 7,33,34
Methane.............................-..... 41
Mica................................-.....- 49
Mill Canyon__________________________________ 13
Miller, W. J., cited...................... 3,6
quoted................................. 7
Milton, C., and Eugster, H. P., cited........ 49
Mineral, identification techniques----------- 43
separation tech niques................... 43
Mineralization.----------------------------- 33
Mineralogy.................................. 43
Mirabilite.............................. 45,48, 50
Montmorillonite............................34,49
Monzonite.................................. 7,8
A55A56
INDEX
Page
Mud, solubles content........................ 24
Mudflows-------------------------------------- 8
Mule Spring Limestone------------------------- 7
Muscovite----------------------------------- 49
N
Nahcolite------------------------------- 46,50
Nelson, C. A., cited------------------------ 6,7
O
Owens Valley-----------—*------------------ 3,12
P
Payson Canyon____________________________ 14,20
Pellisier Granite----------------------------- 7
PH._..........................................38
Phase equilibria--------------------------- 49
Piper, A. M., cited-------------------------- 31
Prissonite.---------------------------------- 48
Plagioclase---------------------------------- 49
Playa, Deep Springs Lake---------------------- 3
sediments________________________________ 24
Poleta Formation------------------------------ 7
Polygonal patterns-------------------------47,48
Potash, production---------------------------- 3
Precipitation------------------------------ 9,22
Pygmy current meter-------------------------- 20
Q
Quartz--------------------------------------- 49
Quartz diorite________________________________ 7
Quaternary geology---------------------------- 8
R
Rainwater, F. H., and Thatcher, L. L., cited. 25
Reed Dolomite---------------------------- 6,7
Response-time equation---------------------- 22
Roberts Ranch------------------------------ 13
S
Sag ponds_____________________________ 17,33,38, 1$,46,51
closed_______________________________________4#, 46
Page
Saline crusts----------------------------- 8,45
Saline mineralogy---------------------------45,50
Saline Valley Formation------------------------ 7
Salinity, calculation------------------------- 43
closed lakes------------------------------ 20
cycle_____________________________________ 25
Salts, capillary----------------------------- 38
precipitation----------------------------- 24
See also Dissolved solids.
Saltpan-------------------------------------- 3
Sampling, water___________________________25,27,43
Sandstone------------------------------------ 6,7
Schist........-........................... 14
Searles Lake--------------------------------- 45
Sediments------------------------------------- 44
Sedimentary rocks---------------------6,7,8,13,33
Seepage------------------------------------ 14,17
Shale__________________________________________ 7
Shorelines__________________________________ 5,21
Sierra Nevada batholith------------------------ 8
Silicate mineralogy--------------------------- 49
Silt, near margin of playa--------------------- 8
Sodium---------------------------------------- 35
Sodium-ion effects---------------------------- 26
Sodium-potassium ratios---------------------- 41
Sodium sulfate________________________________ 25
Soldiers Pass__________________________________ 6
Springs------------------------------------- 15
analyses of waters______________________29,30
Staff gage.................................... 20
Stock wells—--------------------------------- 18
Storm patterns------------------------------- 12
Stratigraphy---------------------------------- 9
Stream channels_______________________________ 18
Streamflow----------------------------------12,29
Sulfate...................................2,35,51
precipitation-------------------------- 25,42
reduction................................ 34
Sulfate-chloride ratios_______________________ 42
Sulfide.....................................27,33
Summary_______________________________________ 48
Page
Surface water, area and volume_____________ 21
discharge----------------------------- 13
Sylvite_____________________________________ 46
T
Talc....................................... 49
Teeple, J. E., cited---------------------- 51
Temperature..-----------------------9,12,37,48
See also Water temperature.
Thatcher, L. L., and Rainwater, F. H., cited. 25
Thenardite__________________ 43,44,45,46,47,49,50
Thermonatrite-----------------------------48,50
Titration_________________________________ 27
Trona_____________________________________46,50
V
Varves____________________________________ 9,45
Vegetation-------------------------------- 6,17
Vents_____________________________________ 40
W
Water, composition_______________ 28,30,31,32,40
movement---------------------------- 18
sampling______________________________25,27
temperature___________________________17,26
See also Ground water; Surface water; Streamflow.
Weather___________________________________ 12
Weathering............................. 8,49
Wells, water analyses_______________________ 29
White Mountain No. 1---------------------- 9,12
White Mountains__________________________3,9,20
Wind, effect.............................. 24
Wyman Canyon_________________________________ 8
Wyman Creek______________________8,13,18,33,34
Wyman-Crooked Creeks system___________ 20,35,38
Wyman Formation______________________________ 7
X
X-ray, analysis_____________________________ 44
powder diffraction______________________ 43
O766-977 0 - 66 (In pocket)
UNITED STATES DEPARTMENT OF THE INTERIOR GEOLOGICAL SURVEY
PROFESSIONAL PAPER 502-A PLATE 2
SCIENCE
EXPLANATION
Bog-mound
springs
Staff gage
Line describing approximate edge of lake water (to the south, shoreline elevations are slightly higher than indicated)
Lines of equal elevation describing receding edge of lake water
Artesian Well
LOCATION OF MAP
DEEP SPRINGS LAKE
Windmill
,L/fe Blanco
Buck ho rn/Springs
MAP SHOWING DETAILED TOPOGRAPHY OF THE BED OF DEEP SPRINGS LAKE, CALIFORNIA, AS DERIVED FROM
SHORELINE POSITIONS AND CORRELATION WITH STAFF GAGE ELEVATIONS
1000
3000
4000 FEET
CONTOUR INTERVAL 80 FEET DOTTED LINES REPRESENT GENERALIZED 20-FOOT INTERVALUNITED STATES DEPARTMENT OF THE INTERIOR GEOLOGICAL SURVEY R
PROFESSIONAL PAPER 502-A PLATE 1
EXPLANATION
Bedrock geology
Granitic rocks mostly quartz monzonite
Sedimentary and metasedimen -tary rocks
Generalized contact of intrusive and sedimentary rocks; dotted where concealed by alluvium
__i_i ' i i i i i i
Normal fault
Hachures on down-thrown side
• DSH
Deep Springs
Maintenance Station
• DSS
Deep Springs School
irrigation well
#44
Sampling site
General areal direction of ground water movement in valley sediments
R. 35 E.
Base map from U.S. Geological Survey Blanco Mtn (1951) and Soldier Pass (1958) quadrangles, California
118°05'
R. 36 E.
118°00' R. 37 E.
INTERIOR—GEOLOGICAL SURVEY. WASHINGTON, D. C. —1964 — W64367
Geology after C. A. Nelson (1963)
GENERALIZED GEOLOGIC MAP OF THE DEEP SPRINGS VALLEY, CALIFORNIA SHOWING SAMPLING SITES OUTSIDE THE DEEP SPRINGS LAKE AREA AND THE GENERAL PATTERN OF WATER MOVEMENT IN THE AREA
SCALE 1:62 500
1 Vz 0 1 2 3 4 5 MILES
I—I I—I I—I 1—I I—1 I--- ----------1 I I ----------,
1 -5 0 1 2 3 4 5 KILOMETERS
I—I M M M M I------------ I I--- I i ... |
CONTOUR INTERVAL 80 FEET DATUM IS MEAN SEA LEVEL§flS Pl
* 0£.-‘
7 DAY
U*arth
SCIENCBS
LlgRARY
Hydrology and Geochemistry of Abert, Summer, and Goose Lakes, and Other Closed-Basin Lakes in South-Central Oregon
GEOLOGICAL SURVEY PROFESSIONAL PAPER 502-B
Hydrology and Geochemistry of Abert, Summer, and Goose Lakes, and Other Closed-Basin Lakes in South-Central Oregon
By KENNETH N. PHILLIPS and A. S. VAN DENBURGH CLOSED-BASIN INVESTIGATIONS
GEOLOGICAL SURVEY PROFESSIONAL PAPER 5 02 -B
A description of several broad, shallow lakes of differing salinity, and an evaluation of factors affecting their hydrologic and chemical character
UNITED STATES GOVERNMENT PRINTING OFFICE, WASHINGTON : 1971UNITED STATES DEPARTMENT OF THE INTERIOR ROGERS C. B. MORTON, Secretary
GEOLOGICAL SURVEY W. A. Radlinski, Acting Director
Library of Congress catalog-card No. 78-610410
For sale by the Superintendent of Documents, U.S. Government Printing Office Washington, D.C. 20402 - Price $1.75 (paper cover)CONTENTS
Page
Abstract_____________________________________________________ B1
Introduction------------------------------------------------- 3
Location of study area___________________________________ 3
Purpose and scope of report______________________________ 3
Acknowledgments---------------------------------------- 3
General features of the lakes and their basins---------------- 3
Physiographic and geologic setting_______________________ 3
Climate__________________________________________________ 4
Inflow to the lakes..............................— 5
Outflow from the lakes___________________________________ 6
Water-level fluctuations_________________________________ 6
Chemical characteristics_________________________________ 6
Effects of man’s activities______________________________ 8
Lake Abert____________________________________________________ 9
Area, volume, and lake-level fluctuations________________ 9
Water supply____________________________________________ 12
Water loss______________________________________________ 13
Water budget for the lake_______________________________ 15
Geochemistry____________________________________________ 17
Summer Lake_________________________________________________ 24
Area, volume, and lake-level fluctuations_______________ 24
Water supply......................................... 24
Water loss______________________________________________ 26
Water budget for the lake_______________________________ 26
Geochemistry____________________________________________ 27
Page
Goose Lake__________________________________________________ B30
Area, volume, and lake-level fluctuations_______________ 31
Water supply______________________________________— 32
Water loss______________________________________________ 33
Water budget for the lake.............................. 33
Geochemistry____________________________________________ 34
Reconnaissance of several other closed-basin lakes___________ 37
Area, volume, and lake-level fluctuations_______________ 37
Water supply........................................ 38
Water loss______________________________________________ 39
Geochemistry___________________________________________ 40
Geochemical trends and processes_____...__________________ 42
Major constituents and properties______________________ 42
Minor constituents.........................._........ 47
Trace elements__________________________________________ 52
Location system_____________________________ .:___________ 53
Records of lake-surface altitude_____________________________ 53
Chemical analyses____________________________________________ 53
Methods of analysis___________________________________ 53
Accuracy of analytical results__________________________ 55
Tables of chemical analyses_____________________________ 56
References___________________________________________________ 56
Index________________________________________________________ 83
ILLUSTRATIONS
Page
Plate 1. Map of Lake Abert, Summer Lake, and vicinity, south-central Oregon, showing basin boundaries, inferred extent of pluvial Lake Chewaucan, and location of stream-gaging, lake-level-measuring, and chemical-quality-sampling
sites_____________________________________________________________________________________________________In pocket
2. Map of Lake Abert and vicinity, south-central Oregon, showing sample-collection sites for chemical analysis and representative water-quality data from the lake survey in 1959._______________________________________________________--In pocket
Figure 1. Index map showing location of lakes discussed________________________________________________________________________ B4
2. Diagram showing relative abundance of major dissolved constituents of the lakes and their tributaries--------------- 7
3. Graph showing relation between dissolved-solids concentration and stage of a lake that occasionally overflows--------- 8
4. Photograph showing southward view from the east shore of Lake Abert___________________________________________________ 9
5. Map showing lakebed contours of Lake Abert_______:________________________________________________________________ 10
6. Hydrograph showing seasonal variations in stage of Lake Abert, 1950-63_______________________________________________ 11
7-9. Photographs of Lake Abert showing:
7. Maximum level attained by lake in June 1958, and stump of drowned tree------------------------------------- 11
8. Beach line at an altitude of 4,269.7 feet near lake gage___________________________________________________ 12
9. Anchored instrument raft__________________________________________________________________________________ 14
IIIIV
CONTENTS
Page
Figure 10. Graph showing relation between the fall in stage and the product uAe at Lake Abert, 1962______________________________ B15
11. Graph showing relation between annual flow of Chewaucan River near Paisley and computed inflow to Lake Abert,
water years 1951-62_____________________________________________________________________________________________ 15
12. Bar graph showing annual flow of Chewaucan River near Paisley and computed inflow to Lake Abert, water years
1913-63...........................................................................................-.............. 16
13. Hydrograph showing computed and observed stages of Lake Abert, 1915-63______________________________________________ 17
14-18. Graphs showing:
14. Relation between dissolved-solids concentration and stage of Lake Abert, 1958-62---------------------------- 18
15. Seasonal variations of dissolved-solids concentration, temperature, silica, and orthophosphate in Lake Abert,
1961-63................................................................................................. 19
16. Dissolved-solids concentration versus discharge of Chewaucan River near Valley Falls-------------------------- 21
17. Relation between dissolved-solids concentration and silica, sodium, and chloride in Chewaucan River near
Valley Falls________________________________________________________-........-..................-....... 22
18. Relation between dissolved-solids concentration and amounts of calcium and magnesium in Chewaucan River
near Valley Falls_______________________________________________________________________________________ 23
19. Hydrograph of seasonal variations in stage of Summer Lake, 1959-62_______________________________________________________ 25
20. Graph showing cycle of dissolved-solids concentration and stage of Summer Lake, June 1961-June 1962-------------------- 27
21. Schematic diagram of hypothetical relation between lake volume and dissolved-solids tonnage in Summer Lake, June
1961-June 1962________________1_________________________________________________________________________________ 27
22. Graph showing relation between dissolved-solids concentration and stage of Summer Lake, 1912, 1959, and parts of
1961 and 1962___________________________________________________________________________________________________ 28
23. Photograph showing clouds of alkali dust rising from surface of dry Summer Lake playa------------------------------- 28
24. Graph showing seasonal changes in dissolved-solids concentration, temperature, silica, and orthophosphate in Summer
Lake, 1961-62.__________________________________________________________—--.............................-....... 29
25. Map showing lakebed contours of Goose Lake__________________________________________________________________________ 31
26. Hydrograph of seasonal variations in stage of Goose Lake, 1958-62____________________ —---------------------------------- 32
27. Graph showing relation between dissolved-solids concentration and stage of Goose Lake, 1912 and 1953-62---------------- 34
28. Graph showing seasonal changes in dissolved-solids concentration, temperature, silica, and orthophosphate in Goose
Lake, 1961-62.............................................................................-..................... 36
29. Graph showing relation between seasonal flow of Silver Creek and computed inflow to Silver Lake--------------------- 39
30-34. Graphs showing relation between dissolved-solids concentration and:
30. Stage of Silver Lake, 1959-61, with and without solute loss_________________________________________________ 41
31. Sodium to potassium ratio in surf ace waters.....................................--------------------------- 43
32. Distribution of major anions in surface waters_______________________________________________________________ 44
33. Carbonate to bicarbonate ratio and pH in lakes_______________________________________________________________ 46
34. Relative amounts of silica in surface waters_______________________________________________________________ 48
35. Graph showing seasonal variations of silica tonnage in Lake Abert, 1961-63------------------------------------------ 48
36. Graph showing relation between dissolved-solids concentration and abundance of calcium and carbonate plus bicar-
bonate in surface waters________________________________________________________________________________________ 49
37. Grid showing rectangular system of land division____________________________________________________________________ 53
TABLES
Page
Table 1. Summary of physical, hydrologic, and chemical data for the lakes....................—------------------------........ B5
2. Monthly temperature and precipitation at Valley Falls____________________________________________________________ 5
3. Relative amounts of major and minor constituents in representative samples from Abert, Summer, Goose, Hart, and
Crump Lakes_________________________________________________________________________________________________
4. Approximate area and volume of Lake Abert------------------------------------------------------------------------
5. Annual discharge of Chewaucan River near Paisley and at Hotchkiss Ford, 1915-20-------------------------------------- 13
6. Monthly wind movement, temperature, relative humidity, and evaporation at Lake Abert, and temperature and pre-
cipitation at Valley Falls, October 1961 and April-October 1962--------------------------------------------- 14
7. Dissolved-solids tonnage in Lake Abert, 1939-63_____________________________________________________________________ 18
8. Relative amounts of five trace elements in Lake Abert and the Chewaucan River--------------------------------------- 19
9. Relative abundance of major dissolved constituents in Lake Abert, 1912 and 1962------------------------------------- 20
10. Changes in chemical character of Chewaucan River between Paisley and Valley Falls----------------------------------- 21
11. Measured discharge of Ana River near Summer Lake between 1904 and 1962---------------------------------------------- 25
00 05CONTENTS
V
Page
Table 12. Changes in the relative amounts of six major constituents in Summer Lake between 1912 and 1961____________________ B30
13. Area and volume of Goose Lake______________________________________________________________________________________ 31
14. Changes in the relative amounts of six major constituents in Goose Lake between 1912 and 1962______________________ 35
15. Average annual flow at gaging stations on principal streams in the Silver Lake, Harney, and Warner Basins__________ 39
16. Theoretical and actual changes in the chemical character of Silver Lake, April 25-June 20, 1961___________________ 41
17. Carbonate-bicarbonate and calcium in stream waters_________________________________________________________________ 45
18. Silica budget for Lake Abert, October 1, 1962-June 30, 1963________________________________________________________ 48
19. Relative amounts of six trace elements in samples collected June 12, 1962, from the Chewaucan River and from Goose,
Summer, and Abert Lakes_____________________________________________________________________________________ 52
20-23. Altitude of water surface in:
20. Lake Abert 1843-1963............................................................................... 60
21. Summer Lake, 1905-63______________________________________________________________________________________ 61
22. Goose Lake, 1832-1963_______________________________________________________________________________ 62
23. Silver Lake, 1861-1963____________________________________________________________________________________ 63
24. Hydrologic references to Malheur and Harney Lakes, 1826-99_______________________________________________________ 64
25. Altitude of water surface at staff gages in Malheur Lake, 1932-62__________________________________________________ 65
26. Altitude of water surface in Harney Lake, 1929-62__________________________________________________________________ 66
27. Altitude of water surface in Crump, Hart, and Bluejoint Lakes, 1867-1963___________________________________________ 67
28. Chemical and physical character of the lakes and their tributary streams___________________________________________ 68
29. Trace-element content of samples from Abert, Summer, and Goose Lakes and the Chewaucan River_______________________ 76
30. Chemical and physical character of sampled spring waters___________________________________________________________ 78
31. Chemical and physical character of sampled well waters_____________________________________________________________ 80CLOSED-BASIN INVESTIGATIONS
HYDROLOGY AND GEOCHEMISTRY OF ABERT, SUMMER, AND GOOSE LAKES, AND OTHER CLOSED-BASIN LAKES
IN SOUTH-CENTRAL OREGON
By Kenneth N. Phillips and A. S. Van Denbukgh
ABSTRACT
Topographically enclosed fault-block basins in the semidesert region of south-central Oregon contain lakes whose volumes have fluctuated in response to changes in water supply (precipitation plus inflow) and water loss. Most of the lakebeds are impermeable, and the water is dissipated only by evaporation or, for some lakes, occasional overflow. Solutes carried into the closed lakes by springs
and small streams tend to be accumulated in solution, or removed from the lake water by organic and inorganic processes to become a part of the lakebed sedimentary deposits. The most soluble salts can be removed only by overflow, by incorporation within the lake-bottom and peripheral muds as interstitial brines, or by wind transport of precipitated evaporite minerals in periods of dryness.
Physical, hydrologic, and chemical data for the lakes studied are summarized in the following table.
Summary of physical, hydrologic, and chemical data for the lakes
Lake Abert Summer Lake Goose Lake Silver Lake Harney Basin Warner basin
Malheur Lake 1 Harney Lake Crump Lake Hart Lake Bluejoint Lake
Drainage area, including lake.sq mi_ _ 860 390 1,140 520 2,150 5,280 700 900 1,920
Altitude (ft):
Overflow _ _______ 4,390 4,390 4,716 4,310.4 4,093.5 + 4,114 + 4,474.4 4,473 4,800 +
Low point in lake bed 4,243.5+ 4,144 + 4,692 4,299 + 4,088.8 4,079.7 4,468 4,463.5 4,455
Highest observed lake level2 4,260.5 4,149 4,707.8 4,309.0 4,095.4 4,087.6 4,476.4 4,474.5 (3)
Lowest observed lake level 2_ _ _ Dry Dry Dry Dry Dry Dry 4,470.6 Dry Dry.
Highest known Pleistocene lake
level _ _____ 4,520 4,520 4,720 + 4,520 4,114 + 4,114 + 4,750 to 4,800
Lake area at high level2 _ _sq mi _ 64 4 71 181 16 + 103 47 12 + 12 (3)
Lake volume at high level2
103 acre-ft._ 496 (3) 1,270 60 + 240 176 (3) 56 (3)
Estimated average annual quantity:
Lake-surface precipitation. _in._ _ 12 11 14 11 9 9 10 8 8
Evaporation . in._ 39 40 42 38 42 41 44 44 44
Inflow _ 103 acre-ft_ 55 65 165 (3) 100 + 60 + 70 + 80 + (3)
Overflow2 __103 acre-ft__ 0 0 0 0 20 + 0 60 + 30+ 0
Dissolved-solids content:
Maximum determined ppm__ 95,000 36,900 2,700 698 5 484 6 22,400 500 5 781 6 3,630
Minimum determined _ _ _ ppm_ _ 18,700 1,800 600 300 6 258 (#) 101 6 230 (6)
Principal constituents 7 Na Na Na (8) CaNa Na CaNa Na Na
CICOa CICOa HCO, HC03 Cl HCO, HCO, HCO,
HCO,
1 Area and volume shown for Malheur Lake at high levels include Mud lake. 6 Only two reliable determinations are available.
2 For period 1924-63. 6 Only one reliable determination is available.
3 Not determined. 7 On basis of equivalents per million. Listed constituents are calcium (Ca),
4 Measured within meander line (est alt 4,152-4,155 ft). sodium (Na), chloride (Cl), carbonate (COa), and bicarbonate (HCOj).
8 Variable.
B1B2
CLOSED-BASIN INVESTIGATIONS
Lake Abert, 25 miles north of Lakeview, is at present (1968) the largest landlocked saline water body in the Pacific Northwest. Inflow to the lake is almost entirely from the Chewaucan River, which drains about 60 percent of the total tributary area. The lake was dry in the summer of 1924 and again during other dry years through 1937. At its highest observed level, in 1958, the lake drowned juniper trees as much as 55 years old.
Summer Lake, 25 miles northwest of Lake Abert, is the shallow evaporating sump for inflow from the spring-fed Ana River and small snow-fed streams that rise on adjacent Winter Ridge. The flow of Ana Springs (believed to be nourished by ground water from nearby Fort Rock basin) has decreased from 160-170 cubic feet per second in 1904 to about 90 cubic feet per second, owing in large part to impoundment of the river and submergence of the spring orifices in 1923. Because of the decreased discharge from springs and the increased use of water for irrigation, the total flow now reaching Summer Lake, including precipitation, is only about 60 percent of that prior to 1923, and the average lake level is now several feet lower. The lakebed has been almost entirely dry late in the summers of many years since 1923.
Goose Lake, which straddles the Oregon-California border south of Lakeview, Oreg., is fed mostly by streams draining areas in Oregon. The lake overflowed briefly into the North Fork Pit River in 1868 and 1881, but storage and diversion of sufficient water for irrigation of about 35,000 acres has subsequently decreased the inflow; thus, future overflow is unlikely. The lakebed was dry in the summers of 1926 and 1929-34.
Silver Lake, in the Fort Rock basin, is fed mostly by Silver Creek and other streams that enter by way of Paulina Marsh. Overflow last occurred in 1905, and use of water for irrigation has decreased the likelihood of future overflow. The lake apparently loses water to unsaturated zones beneath its bed. Silver Lake was dry from 1922 to 1950, refilled to its highest recent level in 1958, and became dry again in August 1961.
Malheur Lake, 20 miles southeast of Burns, receives most of its inflow from the Silvies and Donner und Blitzen Rivers. In wet years, the lake merges with small Mud Lake and, at a slightly higher level, overflows into Harney Lake, the lowest point in Harney Basin. Malheur Lake was dry in 1934, and nearly dry in 1926, 1931-33, 1961, and 1962. A single wet season after a drought produces more inflow than the capacity of Malheur and Mud Lakes, causing overflow. In addition to this occasional overflow, Harney Lake receives intermittent runoff from several small streams.
Warner basin, northeast of Lakeview, contains a 40-mile chain of lakes that includes Crump Lake at the south (upstream) end and Bluejoint Lake at the north low point. Most runoff occurs in the south end of the basin, and much of it is used for irrigation of about 40,000 acres, so that lakes downstream from Hart Lake (below Crump Lake) remain dry or almost dry except in wet years.
For all but the freshest closed lakes, sodium and potassium are by far the most abundant cations. Measured epm (equivalents per million) percentages for the two ions together range from 46 percent of the cations in Crump Lake at 322 ppm (parts per million) of dissolved solids to 99.99 percent in Lake Abert at 20,000-80,000 ppm. The sodium to potassium epm ratio also increases with average solute concentration from 15:1 and less in Crump Lake to 50:1 in Lake Abert.
Among the anions, carbonate bicarbonate ranks either first or a close second in all the lakes, with the relative (percentage) amounts decreasing as dissolved-solids concentration increases (the epm-percentage ranges from about 80 in Crump Lake to about 40 in Lake Abert). The decrease is largely the result of calcium carbonate precipitation and differences in the chemical character of inflow to the more dilute lakes, and the result of several factors including
mineral reaction and wind removal of efflorescent playa salts from the more saline lakes. The abundance of chloride is opposite to that of carbonate bicarbonate (from about 10 epm-percent of the anions in dilute Crump Lake to almost 60 percent in Abert), and most of the same factors that account for carbon depletion with increasing concentration also explain the relative enrichment of chloride. Sulfate is not a principal component in any of the lakes (it makes up a mere 2 epm-percent of the anions in Lake Abert), even though its abundance is comparable to that of chloride in most inflow. Sulfate depletion seems largely the result of bacterial reduction within the lake-bottom muds.
Among the minor constituents, silica is generally the most abundant, at solute concentrations of less than 50,000-100,000 ppm. As much as 200 ppm of silica has been measured in Abert and Summer Lakes, with the upper limit being imposed by mineral equilibria and, perhaps, biologic controls. Calcium and magnesium are among the most abundant constituents of stream and dilute lake waters, but they are nearly absent from the more saline lakes, again because of mineral reactions and possible biologic controls.
Bromide is more abundant than fluoride in the most saline lakes, despite the opposite situation in inflow. In Lake Abert, the epm-percentage for bromide relative to total anions is 0.12 (103 ppm measured at a total solute concentration of about 60,000 ppm), in contrast to only 0.04 percent for fluoride. The amount of bromide relative to chloride in Lake Abert is greater than in any other Western closed-basin lake.
The biologic nutrient orthophosphate is plentiful in the more saline closed-basin lakes, owing mostly to the absence of calcium. Maximum measured concentrations of orthophosphate in Goose, Summer, and Abert Lakes have been 12, 41, and 101 ppm, which represent 0.54, 0.25, and 0.14 percent of the total solute concentrations, respectively. The amounts present in these lakes seem little affected by biologic controls.
Boron also is an abundant minor element in the saltier lakes (its relative amount (or percentage) in Lake Abert is about the same as that of orthophosphate), though some other Western closed-basin lakes contain even higher percentages.
Of the 17 trace elements looked for in samples from Goose, Summer, and Abert Lakes, only aluminum, iron, molybdenum, nickel, and vanadium consistently occurred in concentrations that exceeded the lower quantitative limits.
Lake Abert contained about 8 million tons of dissolved solids in 1939, and almost 14 million tons in 1961. The increment was derived mostly from salts stored in peripheral and lake-bottom sedimentary deposits that were inundated as the lake level rose from near-dryness in 1938 to the record high level in 1958. The average solute contribution of the Chewaucan River, the principal source of inflow, is about 120 ppm, made up mostly of silica, calcium, sodium, and bicarbonate. Peripheral seeps and springs are sources of sodium, bicarbonate, and chloride, but much of their contribution may have been recycled.
Ana Springs, the principal tributaries to Summer Lake, yield about 160 ppm of dissolved solids—mostly silica, sodium, and bicarbonate. During periods of low lake level or dryness since 1923, large amounts of efflorescent salts have been removed from the lake-bed and peripheral mud flats by gusty winds. Concurrently, chloride and sulfate have been enriched in the lake relative to carbonate-bicarbonate, due at least in part to selective windborne removal of the more powdery efflorescent evaporite salts of the latter. Goose Lake underwent a similar chemical change while losing two-thirds of its initial (1912) 3-million-ton solute load during periods of desiccation in the 1920’s and 1930’s.
As Silver Lake receded to near dryness between May 1959 and June 1961, the dissolved salts increased only twofold—from 370 toHYDROLOGY AND GEOCHEMISTRY, CLOSED-BASIN LAKES, OREGON
B3
698 ppm. In contrast, sodium and chloride were enriched at least tenfold, owing to depletion of the other components through organic and inorganic reactions as the lake contracted. These reactions, coupled with the escape of solutes by leakage, caused an estimated 97-percent depletion of dissolved solids during the 25-month period.
INTRODUCTION LOCATION OF STUDY AREA
This study concerns several lakes that occupy topographically closed basins within the northwestern lobe of the Great Basin in south-central Oregon, and one (Goose Lake), in the usually closed northeastern extremity of the adjoining Sacramento River basin, astride the Oregon-California boundary (fig. 1). At high-water levels, the lakes have a combined surface area of about 490 square miles; their drainage basins cover about 10,000 square miles.
PURPOSE AND SCOPE OF REPORT
The study, culminating in this report, was made to bring together all available hydrologic and geochemical data on several topographically enclosed lakes and their tributaries; to show the variations of water level, area, volume, and chemical character of the lakes within historic time; and, insofar as practicable, to correlate such variations with climatic or other causal factors. The report, then, represents a reconnaissance effort to evaluate the generalized hydrologic and geochemical regimen of the selected lakes.
The hydrologic study included determination of the approximate magnitudes of the water-budget elements for each lake—precipitation, runoff, and spring flow as sources of supply, and evaporation, overflow, and possible leakage as means of disposition. Some of the quantities were measured directly; some were ascertained by correlation with published records of precipitation or runoff; some were computed as residuals in the water budget; and the others could only be estimated.
Geochemical aspects of the study included determination of the chemical character of the lakes and their inflow, and evaluation of the possible sources of dissolved solids and the mechanisms of geochemical evolution.
The hydrologic part of the study was made by Kenneth N. Phillips, assisted by Roy B. Sanderson, then district engineer for the Surface Water Branch of the Water Resources Division, and members of his staff in Portland, Oreg. Hydrologic data through September 30, 1963, were used. The geochemical part of the study, including chemical analysis of most water samples, was conducted by A. S. Van Denburgh. Field sampling was done between April 1961 and June 1963.
ACKNOWLEDGMENTS
Many individuals were helpful in the study. The writers are especially grateful to Glenn E. Tyler and P. M. Smith, watermasters for the Oregon State Engineer, who rendered valuable assistance in the field; A. B. Claggett, superintendent of the Summer Lake State Game Management Area, for information on water levels of Silver and Summer Lakes; S. T. Harding, consulting engineer, for supplying the results from his field observations and his unpublished notes on the hydrology of Abert and Goose Lakes; and Dr. D. B. Lawrence, for results from his observations of tree rings and water levels at Lake Abert.
Within the U.S. Geological Survey, George T. Hira-shima checked most of the hydrologic computations; Keith V. Slack identified and commented upon specimens of Lake Abert biota, and many other members of the staff contributed valuable suggestions. Allen G. Hely and Blair F. Jones were principal reviewers of the manuscript.
GENERAL FEATURES OF THE LAKES AND THEIR BASINS
Among the lakes studied, only four—Abert, Summer, Harney, and Bluejoint—occupy impermeable basins deep enough to be truly closed. Such water bodies neither leak nor overflow; consequently, they lose water only by evaporation. All the other lakes studied—Goose, Silver, Malheur, Crump, and Hart—are freshened by rare to frequent overflow. The hydrologic and chemical data for Silver Lake suggest that it leaks as well.
General features of the lakes and their basins are summarized in table 1.
PHYSIOGRAPHIC AND GEOLOGIC SETTING
All the lakes lie within a semidesert, volcanic terrane, whose skyline is dominated by many fault-scarp ridges. Prominent among the scarps are Winter Ridge, overlooking Summer Lake on the west; Abert Rim, bordering Lake Abert to the east; the Warner Mountains, east of Goose Lake; and Hart Mountain, east of Warner basin (pi. 1). These elevated blocks form small mountain ranges whose axes trend generally north south. Between these ranges lie the lake basins.
The lakes themselves are characteristically shallow, and they occupy broad, extremely flat basin floors that are underlain by thick accumulations of generally fine grained sedimentary material. Several lakebeds are so flat that a sudden, strong wind can cause a large part of the lake water to be temporarily spread out over many acres of peripheral playa surface at altitudes slightly higher than the undisturbed lake level.B4
CLOSED-BASIN INVESTIGATIONS
Figure 1.—Location of lakes discussed in this report.
CLIMATE
The study area lies in a belt of prevailing westerly winds. Conventional air movements give rise to local gusts at times. Air temperatures in the area are like those of other high plateaus in the Western United States, ranging
from more than 90°F in summer to less than — 10°F in winter. Table 2 shows the seasonal variation in average monthly temperature at Valley Falls, which is generally representative of conditions on the valley floors. In clear weather, the nighttime loss of heat by radiation is rapid,HYDROLOGY AND GEOCHEMISTRY, CLOSED-BASIN LAKES, OREGON Table 1.—Summary of physical, hydrologic, and chemical data for the lakes
B5
Lake Abert Summer Lake Goose Lake Silver Lake Harney Basin Warner basin
Malheur Lake 1 Harney Lake Crump Lake Hart Lake Bluejoint Lake
Drainage area, including lake.sq mi.. 860 390 1,140 520 2,150 5,280 700 900 1,920
Altitude (ft):
Overflow . . 4,390 4,390 4,716 4,310.4 4,093.5± 4,114± 4,474.4 4,473 4,800 +
Low point in lake bed.. .... 4,243.5+ 4,144+ 4,692 4,299 + 4,088.8 4,079.7 4,468 4,463.5 4,455
Highest observed lake level2 4,260.5 4,149 4,707.8 4,309.0 4,095.4 4,087.6 4,476.4 4,474.5 (3)
Lowest observed lake level2. .. Dry Dry Dry Dry Dry Dry 4,470.6 Dry Dry
Highest known Pleistocene lake
level . . ... 4,520 4,520 4,720± 4,520 4,114 + 4,114 + 4,750 to 4,800
Lake area at high level2 sq mi . 64 4 71 181 16± 103 47 12+ 12 (3)
Lake volume at high level2
103 acre-ft.. 496 (3) 1,270 60 ± 240 176 (3) 56 (3)
Estimated average annual quantity:
Lake-surface precipitation..in... 12 11 14 11 9 9 10 8 8
Evaporation in... 39 40 42 38 42 41 44 44 44
Inflow... . _ _ _ 103 acre-ft . 55 65 165 (3) 100± 60± 70± 80± (3)
Overflow 2 . . 10s acre-ft . 0 0 0 0 20 ± 0 60 ± 30 ± 0
Dissolved-solids content:
Maximum determined ppm_. 95,000 36,900 2,700 698 5 484 622,400 500 5 781 6 3,630
Minimum determined ppm__ 18,700 1,800 600 300 5 258 (6) 101 6 230 (•)
Principal constituents 7 Na Na Na (8) CaNa Na CaNa Na Na
C1C03 CICO3 IICO3 hco3 Cl hco3 HCOs HCO3
HCOs
1 Area and volume shown for Malheur Lake at high levels include Mud Lake.
2 For period 1924-63.
8 Not determined.
* Measured within meander line (est alt 4,152-4,155 ft).
6 Only two reliable determinations are available.
6 Only one reliable determination is available.
7 On basis of equivalents per million. Listed constituents are calcium (Ca), sodium (Na), chloride (Cl), carbonate (COa), and bicarbonate (HCOj).
8 Variable.
and the daily range in temperature often exceeds 50°F. Frost may occur on any clear night in summer.
Relative humidity in summer also has a high diurnal range—typically 10 to 30 percent at midday, and approaching 100 percent at night. Evaporation is high on the clear, windy days typical of the dry period from May to October, but is low at night because of the marked drop in temperature and corresponding rise in relative humidity. The evaporation loss from lake surfaces is increased by occasional strong winds that spread the water over wide areas of flat beach. Table 1 summarizes estimated lake-surface evaporation rates.
Table 2.—Monthly temperature and precipitation at Valley Falls, Or eg.
[Data from records of the U.S. Weather Bur.]
Precipitation (in.)
temperature 1914- -63 Average for 1931-60
Maximum Minimum
January 29.7 3.51 0.20 1.41
February 33.1 3.07 .10 1.31
March 38.1 2.23 .11 1.12
April 45.1 2.90 .13 .96
May 51.6 3.91 .00 1.43
June 57.7 5.53 .00 1.38
July 65.9 2.28 .00 .30
August 63.8 2.61 .00 .29
September 56.7 2.71 .00 .57
October 47.9 6.13 .00 1.11
November 37.6 2.71 .00 .99
December 32.6 3.76 .20 1.50
Year average 46.7 0) (2) 12.37
1 Maximum annual precipitation during the period 1914-63, recorded in 1963, was 18.46 in.
8 Minimum annual precipitation during the period 1914-63, recorded in 1924, was 6.43 in.
Average annual precipitation in the study area ranges from more than 20 inches at high altitudes within several of the tributary areas to 8 inches or less in the more arid parts. The amount may vary greatly in short distances because of changes i;n altitude or topographic orientation with respect to prevailing winds. Table 2 shows the seasonal pattern at Valley Falls, Oreg. Much of the precipitation occurs as snow, and runoff is delayed until melting takes place, commonly in the spring. Lake-surface precipitation ranges from 8 to 14 inches in an average year (table 1), and is a significant part of the total water supply.
The frost-free season is short, which limits agricultural activity. Livestock is pastured on open desert or mountain range during the summer and is fed locally grown hay in winter. The hay crops are produced mostly on lands irrigated with water supplied from surface streams, but ground water is also used in places.
INFLOW TO THE LAKES
The quantity of surface runoff into the lakes differs widely. Over a long period, most of the water entering Abert, Crump, Goose, Malheur, and Silver Lakes is runoff from mountainous areas that are 1,000 to 5,000 feet higher than the lakes. In dry years, however, the streams dwindle to small flows and (or) the water is diverted for irrigation before reaching the lakes. Bluejoint, Harney, and Hart Lakes are indirectly dependent on surface runoff because most of their supply consists of overflow from upstream lakes.
In contrast, most of the inflow to Summer Lake is supplied by a prolific spring system. Much smaller springsB6
CLOSED-BASIN INVESTIGATIONS
and seeps occur along or near fault zones beside Abert, Goose, Harney, and Summer Lakes. Their temperatures are well above the mean annual air temperature. Some of these springs, though too small to be important sources of water supply to the lakes, may nevertheless provide a significant amount of the incoming dissolved-solids load.
The surface- and ground-water flow entering the lakes cannot be measured accurately; therefore, records of precipitation, runoff, and lake level were used to estimate the approximate inflow (table 1). The probable accuracy of such estimates and the relative importance of such inflow as a part of the gross water supply for the lakes are summarized below:
Average annual inflow Probable
Lake (percentage of gross accuracy
water supply) (percent =fc)
Abert______________________ 65 10
Crump______________________ 95 25
Goose______________________ 65 20
Hart_______________________ 95 25
Malheur____________________ 60 25
Summer_____________________ 70 15
OUTFLOW FROM THE LAKES
At present, evapotranspiration dissipates the entire inflow to Abert, Bluejoint, Goose, Harney, and Summer Lakes and a large proportion of the inflow to the other lakes (the proportion is least for Crump Lake and is greatest for Silver Lake).
No leakage occurs from Abert, Goose, Harney, and Summer Lakes, as all these water bodies overlie beds of fine-grained, virtually impermeable lacustrine deposits. Also, they are bordered by many peripheral springs and seeps—evidences of ground-water gradients that slope toward the lakes, rather than away from them.
Silver Lake apparently loses some water by leakage. The lack of peripheral springs and the absence of excessive mineralization in the lakebed soil are consistent with the indications of loss by seepage. Crump, Hart, and Malheur Lakes may lose some water by lateral seepage into peat deposits that border parts of each lake.
Some lakes never approach the level of surface overflow under present-day conditions. For example, the divide between Abert and Summer Lakes is more than 100 feet higher than either lake has been in historic time (table 1). At the other extreme are the lakes, such as Crump and Malheur, which overflow during most wet years. The other lakes studied fall somewhere between the two extremes. Goose and Silver Lakes have had virtually no overflow since 1868 and 1905, respectively, and future occurrences are doubtful under present conditions of climate and agricultural land use.
WATER-LEVEL FLUCTUATIONS
The water levels of the lakes fluctuate in response to changes in water supply and water loss. The peak of the annual cycle typically occurs in the spring, near the end of the period of greatest precipitation and runoff. During the following dry, warm summer, lake levels recede, reaching their annual minimum by the late summer or early autumn. In periods of greater than average precipitation and runoff, lake levels gradually rise until evaporation losses from the increased surface areas equal the augmented supply. In protracted periods of dryness, the water levels fall until some lakebeds are completely dry.
In this century the lake levels, with but few exceptions, have been highest during the early 1900’s and the 1950’s, j and lowest (many lakes went dry) in the 1920’s and 1930’s.
| A general indication of extremes of wetness and dryness during the last few centuries is provided by studies of tree growth. Records of tree growth since 1455 were published by Antevs (1938, p. 66-85), along with historical records of precipitation, runoff, and lake level. Periods of maximum and minimum tree growth and of precipitation match well, except for some periods of retarded tree growth that may have resulted from insect attack, disease, or other unusual factors. Since 1455 the greatest rates of tree growth through 1934 occurred in the periods 1790-92, 1805-25, and 1907-09; the lowest growth rates were in 1842-49 and 1918-34. According to a similar study by Keen (1937), the period 1918-34 was the driest in at least 650 years.
During Pleistocene time, large lakes occupied the deep Abert-Summer, Warner, and Fort Rock basins, where they left shoreline deposits'more than 200 feet above the present-day lake beds. During the same period, lakes in the much shallower Goose and Harney basins doubtless overflowed into the Pit and Malheur Rivers, respectively.
CHEMICAL CHARACTERISTICS
Chemically, the lake waters range from highly dilute— fresher than water used for irrigation and municipal supply throughout much of the United States—to highly saline. In general, the dilute waters are those of lakes freshened by an overflow that ranges from rare to almost seasonal, whereas the most saline waters are those of true closed lakes, where only the water, but not its solute content, is lost by evaporation.
The character of components dissolved in the lakes and their inflow is also wide in range. The more dilute water bodies contain mostly silica, calcium, sodium, and bicarbonate, whereas the highly saline accumulations consist mostly of sodium, carbonate bicarbonate, and chloride. The wide range in chemical character is recorded in table 3 and is shown graphically in figure 2, in which diagrams depict (1) the distribution of major dissolved constituents,HYDROLOGY AND GEOCHEMISTRY, CLOSED-BASIN LAKES, OREGON
B7
Dissolved No. Sample site solids (parts per million) Dissolved No. Sample site solids (parts per million)
LAKE ABERT BASIN l__Chewaucan River near Valley Falls _ _ 94 2 do 193 3 Lake Abert 55,900 SUMMER LAKE BASIN 4 Ana River at gage - 158 5 Summer Lake 7,200 GOOSE LAKE BASIN 6 Goose Lake 1,270 WARNER BASIN 7 Deep Creek at gage _ 85 8 Honey Creek at gage _ _ 135 9 Crump Lake 322 10 Hart Lake 781 11 Bluejoint Lake 3,630 SW Sea water 34,500
Figure 2.—Relative abundance of major dissolved constituents of lakes and their tributaries in south-central Oregon. Solid symbols represent lakes; matching-shaped open symbols represent the respective tributary streams.
by showing points on two trilinear plots, and (2) the overall chemical character of the water, by showing a single point on a diamond-shaped grid.
The characteristic seasonal pattern of changes in dis-solved-solids concentration is approximately as follows: Snowmelt runoff during late winter and early spring increases the lake volume and dilutes the more saline lake water, while adding a small increment of dissolved solids (for several of the lakes, the runoff during some years is sufficient to cause overflow, which removes solutes from the lake). During the hot summer and early fall, high evaporation rates and a lack of significant inflow cause a
lowering of lake level and an increase in dissolved-solids concentration—that is, water is lost, but the salts remain. The evaporation rate decreases greatly during the cold, late fall and winter. This decrease, in combination with small amounts of inflow and often significant amounts of precipitation on the lake, causes a stationary or a rising lake level, as well as a relatively constant or a slightly decreasing dissolved-solids concentration.
For a shallow lake that neither overflows nor dries frequently, the short-term (several-year) relation between the dissolved-solids concentration and the lake level is nearly constant at most times because the shallow waterB8
CLOSED-BASIN INVESTIGATIONS
Table 3.—Relative amounts of major and minor constituents in representative samples from Abert, Summer, Goose, Hart, and Crump Lakes
[Amounts of dissolved solids and of hardness reported in parts per million. Amounts of all other constituents reported as a percentage of total dissolved solids; maximum and minimum percentages for each are in italic]
Lake Summer Goose Hart Crump
Abert Lake Lake Lake Lake
Dissolved-solids content____ 40,800 7,200 1,270 781 322
Silica (Si02)............... 0.88 1.5 4.6 4.5 12
Calcium (Ca)_............... <.005 .04 1.0 6.7 11
Magnesium (Mg)______________ <.005 . 004 -32 3.8 4.8
Sodium (Na)_________________ 40 39 35 25 17
Potassium (K)_______________ 1.8 1.6 2.8 2.0 2.0
Bicarbonate (HC03)*......... 5.8 13 29 33 38
Carbonate (COs)-............ 16 17 7.4 5.5 1.9
Sulfate (S04)............... 1.8 4.8 6.5 12 8.1
Chloride (Cl)............... 35 22 12 7.9 6.2
Fluoride (F)................ .01 .08 .07 .27 .25
Bromide (Br)_............... .17 .11 .07 ................
Orthophosphate (PO4)-------- .14 -25 .54 ------- *31
Boron (B)___________________ .14 '47 -30 ------- --------
Hardness as CaCOa........... 8 8 49 252 144
pH.......................... 9.7 9.6 9.1 8.8 8.4
Sampling date............. 4-26-61 4-25-61 6-12-62 4-27-61 4-27-61
* Calculated as carbonate.
body remains almost homogeneous, and the dissolved tonnage is virtually constant. The relationship is temporarily destroyed during periods of abundant inflow and precipitation on the lake surface, when the dilute incoming water is stratified above the more saline lake water. This stratification remains until wind action again homogenizes the lake-water body.
In lakes that frequently overflow or are dry, the relation between lake level and dissolved-solids concentration changes with time because the salts are depleted during overflow and are either lost or temporarily stored during desiccation. Figure 3 shows schematically the effect of occasional overflow: the accumulation of salts in a lake during periods when the basin remains closed is offset by loss of salts during overflow, and the lake thus tends to remain dilute. A similar illustration could be used to show the solute-depletion effect of occasional desiccation.
EFFECTS OF MAN’S ACTIVITIES
The hydrologic and chemical character of a lake may be artificially altered in several ways by the activities of man. For the lakes in the study area, some activities can be eliminated from consideration. For example, no attempts at weather modification were made, and there are no transbasin diversions of streamflow. Water is diverted only from Hart Lake, and no sewers or mineralized industrial effluents discharge into any of the lakes or their tributaries. The use of saltlicks, fertilizers, insecticides, and herbicides is confined to small areas and probably has had a minor effect on the total tonnage of dissolved salts in the lakes. The activities to be considered, therefore, are storage on tributaries, diversion for irrigation, and drainage of swampland.
In the basin of Lake Abert there is, as yet (1968), no upstream storage reservoir. Drainage of the Chewaucan Marsh during the period 1901-15 released some water which previously would have been lost by evapotranspira-tion from the swampy area, but the concurrent development of irrigation probably resulted in consumptive use of about an equal amount, so that the average inflow to Lake Abert was not changed significantly. In a series of dry years, however, Lake Abert might now receive less inflow than it did under natural conditions; therefore, it is more likely to become dry than it was before the marsh was drained and irrigation was practiced. The same general situation is true of the Harney and Warner basins, as well.
In Goose Lake basin, Drews Reservoir (storage capacity, 62,500 acre-ft) and Cottonwood Reservoir (storage capacity, 7,540 acre-ft) were built in 1911 and 1921, respectively. Several small reservoirs and many stock ponds also exist. Evaporation from these reservoirs and diversion of streamflow may deplete the average inflow to Goose Lake by about one-third. Because of this depletion, the lake is not likely to repeat the brief overflows to the Pit River that occurred in 1868 and 1881; thus, it may now be a true closed-basin lake.
\ \
o
z
CD
<
L±J
cr
o
z
o
<
\
\
X
DISSOLVED-SOLIDS CONCENTRATION INCREASING
Figure 3.—Relation between dissolved-solids concentration and stage of a lake that occasionally overflows. A-B-C, seasonal cycle; D, lake reaches overflow stage; D-E, lake overflows, losing dissolved solids; E, termination of overflow, and start of normal cycle. X, dissolved-solids increment, gained prior to overflow; and Y, dissolved-solids decrement, lost during overflow.HYDROLOGY AND GEOCHEMISTRY, CLOSED-BASIN LAKES, OREGON
B9
The inflow to Silver Lake is materially reduced in most years by upstream storage and irrigation. Although Silver Lake has not overflowed to Thorn Lake since 1905, a slight extension of the wet period 1952-58 would have raised the lake to the level of its former overflow channel. Hence, the depletion of potential inflow has been effective in maintaining a dry lakebed in a period of drought (1922-50), but was much less effective in limiting the rise of the lake in a period of great runoff (1952-58) simply because the reduction in wet years is a smaller part of the total runoff, and because the lake volume at overflow is small (about 75,000 acre-ft).
Normally, the level of Summer Lake is now several feet lower than it was 40 to 50 years ago. Part of that reduction is apparently due to a natural decrease in flow from the springs that feed the Ana River, but most of the change seems to be the result of man’s activities, including the increased agricultural irrigation, the flooding of a wildfowl-management area, the drilling of artesian wells that tap the aquifer feeding the Ana River, and back pressure on the large springs that were flooded by completion of the irrigation dam in 1923.
In such basins as those of Summer, Goose, and Silver Lakes, where development and construction by man have reduced the amount of streamflow runoff, the lakes characteristically contain larger concentrations of dissolved solids than they would under natural conditions.
LAKE ABERT
Lake Abert is a large, shallow body of water that occupies the lowest part of an 860-square-mile closed basin
Figure 4.—Southward view from the east shore of Lake Abert, January 12, 1962. Snow-covered Abert Rim in background. Vegetation boundary indicates high water level of 1958. Dark areas near lakeshore show peripheral seeps melting recently fallen snow.
Table 4.—Approximate area and volume of Lake Abert
[Data are based chiefly on soundings made May 19 and 20, 1959, when lake level was at an altitude of 4,258.9 ft, on outline of the lake at lower levels, as shown by aerial photographs, and on recent topographic mapping]
Altitude (ft) Area (acres) Volume (acre-ft) Altitude (ft) Area (acres) Volume (acre-ft)
4,243.5 + 0 0 4,253 35,300 206,000
4,244 240 100 4,254 36,300 242,000
4,245 8,000 3,200 4,255 37,300 279,000
4,246 12,400 13,300 4,256 38,200 316,000
4,247 19,100 28,900 4,257 39,000 355,000
4,248 24,800 50,900 4,258 39,700 394,000
4,249 28,100 77,300 4,259 40,300 434,000
4,250 30,500 107,000 4,260 40,800 475,000
4,251 32 ,400 138,000 4,261 41,300 516,000
4,252 33,900 171,000
about 25 miles north of Lakeview (pis. 1, 2). Slopes around the lake range from nearly flat to precipitous. At the southern end the Chewaucan River, the lake’s principal tributary, leaves the plain of Chewaucan Marsh and drops about 12 feet over a fault scarp to the lake level. West of the lake the surface of a tilted fault block slopes upward to Coglan Buttes, whereas to the north the mud flats near the lake merge into a hilly terrane. Along the entire east shore, steep talus slopes and near-vertical volcanic rock faces rise to the nearly flat crest of Abert Rim, 1,500 to 2,200 feet above the lake and only 1 mile distant from it (fig. 4). The spectacular cliffs and steep slopes of the narrow tributary belt fringing the east shore are an expression of faulting on a grand scale.
AREA, VOLUME, AND LAKE-LEVEL FLUCTUATIONS
At high stages, Lake Abert is about 16 miles long and., 6 miles wide, and has a maximum depth of more than 15 feet. The lakebed configuration is shown in figure 5, and the area and volume of the lake at various altitudes are listed in table 4.
Lake-level records (table 20) are fragmentary prior to 1950. Since 1950 the level has been observed several times each year (fig. 6), and from September 1961 to May 1963, a continuous record of stage was made. Some levels were deduced from general descriptions of the areal extent and from photographs of the lake. The extreme range since 1843 is well substantiated, as follows:
1. The lake was completely dry in the summer of 1924 for
the first time since the area was settled in the 1870’s, and it was nearly or completely dry in 1926, 1930, 1931, 1933, 1934, and 1937.
2. The lake level rose to an altitude of 4,260.5 feet above
mean sea level in June 1958—the highest level attained since the area was settled by permanent residents. All herbs, shrubs, and trees with root crowns at or below that level were killed by flooding (%• 4).
The shores of Lake Abert were searched for trees whose growth records might give some indication of the lakeBIO
CLOSED-BASIN INVESTIGATIONS
120° 15' 120°10' 120°05'
Figure 5.—Lakebed contours of Lake Abert. Dotted contour, indicated as inferred in explanation, is from preliminary topographic mapping.HYDROLOGY AND GEOCHEMISTRY, CLOSED-BASIN LAKES, OREGON
Bll
Figure 6.—Seasonal variations in stage of Lake Abert, June 1950-June 1963. (Altitude of the lowest point in the basin is 4,243.5+ feet above mean sea level.) Arrows indicate times when samples were collected for chemical analysis (table 28). Dashed parts of hydrograph indicate periods for which lake levels are inferred.
levels prior to the first regular observations. The only such trees are a few isolated junipers along the southeast shore. Those with root crowns at or below an altitude of 4,260.5 feet were drowned by the saline water when the lake rose to that level in 1958. The largest drowned tree— whose root crown was at an altitude of 4,260.3 feet—was cut down, and the annual growth rings on the stump were counted. The ring count showed that the tree had germinated in the year 1903, or shortly before. Cores taken
Figure 7.—Northeastward view from the southeast shore of Lake Abert, July 3, 1963. Dashed line represents maximum lake level, at 4,260.5 feet, which was attained in June 1958. Herbs and shrubs (but no trees) below (or left of) that line have germinated and grown since the saline water receded. Stump in foreground is that of an older generation tree (juniper?) that died 250 (±200) years ago.
from other nearby, slightly smaller trees, at a level 0.4 foot higher, showed that they had germinated about 1908 to 1910. The largest, oldest living tree near the lake measured 21 inches in diameter in 1959, and its base was at an altitude of 4,264.3 feet. A core taken 1 foot above the tree base showed 64 annual rings; hence, that tree germinated before 1895. Certainly, the lake level was lower than the bases of these trees during, and immediately preceding, their lifetimes.
A solitary, older generation tree stump 3 feet high was found adjacent to the lake, with its root crown at an altitude of 4,260.2 feet. The stump, which was measured 11 inches above the ground surface, was 13 to 14 inches in diameter (fig. 7). A cross section of the stump showed 2 inches of decay in the center, as well as 26 wide, uniform growth rings outside the decayed area. The tree was estimated to have been only about 35 years old when it died. The tree apparently grew in a period when ground moisture was adequate for its needs, and it may have been drowned by submergence in the early 1800’s or in the 1860’s. A carbon-14 age determination, made in 1964, showed that it had died about 250 (±200) years ago (U.S. Geol. Survey lab. sample W-1398). The presence of this stump at the altitude of 4,260.2 feet and the similarity of its annual growth rings to those of present-generation trees growing nearby indicate that Lake Abert was below that level for a period of at least 35 years when the inflow was not affected by man’s activities.
A distinct beach line of rounded coarse gravel and cobbles at an altitude of 4,269.7 feet marks a still-stand of Lake Abert that is 9.2 feet higher than the highest observed level. The beach line occurs at many places along the eastB12
CLOSED-BASIN INVESTIGATIONS
Figure 8.—Northward view from east shore of Lake Abert near site of lake gage. Rod is held at edge of the well-defined beach line (alt 4,269.7 ft), formed sometime prior to 1895.
shore (fig. 8). Notes of the Eastern Oregon Railroad survey, in 1907, along the west shore refer to a high water level at an altitude of 4,269 feet (railroad survey adjusted + 10.7 ft to 1929 standard datum), presumed to be the same strand line.
Sage, greasewood, other desert vegetation, and a few juniper trees have become established below the high beach. The oldest cored trees below the line germinated about 1895, but the beach line was probably formed long before that. The cobbles are not weathered, and in many places no soil has accumulated over them. Their appearance suggests youth, and the high lake level they mark may well have occurred within the last two centuries. Unusually rapid tree growth near Lakeview during the period 1805-25 indicates that precipitation in that period was greater than any period of equal length in the last four centuries (Antevs, 1938, p. 66), and Lake Abert may then have formed the prominent beach line.
Fragments of beaches or strand lines, not yet covered by windblown dust or by talus, are present at levels several feet higher than the well-defined beach at an altitude of 4,269.7 feet, the higher beaches being less well defined and apparently somewhat older than the lowest one. Such beaches have been noted on the east shore near the present lake gage at altitudes of 4,273.8 and 4,276.3 feet, and 3 miles farther north, at 4,271.2 and 4,273.2 feet. Their age is not known, but it is probably not more than a few thousand years.
Beach deposits far above the present-day lake attest to the existence of a much larger lake in this basin in Pleistocene time (pi. 1). Allison (1946, p. 64) referred to that body of water at its high level as Lake Chewaucan, and stated (1954, p. 1331) that it attained a maximum altitude of 4,520 feet. At that level, the present Chewaucan
Marsh, near Paisley, was covered by water more than 200 feet deep, and the water surface was almost 280 feet above the present bed of Lake Abert and 370 feet above that of Summer Lake. Lake Chewaucan at its maximum level covered 480 square miles, or 38 percent of the combined drainage basins (1,250 sq mi at present, and probably about the same then). Even at its maximum altitude of 4,520 feet, Lake Chewaucan did not overflow.
WATER SUPPLY
The water of Lake Abert is derived from four sources— precipitation on the lake, small peripheral springs, ephemeral streams that drain arid areas fringing the lake, and the Chewaucan River. None of these sources can be measured precisely, but the total water supply can be approximated fairly accurately.
LAKE SURFACE PRECIPITATION
Extended records of precipitation near the lake are available only at Valley Falls (alt 4,326 ft), which is 10 miles south of midlake (pi. 2). Annual precipitation at Valley Falls has averaged about 12 inches during the period 1915-63. Precipitation on the lake itself may be a little less than at Valley Falls, but for the studies herein, it is assumed to be about the same.
SPRINGS
Many springs emerge along the shores on all sides of the lake (figs. 4, 5). Most of them are mere seeps that maintain small spots of green growth or saturated ground, making no material contribution to the water supply of the lake; others discharge up to about 1 cfs (cubic feet per second). The largest spring is about 4 miles north of Lake Abert (32/21-26D). Its measured flow was 1.04 cfs on July 23, 1950, and 1.96 cfs on November 13, 1962. Its water and that of other seeps and springs irrigate hay crops on lands bordering the lake, and a part of the flow is thus consumptively used. The total combined flow of all seeps and springs reaching Lake Abert, along with that of the minor local surface streams, averages an estimated 10 cfs, or 7,000 acre-feet per year. That estimate is based on an inspection of the springs in 1962 (total observed spring flow, 4.9 cfs), when the lake level was about 4,251 feet in altitude, coupled with a study of inflow data and changes in lake level.
No seeps of significant size were thought to enter the lake below an altitude of 4,251 feet at the time of inspection in 1962. No such seeps were reported by local residents, who had seen the lakebed dry in the late 1920’s and early 1930’s. However, one such observer—Bert Harber of Lakeview—did describe the flow of springs at the Pike Ranch, along the northeast shore, as a wide wet streakHYDROLOGY AND GEOCHEMISTRY, CLOSED-BASIN LAKES, OREGON
B13
that meandered southward for several miles across the lakebed until dissipated by evaporation and seepage (oral commun., 1962).
The many springs that rise along the great fault zone near the northeast shore have, within a distance of 8 miles, a combined flow of about 3 cfs (1962), far more than is to be expected from the small semiarid area topographically tributary to them. At sites 102 and 103 (pi. 2; table 30), the spring flow temperatures were, respectively, 66° and 71° F, about 20° F higher than mean air temperature. On the opposite shore, other springs (sites 104, 106) in a lesser fault zone flow at temperatures of 68° and 66° F. The discharge of all these springs apparently rises from considerable depth. The similarity in temperature and the lack of adequate drainage area to account for those near the northeast shore suggest a common source, perhaps outside the area topographically tributary to the lake.
STREAMS
In most years the Chewaucan River supplies most of the water reaching Lake Abert. At the gaging station near Paisley, where the river leaves the forested mountains, the flow was measured during water years 1913-21 (Oct. 1912 to Sept. 1921) and water years 1925-63. The average flow for those 48 years of record is 136 cfs (98,500 acre-ft per yr). For this study, the average annual flows for water years 1922-24 were estimated, on the basis of records for Silvies River near Burns and Camas Creek near Lakeview, as 80, 60, and 25 cfs, respectively. The average flow for the 51-year period 1913-63, determined from these estimates, is 131 cfs (94,900 acre-ft per yr).
The drainage area of the Chewaucan River is 275 square miles at the gaging station near Paisley, 430 square miles at the former gaging station at Hotchkiss Ford, and 490 square miles at the mouth. As of 1963, diversions from the reach between gaging stations supplied irrigation water to 32,000 acres, and diversions below Hotchkiss Ford, to about 8,700 acres. The principal crop is wild hay.
Table 5 shows the depletion of streamflow between the gaging stations during 1915-18 and 1920, caused chiefly by irrigation diversions.
Table 5.—Annual discharge of Chewaucan River near Paisley and Chewaucan River at Hotchkiss Ford, 1915-18 and 1920
[Drainage area: Chewaucan River near Paisley, 275 sq mi; Chewaucan River at Hotchkiss Ford, 430 sq mi]
Discharge of Chewaucan River
(acre-ft) Depletion
Water year ------------------------------------- (acre-ft)
Near At Hotchkiss
Paisley Ford
1915 ______________________ 63,800 15,300 48,500
1916 ______________________ 106,000 51,700 54,300
1917 _____________________ 118,000 65,600 52,400
1918-______________________ 57,700 23,100 34,600
1920________________________ 45,400 7,990 37,400
The use of water from the Chewaucan River for irrigation began about 1884; drainage of marshes and irrigation development continued until about 1915. As yet, no upstream storage reservoirs have been built on the river, but one is now (1968) under serious consideration. Such storage development would decrease the average annual inflow to Lake Abert.
Streamflow into Lake Abert was not directly measured. Records of Chewaucan River near Paisley are the only long-term data in the basin. Between that gaging station and the lake, evapotranspiration from Chewaucan Marsh significantly reduced the inflow to Lake Abert prior to the beginning of irrigation. The amount of that natural depletion is not known, but over a period of years it was probably similar to the average depletion for the presently irrigated 41,000 acres. The present net depletion may be about 1.6 acre-feet per acre irrigated. In many dry years the supply available is not adequate to irrigate the entire 41,000 acres, just as the supply was in many years not adequate to flood the entire marsh under natural conditions. Thus, the average annual discharge now reaching Lake Abert probably is about as great as it was under natural conditions. In periods of drought, however, the small available flows may now be controlled and consumed more effectively by irrigation than they were by dissipation in the natural marsh, thereby increasing the frequency with which the lake dries during such periods.
The average annual inflow to Lake Abert from the Chewaucan River may be about 48,000 acre-feet (51 percent of the quantity measured 32 miles upstream, near Paisley), on the basis of water-budget computations (p. B15).
WATER LOSS
EVAPORATION
Evaporation from Lake Abert was computed by the mass-transfer method described by Harbeck (1962). The basic equation is
E = Nu(e o—e„),
in which
E = evaporation rate, in inches per day;
N = coefficient of proportionality (termed “the mass-transfer coefficient”); tt = windspeed, in miles per hour; e0 = saturation vapor pressure, in millibars, for the water-surface temperature; and ea = vapor pressure of the airmass, in millibars.
The difference in vapor pressure (eo—ea) is commonly expressed as Ae. The product uAe for any period is an index of evaporation loss.
For periods with no precipitation and only minimal inflow, the evaporation loss from the lake (E) is equal toB14
CLOSED-BASIN INVESTIGATIONS
Table 6.—Monthly wind movement, temperature, relative humidity, and evaporation losses at Lake Abert, and temperature and precipitation
at Valley Falls, October 1961 and April to October 1962
[Wind movement measured at midlake 2 meters above water surface; temperature of lake at surface measured at midlake. Air temperature and relative humidity
measured 1 mile south of lake]
Month Wind movement (mi) Average temperature (°F) Relative humidity (percent) Computed Valley Falls (3 mi SSW of lake)
Total Average per hour At lake surface Air loss (in.)1 Temperature (°F) Precipitation (in.)
1961
October 2_ 6,750 10.4 50.2 47.5 61.7 3.53 46.0 0.66
1962
April3 3,390 11.8 54.0 48.9 60.2 4.66 49.8 .64
May - . - 6,760 9.1 54.0 51.2 67.2 3.50 50.8 3.91
June_ - 5,500 7.6 65.3 61.3 57.5 5.44 59.2 .08
July 4,820 6.5 71.5 63.4 52.6 7.00 64.3 .00
August -- - -- 6,420 8.6 67.6 62.2 51.6 7.69 64.6 .23
September 5,050 7.0 63.9 58.4 52.4 5.19 59.9 .16
October . 5,960 8.0 50.8 46.1 79.6 2.09 48.8 6.13
1 Not adjusted for lake-water density. 3 Data except those for Valley Falls cover period April 19-30 only; values are
2 Data except those for Valley Falls cover period October 1-27 only; values are estimated for remainder of month, estimated for remainder of month.
the fall in lake level, adjusted slightly for the effects of inflow. If windspeeds (u) and vapor-pressure differences (Ae) for the same periods are known, the relation between E and uAe can be defined and the coefficient N determined. Then, the evaporation can be computed for any period for which windspeed and vapor-pressure data are available.
The following data w'ere collected for Lake Abert:
1. A continuous record of lake stage at a point on the east
shore 10 miles northeast of Valley Falls.
2. Computed or estimated inflow from the Chewaucan
River and from springs.
3. A continuous record of wind movement 2 meters above
the lake surface at the midlake raft (fig. 9; data summarized in table 6).
4. A continuous record of lake-surface temperature at the
raft (summarized in table 6).
5. Continuous records of temperature and relative
humidity of the air 1 mile south of the lake and 2 miles northeast of Valley Falls (table 6).
To ascertain the mass-transfer coefficient, N, for Lake Abert, the average daily fall in stage (AH) was taken from the recorder trace for selected periods when there was little or no precipitation and comparatively little inflow to the lake. These values of AH, adjusted for effects of inflow, are equivalent to the loss by evaporation and were plotted against the product of average windspeed and difference in vapor pressure, uAe, for concurrent periods. The resulting plot (fig. 10) defines a straight line on arithmetic coordinate paper that passes practically through the origin, with a slope (N) of 0.00216. This value of N was rounded to 0.0022 for the computing of the monthly evaporation losses listed in table 6.
For the 6-month period of complete record, May-October 1962, the total computed evaporation loss was 30.9 inches. That value must be corrected for effects of lake-water density, which averaged about 1.05 g/ml (grams per milliliter) at 20° C during the period. Assuming about a 1 percent evaporation reduction for each 0.01 density unit above 1.00 (C. H. Lee, in Harbeck, 1955, p. 1), the 30.9-inch computed value would be equivalent to 32.5 inches from a fresh-water body. The 6-month total for 1962 was then adjusted to an average full-year value by comparison with U.S. Weather Bureau records for the Medford Experiment Station, 135 miles to the west—the nearest site with a long-term year-round record (1943-63). There, the measured evaporation for May-October 1962 was 78.6 percent of the full-year value, which
Figure 9.—Anchored instrument raft on Lake Abert, September 17, 1962. Abert Rim is in background. Midlake raft housed equipment for measuring wind velocity and lake-surface temperature, as part of evaporation study.HYDROLOGY AND GEOCHEMISTRY, CLOSED-BASIN LAKES, OREGON
B15
in turn was 102.3 percent of the 21-year average. On the basis of these data, the estimated long-term fresh-water evaporation rate at Lake Abert would be about 40.3 inches (3.36 ft) per year. Assuming the average density to be about 1.04 grams per milliliter at 20° C, the actual lake-surface evaporation rate is probably about 38.7 inches (3.22 ft) per year.
LEAKAGE
Lake Abert does not leak. The lakebed is lower than any other adjacent valley floor except that of Summer Lake. Furthermore, Lake Abert is surrounded by springs and seeps (fig. 5) that indicate a ground-water gradient toward, rather than away from, the lake.
OVERFLOW
The lowest topographic divide, which separates the Abert and Summer Lake basins 2 miles north of Paisley, is about 140 feet higher than the bed of Lake Abert. No overflow has occurred there within historic time. During the Pleistocene period, however, inflow to Lake Abert was sufficient to cause overflow, which filled the adjacent Summer Lake basin, forming Lake Chewaucan. At its maximum level, the large Pleistocene lake covered the present-day divide to a depth of about 130 feet. Overflow from Lake Abert has occurred since that time, as evidenced by a channelway that meanders generally northwestward from the divide to an altitude of about 4,330 feet. The
o 200 -
^ 7
/
/
/^1956
/
1958"
//>'
1952.
/■{?
A?
1953.
5 100 -
50 -
19S4w
/*1951
/
/
/
/ .1957
/
/
1955. I!**/
1959./1961 I //.1960 I
50 100 150 200 250
ANNUAL STREAMFLOW (S), IN THOUSANDS OF ACRE-FEET
300
Figure 11.—Relation between annual streamflow of Chewaucan River near Paisley and computed inflow to Lake Abert, water years 1951-62.
time and quantity of the most recent overflow to Summer Lake by way of this channelway are unknown.
WATER BUDGET FOR THE LAKE
A generalized relationship between the annual stream-flow, or discharge, of Chewaucan River near Paisley and the total annual inflow to Lake Abert has been defined by a study of the water budget of the lake for the water years 1951-62. For those years, the altitude of the lake surface on September 30 is known or has been closely approximated by interpolation between observations. The annual inflow was computed as equal to the change in volume of the lake (computed from water-level data and table 4) plus the volume of water evaporated (3.22 ft, multiplied by the average surface area) minus the volume of precipitation on the lake (observed precipitation at Valley Falls, multiplied by the average area).
The net yearly inflow so computed for water years 1951-62 is plotted against the concurrent yearly discharge at the station near Paisley in figure 11. The relationship in that figure is expressed by a straight-line equation:
Annual inflow (/), in acre-feet
= (1.25) [Paisley streamflow (S) —55,000 acre-ftj.
Figure 10.—Relation between fall in stage (AH) and the product uAe at Lake Abert near Valley Falls during May-October 1962. Symbols: u = average wind velocity, in miles per hour; Ae = difference, in millibars, between vapor pressure of saturated air over lake water and vapor pressure of incoming air. Slope of line (AH/uAe) is mass-transfer coefficient (N).
However, the inflow is assumed to be never less than 7,000 acre-feet per year because of the peripheral springs and local runoff. This relationship may not give accurate results for any given year, but it probably provides a fairly accurate estimate over a period of several years.B16
CLOSED-BASIN INVESTIGATIONS
Figure 12.—Annual flow of Chewaucan River near Paisley and computed inflow to Lake Abert, water years 1913-63.
The equation was used to compute the annual inflow to Lake Abert in the period 1913-63 (fig. 12). For the entire period, the computed annual inflow averaged about 55,000 acre-feet. Of this amount, about 48,000 acre-feet per year may have been contributed by the Chewaucan River.
This generalized water budget was tested for validity by synthesizing a record of the stage of Lake Abert for May 31 (near-maximum) and September 30 (nearminimum) of each year from the first accurately known lake stage, at an altitude of 4,154.2 feet on October 4, 1915, to September 30, 1963. In this computation, the lake was assumed to have been dry whenever the stage fell to an altitude of 4,243.5 feet. No additional adjustments were applied to make the synthesized record agree with later observations. According to this computation, the lake should have been dry or practically dry in 1924, 1926, 1930, 1931, 1933, 1934, and 1937, which is in complete agreement with a news account of 1924 and with the oral statements of reliable local witnesses. The computed annual low stage for September 30 was compared with the known lake stages in 18 years of the 48-year period,
and the greatest difference was 1 foot, in 1961. The good agreement between computed and observed lake stages warrants the conclusion that the computed values are probably within 1 foot of actual lake stages in years when no observation was made. Extension of the estimates to periods before 1915 is not possible, however, because earlier records of precipitation at Valley Falls and runoff at Paisley are lacking. The computed record of lake-level fluctuation since 1915 is shown graphically in figure 13.
The budget equation also shows that Lake Abert could again reach the level at an altitude of 4,269.7 feet indicated by distinct beach deposits (p. Bll), provided that there is no large additional upstream diversion or storage. At the high level, the lake covers about 50,000 acres, and its volume is about 1 million acre-feet. The net evaporation loss would offset an average annual inflow of 110,000 acre-feet, which would be attained with an annual discharge at Paisley of 143,000 acre-feet—about 50 percent more than the average flow in the period 1912-63. If the water year 1956 ( 250,000 acre-ft of discharge at Paisley) had been followed by 4 more years of equally heavy dis-HYDROLOGY AND GEOCHEMISTRY, CLOSED-BASIN LAKES, OREGON
B17
£ 1915
o
1920
1925
1930
1935
1939
LU
Figure 13.—Variations in the stage of Lake Abert, 1915-63. Dashed line connects water levels on May 31 and September 30, computed by annual water-budget method and projected from the observed water level on October 4, 1915. Dots represent observed water levels, obtained from gage reading or by leveling; small circled crosses represent less accurate observations.
charge, Lake Abert would have risen above the old high level. The average annual discharge at Paisley in the period 1951-58 (156,000 acre-ft) would also have raised the lake to that level if continued for an additional 15 years.
GEOCHEMISTRY
GENERAL CHEMICAL CHARACTER OF THE LAKE
In 1963 Lake Abert contained about 13 million tons of dissolved solids, covered an area of about 55 square miles, and was the largest landlocked saline water body in the Pacific Northwest. The dissolved-solids content of the lake fluctuates considerably, but generally ranges from 20,000 to 80,000 ppm. The three most abundant dissolved constituents—sodium, carbonate, and chloride—make up about 90 percent of the dissolved solids. Potassium, bicarbonate, and sulfate account for more than 9 of the remaining 10 percent. Although silica, bromide, orthophosphate, and boron constitute less than 1 percent, these constituents occur in large parts-pet-million concentrations. (See table 3.) Concentrations of calcium and magnesium in the lake are unusually small; combined, they probably remain less than 5 ppm at most times.
Among the trace elements, iron, aluminum, vanadium, molybdenum, nickel, lead, and cobalt are present in measurable parts-per-billion concentrations, generally in the above-listed order of abundance (table 29).
VARIATIONS IN DISSOLVED-SOLIDS CONCENTRATION
The dissolved-solids content of samples collected from Lake Abert since 1882 has ranged from 18,700 ppm (July 8, 1958) to 95,000 ppm (July 21, 1939). The estimated (unrecorded) minimum salt content during the entire 80-year period—about 18,000 ppm—occurred in 1958 during the highest historically observed lake stage (alt 4,260.5 ft above mean sea level). In contrast, amounts greater than 100,000 ppm doubtless occurred during periods of near dryness between 1924 and 1937.
Except at near-dryness stage, the relation between dissolved-solids concentration and lake level is almost constant for Lake Abert over periods of several years, such as 1958-62 (fig. 14). This is because of the nearly unchanging solute tonnage in the lake during such periods (table 7) and because the shallow water body remains virtually homogeneous when inflow is negligible. (For example, note data on pi. 2.)
Concurrent records of lake stage and dissolved-solids concentration for Lake Abert prior to 1939 are not available. Hence, accurate comparisons of dissolved-solids tonnages in the lake before and after the long period of near dryness (1924-37) are not possible. Table 7 summarizes available tonnage information subsequent to 1938. The data show that the quantity of salts in the lake increased by about 75 percent—from about 8 million tons to almost 14 million tons—during the 22-year period 1939-61. Apparently, most of the accumulation occurred betweenB18
CLOSED-BASIN INVESTIGATIONS
Table 7.—Dissolved-solids tonnage in Lake Abert, 1939-68
Lake level Lake Dissolved-
Date Analysis No. (ft above volume solids Load
(table 28) mean sea (103 content (10® tons)1
level) acre-ft) (ppm)
7- 21-39........... 5a 4,248.2 56 95,000 7.8
8- -52............. 5b 4,252.7 196 35,000 9.6
8- 12-55........... 2a 4,252.8 198 40,400 11.2
7- 9-58............ la 4,260.5 496 18,700 12.8
1-29-59............ lb 4,259.2 442 20,500 12.5
5- 20-59........... 6a 4,258.9 430 21,700 12.8
4-26-61____________ lc 4,253.86 237 40,800 13.6
6- 27-61............ Id 4,253.31 217 45,000 13.7
9- 7-61____________ le 4,252.04 174 55,400 13.7
1-12-62............ If 4,251.81 165 55,900 13.1
3- 13-62___________ lg 4,252.14 177 52,000 13.0
4- 18-62___________ lh 4,252.30 181 52,000 13.3
9-17-62____________ 8a 4,250.70 129 71,000 13.1
10-30-62____________ lj 4,251.19 145 63,000 13.0
3-19-63____________ lk 4,253.08 208 45,000 13.1
1 Error probably less than about 0.5 million tons. Additional analyses are available (table 28), but accuracy of analytical results is questionable because sample-collection sites were subject to dilution by inflow from the Chewaucan River or from peripheral seeps.
1952 and 1958, when the lake rose from an altitude of about 4,250 feet to almost 4,261 feet (equivalent to a fivefold volume increase and an almost 45-percent lake-area expansion). During the 13-year period prior to 1952, the lake stage had only fluctuated from an altitude of about 4,246 feet to about 4,252 feet. During the present study (Apr. 1961-Mar. 1963), the salt tonnage in solution has remained almost constant, averaging 13.3 million tons (table 7). Possible reasons for accumulation of the large dissolved-solids increment (about 4 million tons) between 1952 and 1961 are discussed in the section on sources of dissolved solids.
Despite a lack of data on solute loss from Lake Abert during times of desiccation between 1924 and 1937, the major solute losses from Summer and Goose Lakes (p. B28, B34) imply a similar occurrence at Lake Abert as well. However, if the changes in chemical character of Summer and Goose Lakes were largely the result of salt depletion during periods of desiccation (p. B28), the lack of chemical change in Lake Abert (table 9) may mean that loss of salts from the Abert playa between 1924 and 1937 was not great. This possibility is further supported by the large tonnage increase in the lake between 1939 and 1961. The great increase implies that much of the salts given up by the lake during desiccation remained accessible for later re-solution, and were not removed from the vicinity of the playa, or even from the basin.
CHANGES IN CHEMICAL CHARACTER
The amounts of many dissolved constituents in Lake Abert have remained virtually unchanged relative to one another during the period of study, despite the large variations in dissolved-solids content. However, the relative (percentage) concentrations of two major constituents, several minor constituents, and all the identified trace elements fluctuated significantly. Among the major
Figure 14.-—Relation between dissolved-solids concentration and stage of Lake Abert, July 1958-October 1962.
constituents, the amounts of carbonate and bicarbonate vary relative to one another, even though their combined concentration (calculated as carbonate) remained about 22 percent of the dissolved-solids content. This trend is characteristic of Summer and Goose Lakes as well, and it is discussed in more detail on page B45.
Another characteristic of the three adjacent lakes is a seasonal variation in the relative concentrations of silica and orthophosphate; the variations are different from lake to lake, however. On the basis of 2 years of record at Lake Abert (during which the measured concentrations of silica and orthophosphate ranged from 130 to 201 ppm and from 55 to 101 ppm, respectively), relative amounts of the two nutrients apparently tend to reach peak values during the summer (June-August) when water temperatures are highest (fig. 15). The smallest relative amounts were found between December and March. The range in fluctuation of silica relative to dissolved-solids content is wide—from 0.248 to 0.385 percent between April 1961 and October 1962, equivalent to a range from 31,000 to 51,000 tons of silica.1 * The larger percentages occurred at times of lesser dissolved-solids content. Fluctuations in the relative
1 Percentages are easily translated into tonnages, because, if the percentage
for any constituent and the total solute tonnage in the lake remain constant regardless of a changing concentration, the constant percentage is equivalent to a
constant tonnage. For example, Lake Abert now contains about 13 million tons of dissolved solids. If silica were to remain at 0.248 percent of the dissolved-solids concentration, even though that concentration changed from 30,000 to 60,000 ppm because of a 50-percent lake-volume decrease, the silica tonnage would remain at 31,000.HYDROLOGY AND GEOCHEMISTRY, CLOSED-BASIN LAKES, OREGON
B19
o
GO
2 0.14
■ 0.13
ciO.40
UJ
>
o
co
2 0.36 o
Li_
o
o 0.32
<
t—
z
UJ
o
£ 0.28 -Q_
CO
<
5 0.24 -
Dissolved solids / _ / -•_______a /
A
/ \
\
\
\
\
\
\
/ \
\
/ \ / •
Temperature
\
/
\
/
/
\
\
/
Orthophosphate^/
/ \
\
/
/
S
/
\
\
4
/.
/
/ J
<
/
\
N.
/
Silica x*"
'•------
80 „ UJ UJ
70 cc o
UJ
60 Q QZ
40 = 30 £
Figure 15.—Changes in dissolved-solids concentration, temperature, and relative amounts of silica and orthophosphate in
Lake Abert, April 1961-July 1963.
amount of orthophosphate in the lake are smaller and more uniform from season to season than those for silica; the estimated seasonal range (fig. 15) was from 0.132 to 0.144 percent orthophosphate, equivalent to a variation from 16,800 to 18,300 tons. The orthophosphate fluctuations closely follow those of water temperature in Lake Abert—a correlation that does not apply for Summer Lake (fig. 24) or Goose Lake (fig. 28).
Seasonal fluctuations of silica and orthophosphate in the three lakes are discussed in more detail on pages B48 and B51.
Changes in the absolute (parts per billion) and relative (percentage) amounts of trace elements in Lake Abert are erratic (tables 8, 29). No consistent pattern of either seasonal fluctuation or variation with changing dissolved-
Table 8.—Relative (percentage) amounts of jive trace elements in Lake Abert and the Chewaucan River
(All data from table 29. Relative amounts expressed as a percentage of dissolved-solids content, multiplied by 105]
Chewaucan
Lake Abert River near
Element Valley Falls
January 12, June 12, September 17, June 12, 1962 1962 1962 1962
Cobalt (Co)__________ <0.4 <1.0 1.2 78
Lead (Pb)_____________ <.4 <1.0 3.2 <75
Molybdenum (Mo)__ >8.6 2.6 4.4 21
Nickel (Ni)___________ g.l 3.5 2.4 150
Vanadium (V)______ >8.6 16 2:6.6 300
Dissolved-solids content, in parts per
million____________ 55,800 49,100 71,200 193B20
CLOSED-BASIN INVESTIGATIONS
Table 9.—Relative abundance of major dissolved constituents in Lake Abert, 1912 and 1962
Percentage of total
Constituent dissolved-solids content
February 1912 1 January 12, 1962
Sodium (Na)__ . . . 38.9 39.3
Potassium (K) 1.7 1.3
Carbonate (CO3) plus bicarbonate (HCOs)2 20.8 22.2
Sulfate (S04) _ 1.9 1.8
Chloride (Cl) 36.1 34.5
Dissolved-solids content, in parts
per million .. 329,600 55,900
1 Data after Van Winkle (1914, p. 119, analysis 4).
2 Calculated as carbonate.
* Corrected for estimated concentrations of fluoride, bromide, orthophosphate, and boron; bicarbonate recalculated as carbonate; dissolved-solids content of sample may have been significantly less than lake average at time of collection, owing to dilution by inflow at south end of lake.
solids concentration is evident. Some of the apparently erratic change may be the result of variable reliability of the data, as discussed on page B56.
No major long-term net changes were noted in the relative abundance of anions at Lake Abert, whereas such changes did occur in Summer Lake (table 12) and Goose Lake (table 14). Analyses of samples collected from Lake Abert in 1912 and 1962 (table 9) show only minor net changes, perhaps within limits of analytical error. The data suggest that sulfate plus chloride may have decreased from 38.0 to 36.3 percent of the dissolved-solids concentration during the 51-year period; this decrease represents only 4 percent of the original value, in contrast to increases of 29 and 21 percent for Goose and Summer Lakes during a comparable period.
BIOLOGIC CONTROLS ON CHEMICAL QUALITY
The biologic assemblage in Lake Abert doubtless exerts controls on, and is also controlled by, the amounts of certain constituents in the lake water, such as silica, calcium, magnesium, sulfate, nitrate, orthophosphate, and probably several of the trace elements. Brine shrimp (Artemia salina) abound in the water during certain periods of the year, and at such times one can hardly dip a single cupful of the lake water without getting several specimens. In early autumn, the lakeshores are lined with small windrows of larval skins discarded by emerging adult brine flies. In summer vast numbers of algae (Claclophora) float freely in the water and cover much of the lake bottom, their filaments clustered into balls formed by the action of waves in the warm shallow water. Several other kinds of plankton, including diatoms and Anabaena, may also be present in significant populations. The important inorganic chemical constituents of three of the above biologic organisms are summarized below (data after Welch, 1952, p. 274, 275, 304). No information is available regarding the major inorganic constituents of the brine shrimp, but such crustaceans probably contain
significant amounts of phosphorus and calcium, plus lesser amounts of silica and magnesium (K. V. Slack, written commun., 1962).
Organism Percentage of dry weight of sample
Ash Silica (SiOs) Phosphorus (as PO<) Calcium (Ca) Magnesium (Mg)
Cladophora 1 26.5 7.1 1.0 2.4 1.0
Diatoms 2 . 39.5 30.8
Anabaena 8 7.2 1.0 1.6 1.0 .4
1 Welch (1952, table 29, p. 304).
2 Welch (1952, table 22, p. 274); collected at Lake Mendota, Wis. 8 Welch (1952, table 23, p. 275); collected at Lake Mendota, Wis.
The chemical data suggest that assimilation of certain constituents during periods of population growth, and at least partial release of these constituents after death, may affect the amounts dissolved in the lakes. However, the variations in orthophosphate tonnage, and to a lesser degree silica tonnage (as shown by relative quantities in fig. 15), seem to contradict results expected on the basis of incomplete knowledge of biologic activity in the lakes. The possible importance of biologic influences on seasonal fluctuations of silica and orthophosphate are discussed in more detail on pages B48 and B51.
The amount of sulfate in Lake Abert may be influenced by anerobic sulfate-reducing bacteria within the lake-bottom muds. This subject is discussed briefly on page B46.
SOURCES OF DISSOLVED SOLIDS
The Chewaucan River and precipitation are, by far, the two most important sources of water for Lake Abert, but they may not be the most important source of dissolved salts. Numerous small springs rise along or near the periphery of the lake (figs. 4, 5), and although their combined discharge is small (an estimated 10 cfs), their total dissolved-solids contribution to the lake and adjacent playa may be more than that from surface inflow and precipitation. A significant amount of salts may also be contributed to the lake by wind transport of alkali dust from other parts of the basin and from outside the basin. In addition, the recovery of salts lost during periods of lake dryness or of near dryness is a significant short-term source.
The several potential sources of dissolved or dissolvable salts are discussed below.
CHEWAUCAN RIVER
Unfortunately, the only available chemical-quality data for the river near its mouth are analyses of 16 samples collected during the period January 1959-November 1962 (analyses 14a-14p, table 28). For these analyses, the relation between specific conductance, in micromhos, andHYDROLOGY AND GEOCHEMISTRY, CLOSED-BASIN LAKES, OREGON
B21
dissolved-solids content, in parts per million, is expressed by the equation:
Dissolved solids = (0.583 X specific conductance)+25.
The dissolved-solids concentration in the Chewaucan River near its mouth is poorly related to discharge (fig. 16). Nonetheless, the data suggest that the average concentration may be near 120 ppm. The principal constituents of the river, regardless of the amount of dissolved solids, are silica, calcium, sodium, and bicarbonate. However, as the concentration of dissolved solids increases, the amounts of sodium, sulfate, and chloride increase relative to calcium and bicarbonate, whereas the relative percentage of silica decreases considerably (fig. 17). The calcium content, in parts per million, is generally three to four times that of magnesium (fig. 18).
Much of the dissolved-solids load carried by the ] Chewaucan River apparently is derived in the mountainous j area upstream from Paisley. Data from three samples j collected in April 1962 (table 10) show that, although dissolved-solids content increased by 67 percent in the 31-mile reach from the gaging station above Paisley to the river mouth near Valley Falls, the discharge decreased by 58 percent; therefore, the load decreased from 71 to 49 tons per day. Most of the decrease was due to loss of silica. In contrast, the amount of chloride increased from 0.2 to 2.6 tons per day in the same reach. Thus, most of the chloride (a major constituent in Lake Abert) in the Chewaucan River accumulates between Paisley and the mouth. Windblown alkali dust, metabolic animal-waste 1 products, and, perhaps, flow from springs are principal sources of the chloride. In addition, residual salts from pluvial Lake Chewaucan may also contribute some chloride to the river.
Table 10.—Changes in chemical character of Chewaucan River \ between Paisley and Valley Falls
Chewaucan River station
Above At The Near Valley
Paisley Narrows Falls
River miles upstream from
mouth1____________________ 32 16 1
Date of collection__________ 4-16-62 4-16-62 4-18-62
Discharge______________cfs_. 419 194 2175
Dissolved solids:3
Parts per million_______ 65 70 105
Tons per day____________ 73 36 49
Silica (Si02):
Parts per million_______ 30 30 32
Tons per day____________ 33 16 15
Sodium (Na):
Parts per million_______ 4.9 6.4 15
Tons per day___________________ 5.5 3.3 7.1
Chloride (Cl):
Parts per million_______________ .2 .8 5.5
Tons per day____________________ .2 .4 2.6
1 Approximate; does not include the numerous meanders between Valley Falls and The Narrows.
2 Estimated discharge; others are measured.
8 Estimated using specific conductance.
Figure 16.—Dissolved-solids concentration versus estimated discharge of Chewaucan River near Valley Falls. Triangle indicates estimated average discharge (66 efs) and dissolved-solids concentration (120 ppm).
The chemical contrast between water of the Chewaucan River and of Lake Abert shows that most of the incoming silica, calcium, and magnesium, as well as lesser proportions of bicarbonate, must be removed from solution soon after mixing with the lake water. For example, the annual contribution of silica to Lake Abert from the river probably is about 2,000 tons. Yet, the entire silica load in the lake has not exceeded 51,000 tons in the period of intensive study, and the river alone seems to deliver that much in less than 30 years. The possible mechanisms for depletion of silica and other constituents are discussed in the section entitled “Geochemical Trends and Processes” (p. B42).
The relations between trace-element concentrations in the Chewaucan River and in the lake (table 8) are uncertain because the analytical results are of questionable accuracy (p. B56) and, also, because only one analysis is available for the river water. Nonetheless, the data in table 8 were recalculated to provide an order-of-magnitude comparison of quantities in the lake and its principal inflow:
Chewaucan Theoretical
Element Lake Abert 1 River2 number of years
(lb, rounded) (lb per yr, to accumulate
rounded) Lake Abert load
Cobalt (Co)______________ 3,000 100 30
Lead (Pb)________________ 8,000 <100 >80
Molybdenum (Mo)_______ >20,000 30 >700
Nickel (Ni)______________ 9,000 200 40
Vanadium (V)____________ 40,000 500 80
1 Based on largest percentage values in table 8.
2 Based on percentage values in table 8, and an estimated average annual load of about 8,000 tons.
The calculations suggest that large amounts of incoming trace elements are removed from solution. This is indicatedB22
CLOSED-BASIN INVESTIGATIONS
20
LlI
o
<
o o:
LlI
0-
00
<
5 -
Figure 17.—Relation between dissolved-solids concentration and the relative amounts of silica, sodium, and chloride in discharge of Chewaucan River near Valley Falls. Solid circles represent silica; solid triangles, sodium; open circles, chloride.
by the very short replenishment period that would theoretically be necessary for accumulation of each constituent in the lake.
PERIPHERAL SPRINGS
Most water from springs adjacent to the lake (fig. 5) chemically resembles a dilute version of the lake water; sodium, bicarbonate, and chloride are the principal constituents, even though the dissolved-solids content is less than 1,000 ppm (analyses 104 and 106, table 30). The resemblance is not as great for peripheral spring water that contains less than about 300 ppm of dissolved solids (analyses 101 and 105, table 30); and the only available analysis of nearby well water shows almost no chemical resemblance to the lake (analysis 114, table 31).
A part of the dissolved solids in peripheral spring flow may represent salts recycled from the lake or from lacustrine deposits underlying the lake. This is suggested by the large dissolved-solids content in the discharge from two of
the sampled springs that are closest to the lakeshore (analyses 104 and 106, table 30), and by the similarity of chloride percentages in these two spring waters and in the lake (32 and 34 percent for the spring waters, compared with 35 percent for Lake Abert; all other peripheral ground-water samples contained less than 25 percent chloride). If an appreciable amount of the dissolved solids contributed by peripheral spring flow does represent recycled salts, then the Chewaucan River may be the primary source of solutes now present in Lake Abert.
The solute contribution of springs adjacent to Lake Abert is difficult to assess because (1) the combined flow rate of the springs has been estimated during only one period (July-November 1962), and the rate may vary considerably with fluctuations of annual precipitation; (2) the overall chemical character of the spring flow can be only approximated, mostly on the basis of analyses of six samples collected June 9, 1962; (3) the amount of dissolved solids in the spring discharge that actuallyHYDROLOGY AND GEOCHEMISTRY, CLOSED-BASIN LAKES, OREGON
B23
Figure 18.—Relation between dissolved-solids concentration and amounts of calcium and magnesium in discharge of Chewaucan River near Valley Falls.
represents recycled salts from the lake and lakebed is uncertain; and (4) the amount of spring-derived soluble material removed from the northern playa by wind before reaching the lake is unknown. Nonetheless, data in table 30 suggest that the annual contribution of major lake constituents (sodium chloride and carbonate-bicarbonate) by the peripheral springs may be greater than that of the Chewaucan River, even though the total spring discharge is far less than the streamflow.
PRECIPITATION
Almost no information is available on the chemical character of precipitation in the Lake Abert basin. Only one example of precipitation, a fresh-snow sample, has been collected within the basin. Because the snow cover was sampled at the same time as the nearby lake, the amount of dissolved solids in the snow sample may reflect slight contamination during handling of the two samples; yet, the snow sample contained only 4 ppm of dissolved solids—principally sodium, bicarbonate, and chloride. The snow presumably fell after the air had been cleared of dust by earlier precipitation; thus, the dissolved-solids content may be similar in magnitude to that of
incoming precipitation unaltered chemically by the alkali environment.
Rain and snowfall within the Lake Abert basin undoubtedly contains more than 4 ppm of dissolved solids at many times during the year, owing to the pickup of soluble substances in atmospheric dust. Although some of these solutes may represent recycled material derived within the basin, much undoubtedly is brought into the basin by air movements. Nonetheless, the direct solute contribution of precipitation on the lake surface certainly must be far less than that of the streams and peripheral springs.
WIND-TRANSPORTED SALTS
The most difficult parameter to evaluate, even semi-quantitively, is the net contribution of windblown salts. Alkali dust clouds have been observed traveling east into the Abert basin from adjacent Summer Lake (R. W. Childreth, U.S. Geol. Survey, oral commun., 1962). However, violent winds also pick up large amounts of salts from the flat playa north of Lake Abert, and, although much of this material is dropped within the basin, an unknown amount is carried beyond the basin. Also, the rate of eolian salt removal undoubtedly is accelerated during periods of low lake stage, when broader lakebed areas are exposed to wind action. Thus, an average annual gain (or loss) of salts by wind action cannot be estimated.
RE-SOLUTION OF LAKEBED SALTS
Most of the 75-percent solute-tonnage increase in Lake Abert during the period 1939-61 (p. B17) probably results from recovery of salts formerly isolated from the lake during the 14-year period of dryness and near-dryness (1924-37). The subject of solute storage within playa deposits peripheral to Abert and Summer Lakes is planned to be covered in considerably greater detail in a paper on the solute balance in the combined basin (Van Denburgh, unpub. data).
OTHER SOURCES OF DISSOLVED SOLIDS
Several other sources of dissolved solids must be considered, even though their average annual contributions are insignificant compared with those from the Chewaucan River and peripheral springs. Poison Creek, which heads on the Abert Rim plateau and courses down the steep scarp to the lake (location 35/21-1; pi. 2), is the only peripheral stream other than the Chewaucan River that supports even a small discharge during much of the year. Surface flow at the U.S. Highway 395 culvert is usually less than 1 cfs, but underflow in the broad alluvial fan augments the surface discharge by rising in an almost continuous line of seeps on the fan toe adjacent to the lakeshore. Chemical analyses of streamflow samplesB24
CLOSED-BASIN INVESTIGATIONS
collected at the culvert in June and October 1962 (analyses 15a and 15b, table 28) show that the chemical character of Poison Creek is similar to that of Chewaucan River above Paisley throughout the year.
Overland flow during short periods of heavy rainfall may be a significant contributor of salts to the lake, especially because of the alkaline soils throughout most of the basin. However, the magnitude of contributed solute tonnage is unknown. The flow of ephemeral streams doubtless contributes salts to the lake either directly as inflow or indirectly by percolation to the ground-water aquifers feeding the peripheral springs.
SUMMER LAKE
Summer Lake is a shallow body of water that lies on the surface of a deep valley fill of alluvial and lacustrine deposits in a closed fault-block basin. The fill is reported to be more than 1,000 feet deep in places (Brown, 1957, p. 4). Southeast of Summer Lake, near Paisley, the basin is separated from Chewaucan Marsh by the lacustral-alluvial fan of the Chewaucan River (pi. 1). To the north it is separated from Fort Rock valley by a fault-scarred ridge that trends east-west. The south and west sides of the basin are formed by a great composite fault scarp that rises steeply along the crest of Winter Ridge (Walker, 1963). Less prominent scarps occur along the east edge. Winter Ridge is timbered; other parts of the basin are arid and support only desert-type vegetation.
AREA, VOLUME, AND LAKE-LEVEL FLUCTUATIONS
Summer Lake is very broad and shallow. At high stages (alt 4,149-4,151 ft) the lake is about 15 miles long, north to south, and about 5 miles wide, covering an area of about 60 square miles, yet its maximum depth is only 5-7 feet. Within the meander line (alt 4,152-4,155 ft) the area of the lake is about 45,000 acres; the corresponding volume may be 300,000 acre-feet or more. At an altitude of 4,146 feet, the lake area is about 25,000 acres, and its estimated volume about 25,000 acre-feet. At any stage, and especially at low levels, the wetted area may suddenly be increased significantly by the gusty winds that frequently sweep over the basin.
The fragmentary records of lake level are given in table 21, and seasonal fluctuations during the period 1959-62 are shown in figure 19. The records have been obtained almost entirely by leveling from reference marks. At times the lake level is affected by wind.
The highest stage of record (alt 4,151.4 ft) occurred in February, March, and April 1905, a time when precipitation and runoff were abundant, and the use of water for irrigation had not been fully developed. The highest stage attained in recent years through 1963 was at an altitude of
about 4,149 feet in the spring of 1952, when the water level was just below the shoulder of State Highway 31 at the “sunken grade” southwest of the lake (A. B. Claggett, Oregon State Game Comm., oral commun., 1961).
At times the lake has been practically dry. The water area was reduced to about 10 acres in the late summer of 1950, and again in August 1961 (A. B. Claggett, oral commun., 1961) and in 1962. The lowest level actually observed was at an altitude of 4,144.86 feet on September 30, 1961; at lower stages, leveling is impracticable because of the wide mudflats around the lake. The lowest part of the lakebed probably is slightly above 4,144 feet in altitude, which is lower than any other point within about 70 miles.
At many sites around the lake, well-preserved shoreline deposits at altitudes up to 4,520 feet (370 ft above present lake level) attest the existence in Pleistocene time of a much larger pluvial lake, known as Lake Chewaucan (p. B12).
WATER SUPPLY
Summer Lake is fed by lake-surface precipitation, small streams that course down the steep slopes west and south of the lake, intermittent runoff from the east, and many springs, both large and small.
LAKE-SURFACE PRECIPITATION
Records of precipitation have been kept since April 1957 by the U.S. Weather Bureau at a station in the shadow of Winter Ridge, 1 mile south of Summer Lake Post Office and about 1 mile northwest of the lake. For the years 1958-62, the average annual precipitation was 12.4 inches. In these same 5 years, the precipitation at Valley Falls was about 3 percent more than the long-term average, suggesting that the average at the Summer Lake station is probably about 12.0 inches per year. The lakewide average is probably a little less—perhaps about 11 inches.
SPRINGS
The largest and least variable source of inflow to Summer Lake is a group of springs called Ana Springs, which rise north of the lake in at least five separate clusters (location 30/17-6E) and feed the Ana River, a stream about 5 miles long. According to Brown (1957, p. 1), “The waters of the springs rise by artesian pressure from lava-rock aquifers through about 150 feet of overlying thinly bedded lake deposits * * The springs may be fed by ground-water flow from the adjacent Silver Lake valley (p. B39).
An unpublished engineering report by Barr and Cunningham, Engineering Consultants, in 1920 indicates that the five large springs have their openings into the channel of the Ana River at the following altitudes: springs 1 and 2, 4,198 feet; springs 3-5, 4,175 feet (theHYDROLOGY AND GEOCHEMISTRY, CLOSED-BASIN LAKES, OREGON
B25
Figtjke 19.—Seasonal variations in the stage of Summer Lake, 1959-62. Lettered points correspond to samples from site 16 (table 28).
Dashed where approximated.
base level used presumably was the sea-level datum prior to the 1929 adjustment, which probably would make the altitudes in error by less than 1 foot relative to present-day datum). Several years prior to 1920, a temporary dam was built below spring No. 2 to raise the water level about 10 feet. That dam was still in existence in 1920, and the flow at the dam was then thought by Barr and Cunningham to be undiminished by the submergence.
In 1922-23 an earthen dam with a metal outlet pipe near
Table 11.—Measured discharge of Ana River near Summer Lake between 1904 and 196%
[Data not available for all years]
Year Month Discharge (cfs) Year Month Discharge (cfs)
1904__ July, November » 164 1929. _ February 100
1905.. March, April, June.. i 148 1930.. February-March. » 96
1906.. April 141 1931 3 89
1909.. January, May i 142 1932 3 91
1910.. September — 141 1933 3 92
1912.. February- 140 1934 3 91
1914.. August, December— » 134 1935 3 88
1915.. May, October i 133 1936 3 93
1917.. 132 1937 3 97
1918.. April 124 1938 3 100
1919.. March, August. » 120 1951 . 3 84-4 94
1920.. March, August i 114 1952 . 3 85-4 91
1921.. March, August, 1953 . 3 86-4 90
October 1 127 1954 . 3 83-4 87
1922.. April 4— 130 1955 _ 3 87-4 95
August 17 2 106 1956 . 3 81-4 91
1923.. January 29 2 109 1957 . 3 86-4 97
1924.. September 124 1958 . 3 90-4 102
1925._ January 126 1959 _ 3 86-4 98
1926.. January— _ 2 109 1960 _ 3 90-4 97
1927.. 91 1961 3 83-4 93
1928.. April 106 1962 . 3 90-4 99
1 Average of two or three gagings.
2 Springs submerged by pool upstream from temporary dam at times in 1922-23, and continuously from permanent dam since 1926.
3 Average summer flow; includes diversion in canal. Springs under full pressure from reservoir.
4 Average flow the following winter; no flow in canal. Reservoir level about 7 feet lower than in summer season.
its base was built across the Ana River below the five springs to divert part of the flow into Summer Lake Canal for irrigation (Brown, 1957, p. 6). Shortly thereafter, slumping of the earthfill caused collapse of the outlet pipe, and a new and higher pipe was installed in 1926. As constructed and now operated, the pool over the springs is maintained at about 4,221 feet in altitude from May to September each year, and at about 4,214 feet from October to April. The pool therefore produces a hydraulic back pressure on the springs of 16 to 39 feet in winter and 23 to 46 feet in summer.
The combined discharge of the five springs, as measured below the dam and including the flow in Summer Lake Canal, is given in table 11. Records are fragmentary through 1950, and are continuous after June 1951. For the 12 water years of complete record, 1952-63, the combined discharge averaged about 91 cfs (66,000 acre-ft per yr).
The data in table 11 indicate that submergence of the springs reduces their flow. A rather abrupt drop of about 20 cfs accompanied first submergences by the pool in 1922 and in 1926. From 1951 to 1962, the flow in November and December was always 4 to 12 cfs (avg, 9 cfs) more than in the preceding summer season. This difference is due to the lowering of the pool level by about 7 feet after the irrigation season each year. The observed decrease in flow of Ana Springs may be partly offset by an increase in flow of other springs, but which springs are thus affected has not been determined.
Downstream from the dam and the gaging station, several other small springs appear in the Ana River channel. The effect of the dam on the flow of these springs is not known.
Flow of the Ana River is not affected by seasonalB26
CLOSED-BASIN INVESTIGATIONS
variations in local precipitation, but it doubtless responds slowly to long-term changes in precipitation. For example, table 11 shows a gradual decrease in discharge from about 160 cfs in 1904 to about 125 cfs 20 years later—presumably due mostly to natural causes.
The records since 1926 in table 11 include flow in a large diversion canal at the dam, several hundred feet upstream from the point where discharge measurements are made. That canal diverts about 11,000 acre-feet per year, mostly in the period May to September. Several smaller agricultural diversions are downstream, the largest one supplies irrigation water to about 1,000 acres. Considerably larger diversions have been made to maintain a suitable marsh habitat for wild ducks and geese in an area reclaimed from the bed of Summer Lake, above the level of the lake in recent years. That wildfowl area is flooded during the entire year, as much as a third of the riverflow being diverted for that purpose (A. B. Claggett, Oregon State Game Comm., written commun., 1964). The fresh water of the Ana River has been used to flush out salts from this portion of the former lakebed since the 1920’s or 1930’s. This use is possible because the lake now occupies a smaller area than it did when less of the tributary flow was used for irrigation.
In addition to those springs which feed the Ana River, there are many smaller springs scattered throughout Summer Lake basin. Many of them occur along the faultline west of the lake. At the Thousand Springs Ranch, 6 miles southeast of Ana Springs, many seeps supply the water used to irrigate fields of wild hay. Most of the smaller springs are used for irrigation; hence, in the summer they do not contribute materially to Summer Lake. South of Summer Lake there is at least one small hot spring, with a flow of about 0.05 cfs, which Waring (1908, p. 55) called Woodward Spring. Now, this spring is better known as Summer Lake Hot Spring (location 33/17-12A).
STREAMS
Summer Lake is fed by many small streams that tumble down the steep slopes west and south of the lake. No attempt was made to estimate the contribution of these streams, partly because their summer flows have been almost entirely diverted for irrigation for many years.
To the east, runoff presumably has reached the lake occasionally following heavy showers. The magnitude of this runoff is unknown.
WATER LOSS
EVAPORATION, LEAKAGE, AND OVERFLOW
Short-term records at the Summer Lake weather station, about 1 mile northwest of the lake, suggest that potential evaporation from a fresh-water body there is
about 40 inches per year. Wind on the lake sometimes increases the wetted area subject to evaporation. On the other hand, the salinity of the lake water tends to reduce evaporation slightly. In the absence of conclusive data, the loss from the lake is estimated to be 40 inches per year.
Summer Lake does not leak. The presence of many springs around the shores of the lake indicates that the water table slopes toward the lake, rather than away from it, and the lakebed is lower than any other point within 70 miles.
The lowest topographic divide in the Summer Lake basin, 2 miles north of Paisley, is about 240 feet above the lakebed, and the volume capacity of the basin below that divide (about 25 million acre-ft) is more than enough to contain all flow generated within the basin.
WATER BUDGET FOR THE LAKE
The observed variations in water level (table 21) permit a rough approximation of the average annual water budget for Summer Lake:
Approximation of annual water budget for Summer Lake for the periods 1904-22 and 1923-63
[Data reported in thousands of acre-feet per year]
Source or disposition of water Period 1904-22 Period 1923-63
a. Inflow from the Ana River _ 100 70
b. Inflow from other sources- 20 20
c. Net depletion by irrigation . . d. Total dissipated by net evaporation 10 25
(a+6 —c) . _ , 110 65
Area of lake corresponding
to item d______________103 acres _ _ 46 27
Approximate lake level corresponding
to item d____________________ft.. >4,150 4,146-4,147
In the above table, the lake-surface areas required to dissipate the estimated inflow were estimated on the basis of (1) an average annual precipitation of 11 inches, and (2) a loss by evaporation of 40 inches (net evaporation, therefore, is 29 in., or 2.4 ft); the comparable lake altitudes are those believed to provide such surface areas. The two estimated lake levels are supported by the few observations. Most of the observed levels prior to 1923 were above an altitude of 4,150 feet; since then, lake-surface altitudes have ranged about equally above and below 4,146^ feet (table 21). The change is ascribed to the reduction in flow of the Ana River (which began even prior to 1923) and to the increased use of water both for irrigation and for flooding areas reserved for waterfowl. At the lower levels during recent decades, the lake area has been accordingly smaller, and the resultant decrease in evaporation volume has offset both the decreased flow into and the precipitation on the lake.HYDROLOGY AND GEOCHEMISTRY, CLOSED-BASIN LAKES, OREGON
B27
GEOCHEMISTRY
GENERAL CHEMICAL CHARACTER OF THE LAKE
Summer Lake is similar chemically to Lake Abert, its neighbor to the east. Sodium, bicarbonate, carbonate, and chloride are the principal constituents, whereas calcium and magnesium are almost lacking (table 3; fig. 2). Concentrations of silica, potassium, sulfate, orthophosphate, and boron are appreciable, but the five constituents make up only about 10 percent of the total dissolved-solids content. Compared with Lake Abert, Summer Lake contains significantly greater proportions of silica, bicarbonate, sulfate, fluoride, orthophosphate, and boron, whereas proportions in Summer Lake of chloride and bromide are significantly less, and the proportions of sodium, potassium, and carbonate are about the same (table 3).
Positively identified trace elements, as indicated by the analysis of a sample collected June 12, 1962, include aluminum, molybdenum, cobalt, iron, nickel, and vanadium, in that order of abundance (table 29). However, the sample was very turbid at the time of collection, and the analytical results, thus, may not be accurate. (See discussion, p. B56.)
VARIATIONS IN DISSOLVED SOLIDS CONCENTRATION
The dissolved-solids concentration in Summer Lake varies considerably with time. The largest recorded value is about 37,000 ppm, for a sample collected in 1901 (Van Winkle, 1914, p. 119), though even greater concentrations are doubtless present at the lower levels near dryness. The smallest measured value is 1,800 ppm for a sample collected October 30, 1962. On a short-term basis, the concentration tends to vary inversely with stage—that is, with lake volume. When the lake is nearly dry or is refilling rapidly, however, the above relation is not consistent. For example, the lake dried almost completely in the late summer of 1961, and during at least the first 4 months after appre-
DISSOLVED SOLIDS. IN PARTS PER MILLION
Figure 20.—Cycle of dissolved-solids concentration and lake stage, in part hypothesized, for Summer Lake between June 1961 (point b) and June 1962 (point f). Lettered points correspond to samples from site 16 (table 28).
Figure 21.—Hypothetical relation between lake volume and dissolved-solids tonnage for Summer Lake between June 1961 (point b) and June 1962 (point f). Lettered points correspond to samples from site 16 (table 28).
ciable inflow resumed in September 1961, the dissolved-solids content of the lake at the sampling site remained less than the predryness content at comparable lake levels (fig. 20). Perhaps part of this difference was because the samples collected along the lake periphery during the first several months of refilling were not chemically representative of the entire lake, owing to areal differences in concentration prior to thorough mixing of the lake water with inflow. More importantly, however, the incoming water probably did not immediately recover all the salts previously lost during desiccation. In reference to figures 19 and 20, the lake dried almost completely after collection of sample 16b and had just started to fill when sample 16c was collected. During the period between collection of samples 16c and 16e, the lake regained much of the dissolved-solids tonnage lost during desiccation.
The theoretical relation between lake volume and dissolved-solids tonnage during the period April 1961 to June 1962 is shown schematically in figure 21. The lake volume decreased with no major loss of salts until it reached hypothetical volume 1. Continued volume decrease below that level was accompanied by solute depletion until, at point 2, the lake basin became dry or nearly so, and depletion was virtually complete. When the basin subsequently began to fill, the depleted salts were slowly regained until, at point 3, the tonnage approachedB28
CLOSED-BASIN INVESTIGATIONS
Figure 22.—Relations between dissolved-solids concentration and stage of Summer Lake in 1912, 1959, and parts of 1961-62.
the predesiccation quantity. (The relations depicted for Summer Lake in figures 20 and 21 resemble in some ways the general relations suggested by Langbein (1961, p. 10) in figure 9 of his report, parts B and C.)
Physical, rather than chemical, reasons probably explain the slow recovery of depleted solutes. Wind action at low lake stages prior to desiccation may have blown the relatively small residual brine over large areas of the flat surrounding playa, isolating parts of it at altitudes as much as several feet above the normal lake level for that volume. Thus, the isolated salts would not be recovered until the lake had again reached the higher levels. Actually, the recovery of solutes entrapped as interstitial brines, even in the lowest central areas of the playa, is probably a slow process because of the low permeability of the finegrained host sediments.
Figure 22 shows that a spectacular net solute loss occurred in the Summer Lake basin between 1912 and 1962. Van Winkle’s February 1912 analysis (1914, p. 120) indicated a dissolved-solids concentration of 16,800 ppm at an estimated lake level of 4,151.5 (±1.0 ft),2 whereas the estimated 1961-62 dissolved-solids concentration at the same lake stage would have been only 1,000-1,500 ppm. The difference represents a 93-percent net solute depletion in the 50-year period (fig. 22).
Most of the solute loss doubtless occurred during periods of basin dryness or near dryness after 1926. Prior to impoundment and extensive use of water from Ana Springs for irrigation in the 1920’s, Summer Lake probably contained water at almost all times, owing to an abundant and consistent inflow. However, after construction of the Ana River dam in 1926, depletion of the potential water supply due to impoundment and irrigation, coupled with a
2 Measured lake level on Nov. 10, 1912, was at an altitude of 4,151.3 ft. Level in Feb. 1912—prior to most spring snowmelt inflow or summer evaporation—is assumed to have been within about 1 ft of the November value.
period of drought, resulted in a shrinking lake volume. Complete or nearly complete desiccation occurred during many of the summers from 1929 to 1951. Although the entrapment of solutes within the playa muds at levels not attained by the lake since the early 1900’s accounts for much of the observed depletion, eolian transport of powdery evaporite salts during these periods of basin dryness or near dryness probably has been a significant means of solute removal from the Summer Lake playa. Wind action is violent within the basin at many times throughout each year. During the summer months, when the lakebed is dry or almost dry, billowing white clouds of alkali dust rise above the playa (fig. 23) and are carried to remote parts of the basin, or even outside the basin.
Wind transport of solutes is not restricted to periods of dryness or near dryness, however. Even at times of moderate lake stage, gusts blow water out over the flat surrounding playa, especially to the east, where some of the water and its dissolved solids are stranded and lost from the main body. Subsequent evaporation of this stranded water leaves a powdery efflorescent deposit that is readily eroded from the playa and blown away. Nonetheless, removal of salts by wind probably has been more commonplace since 1926 than during any other period within the last several thousand years.
CHANGES IN CHEMICAL CHARACTER
The relative amounts of most constituents, expressed as percentage of the total dissolved-solids content, apparently remained almost unchanged during the 2-year period of intensive study, despite the large fluctuations in the amount of dissolved solids. However, silica and orthophosphate exhibited significant seasonal fluctuations in
Figure 23.—Southward view toward dry Summer Lake playa, September 8, 1961. Clouds of wind-transported alkali dust are shown rising from the playa surface and being carried south-westward. Dashed line marks crest of 3,000- to 3,700-ft ridge obscured by dust clouds.HYDROLOGY AND GEOCHEMISTRY, CLOSED-BASIN LAKES, OREGON
B29
0.30
< ° < «
0.28
0.26
* 5
"T I 17 1 rr / \
/
/
• a
\
d_______
Temperature
/
/
/
oo 3 -
< 2
/
“I--------T
1----1---1---1---r
?
/' \ / \
/ \ / \
V
Dissolved solids
/
Orthophosphate
Silica
\
\
- 90 £
X
<
- 80 u-
(/>
UJ
UJ
- 70 g
UJ
o
60 z
UJ
50 3
5
cr
40 £ 2 UJ
“ 30 ^
oo 2 o Q H ° □ —
"a:
Q UJ
uj Q-
N
A
1961
J____L
J_____L
M J
1962
Figure 24.—Variations in dissolved-solids concentration, temperature, and relative amounts of silica and orthophosphate in Summer Lake, April 1961-October 1962. Lettered points correspond to samples from site 16 (table 28).
relative concentration (fig. 24). As in Goose and Abert Lakes, the relative amount of silica in Summer Lake varies more than that of orthophosphate. On the basis of seven samples collected between April 1961 and October 1962, the extremes, expressed as percentages of the dissolved-solids content, are 1.14 and 5.53 percent for silica, but only
0.25 and 0.30 percent for orthophosphate (absolute extremes ranged from 86 to 207 ppm and from 5.2 to 41 ppm, respectively). During the brief study, maximum relative concentrations of both constituents occurred in January, whereas minimum amount of silica occurred about July, and that of orthophosphate, about March.B30
CLOSED-BASIN INVESTIGATIONS
The amount of silica in Summer Lake tends to vary inversely with the dissolved-solids content (fig. 24), whereas orthophosphate shows little if any relation to dissolved solids. Seasonal fluctuations of silica and orthophosphate are discussed in more detail on pages B48 and B51.
The concentrations of several other minor constituents may also have varied slightly relative to total dissolved-solids content during the 2-year period of study, out the data available and the analytical methods used did not define such fluctuations.
Changes in the amount of carbonate relative to bicarbonate apparently accompany changing concentration of dissolved solids in Summer Lake, even though the total amount of carbonate plus bicarbonate (expressed as carbonate) remains at about 30 percent of the total dissolved solids. This characteristic was also observed in other lakes, and the probable causes are discussed on page B46.
A long-term net increase in the amounts of sulfate and chloride relative to those of carbonate and bicarbonate occurred during the 49-year period 1912-61. (See table 12.) The amount of sulfate plus chloride increased from 22 to 27 percent of the total dissolved-solids content, whereas the percentage of carbonate plus bicarbonate decreased by a like amount during the same period. The percentages of sulfate and chloride each increased by | about one-fifth of their 1912 values.
The relative enrichment of sulfate and chloride is attributed, at least in part, to the wind removal of efflorescent salts from the lake playa. During final stages of desiccation, winds probably blew the saline water out over the surrounding playa, and spread it in thin sheets that facilitated rapid evaporation. Some brine was undoubtedly trapped on the playa at altitudes as much as several feet higher than the lowest point in the basin. These isolated parts of the brine would tend to evaporate, leaving an efflorescent deposit of salts underlain by moist playa sediment. The thin layer of salts may have contained
Table 12.—Changes in the relative amounts of six major constituents in Summer Lake between 1912 and 1961
Constituent Percentage of total dissolved-solids content February 1912 1 April 25, 1961
Sodium (Na)_ 39.2 39.3
Potassium (K) _ . . 1 .0 1.6
Carbonate (CO3) plus bicarbonate
(IICOs)2 35.2 30.0
Sulfate (SO4)-- . _ _ . . 4.1 4.8
Chloride (Cl) - . 18.1 22.1
Dissolved-solids content, in parts
per million __ _ _ . _ _ 316,800 7,200
1 Data after Van Winkle (1914, p. 120, analysis 7).
2 Calculated as carbonate.
* Corrected for estimated concentrations of fluoride, bromide, orthophosphate, and boron; bicarbonate recalculated as carbonate.
a large percentage of powdery, less hygroscopic sodium carbonate bicarbonate minerals, whereas much of the residual sulfate and chloride (plus accompanying sodium) would form the crustier parts of the surface deposit, or would remain near surface within the moist sediment. Wind action could then remove the powdery efflorescenses rich in sodium, carbonate, and bicarbonate, leaving the surface and near-surface of the playa slightly enriched in sulfate and chloride relative to carbonate and bicarbonate. Thus, when the lake basin filled again, and the remaining solutes were redissolved, the resulting lake water would become slightly enriched in sulfate and chloride.
The depletion of silica in Summer Lake is almost as pronounced as that in neighboring Lake Abert. Silica comprises 23 percent of the dissolved-solids content in the Ana River, whereas it makes up only about 2.5 percent (measured range, 1.5 to 5.5 percent) of the dissolved-salt tonnage in Summer Lake. Silica in the several lakes is discussed in more detail on page B47.
SOURCES OF DISSOLVED SOLIDS
Most of the salts dissolved in Summer Lake have, at least during postpluvial time, come from the Ana Springs via the Ana River (analyses 20a-20f, table 28). Because of the relatively consistent discharge from these springs, a large amount of dissolved material presumably has been added to the lake during the last few thousand years. More than half of the 160 ppm increment consists of the major dissolved constituents of Summer Lake—sodium plus equivalent chloride and bicarbonate.
The amount of incoming calcium and magnesium that is available for removal from solution by organic and inorganic processes is somewhat less in Summer Lake than in Lake Abert because most of the flow into Summer Lake contains only small concentrations of these two constituents. In the Ana River, for example, calcium plus magnesium make up only about 5 percent of the total load, in contrast to about 14 percent for the Chewaucan River.
Other, lesser sources of dissolved solids include small springs north, west, and south of the lake, streams draining the relatively small area west and south of the lake, and occasional thundershower runoff from the east. The quantity and character of solutes derived from these sources have not been evaluated.
GOOSE LAKE
Goose Lake is situated immediately west of the Basin and Range physiographic province, in Modoc County, Calif., and Lake County, Oreg. (fig. 1). Its rarely occurring overflow is tributary to the North Fork Pit River in the Sacramento River basin. Although most of the lake itself lies in California, most of its valley and almost two-thirds of the *otal drainage basin are in Oregon. Among lakes inHYDROLOGY AND GEOCHEMISTRY, CLOSED-BASIN LAKES, OREGON
B31
semiarid zones, Goose Lake is noteworthy in that it currently occupies an unusually large part, 13 percent, of its total drainage basin.
The lake lies in a semiarid plain, almost completely surrounded by mountains and wooded ridges. Immediately to the east, the Warner Mountains in California and their northern extension include peaks in each State that rise to altitudes of more than 8,000 feet within 5 or 6 miles of the 4,700-foot lake. To the north, a rather broad plain rises gradually through rolling slopes that are referred to locally as the Fremont Mountains (pi. 1), which reach an altitude of about 7,000 feet. Similar but slightly lower wooded ridges lie within 10 miles of the west shore. At the south end of Goose Lake, a low gravelly terrace or old lake beach separates the lake from a marshy meadow in an old river channel that winds southward to the headwaters of the North Fork Pit River.
AREA, VOLUME, AND LAKE-LEVEL FLUCTUATIONS
At overflow stage (alt 4,716 ft), Goose Lake covers 194 square miles; it has a maximum depth of 24 feet, a length of 28 miles, and a maximum width of 10 miles. Surveys of the bed and outlet of Goose Lake were made in 1959 by the California Department of Water Resources. The results are shown in figure 25, and the area and volume are given in table 13. Goose Lake has varied greatly in area and volume within historic time, as shown by the data in table 22 and by the pattern of fluctuations since 1958 (fig. 26). Similar variations have undoubtedly occurred for thousands of years. Even before any water was stored or diverted for irrigation, the historical records attest to conditions ranging from near desiccation to overflow. For example, on a visit in 1882, Russell (1884, p. 456) found the lake to be 15 feet deep over a former road crossing 4 or 5 miles from the south end. That road was undoubtedly at the site of the present causeway
Table 13.—Area and volume of Goose Lake
[Area and volume data are rounded to three significant figures]
Altitude (ft) Area (acres) - Volume (acre-ft)
Oregon California Total
4,692.0 0 0 0 0
4,692.5 210 6,180 6,390 1,060
4,693.0 3,290 12,500 15,800 6,620
4,693.5 5,620 19,200 24,800 16,800
4,694.0 8,960 29,400 38,400 32,600
4,694.5 12,800 34,200 47,000 54,000
4,695.0 17,100 38,000 55,100 79,500
4,696.0 21,300 46,100 67,400 141,000
4,697.0 . 25,500 52,700 78,200 214,000
4,698.0 27,400 56,100 83,500 294,000
4,699.0 28,400 59,000 87,400 380,000
4,700.0 . 29,900 62,000 91,900 470,000
4,7041 33,900 73,700 108,000 862,000
4,716 43,600 80,600 124,000 2,250,000
1 Lake at an altitude of 4,704 feet during survey made by the California Department of Water Resources in 1959.
Figure 25.—Lakebed contours of Goose Lake. Map furnished by California Department of Water Resources.B32
CLOSED-BASIN INVESTIGATIONS
Figure 26.—Seasonal variations in the stage of Goose Lake, October 1958 to November 1962. Dashed where approximated. Lettered
points correspond to samples from site 21 (table 28).
across the lake, where the lowest lakebed altitude was about 4,699.5 feet. Russell recognized that the high level he observed was not permanent, and predicted—though the lake had no historical record of dryness—that in a series of arid seasons the lake would “unquestionably evaporate completely * * It did indeed become dry in 1926 and was virtually dry in the late summer of each year during the period 1929-34. The desiccation in 1926 exposed ruts of another pioneer road that crossed the lake at an altitude of about 4,694 feet several miles farther north of the present causeway.
Some early records of lake level (table 22) are based on descriptive narratives and diaries of travelers that give estimates of the size of the lake or of its stage in reference to the outlet level. (See Harding, 1965, p. 24, 26, 27, 32.) Accuracy of these descriptions depends on the ability of the observers to estimate those factors.
WATER SUPPLY
Goose Lake derives its water supply from lake-surface precipitation, stream inflow, and small peripheral springs.
LAKE-SURFACE PRECIPITATION AND SPRINGS
Rain and snow at Lake view, about 12 miles north of midlake, averaged 14.1 inches per year during the nearly complete period of record 1885-1963. Records of precipitation at Davis Creek, Calif., and at a site west of Goose Lake, obtained during the period 1959-61 by California State Department of Water Resources, correspond closely to records at Lakeview; hence, the average annual precipitation on Goose Lake is assumed to be about 14 inches.
No large springs emerge from the bed of Goose Lake,
but several small springs are said to maintain soggy areas (Harding, 1965, p. 27, 47). Many small seeps and springs line the east shore of the lake; some of them are thermal, and others are fed by percolating irrigation water. Their total flow is a very small part of the water budget for the lake.
STREAMS
The largest streams in the Goose Lake basin are Drews, Cottonwood, and Thomas Creeks, all in Oregon. Their summer flows are now diverted for irrigation, and in some years very little of their flows reach Goose Lake. The annual flow of these streams is extremely variable. For example, on Drews Creek, during the discontinuous-record period of 32 water years, the flow ranged from about 5,000 acre-feet in 1931 to 163,000 acre-feet in 1956.
The available records for Drews, Cottonwood, and Thomas Creeks, summarized below, have been obtained
Average annual flow at gaging stations on streams tributary to Goose Lake
Stream Drainage Altitude of gage ■ (ft) Total record Concurrent record, water years 1913-17, 1928-30 (acre-ft)
ai ca (sq mi) Period (water yr) Acre- feet
Drews Creek 1 212 4,835 1913-30, 1938, 1940-41, 1947, 1954-63. 49,400 47,600
Cottonwood
Creek 2 32.9 5,007 1910-19, 1925-35, 1939-42, 1947-63. 15,300 16,400
Thomas Creek 30 4,800 1913-17, 1928-31.. 9,480 10,400
Total
(rounded).. 275 74,200 74,400
1 Adjusted for diversion by North Drews Canal and change in contents of Drews Reservoir.
2 Not adjusted for change in volume of Cottonwood Reservoir (ordinarily empty on Sept. 30).HYDROLOGY AND GEOCHEMISTRY, CLOSED-BASIN LAKES, OREGON
B33
near the points where those streams leave the wooded hills and enter the alluvial plain; hence, the data are not an index of flow into Goose Lake, except that in very wet years much of the observed streamflow does reach the lake because it exceeds the needs for storage and irrigation. The runoff per square mile from the lower, drier plain is believed to be less, and that from the Warner Mountains greater, than the quantities measured at these stations.
The average annual discharge of all streams tributary to Goose Lake is estimated as 250,000 acre-feet, unadjusted for the effects of irrigation.
Irrigation began in a small way when the valley was first settled in the 1870’s. Water has been stored since 1911 in Drews Reservoir (capacity 62,500 acre-ft) and since 1921 in Cottonwood Reservoir (capacity 4,160 acre-ft, until enlarged, in 1962, to capacity of 7,540 acre-ft), upstream from the two gaging stations. Harding (1965, p. 39) estimated the acreage irrigated and the resulting average annual streamflow depletion (including loss by evaporation from surface-storage reservoirs) as follows:
Estimated effect of irrigation on inflow to Goose Lake during years of adequate streamflow
Average Average net annual
Period area irrigated streamflow depletion
(acres) (acre-ft)
1882-89__________________ 5,000 10,000
1890-1904_______________ 10,000 20,000
1905-11_________________ 15,000 30,000
1912-19_________________ 20,000 45,000
1920-25_________________ 25,000 62,500
1926-35_________________ 30,000 68,500
1936-60_________________ 35,000 85,000
Thus, for the period since 1935, assuming that the average total streamflow has been 250,000 acre-feet per year, annual inflow to the lake has averaged about 85,000 acre-feet less, or about 165,000 acre feet.
WATER LOSS
EVAPORATION AND LEAKAGE
The average annual evaporation loss from Goose Lake is believed to be about 42 inches. That estimate is based on observed seasonal fluctuations of the lake level in dry years (fig. 26), and on studies of evaporation at Eagle Lake in California and at Klamath and Abert Lakes in Oregon.
Goose Lake was at one time suspected of leaking into the headwaters of the North Fork Pit River (Van Winkle, 1914, p. 38). When the lake was being reduced to dryness in the 1920’s and was refilling in the 1930’s, no evidence of any such leakage was noted, and no known variations in the base flow of streams in the Pit River basin can be correlated with changes in the level of Goose Lake. In dry seasons, the rate of fall of the lake level is no more than would be expected from evaporation alone. There is good reason to believe, then, that the lakebed is water-
tight, and that all inflow is dissipated by evaporation, except for small volumes of overflow at rare intervals.
OVERFLOW
Loss by overflow within the last 100-150 years has been rare and small. Goose Lake overflowed in 1868, after two or more very wet years. The quantity of water lost is not known, but it probably was only a small part of the stream discharge reaching the lake in that year. The lake overflowed again in 1881 for about 2 hours because of the temporary effect of a strong wind from the north. The volume of overflow in that year was insignificant. Other reports of overflow in 1862, 1869, 1871, 1875, 1898, and 1910 cannot be verified and are thought to be erroneous (Harding, 1965, p. 31-33); the events of 1868 and 1881 are believed to be the only instances of overflow since 1832.
Earlier overflow probably was more frequent and of greater volume than that in the last hundred years. Terraces or strand lines many miles long have been formed by the action of waves and ice at altitudes ranging from 4,716 to 4,720 feet. (The railroad grade along the east shore is built on such a terrace at the State line.) Strand lines of such magnitude are evidence that the lake attained overflow level for an extended period. Likewise, the relatively low concentration of dissolved salts in Goose Lake, as compared with those in waters of true closed lakes, also suggests that part of the inflow has in the past been dissipated by outflow, rather than by evaporation.
The probability of overflow in the future has been greatly reduced by the consumptive use of about 85,000 acre-feet of wTater per year from tributary streams. In the past, the lake has approached overflow only when two or three very wet years occurred in succession after the lake had previously attained at least a moderate level. In 1958, when other closed lakes reached or approached record high levels, about 1 million additional acre-feet (six times the average annual inflow) would have been required to cause overflow. Under present-day climatic and agricultural conditions, then, Goose Lake can be considered a true closed lake, rather than one that occasionally overflows.
WATER BUDGET FOR THE LAKE
Goose Lake receives an average annual precipitation of about 14 inches, but it loses about 42 inches annually by evaporation. Its tributary springs and streams provide an average annual discharge of about 250,000 acre-feet, of which about 165,000 acre-feet is inflow to the lake.
For the present degree of streamflow depletion by irrigation, the average calculated lake level is at an altitude of about 4,697 feet (lake volume, 214,000 acre-ft), a stage at which the lake area provides an average net evaporation loss of about 165,000 acre-feet per year, which offsets the average net inflow. Similar climatic conditions prior to any depletion for irrigation would have providedB34
CLOSED-BASIN INVESTIGATIONS
runoff to maintain the lake at a calculated average level at an altitude of about 4,705 feet (lake volume, 950,000 acre-ft).
GEOCHEMISTRY
GENERAL CHEMICAL CHARACTER OF THE LAKE
Goose Lake is chemically similar to Abert and Summer Lakes, but is characteristically much more dilute. The principal dissolved constituents in Goose Lake are sodium, carbonate plus bicarbonate, and chloride (table 3; fig. 2). Silica, potassium, and sulfate are less abundant, but collectively they are about 14 percent of the total dissolved-solids content. Orthophosphate and boron occur in significant concentrations. Only small amounts of calcium and magnesium are present in the lake: their combined concentration usually makes up only 1 or 2 percent of the dissolved-solids content. The calcium to magnesium ratio ranges from 3:1 to 4:1.
Positively identified trace elements present in a lake sample collected June 12, 1962, include vanadium, aluminum, mr .ybdenum, iron, and nickel, in that order of abundance (table 29). (See remarks regarding the reliability of trace-element data, p. B56.)
4' V f
VARIATIONS IN DISSOLVED SOLIDS CONCENTRATION
The measured or estimated dissolved-solids concentration in 16 samples collected from Goose Lake during the period 1953-62 has ranged from about 600 to about 2,800 ppm. A plot of dissolved-solids concentration versus lake stage (fig. 27) indicates a reasonably consistent relation. The moderate scatter of points is due to small amounts of dilute inflow near sampling sites and to other slight chemical heterogeneities throughout the lake. For example, samples collected by Harding on July 9, 1958, at three different lakeshore sites contained 600 to 740 ppm of dissolved solids (analyses 21a, 22e, and 23d, table 28), whereas the estimated average solute content at that lake level was 650-700 ppm.
Figure 27 shows that the rate of change of concentration with changing lake stage is slow above altitudes of about 4,702 feet, and rapid at altitudes below that level, as the lake volume diminishes. Because Goose Lake is nearly homogeneous except during periods of appreciable inflow, seasonal fluctuations in the dissolved-solids concentration at most sites are related to changes in lake volume.
An average value of about 1 million tons for the solute load in Goose Lake during the period 1953-62 is indicated by the relations between lake stage, volume, and dissolved-solids concentration (tables 13, 28). A long-term decrease in dissolved-solids tonnage is indicated by comparing the relation between lake stage and solute concentration for recent samples with that calculated for a 1912 sample
Figure 27.—Relation between dissolved-solids concentration and stage of Goose Lake, 1912 and 1953-62. Lettered points correspond to samples from site 21 (table 28).
analyzed by Van Winkle (1914, p. 39). Figure 27 shows that 1,010 ppm of dissolved solids, which corresponded to an estimated lake stage at an altitude of about 4,714 feet in 1912, would also have occurred at a stage at an altitude of about 4,702 feet during the period 1953-62. If the 1912 sample is assumed to represent the average lakewide dissolved-solids concentration at that time,3 it indicates a tonnage of about 3 million tons—three times the calculated tonnage for 1953-62.
The spectacular loss of dissolvable solids can probably be attributed in part to removal of salts from the lakebed by winds during periods when the basin dried completely or almost completely in the summers of 1926 and 1929-34. This theory is supported by newspaper accounts of alkali duststorms on August 26, 1926, and July 16, 1931 (S. T. Harding, written commun., 1962). Additional unrecorded periods of less spectacular wind undoubtedly occurred during the seven dry years. A large amount of windblown salts, if not actually removed from the basin, has at least been isolated from rapid return to the lake as part of surface runoff. Entrapment of salts within the broad mudflats surrounding the lake undoubtedly also accounts for some of the depletion. Recovery of the lost salts in Goose Lake will be a slow and incomplete process, as indicated by the lack of any noticeable increase in solute tonnage during the period 1953-62.
3 The dissolved-solids concentration and the lake stage in 1912 compare favorably with similar values—about 1,100 ppm at an altitude of about 4,715 ft—reported for a sample collected in August 1904 (Harding, 1965, p. 50). Even though the accuracy of certain individual determinations in the 1904 analysis is questionable, the reported value for total dissolved-solids concentration probably is of sufficient accuracy to verify the relation in 1912.HYDROLOGY AND GEOCHEMISTRY, CLOSED-BASIN LAKES, OREGON
B35
CHANGES IN CHEMICAL CHARACTER
The relative (percentage) amounts of most dissolved constituents in Goose Lake have not changed significantly since 1953. Only the ratio of carbonate to bicarbonate and the relative quantities of silica, calcium plus magnesium, and orthophosphate have fluctuated with any certainty during the 10-year period. The relation of carbonate to bicarbonate is discussed on page B45.
The abundance of calcium and, perhaps to a lesser extent, magnesium is controlled largely by the solubility of alkaline-earth carbonate minerals. The relative quantities also may be influenced by the actions of calciumutilizing biota, and by reactions that may occur between the alkaline-earth ions—especially magnesium—and certain clay minerals present on the lake bottom. The several processes are discussed in more detail on pages B49-B50.
Amounts of the biologic nutrients silica and orthophosphate vary appreciably relative to the dissolved-solids content in Goose Lake. (See fig. 28.) Silica varies the most. For six samples collected between April 1961 and September 1962, the absolute concentration varied from 50 to 80 ppm, while the relative amount, expressed as a percentage of the dissolved-solids content, ranged from 2.9 to 5.6 percent. This is equivalent to a range in silica content of 29,000 to 56,000 tons. (See footnote 1, p. B18.) Variations in orthophosphate content are less pronounced. Relative amounts in five samples ranged from 0.44 to 0.54 percent of the dissolved-salt content (absolute range, from 5.3 to 12 ppm). On the basis of only 16 months of record, the largest relative amounts of both constituents characteristically seem to occur during the spring, and the smallest amounts occur sometime between August and November.
Figure 28 shows that the relative amounts of silica and orthophosphate in Goose Lake tend to vary inversely with dissolved-solids concentration but seem unrelated to water temperature. A more thorough discussion of silica and orthophosphate is presented on pages B48 and B51, respectively.
A significant net change in the distribution of major anions occurred in Goose Lake during the period 1912-62. (See table 14.) The amount of sulfate plus chloride increased from 15 percent to 20 percent of the sum of the six most abundant constituents (sodium, potassium, carbonate, bicarbonate, sulfate, and chloride), whereas the relative amount of carbonate plus bicarbonate (calculated as carbonate) decreased by a similar percentage; sodium plus potassium remained at about 41 percent during the 50-year period.
As at Summer Lake, the change may have, occurred during periods of desiccation (in 1926 and 1929-34 at Goose Lake), owing largely to eolian removal of efflorescent salts from the lake playa (p. B28).
Table 14.—Changes in the relative amounts of six major constituents in Goose Lake between 1912 and 1962
Percentage of the six-constituent
Constituent parts-per-million sum
May 12, 1912 i June 12, 1962
Sodium (Na) — . 37.2 38.0
Potassium (K) Carbonate (CO3) plus bicarbonate (IICO3) * 1 3.6 3.0
43.8 39.0
Sulfate (S04) - - - 4.8 7.0
Chloride (Cl)_ 10.6 12.9
Dissolved-solids content, in parts per
million _ 1,060 1,270
1 Data after Van Winkle (1914, p. 39).
2 Calculated as carbonate.
It may at first seem anomalous that the relative amounts of sulfate and chloride together have increased by almost one-third in just 50 years, especially when they amounted to only about 15 percent of the major-constituent sum at the start of the 50-year period (table 14). The principal reason may be as follows: The lake probably dried only rarely during the last few thousand years, but in this century storage and diversion of runoff for irrigation have reduced inflow almost 35 percent. Thus, the extensive diversion of runoff, together with several years of abnormally low precipitation, caused the unusual periods of desiccation and accompanying removal of salts during 1926 and 1929-34.
Increased use of ground water during the last 50 years might be suspected as a secondary cause of the marked increases in sulfate and chloride. However, the available chemical analyses (tables 30, 31) indicate that most ground water in the basin is not a source of significant amounts of sulfate and chloride relative to bicarbonate (even assuming the elimination of some bicarbonate due to precipitation of alkaline-earth carbonate minerals). Exceptions include certain hot-spring and well waters located within, and adjacent to, fault zones along the eastern limits of the valley. Although this ground water contains large relative concentrations of sulfate and chloride (for example, analyses 109 and 110, table 30), the aggregate discharge probably is too small for the solute contribution to have a significant short-term (50-year) effect on the chemical character of Goose Lake.
SOURCES OF DISSOLVED SOLIDS
Surface inflow probably has been the largest long-term source of the dissolved solids in Goose Lake. Unfortunately, little information on the chemical character of this inflow is available (table 28). Presumably, though, most of the streamflow reaching Goose Lake chemically resembles the flow of the Chewaucan River in the adjacent Lake Abert basin (p. B20). Streams such as ThomasB36
CLOSED-BASIN INVESTIGATIONS
ORTHOPHOSPHATE, AS PERCENTAGE OF DISSOLVED-SOLIDS CONCENTRATION
o
o
—
o,
o o
^ (Jl
oo o
© o
© or
NJ -C*
t--------r
SILICA, AS PERCENTAGE OF DISSOLVED-SOLIDS CONCENTRATION
INO CO ^ CJ1 ©
IN PARTS PER MILLION
Figure 28.—Changes in dissolved-solids concentration, water temperature, and relative amounts of silica and orthophosphate in Goose Lake, April 1961-September 1962. Lettered points correspond to samples from site 21 (table 28).
Creek, which pass through, and are used in, sizable irrigated areas, probably contribute water to Goose Lake that is not much more dilute than the Chewaucan River inflow to Lake Abert. Other streams, which are not influenced as much by irrigation, probably contribute a characteristically more dilute inflow.
Much ground water in the basin contains considerably greater concentrations of dissolved solids than those
assumed to be present in the streams (tables 30, 31).4 Ground water may have contributed somewhat larger than usual amounts of dissolved solids to the lake during the last 50 years, owing to accelerated use for irrigation, but the long-term contribution, as spring flow, probably
4 Additional analyses for the Goose Lake basin in California have been published by the California Department of Water Resources (1965, p. 157, and subsequent annual reports in the same series; also, 1966, table 44).HYDROLOGY AND GEOCHEMISTRY, CLOSED-BASIN LAKES, OREGON
B37
is a relatively minor part of total dissolved-solids influx to the lake.
RECONNAISSANCE OF SEVERAL OTHER CLOSED-BASIN LAKES
Hydrologic and chemical data of varying quantity are available for several other lakes in south-central Oregon: Silver Lake in the Fort Rock basin north of Summer Lake; Malheur and Harney Lakes in the Harney Basin near Burns; and Crump, Hart, and Bluejoint Lakes in the Warner basin east of Lake Abert (pi. 1; fig. 1).
Hydrologic data for the lakes are summarized in table 1. Additional pertinent information is presented in this section.
AREA, VOLUME, AND LAKE-LEVEL FLUCTUATIONS
The approximate area and volume of Silver Lake at three altitudes are estimated below (the lake dries between altitudes of 4,299 and 4,300 ft). At an altitude of about
4.310.4 feet (area about 11,000 acres, volume about 75,000 acre-ft), the lake overflows into adjacent Thorn Lake. Records of lake level for Silver Lake are listed in table 23. The highest level observed, noted from a high-water mark made in the spring of 1904, was at an altitude of 4,311.8 feet. The highest level in recent years was at an altitude of
4.309.05 feet on June 26, 1958; after that date, the water level fell steadily until the last pools disappeared about August 11, 1961 (A. B. Claggett, Oregon State Game Comm., oral commun., 1961).
Altitude (ft) Area (acres) Volume (acre-ft)
4,303 4,306 4,309 6,000 8,500 10,000 10,000 32.000 60.000
Silver Lake is the remnant of a Pleistocene lake that at one time covered more than 500 square miles, including not only the present lake, but also Paulina Marsh, Thorn Lake, Fossil Lake, Christmas Lake, and Fort Rock Valley. Well-developed lake terraces were formed at several levels north of the town of Silver Lake, and waterworn notches in the lava cliffs on Fort Rock are attributed to wave action (Hubbs and Miller, 1948, p. 73). According to Allison (1940, p. 300), the lake at one time overflowed to Deschutes River, in the Columbia River basin, until the outlet was buried by more than 100 feet of lava. Allison (written commun., 1963) proposed “Fort Rock Lake” as the name of this pluvial body of water, and found the highest beach line to be at an altitude of about 4,520 feet (220 ft above the bed of Silver Lake).
Malheur and Harney Lakes occupy shallow depressions in the central plain of Harney Basin (fig. 1) and are separated by a smaller body of water known as Mud Lake. When Malheur Lake rises above an altitude of about
4,091.5 feet (maximum depth, about 2.5 ft), water begins to flow through a winding, reed-grown channel (The
Narrows) westward into Mud Lake. At a somewhat higher stage, Mud Lake overflows westward through a gap in a bar locally known as the Sand Reef into Harney Lake. The bar was apparently formed by wind, waves, and ice on Harney Lake during periods of no inflow; its altitude at the inlet varies. In 1881, for instance, the Sand Reef was observed to be a temporary dam about 10 or 12 feet high, according to testimony before the U.S. Supreme Court (1932, p. 841), but in April 1942, John C. Scharff of the U.S. Fish and Wildlife Service (written commun., 1942) observed it to be 1 foot (or less) high and to be lower than the level of overflow from Mud Lake. The approximate areas and volumes of the lakes at various levels were listed by Piper, Robinson, and Park (1939, p. 20).
Malheur Lake has ranged from near dryness to a maximum observed level of 4,095.39 feet on June 9, 1952. The lake was nearly dry in 1889, 1926, 1931-35, and 1960-62; in 1934 the lakebed was planted to grain.
Harney Lake is the final evaporating sump for the basin. At times, its lakebed has been practically dry for long periods, including 1929-34, 1962, and probably other years for which no observations were reported. The highest water level of record is an altitude of 4,087.60 feet on June 16, 1943, but higher stages probably occurred in the 1950’s.
Historical references abstracted from records of the U.S. Supreme Court (1932) provide a general indication of lake-level fluctuations in the Harney Basin prior to 1900; the references are summarized in table 24. Records of stage for Malheur and Harney Lakes during 1903-31 were listed by Piper, Robinson, and Park (1939, p. 129-131). In their tabulation, reported low levels at The Narrows gage do not, as implied, represent the stage of Malheur Lake, because the hydrologic connection between the lake and gage is doubtful below an altitude of about 4,092 feet, and is nonexistent below about 4,091.5 feet. At the gage near Voltage, in the channel of the Donner und Blitzen River, stages are at times somewhat higher than the level of Malheur Lake because of the effect of streamflow or wind. In addition, comparison of data at the two gages during periods of falling lake stage indicated that readings at the Voltage gage are consistently about half a foot higher than those at The Narrows, suggesting a difference in datum.
Records of stage subsequent to those reported by Piper, Robinson, and Park (1939) for Malheur and Harney Lakes are listed in tables 25 and 26, respectively.
During Pleistocene time, a large but shallow water body (maximum area, about 900 sq mi; maximum depth, 35 ft) covered present-day Malheur and Harney Lakes and overflowed occasionally into the Snake River basin via the Malheur River (information largely from Snyder and others, 1964).
The floor of Warner basin is virtually a chain of marshes, shallow intermittent lakes, and drained meadows. InB38
CLOSED-BASIN INVESTIGATIONS
downstream order, they are Pelican, Crump, Hart, Anderson, Swamp, Mugwump, Flagstaff, Campbell, Lower Campbell, Stone Corral, and Bluejoint Lakes. They vary greatly in size; all the lakes have been dry or practically dry at times (the lower ones have been dry for periods of several years), whereas during periods of high water, many of the lakes merge. From Crump Lake to Bluejoint Lake (about 30 miles), the total drop in lake-bed surface is only about 13 feet. In this report, the discussion is restricted almost entirely to Crump, Hart, and Bluejoint Lakes—-the only ones for which appreciable data are available.
The area and volume of Hart Lake at several altitudes are given below (Oregon State Engineer, written commun., 1961). Similar data are not available for Crump or Blue-joint Lakes.
Altitude (ft) Area (acres) Volume (acre-ft)
4,463.5 4.465.0 4.467.0 4.469.0 4.471.0 1 4,473.0 0 2,600 4,200 5,800 6,900 7,400 0 2,100 8,400 17,800 30,400 44,600
1 Spillway altitude (lake is diked).
Records of stage for Crump, Hart, and Bluejoint Lakes are listed in table 27. In some recent years, Hart Lake has been pumped dry for irrigation before the end of summer. Large parts of it were naturally dry in 1889—when the use for irrigation was slight—and teams were driven across the lakebed from the east side of the valley to Plush (Whistler and Lewis, 1916, p. 32).
A distinct beach line was observed in 1915 around the north end of Bluejoint Lake, about 18 feet higher than the lowest point in the lakebed. The beach is thought to have been formed about 1868, on the basis of the following evidence:
1. Precipitation received at Fort Bidwell, Calif., about 60 miles
southwest of the lake, in the period July 1, 1866, to June 30, 1868, was by far the greatest 2-year precipitation at that station in 75 years of record.
2. Sage brush growing below the beach line was not more than 40
years old when observed in 1914, and was distinctly younger than that above the beach line.
3. A. M. Hammersley of Lakeview stated (written commun., 1912)
that the lakes were very high when his father came to the valley in 1869, which was “probably when the beach line was made.”
Although Waring (1908, p. 38) cited litigation as evidence of recession of the lakes in Warner basin since about 1860, the lake outlines shown on his reconnaissance maps (pis. II, VI), made on the basis of inspection in the autumn of 1906, suggest that lake levels were not low then (Crump Lake was continuous with Hart Lake, and Flagstaff Lake covered a large area).
Extensive remnants of another prominent beach line at
4,750-4,800 feet—some 300-350 feet above the bed of Bluejoint Lake—attest to the existence of a very large lake in Warner basin in pluvial time (approximate shoreline altitude estimated from U.S. Geol. Survey 7|-minute Bluejoint Lake East and Crump Lake quadrangles). At that level, the lake covered about 500 square miles, or 26 percent of its entire drainage basin. Russell (1884, p. 459) and Free (1914, p. 26) concluded that, even at that high level, Lake Warner did not overflow—a conclusion that was later questioned by Van Winkle (1914, p. 118). Hubbs and Miller (1948, p. 65, 151) found the faunal evidence to be confusing, but they concluded that, at its highest level, pluvial Lake Warner was probably closed. Information presented by Van Winkle (1914, p. Ill), adjusted to the present-day datum (table 27, headnote), suggests that the lowest topographic divide is above an altitude of 4,800 feet, and presumably above the high shoreline.
WATER SUPPLY
Estimates of average lake-surface precipitation are listed in table 1.
Ground water contributes little or no inflow to any of the lakes except Harney Lake. Springs in the Warm Spring Valley, west of Harney Lake, discharge 30 to 40 cfs, which reaches the lake when not used for irrigation.
Principal sources of inflow to the several lakes are as follows:
Lake Source of inflow
Silver Lake basin
Silver Lake _ Paulina Marsh.
Harney Basin
Malheur Lake Harney Lake- The Silvies and Donner und Blitzen Rivers, plus other smaller streams northeast of Burns. Silver and Warm Spring Creeks, plus over- flow from Malheur Lake.
Warner basin
Crump Lake Hart Lake Bluejoint Lake Lake.
Table 15 summarizes the records of discharge for those of the streams that are gaged. The gaging stations are upstream from irrigated and (or) marshy areas, as well as reservoirs, that deplete the flow considerably. In fact, in dry years these areas consume all, or almost all, the potential inflow.
At Silver Lake the approximate annual inflow is related in a general way to the flow of Silver Creek, as shown in figure 29. The figure suggests that negligible surface runoff reaches the lake in water years when the measured flow ofHYDROLOGY AND GEOCHEMISTRY, CLOSED-BASIN LAKES, OREGON
B39
Silver Creek during December-June totals less than about 30,000 acre-feet. In wetter years the inflow is about twice the amount by which the 7-month total exceeds 30,000 acre-feet. On the basis of this general relation, the longterm record for Silver Creek suggests that appreciable flow has reached the lake during only a few years (probably less than 12 of the 53 years of record), which makes any estimate of average annual inflow almost meaningless.
At Malheur Lake the inflow from all streams probably averages about 100,000 acre-feet per year, though it varies so much from year to year that, as for Silver Lake, an average figure has little meaning. At Harney Lake the annual inflow in addition to periodic overflow from Malheur Lake may average about 40,000 acre-feet—also highly variable from year to year.
Estimates of average yearly inflow to Crump and Hart Lakes in the Warner basin are listed in table 1. Observations at Bluejoint Lake and streamflow data for Twenty-mile, Deep, and Honey Creeks suggest that the lake generally receives inflow when the gaged discharge exceeds about 200,000 acre-feet per year.
WATER LOSS
Estimated evaporation rates at the lakes are listed in table 1. The data are based largely on the generalized map of Kohler, Nordenson, and Baker (1959, pi. 2).
As early as 1908, Waring (p. 41) suggested that the bed of Silver Lake may leak, and hydrologic evidence supports this suggestion. Ground water may move from the Silver Lake basin toward Summer Lake (Hampton, 1964, p. 18), perhaps along southeast-trending fault zones, such as that marked by the spectacular scarp east of Silver Lake. (Much of this interbasin ground-water flow may feed the
Table 15.—Average annual discharge at gaging stations on principal streams in Silver Lake, Harney, and Warner basins
Stream Drainage Total full-year record Concurrent record, by basin (acre-ft)
gage Period (sq mi) 1 (water yr) Acre- feet
Silver Lake basin
Silver Creek 180 1906, 1910-27, 1930-63. 19,400
Harney Basin
Silvies River Donner und Blitzen River Silver Creek _ 934 1904-5, 1910-12, 1918-21, 1923-63. 200 1912-13, 1915-16, 1918-21, 1939-63. 228 1952-63 117,000 87,900 32,400 140,000 85,300 32,4u0
Total (rounded). 1,360 237,000 258,000
Warner basin
Twentymile Creek Deep Creek. Honey Creek 194 1911-15, 1919, 1941-44, 1946-63. 249 1923, 1930-63 170 1911-14, 1931-63 35.800 86,200 19.800 37,800 98.900 21.900
Total (rounded). 610 .................... 142,000 159,000
DECEMBER 1-JUNE 30 STREAMFLOW (S), IN THOUSANDS OF ACRE-FEET
Figure 29.—Relation between seasonal (Dec. 1-June 30) stream-flow of Silver Creek near the town of Silver Lake and computed inflow to Silver Lake, listed by water year.
prolific Ana Springs, which rise 7 miles southeast of the lake, at a point on a projection of the scarp mentioned above. See pi. 1.) Almost all surface inflow enters Silver Lake from the northwest, and peripheral springs— indicators of ground-water gradients toward the lake—are absent from the south and east sides of the lake basin. In 1948, following a long period of lake dryness, the ground-water level was at least 10 feet below land surface in wells on the eastern part of the lakebed. By December 1954, after the lake had filled to a maximum depth of at least 4 feet, the water levels in two such wells had risen more than 10 feet (E. It. Hampton, U.S. Geol. Survey, oral commun., 1963). Because the amount of recharge from areas south of Silver Lake seems inadequate to have produced such a pronounced rise in water level, seepage from the lake itself is considered to have been the principal source (other contributing factors may have been loading effects of the lake-water body, pressure response to recharge northwest of the lake, and leakage into the wells from the lake).
That leakage occurs is further supported by records of lake level and precipitation for the periods from November 1 to April 30 in 1958-59 and 1959-60. No inflow occurred in these two periods, and the lake declined from its high level of 1958. Evaporation during each 6-month period was estimated to have been about 20 percent of the annual total (Kohler and others, 1959, pi. 4). At Medford Weather Station, about 110 miles southwest of Silver Lake, pan evaporation for these same two periods was 19 and 20 percent of the annual totals for water years 1959 and
Gage locations are shown in figure 1 and on plate 1.B40
CLOSED-BASIN INVESTIGATIONS
1960, respectively. Thus, assuming the average annual evaporation at the lake to have been about 38 inches, unaffected by local or temporary abnormalities, the total for each November-April period would be about 7.6 inches. The following calculations show that lake-level declines during the two 6-month periods—including
Estimate of lake-level decline due to seepage loss from Silver Lake for the periods November 1 to April 30 in 1958-69 and 1959-60
[Note.—Silver Lake received no inflow during either given period]
Quantity (inches)
Water budget for Silver Lake ------------------------------------
Nov. 1, 1958- Nov. 1, 1959-Apr. 30, 1959 Apr. 30, 1960
a. Precipitation 1 4.3 C.l
b. Drop in lake level . _ _ c. Lake-level drop plus precipita- 8.0 4.6
tion (a+6) 12.3 10.7
d. Estimated evaporation _ _ _ e. Estimated lake-level decline due 7.6 7.6
to seepage loss (c — d).. - _ _ 4.7 3.1
1 Average of precipitation totals recorded at The Poplars and Summer Lake weather stations, 20 miles north and 13 miles southwest of Silver Lake, respectively.
allowances for lake-surface precipitation—were, respectively, about 4.7 and 3.1 inches greater than the decrements ascribed to evaporation alone. The two quantities are considered to represent seepage, and the apparent difference between them presumably is due, at least in part, to decreased head as the lake level declined about ‘2\ feet between April and November 1959. The two calculated rates of seepage, projected over the entire period between the modern-day lake-level maximum in May 1958 and dryness in August 1961, suggest that total leakage during the 37-month period may have been about 10,000 acre-feet. This quantity represents almost 20 percent of the net lake-volume decrease during the same period (from about 60,000 acre-ft in May 1958 to dryness in Aug. 1961).
Additional evidence for leakage from Silver Lake is provided by chemical data (p. B41).
The shallow depression occupied by Malheur Lake is bordered by a flat plain that in places has extensive surface and near-surface deposits of peat derived from the decomposition of tules and other marsh-loving vegetation (Piper and others, 1939, p. 30). Crump and Hart Lakes occupy a similar environment. Although no direct evidence of lateral leakage from any of these lakes exists, seepage into the peat beds may occur under some conditions as the lakes rise, and much of the seepage probably does not return to the water bodies as their levels fall.
Harney Lake certainly does not leak, and Bluejoint Lake probably does not.
All the lakes except Harney and Bluejoint have overflowed within the last century, although the overflow frequency is wide in range. Silver Lake has not overflowed since 1905, presumably due in large part to the diversion of potential inflow for consumptive agricultural use. In
contrast, overflow from Malheur, Crump, and Hart Lakes is frequent. At Malheur Lake, for example, the average annual inflow is about 100,000 acre-feet, yet the lake’s capacity (including Mud Lake) at overflow is only about 120,000 acre-feet. Thus, overflow to Harney Lake can occur in a somewhat wetter than average year, even following a dry year.
Crump Lake overflows during many years, whereas outflow from Hart Lake and other downstream water bodies in the Warner basin is less common. Comparison of records for Hart Lake with streamflow data for Twenty-mile, Deep, and Honey Creeks suggests that overflow from the lake generally is small or nonexistent when the combined discharge at the three stream gages is less than about 100,000 acre-feet per year (somewhat below average). In relatively dry years, however, the water that reaches Hart Lake is pumped to irrigate nearby marshlands for crops of wild hay, and in some years the lake is pumped dry by late summer. Such pumpage is equivalent, in effect, to outflow in removing the water and its dissolved salts from the lake.
GEOCHEMISTRY
Analyses of lakes, streams, and ground water in the Silver Lake, Harney, and Warner basins are listed in tables 28, 30, and 31 (for the Harney Basin, many ground-water analyses, plus five additional stream analyses, were listed by Piper and others, 1939, p. 114-119). The small number of available lake-water analyses permit only a few specific comments.
At Silver Lake, the only available chemical-quality data, except for a single partial analysis by Van Winkle (1914, p. 117), were obtained during 1959-61, a period of virtually uninterrupted lake-level decline. The chemical character of the lake varied considerably during that period. In January and May of 1959, at a relatively high stage (alt about 4,307 ft), the lake contained 300 and 370 ppm of dissolved solids, respectively (analyses 31b and 31c, table 28), which consisted almost entirely of calcium, magnesium, sodium (in that order of abundance), bicarbonate, and probably silica (amount not determined). Concentrations of chloride were small—1.0 and 4.0 ppm, respectively. In contrast, on June 20, 1961, just prior to desiccation, the lake contained almost 700 ppm of dissolved solids, principally sodium, carbonate, and bicarbonate, with 40 ppm of chloride (analysis 31e, table 28).
Between April 25 and June 20, 1961, a marked change in chemical character accompanied the drop in lake level from an altitude of 4,300.85 to 4,300.32 feet (analyses 31d and 31e, table 28). In April (analyses 31d), principal components of the 686-ppm solute load were sodium, calcium, magnesium, and bicarbonate; the pH was 8.5. Although the dissolved-solids content had increased slightly to 698 ppm (analyses 31e) by June 20, the pH hadHYDROLOGY AND GEOCHEMISTRY, CLOSED-BASIN LAKES, OREGON
B41
increased to 9.7, and sodium plus chloride were enriched appreciably, while most other constituents were depleted (table 16). The amount of change can be estimated by comparing the pairs of sodium and chloride values. Chloride and, to a lesser extent, sodium can be used as indicators of concentration or dilution, because under many circumstances these two ions (especially chloride) do not enter appreciably into short-term reactions with other constituents of their environment. During the 2-month period in 1961, sodium in Silver Lake increased from 137 to 234 ppm (about 70 percent), and chloride increased from 21 to 40 ppm (about 90 percent). Thus, assuming an average increase of about 80 percent, the dissolved-solids content on June 20 should theoretically have been 1,234 rather than 698 ppm. (See table 16.) Much of the difference (about 220 ppm) apparently represents precipitation of calcium carbonate when the lake water became saturated with respect to the carbonate mineral as a result of (1) an increased pH due to lake-bottom plant growth, and (2) concentration of the lake water due to evapotranspiration (the more detailed aspects of carbonate-mineral equilibria are discussed on p. B49). However, significant amounts of silica, magnesium, and potassium also were lost during the same period.
An even greater loss of dissolved solids occurred at Silver Lake between May 1959 and June 1961. Again using sodium and chloride as indicators, the estimated dissolved-solids content would have increased from about 370 to about 3,400 ppm during the 25-month period if no solutes had been lost because of mineral precipitation or other chemical reactions. Yet, the actual value in June 1961 was only 698 ppm.
Without solute loss by any means, including leakage, the dissolved-solids content in June 1961 would have been 20,000-25,000 ppm, instead of about 700 ppm (fig. 30).
Table 16.—Theoretical and actual changes in the chemical character of Silver Lake, April 25-June 20, 1961
[«, about, or approx.]
Concentrations, in parts per million
Theoretical Per-
Constituent Initial new values Actual Loss centage
values, based on new loss
4-25-61 80-percent values, increasea 6-20-61
Silica (SiOz)............. 22 40 19 21 52
Calcium (Ca)______________ 59 106 16 90 85
Magnesium (Mg)------------ 44 79 18 61 77
Sodium (Na)_______________ 137 246 234 «0 «0
Potassium (K)_____________ 28 50 32 18 36
Carbonate (COi) plus bicarbonate (HCOa)h______ 373 671 335 336 50
Chloride (Cl).......— 21 38 40 «0 «0
Sulfate (SO4) plus fluoride (F) plus orthophosphate (PO4)............ 2.3 4.1 4.2 «0 «0
Dissolved-solids content-_ 686 1,234 698 530 43
pH_..................... 8.5 ............ 9.7 ..........................
0 80 percent of initial value. Does not allow for solutes lost by lake-bottom leakage.
6 Expressed as carbonate.
Figure 30.—Comparison of estimated relation between dissolved-solids concentration and stage of Silver Lake, 1959-61,
with and without solute loss due to mineral precipitation, other
chemical reactions, and lake-bottom leakage.
Thus, the various mechanisms of depletion caused an estimated 97-percent loss of solutes during the 25-month period.
Malheur Lake is characteristically dilute because of solute depletion by overflow, whereas Harney Lake is saline because it is the final accumulation sump for salts in the basin. The two analyses of Malheur Lake (samples collected in 1912 and 1961) are significantly different. Most of the difference is caused by (1) year-to-year changes in the lake’s stored-solute tonnage due to varying income and depletion by overflow, and (2) controls on the amount of calcium by calcium carbonate solubility. However, these factors do not explain the pronounced increase in relative amounts of sulfate (from two times the chloride content in 1912 to eleven times in 1961).., This increase may be due instead to agricultural application of gypsum -(calcium sulfate) as -an alkali-soil conditioner, whereas it probably is not due to increased ground-water use in the basin (the sulfate to chloride ratio in sampled ground water characteristically is 2:1 or less, according to the data of Piper, Robinson, and Park, 1939, p. 114-118).
In Warner basin, the chemical contrasts among Crump, Hart, and Bluejoint Lakes during infrequent periods when all contain water probably are similar to those shown by analyses made in 1912 (Van Winkle, 1914, p. 113). Each lake at a successively lower level has a successively greater concentration of dissolved solids because (1) the major streams in the valley feed the upstream lakes; (2) subsequent evaporation causes an increase in solute concentration; and (3) the lake-to-lake passage of water carries salts to, and accumulates them in, the lowest (northernmost) basins in the chain of lakes. Bluejoint Lake is the lowest point—the accumulation sump—in Warner basin. By assuming at least a modest long-term supply of dissolved solids to the lake, especially prior to extensive diversion and impoundment of upstream flowB42
CLOSED-BASIN INVESTIGATIONS
for irrigation, the relatively small solute concentration in 1912 (3,630 ppm) suggests prior depletion of solutes. Most of the salts may have been dissipated by wind during periods of desiccation, and some salt reportedly was removed for use by livestock (Van Winkle, 1914, p. 112). The effectiveness of depletion is evident by the ability of the dry lakebed to support crops of salt-tolerant wild hay in years when little water reaches the northern parts of Warner basin.
GEOCHEMICAL TRENDS AND PROCESSES
This section discusses the lake-to-lake similarities and contrasts in chemical character and the governing geochemical processes. The discussion includes consideration of the major dissolved constituents and properties, the minor components, and the trace elements.
MAJOR CONSTITUENTS AND PROPERTIES
DISSOLVED-SOLIDS CONCENTRATION
A wide range of dissolved-solids concentration occurs in j the lakes studied. For example, Crump Lake contained about 120 ppm of dissolved solids during a period of overflow in 1962, whereas true closed-basin lakes, such as Lake Abert, contain more than 100,000 ppm during periods of near dryness. The tonnages of stored solutes are similarly wide in range.
Five major factors determine the amount of dissolved solids in lakes that occupy arid,and semiarid, topographically enclosed basins:
1. The amount of solutes contributed to the lake by
inflow, '
2. The relative abundance of “stable” incbmitig con-
stituents, such as sodium, chloride,- and much of the carbonate plus bicarbonate. These'constituents tend to be retained in the lakes, rather than lost through organic and inorganic chemichl reactions. (This factor is important in all but the most dilute lakes.)
3. The amount of overflow or leakage. Regardless of how
many tons of dissolved solids arb added to a lake annually, the long-term accumulation of salts is • either diminished or prevented because of loss during overflow of leakage. • ‘
4. The frequency of lake desiccation, which, in turn, is
related to both1 2 3 4 5 the quantity and the year-to-year t■ regularity of inflow. The amount of dissolved l shits fe dependent; on the frequency of desiccation because deflation during the diy periods can remove large amounts of salts from a saline-lake basin.
5. The areal extent and the solute content of peripheral
mudflats, which were at one time inundated, but
which now lie outside the reach of lake water. If the marginal area at a particular time is sizable, a large solute tonnage can be entrapped within the playa muds. This last factor is most applicable to true closed-basin lakes.
The relative importance of each of the above factors is different in every lake; consequently, no two lakes have the same geochemical characteristics. For example, Lake Abert never overflows and rarely dries completely. Because of this, its solute loss is relatively slight, and the lake contains a large tonnage of dissolved solids, even though the principal lake constituents probably constitute less than one-half of the total income. Summer Lake, by contrast, does not overflow, but it has dried frequently in recent years. The resultant dissolved-solids tonnage in Summer Lake is small, as compared with that in Lake Abert, despite a moderate amount of inflow that contributes salts consisting largely of the principal constituents of the lake. The effect of recurrent desiccation at Summer Lake thus overshadows the factors favoring solute accumulation.
A different geochemical situation is characteristic of Goose Lake; it dries about as frequently as Lake Abert but has overflowed on rare occasions (the topographic divide is only 24 feet above the lowest point in the basin). Although the overflow has not removed significant amounts of dissolved solids during the last 100 years, it certainly must have done so prehistorically, before agricultural development of the valley began depleting potential inflow. In addition, large quantities of salts were blown off the central basin floor during desiccation in 1926 and in 1929-34. Surface-water flow into the lake doubtless contains only modest amounts of Goose Lake’s major constituents (sodium, bicarbonate, and chloride). The relatively small amount of dissolved solids in Goose Lake, therefore, has resulted from desiccation, from a presumably significant prehistoric overflow, and from a deficiency of stable incoming constituents.
Whereas Lake Abert provides an example of effective salt accumulation, Crump Lake represents the opposite extreme. Crump Lake occupies a very shallow basin that frequently overflows for several months during normal and wet years. Thus, the lake is not closed in the usual sense, but represents, instead, a transition between closed lakes and those that overflow every year. As a result, the lake contains only a small amount of solutes in spite of a rather large annual increment. The solute balance in Crump Lake illustrates the potential dominance of overflow.
SPECIFIC CONDUCTANCE
Specific conductance is a measure of the ability of a solution to conduct electric current, which, in turn,HYDROLOGY AND GEOCHEMISTRY, CLOSED-BASIN LAKES, OREGON
B43
depends on the number of ions in solution (the dissolved-solids concentration). A consistent relation exists between the specific conductance and the dissolved-solids concentration of most lakes and streams throughout the semi-desert area of south-central Oregon. In the specific-conductance range from 300 to 20,000 micromhos, the dissolved-solids content, in parts per million, is generally about 65 percent of the specific conductance. Above about 20,000 micromhos, the dissolved-solids content increases more rapidly than specific conductance, owing to decreased ionic mobility. Below 300 micromhos, the dissolved-solids value approaches that of specific conductance because of the increasing relative abundance of undissociated (nonionic) silica (fig. 34); the two values are equal at about 60 micromhos.
DENSITY
The relation between increasing dissolved-solids concentration and increasing density for saline lake waters in south-central Oregon is consistent throughout the range of available values, and can be described by the following equations, in which D is the density (in grams per milliliter at 20° C), and X and Xi are the dissolved-solids content, in parts per million and milligrams per liter, respectively:
D= (8.3X10-7)A+0.998;
D = 0.499+ V/0.249+(8.3X10-7)Xi.
SODIUM AND POTASSIUM
Sodium and potassium are by far the most abundant cations in almost all closed lakes of south-central Oregon. In Lake Abert, for example, the two alkali metals account for 99.99 percent of the positive ions, on the basis of equivalents per million. (By comparison, the two constituents make up 79 epm-percent of the cations in sea water.) Only in dilute lakes does sodium plus potassium amount to less than 50 epm-percent of the cations. For instance, on April 27, 1961, Crump Lake contained only 322 ppm of dissolved solids, and the value for sodium plus potassium was 46 epm-percent.
The increased importance of sodium and potassium relative to calcium and magnesium in the more concentrated lake waters reflects the loss of alkaline-earth cations (p. B49 -B50) and the resultant alkali enrichment.
The sodium and potassium content in streams tributary to the lakes is variable. In mountainous upstream reaches the two ions generally account for only 25 to 35 epm-percent of total cations (dissolved-solids content is characteristically less than 100 ppm). However, many of the streams pour forth from the mountainous headwaters onto marshy flatlands before flowing into the lakes. In these flat downstream reaches, the dissolved-solids content
o ; i= < <:
40
2g20
Abert <
Harney / /
/
/
/
Summer
/
•/
/
Cr
J____L
Hart t /
Crump y/
o°
■o
/
/
/
/
/» Goose
10 20
50 100 1000 10,000
DISSOLVED-SOLIDS, IN PARTS PER MILLION
100,000
Figure 31.—Relation between sodium to potassium ratio and dissolved-solids concentration in surface waters of south-central Oregon. Triangle indicates ratio in sea water. Open circles represent the Chewaucan River; solid circles represent lakes as indicated.
and the relative amount of sodium plus potassium increase, especially during periods of low flow. As an example, data for the Chewaucan River, the principal tributary to Lake Abert, indicate that the average dissolved-solids content at the river mouth is about 120 ppm, and the epm-percentage for sodium plus potassium relative to total cations probably averages about 40 (observed range, 35-54 percent). In contrast, values for the river about 30 miles upstream, near Paisley, are about 65 ppm of dissolved solids and only 35 epm-percent sodium plus potassium.
Widespread silicic extrusive rocks in the various basins are an ample source of sodium and potassium. (See Walker, 1963.) The two constituents are among the major elements released to solution as a result of alteration of silicate minerals (the chemical equation on p. B50 is an example of this type of alteration).
The relation between sodium and potassium in the surface waters of south-central Oregon is surprisingly consistent. (See fig. 31.) The amount of potassium relative to sodium decreases with the increase in dissolved-solids content. In dilute samples from the Chewaucan River, the epm-percentage of sodium is only about five times that of potassium, whereas in Goose Lake (dissolved-solids content, about 1,000 ppm) sodium is 20 times as abundant as potassium, and in Lake Abert (about 50,000 ppm) it is 50 times as abundant.
The decrease in potassium relative to sodium with an increase in dissolved-solids content reflects an actual rather than apparent loss of potassium. The potassium probably is depleted by chemical reaction with clay minerals on the lake bottom.B44
CLOSED-BASIN INVESTIGATIONS
100
80
O
I-
z
o
cr
UJ
Q.
<
>
D
O'
z
o
Z 40 <
o
<
o
K
UJ
20
> A,
L^e
N /
X _
tr
03
X
O
o
\6V
■ X
JQ
<
0L_O_ 50
O __ □
____L
100
—
"b O
200
V--
--------------------Sulfate
------------__------
~~ —•
500 1000 10,000
DISSOLVED SOLIDS, IN PARTS PER MILLION
100,000
Figure 32.—Changes in the distribution of major anions with increase in dissolved-solids concentration in surface waters of south-central Oregon. Open symbols represent the Chewaucan River; solid symbols represent lakes.
CARBONATE, BICARBONATE, AND pH
The combined amount of carbonate plus bicarbonate ranks either first or a close second among the anions of almost all closed lakes in south-central Oregon. However, the relative amount tends to be least in those lakes having the greatest dissolved-solids concentrations (fig. 32).
The straight-line relation between carbonate bicarbonate abundance and dissolved-solids content, as shown in figure 32, is in part fortuitous. At dissolved-solids content greater than 1,000 ppm, this trend may be explained partly by the varied effectiveness of long-term eolian removal of efflorescent carbonate bicarbonate evaporite salts from the lake playas in preference to the generally more hygroscopic salts of sulfate and chloride (p. B30; fig. 23). However, another significant reason probably is a basin-to-basin difference in the character of chemically important inflow.
The amount of calcium relative to that of carbonate bicarbonate in the inflow exerts an equally important
influence on the anionic character in the lakes. Calcium contributed to the lakes by inflow is depleted by precipitation of carbonate minerals. Therefore, if the incoming amount of calcium is comparable to that of carbonate plus bicarbonate, the loss of calcium and equivalent carbonate will leave a lake water with only small amounts of residual alkalinity. This, in turn, will result in a pronounced enrichment in chloride and sulfate relative to carbonate plus bicarbonate. Conversely, if calcium is a minor constituent in inflow, whereas bicarbonate is by far the most abundant anion, as in the Summer Lake basin, for example, carbon species will remain as the principal anionic constituents in the lake water.
Bicarbonate is by far the most abundant anion in nearly all stream waters of the region. In the analyses of 20 samples from 15 streams, bicarbonate (plus carbonate, where present) ranged from 35 to 97 percent and averaged 87 percent of the negative ions, on the basis of equivalentsHYDROLOGY AND GEOCHEMISTRY, CLOSED-BASIN LAKES, OREGON
B45
Table 17.—Carbonate-bicarbonate and calcium in stream waters
[Data from table 28]
Stream Location Discharge (cfs) Dissolved-solids content (ppm) Percentage of anion or cation epm total
Carbonate plus bicarbonate Calcium Carbonate plus bicarbonate in excess of calcium
Lake Abert basin
Chewaucan River 81 95 39 56
210 58 93 43 50
144 66 97 43 54
Do... 255 82 34 48
22 94 92 46 46
28 193 85 32 53
Poison Creek 35/21-1C 2 73 85 43 42
Summer Lake basin
Ana River 30/17-6SEM ~ 92 158 79 11 68
Goose Lake basin
Warner Creek 39/20-8R 3 525 35 15 20
Drews Creek 40/19-18P 15 94 95 38 57
Dry Creek . 41/19-7B 12 115 93 53 40
Davis Creek __ __ 45N/14E-16J 25 60 92 54 38
Willow Creek 47N/14E-22Q 5 111 91 40 51
Lassen Creek . 47N/14E-27L 15 65 81 36 45
Pine Creek 48N/15E-32L- 52 56 79 46 33
Silver Lake basin
Silver Creek.
28/14-28SW Y\.
23 66 94 47
47
Malheur Lake basin
Donner und Blitzen River.......... 32/32^-8SW^........ 43 85 90 36 54
Silvies River. ................... 22/30-8SW^......... 9 120 93 42 51
Warner basin
Deep Creek.. Honey Creek.
39/23-15NW ........ 20 90 94 36 58
36/24-29NEK-........ 1 135 93 43 50
Average.
125 87 39
48
per million (table 17). Calcium is also an important inflow constituent. For the same 20 samples, the alkaline earth ranged from 11 to 54 epm-percent and averaged 39 percent of the cations. However, the abundant carbon species overshadow the calcium (last column, table 17), and an appreciable amount of carbonate bicarbonate therefore remains in the lakes even after great losses due to the precipitation of alkaline-earth carbonate minerals.
Several reactions in addition to the depletion of incoming calcium tend to affect the carbonate plus bicarbonate content of closed-basin lakes. Ion-exchange mechanisms involving clay minerals within the lacustrine sediments can indirectly deplete carbonate. When sodium from the lake water replaces calcium on the clay-mineral exchange site, the released calcium combines with carbonate, as shown in the following idealized reaction:
Ca clay+2Na+1+C03~2—»Na clay+CaC03.
Two other reactions tend to add rather than deplete carbonate plus bicarbonate. Anaerobic decomposition of organic matter by sulfate-reducing bacteria, according to
the type of reaction discussed on page B46, yields bicarbonate. Similarly, the inorganic breakdown of silicate minerals (p. B50) produces bicarbonate, rather than hydroxyl ions (OH-1), if carbon dioxide enters the reaction along with water.
The relative importance of these several reactions is unknown, but all of them probably occur within or beneath the saline lakes. The trend of decreasing carbonate plus bicarbonate abundance with increasing average salinity (fig. 32) suggests, however, that the depletion mechanisms as a group are more effective overall than the bicarbonate-producing reactions.
The amount of carbonate relative to bicarbonate in the lakes of south-central Oregon differs widely. Compared on the basis of equivalents per million, the measured carbonate to bicarbonate ratio has ranged from less than 0.05:1 in Crump Lake to about 4:1 in Lake Abert. Relative amounts of the two carbon species are closely related to pH, a property that also varies among the lakes (from 8.4 in Crump Lake to 9.7 in Lake Abert). Both pH and the carbonate to bicarbonate ratio are invariablyB46
CLOSED-BASIN INVESTIGATIONS
Figure 33.—Variations in pH and carbonate to bicarbonate ratio relative to dissolved-solids concentration in lakes of south-central Oregon. Numbers indicate pH. Letters indicate: C, Crump Lake; H, Hart Lake; S, Silver Lake. Ratio and pH values are based on laboratory determinations.
greater in the lakes that contain larger concentrations of dissolved solids, as shown in figure 33. (See remarks regarding the reliability of carbonate plus bicarbonate and pH data on p. B56.)
This tendency is associated with decreased hydrolysis of the carbonate ion (C03-2) in solutions containing greater concentrations of carbonate and bicarbonate (Glasstone, 1942, p. 375), and therefore dissolved solids.
SULFATE
Sulfate is not a principal constituent of closed lakes in south-central Oregon; relative amounts rarely exceed 10 epm-percent of the negative ions, regardless of dissolved-solids content. In fact, the percentage of sulfate characteristically is smallest in lakes that contain the largest concentrations of dissolved solids (fig. 32).
Measured concentrations of sulfate have ranged from 26 ppm in Crump Lake (Apr. 1961; 322 ppm of dissolved solids) to 1,000 ppm in Lake Abert (July 1939; 95,000 ppm of dissolved solids). Of course, the maximum range has been much greater; Crump Lake doubtless contains less than 10 ppm of sulfate during periods of overflow, whereas concentrations greater than 10,000 ppm may occur in some lakes immediately before and after periods of desiccation.
Adequate chemical-quality data are not available for principal tributary streams near their mouths. Available data suggest, however, that most of the streams probably
contribute sulfate to the lakes in concentrations about equal to or greater than those of chloride (table 28).
The amount of sulfate in spring flow ranges widely, in terms of concentration and relation to accompanying amounts of chloride. Discharge of the Ana Springs system, principal source of dissolved solids for Summer Lake, contains about 6 ppm of sulfate and has a chloride to sulfate epm ratio of approximately 3:1, compared to about 6:1 for the lake itself. At Lake Abert, sampled peripheral springs yield sulfate in the 25- to 50-ppm range, with chloride to sulfate ratios from 3:1 to 9:1. In contrast, the lake itself contains 26 times more chloride than sulfate.
The data for streams and springs indicate a discrepancy between the relative amounts of sulfate and chloride contributed to the lakes and the relative amounts in the lakes themselves, with the discrepancy greatest in the most saline lakes. Thus, incoming sulfate apparently is depleted, rather than accumulated within the lakes. The most probable mechanism of depletion is bacterial sulfate reduction within the lake-bottom muds. Anaerobic decomposition of organic matter by bacteria generally is accompanied by reduction of sulfate, according to a reaction of the following type:
2CH20+S04-2+C03-2->HS-1+3HC03-1,
where organic matter such as carbohydrates is generalized as CH20 (Berner, 1966, p. 22). In shallow lakes, such as Summer and Abert, some of the reduced sulphur species produced by such reactions may be lost to the atmosphere, while other increments may combine with iron to form iron sulfide minerals within the lake-bottom muds.
CHLORIDE
Chloride is the most abundant or second most abundant anion in nearly all closed-basin lakes in south-central Oregon, even though it is a minor constituent of stream waters in the area (fig. 32). In lakes that rarely if ever overflow, chloride invariably makes up at least 20 epm-percent and in some lakes as much as 60 epm-percent of the negative ions. The greatest relative (percentage) amounts occur in lakes that characteristically contain more than 5,000 ppm of dissolved solids (Abert, Harney, and Summer Lakes among those studied). This increase with increasing dissolved-solids concentration is comparable to that suggested by Langbein (1961, p. 12) for closed lakes in general.
The relative amounts of chloride in closed-basin lakes can be used as a clue to the hydrochemical history of the lakes and their basins because chloride, more than almost any other constituent, is immune to the influence of most biologic, oxidation-reduction, and mineral-alteration phenomena, and it does not enter into mineral-precipitation reactions, except during the final stages of lake desiccation.HYDROLOGY AND GEOCHEMISTRY, CLOSED-BASIN LAKES, OREGON
B47
Relative chloride concentrations indicate the following with regard to the closed-basin lakes of south-central Oregon:
1. For lakes that characteristically contain more than about 1,000
ppm of dissolved solids:
a. If chloride is more than about 50 epm-percent of the
anions, the principal source of dissolved solids may be a chloride-rich ground-water supply; or, chloride may have undergone a long-term enrichment relative to carbonate plus bicarbonate; or, the amount of incoming calcium may be adequate to remove much of the incoming carbonate plus bicarbonate, thereby enriching chloride. The abundance of chloride can generally be attributed to a combination of these factors, with the importance of each differing from lake to lake.
b. If the relative amount of chloride in the lake is small,
streamflow containing only minor amounts of chloride may be the most important source of dissolved solids; or, most inflow may contain only small amounts of calcium relative to sodium, thereby diminishing the importance of chloride enrichment by the precipitation of alkaline-earth carbonates.
2. For closed lakes that characteristically contain less than about
1,000 ppm of dissolved solids, the conclusions that can be drawn from relative chloride content are, in general, similar to those above, but not as definitive. In such lake waters, calcium can be a major constituent, and if so, the relative amount of chloride is less than it would be in the same lake if the dissolved-solids content were greater. For example, on April 27, 1961, Crump Lake contained 322 ppm of dissolved solids, and calcium amounted to 32 epm-percent of the cations, whereas chloride totaled only 10 epm-percent of the anions. If the calcium and equivalent carbonate were removed from Crump Lake by mineral precipitation, chloride would increase to 15 epm-percent of the anions.
The concentration of chloride in stream waters varies with dissolved-solids content and downstream location. Except where chloride-rich ground water contributes significantly to surface flow, the amount of chloride present upstream from the flat marshy basin floors characteristically is 2 ppm or less, which in turn represents about 5 epm-percent or less of the total anions. Little is known about the chloride content of stream water as it enters the lakes after flowing through the marshy lowland areas. The chloride increment derived from these lowland areas is uncertain, but may be significant in several, if not all, of the valleys. In the Lake Abert basin, for example, the amount of chloride in the Chewaucan River near Paisley, upstream from Chewaucan .Marsh, rarely exceeds about 1 ppm, but concentrations of as much as 17 ppm (11 epm-percent of the anions) have been measured at the mouth near Valley Falls.
Even though the proportion of incoming chloride is increased in the lake, owing to the precipitation of alkaline-earth carbonate minerals, the chloride tonnages added to the lakes annually by stream inflow are, nonetheless, small compared with the amounts that have accumulated in the more concentrated lake waters.
Ground-water inflow undoubtedly is a significant chloride contributor in several of the closed basins, but its quantitative importance relative to streamflow is difficult to assess.
The ultimate source of chloride now present in the lakes is uncertain. Chloride derived from outside the study area is atmospherically transported into the region and contributed to the basins as dry fallout or as a constituent of precipitation. However, the amount of chloride thus introduced may be small, compared with the amount originating within the region.
The most important long-term source of “new” chloride may be related to volcanism, which was prevalent within, and adjacent to, the closed-basin area throughout most of Tertiary and Quaternary time.
Much of the chloride presently contributed to the lakes by ground and surface water has been recycled from within, or adjacent to, the specific basin. Recycled chloride in surface water is derived from wind-deposited dust on the land surface and from evaporite salts leached at and near the land surface, whereas part of the chloride contributed by spring flow is acquired from soluble salts in buried lakebed deposits and from salts leached within and below the soil zone by downward-percolating water.
MINOR CONSTITUENTS
Minor constituents studied include silica; calcium and magnesium; fluoride; bromide; nitrate; orthophosphate; and boron. In waters of more dilute lakes (those containing less than about 1,000 ppm of dissolved solids), calcium, magnesium, and silica constitute a large part of the dissolved-solids content. However, these constituents are of only minor abundance in the true closed-basin lakes, and their occurrence is thus considered along with that of the consistently minor components.
SILICA
Unusually large concentrations of silica occur in the closed-basin lakes of south-central Oregon. Most lake waters throughout the world contain 5 to 35 ppm of silica, and concentrations greater than 50 ppm are rare. In contrast, Abert and Summer Lakes are known to have contained more than 200 ppm of silica at times (table 28).
The characteristically large amounts of accumulated silica in the more saline, closed lakes of the study area might be expected simply on the basis of plentiful supplies—25 to 40 ppm—provided by tributary streams. Yet the amount of silica now present in the lakes is only a small fraction of the long-term amount introduced by surface and ground water. For example, the average annual increment of silica contributed to Lake Abert by the Chewaucan River is about 2,000 tons, meaning that the present silica load in Lake Abert (30,000-50,000 tons)B48
CLOSED-BASIN INVESTIGATIONS
Figure 34.—Relation between silica and dissolved-solids concentration in surface waters of south-central Oregon. A, Lake Abert; Cr, Crump Lake; G, Goose Lake; H, Harney Lake; Ht, Hart Lake; M, Malheur Lake; S, Summer Lake; open circles, Chewaucan River.
would theoretically accumulate in less than 30 years, assuming that no silica depletion occurs. The large concentrations of silica in the saline lakes are therefore not related to accumulation time. Instead, most of the incoming silica has been removed from the lake waters, as shown in figure 34. With increasing dissolved-solids content, silica constitutes a smaller and smaller percentage of the dis-
Figure 35.—Seasonal variations of silica tonnage in Lake Abert, 1961-63. Numbers above circles indicate dissolved-solids concentration, in thousands of parts per million; those below indicate water temperature, in degrees Fahrenheit. Horizontal bars indicate periods of appreciable inflow (more than about 50 cfs).
solved material, even though the absolute silica concentration increases from about 30 ppm in the dilute stream waters to 200 ppm or more in the most saline lakes. In addition to the variations from lake to lake, the silica tonnage in true closed-basin lakes varies seasonally.
Two factors can be important in controlling the amount of dissolved silica: the growth and decay of silica-using organisms and the reactions between silica-rich minerals and lake water. The importance of biologic controls on the | amount of silica in the lakes is uncertain. Diatoms, the j potentially most important silica users, are present in Lake Abert at least (p. B20), but their abundance, bloom | periods, silica requirements, and decomposition rates are unknown. Seasonal fluctuations in dissolved-silica tonnage i at the lake might be expected to mirror the activities of silica-rich biota. However, the data (fig. 35) are inconclusive—no consistent seasonal pattern is shown. During the 9-month period October 1962 to June 1963, the amount of silica in Lake Abert increased by 14,000 tons (a remarkable 40 percent of the initial tonnage), whereas only a 6,600-ton increment is attributable to inflow during that
Table 18.—Silica budget for Lake Abert, October 1, 1962-June 30, 1963
a. Lake-volume increase
(tables 4, 20)____________________________acre-ft__ 153,000
b. Evaporation (est. 22 in.)__________________aere-ft__ 64,000
c. Direct precipitation (est. 19 in.)_________acre-ft__ 55,000
d. Net volume increase attributable to inflow
(a+6 — c)_________________________________acre-ft-_ 162,000
e. Silica increment (fig. 35)____________________tons.. 14,000
/. Assume all inflow is from the Chewaucan River,
containing 30 ppm HiO2; therefore, silica
increment___________________________________tons- _ 6,600
g. Increment necessarily derived from organic or inorganic lake-bottom sediment (e—/)_________________tons-. 7,400
period (table 18). The difference, about 7,400 tons, must have been derived mostly from decomposition of silica-rich organic material in the lake, or perhaps from dissolution or alteration of inorganic deposits on the lake bottom. The 7,400-ton increment, accumulated during only 9 months, seems too large to be ascribed to the short-term breakdown or dissolution of silicate minerals. An analogous net loss of dissolved silica, about 20,000 tons, which occurred during the 9-month period July 1961 to March 1962 (fig. 35), also seems too great to have resulted from formation or alteration of silicates. Diatoms, which are the most important silica users, often become dominant in the late fall, winter, and early spring (K. V. Slack, U.S. Geol. Survey, written commun., 1964). Yet, biologic extraction of such a large amount of silica during the frigid, latter half of the period July 1961 to March 1962 (average air temperature 31° F at Valley Falls) seems uncertain. Thus, although organic processes apparently help to control the amount of dissolved silica in the lakes, their exact nature and relative importance are as yet unknown.HYDROLOGY AND GEOCHEMISTRY, CLOSED-BASIN LAKES, OREGON
B49
Quartz exerts no control on silica in the lakes, primarily due to the sluggishness of its equilibrium reactions, compared to those of amorphous silica. Calculations by Jones, Rettig, and Eugster (1967, fig. 1) indicate that amorphous forms of silica at 77° F should be soluble to the extent of about 110 ppm when the pH is less than 8.5. At higher pH, however, the dissociation of silicic acid causes a pronounced solubility increase: about 300 ppm of silica should be soluble at pH 9.7. Thus, the solubility controls exerted by amorphous silica should permit concentrations in the 150- to 200-ppm range in Abert and Summer Lakes (pH, 9.5-9.8) except at low temperatures.
The exact nature and extent of interactions between the lake water and silica-rich lake-bottom materials are unknown. At Lake Abert, bottom sediments include large amounts of an amorphous to montmorillonoid material which may have formed authigenically in contact with the lake water (B. F. Jones, U.S. Geol. Survey, oral commun., 1965), thereby providing a means of silica depletion.
At Summer and Goose Lakes, but not at Lake Abert, the seasonal fluctuations of silica tonnage are inverse to those of dissolved-solids concentration (figs. 15, 24, 28; footnote 1 on p. B18). This relation may result in part from restraints imposed on the concentration of silica by mineral equilibria.
CALCIUM AND MAGNESIUM
Calcium and magnesium, two of the most abundant constituents of most waters and rock-forming minerals, are present in only minute quantities in the saline lakes of the study area. In Lake Abert, for example, the two components usually total less than 5 ppm and constitute less than 0.02 epm-percent of the cations. The dearth of calcium and magnesium might seem strange in view of the great abundance of the two constituents in most stream waters of the study area. Excluding the Ana River, which is fed almost entirely by the Ana Springs, the alkaline-earth content of sampled streams in the study area ranged from 20 to 72 epm-percent of total cations. The average was about 60 percent, two-thirds of which was calcium.
Perhaps the most important reason for calcium depletion is the precipitation of calcium carbonate minerals. The chain of events that cause the loss of calcium is illustrated by assuming that a lake water originally contains about 200 ppm of dissolved salts (including major proportions of calcium and bicarbonate). As the calcium and dissolved-solids concentrations increase owing to accumulation of incoming salts and evaporation, the concentration of carbonate and its abundance relative to bicarbonate also increase (fig. 33). Finally, the solubility limits of alkaline-earth carbonates are exceeded, and the calcium plus the accompanying carbonate begin to drop out of solution. This decreased tolerance for calcium in solution as the dissolved-solids concentration, pH, and relative
Figure 36.—Changes in the relative amounts of calcium and carbonate plus bicarbonate with increasing dissolved-solids concentration in surface waters of south-central Oregon. Open symbols represent the Chewaucan River; solid symbols represent lakes. Numbers indicate carbonate abundance, as percents age of carbonate plus bicarbonate on the basis of equivalents per million.
proportion of carbonate increase is shown in general terms by the following data for Goose Lake (from table 28) .s
Dissolved solids (ppm) pH Relative abundance, as percentage of dissolved-solids concentration
Carbonate Calcium
823 8.9 6.0 1.7
868 8.8 3.5 1.5
1,040 9.0 7.6 1.7
1.270 9.1 7.4 1.0
1,340 9.2 11 .9
With continued salinity increase, the losses also continue until almost all calcium has been removed; this is possible because during the final stages of depletion, the carbon species are present in concentrations far greater than those of calcium (fig. 36). The last stages of depletion occur at a dissolved-solids content of about 1,500 ppm or less. Figure 36 also shows, by inference, the characteristic “path” of calcium depletion with increasing dissolved-solids content in lakes of the study area.
Depletion of calcium and carbonate is accelerated considerably at a high pH, such as that of Abert and Summer Lakes (9.5-9.8). Under these conditions, however, the actual precipitation of carbonate minerals (represented by reaction 1, below) tends to lower the pH, because hydroxyl ions (OH-1) are consumed during the
5 See remarks regarding the reliability of carbonate, bicarbonate, and pH data on p. B56.B50
CLOSED-BASIN INVESTIGATIONS
resulting replenishment of carbonate (reaction 2);
C03-2+Ca+2—»CaC03, and (1)
HC03+0H-1->C03-2+H20. (2)
Hence, the high pH must be maintained by chemical reactions other than carbonate precipitation. The two most important phenomena that can increase the pH of a lake water are hydrochemical alteration of silicate minerals and photosynthetic activities of aquatic plant life— especially algae. Both of these actions can be considered to produce hydroxyl ions, particularly where the counterbalancing uptake of carbon dioxide is sluggish. Silicate alteration is exemplified by the transformation of feldspar to clay minerals, such as albite to kaolinite:
2NaAlSi308+llH20
—»Al2Si205 (OH) 4+4H4Si04+2N a+1+20H-1.
Photosynthesis can be characterized by the following reaction, in which carbon is extracted from the lake water to produce cell material (here generalized as carbohydrate) :
HC03-1+H20—>CH20 + 02+OH-1.
Depletion of magnesium is not a simple matter of carbonate precipitation. Recent work by B. F. Jones (Jones and Van Denburgh, 1966, p. 443) on the mineralogy of lake-bottom sediments at Lake Abert indicates that much of the magnesium apparently combines with silica to form an amorphous or poorly defined montmorillonoid magnesian silicate.
Inorganic reactions certainly are not the only means of alkaline-earth depletion. Biota in the lakes undoubtedly require and utilize significant amounts of calcium and magnesium as nutrients. For example, the several major biologic species at Lake Abert use the alkaline earths (p. B20); similarly, some of the calcium that enters Summer Lake is assimilated as shell-building material by tiny ostracoda (a subclass of Crustacea), which are abundant in the lake-bottom deposits according to William Taft (oral commun., 1962). Although much of the biologid depletion is only temporary, some of it, at least, is permanent, as indicated by organic components of the lake-bottom sediment.
A striking example of alkaline-earth removal from a south-central Oregon lake is provided by two samples from Silver Lake (analyses 31c, d, table 28). Between April 25 and June 20, 1961, at least three-fourths of the calcium and magnesium and one-half of the carbonate and bicarbonate were lost from the lake during a period of evaporation and decreasing lake volume. (See p. B41; table 16.)
FLUORIDE
The minor-halide content of the lakes and their inflow was not studied in detail. Only fluoride was determined consistently in samples of lake, stream, and ground
waters. Concentrations of fluoride showed an unusually small variation compared with those of chloride, the major halogen. Dilute stream and ground waters characteristically contain from 0.7 ppm to less than 0.1 ppm of fluoride, yet concentrations ranging from about 1 ppm to not much more than 5 ppm are characteristic of the closed lakes. The largest measured amount—13 ppm—was from Harney Lake, sampled at a stage of near-dryness (analysis 34b, table 28).
The relative quantities of fluoride in saline lakes of the study area are thus less than might be expected, compared with the relative amounts contributed by inflow. Specifically, the Chewaucan and Ana Rivers—the principal sources of flow into Abert and Summer Lakes, respectively—have chloride to fluoride epm ratios ranging from 25:1 to 3:1, whereas ratios for the two lakes are about 1,400:1 and 150:1, respectively. The disposition of fluoride is similar to that of many other constituents (silica, calcium, magnesium, potassium, the carbon species, and sulfate) that are removed from solution as components of organic and inorganic material in the lakes. Fluoride is a trace constituent of biota (Rankama and Sahama, 1950, p. 761), and it also is sorbed, or actually incorporated, in certain authigenic minerals—both silicates and nonsilicates.
Sources of fluoride doubtless include chemical alteration of hydrous silicate minerals, deep-circulating ground water, and, in the past, volcanic emanations.
BROMIDE
Bromide is undoubtedly far less plentiful than fluoride in dilute waters (Livingstone, 1963, p. 41-42), but it is far more abundant in the most saline lakes. Measured bromide content ranged from about 1 ppm in Goose Lake to 103 ppm in Lake Abert (table 28).
Bromide is hydrologically one of the most mobile ions—that is, like chloride, it is immune to the influence of most organic and inorganic chemical reactions. Although many other dissolved constituents may be waylaid at various stages in a hydrologic sequence, bromide moves largely unhindered to the endpoint, or accumulation sump, of the sequence. Thus, bromide is amassed at the lowest parts of true closed basins, and the lake waters that occupy such depressions are enriched in bromide. As with chloride, the relative amount of enrichment is a reasonably good measure of the long-term effectiveness of solute accumulation. For example, bromide and chloride, respectively, make up 0.17 and 35 percent of the dissolved-solids content in Lake Abert (table 3), but only about 0.04 and 12 percent in Goose Lake. This reflects, at least in part, the greater efficiency of long-term accumulation in Lake Abert.
Because bromide and chloride are extremely mobile in almost any hydrologic environment, the relation between amounts of the two halides can be used to characterize aHYDROLOGY AND GEOCHEMISTRY, CLOSED-BASIN LAKES, OREGON
B51
particular lake water and to evaluate the potential sources of dissolved solids. For example, the chloride to bromide ratios (in equivalents per million) for Abert and Summer Lakes are 470:1 and 500:1, respectively. In contrast, other closed-basin lakes in the Western United States contain considerably smaller amounts of bromide relative to chloride, as shown by the following ratios (from data of Whitehead and Feth, 1961, table 1):
Chloride
Closed-basin lake to bromide
ratio (epm)
Big Borax Lake, Calif_______________________ 680:1
Pyramid Lake, Nev________________________________ 800:1
Mono Lake, Calif_________________________________ 920:1
Big Soda Lake, Nev_______________________________ 950:1
Walker Lake, Nev_______________________________ 1,000:1
Great Salt Lake, Utah__________________________ 3,700:1
Ocean water also contains less bromide relative to chloride than do the two Oregon lakes; the average ratio for sea water is about 660:1 (from data of Sverdrup and others, 1942, p. 176). On the basis of limited data, the chloride to bromide ratio of most volcanic emanations apparently ranges from about that of sea water to more than 10 times that of sea water (White and Waring, 1963, p. K5). At the other extreme, the ratio for igneous and sedimentary rocks is less than 350:1, according to data of Turekian and Wedepohl (1961, table 2).
These data suggest that rock alteration may be a significant source of bromide in Abert and Summer Lakes, despite the fact that bromide is such a minor component of most rock-forming minerals. For example, its average abundance in igneous rocks is less than one-hundredth of that of fluoride (Turekian and Wedepohl, 1961, table 2).
NITRATE
The nitrate content of most surface waters throughout the world is less than 5 ppm (Livingstone, 1963), and lakes in the study area are no exception. Aleasured concentrations ranged from 0.9 ppm in Silver Lake to 7.8 ppm in Goose Lake (table 28). The nitrate content of streams draining into the lakes is generally less than 1.0 ppm, as is that of most spring waters in the study area. (See comments on the reliability of nitrate data, p. B56.)
ORTHOPHOSPHATE
Most surface waters throughout the world contain less than 1 ppm of orthophosphate, as indicated by the data of Livingstone (1963). In contrast, the more saline lakes of south-central Oregon characteristically contain much more than 1 ppm; maximum measured amounts of orthophosphate in Goose, Summer, and Abert Lakes are 12, 41, and 101 ppm, respectively. (Table 3 shows, however, that the relative amount decreases, rather than increases, with increasing solute concentration.)
The data available on the orthophosphate content of streams draining into the lakes are only meager. The Chewaucan River, which chemically may resemble most
stream inflow, averages 0.2-0.3 ppm. Thus, orthophosphate apparently is enriched (accumulated) in the closed-basin lakes to a considerably greater degree than are silica and nitrate, the two other most abundant biologic nutrients. For example, the orthophosphate tonnage in Lake Abert (about 18,000 tons) is more than 1,000 times the 17-ton average annual contribution estimated for the Chewaucan River (assumed on the basis of a stream input of almost 48,000 acre-feet per year, with an orthophosphate content of 0.25 ppm), whereas the silica and nitrate tonnages in the lake are less than 50 times the annual stream input.
Large concentrations of orthophosphate are made possible in the true closed-basin lakes by a dearth of calcium. Calcium phosphate minerals are even less soluble than the calcium carbonates, but a great abundance of carbonate ion in the lake waters prohibits all but trace amounts of calcium, and the near absence of calcium, in turn, permits the accumulation of orthophosphate.
Seasonal fluctuations in relative amounts of orthophosphate were noted in Abert, Summer, and Goose Lakes, the only lakes for which seasonal data are available (figs. 15, 24, 28). However, the pattern of variation is different from lake to lake. Greatest and smallest amounts of orthophosphate occurred in Lake Abert during summer and winter, respectively, whereas in Summer and Goose Lakes they occurred erratically during the 1^-year study period. Because of the importance of orthophosphate as a biologic nutrient, its fluctuations may be due primarily to seasonal variations in the rates of biotic growth in the lakes. However, fluctuations within Lake Abert seem to contradict those to be expected on the basis of observed biotic blooms. For example, amounts of dissolved orthophosphate increased during the period March-July, yet during the same period a large biologic-population increase was observed in the lake. In 1962, the March-July increase was about 1,500 tons—only 10 percent of the initial 15,000-ton load, but far more than the amount contributed by ground water and surface inflow (estimated to be only 10 tons). For the average lake volume during the period (about 170,000 acre-ft), the orthophosphate increment represented a sizable increase in concentration (6-7 ppm). The apparent conflict between orthophosphate gains and biologic blooms at Lake Abert may be explained at least in part by the accelerated liberation of orthophosphate by decomposing organisms on the lake bottom as the water temperature increases during the spring and early summer (fig. 15).
At Summer and Goose Lakes, not enough is known about the nature, extent, and timing of biologic activity to evaluate its influences on orthophosphate. The seasonal variations at Goose Lake (fig. 28) show no apparent relation to water temperature; rather, a crude inverse relation apparently exists between orthophosphate tonnage and dissolved-solids content. Because Goose Lake contains calcium in more than trace amounts (table 28), theB52
CLOSED-BASIN INVESTIGATIONS
controls imposed by calcium phosphate insolubility may help to govern the orthophosphate content.
BORON
Although boron comprises less than one-half of 1 percent of the dissolved-solids content in the more saline Oregon lakes, the concentrations of this constituent are nonetheless large. Maximum measured amounts of boron in Goose, Summer, and Abert Lakes are 4.8, 34, and 85 ppm, respectively. Large concentrations are not restricted to lakes of the study area, however. Boron is an abundant minor-element component of closed-basin lakes throughout the Western United States (Whitehead and Feth, 1961, table 1); several lakes in the Western States contain even greater relative and absolute amounts than do the lakes in south-central Oregon.
The only available information on the boron content of streams in the study area is the analysis of a sample collected August 7, 1963, from Chewaucan River at Valley Falls (dissolved-solids content, 102 ppm; discharge, about 20 cfs), which shows only 0.08 ppm of boron.
The boron content of ground water is not much better known. Hot springs in the study area contribute some boron; measured concentrations in four samples ranged from 1.0 to 9.9 ppm (table 30). No data are available for nonthermal and warm springs—except Ana Springs, which yields about 0.5 ppm boron (analysis 20b, table 28). Sampled well waters not associated with hot-spring activity characteristically contained less than 1 ppm, and some contained almost none (table 31).
The abundance of boron in the saline lakes of south-central Oregon is attributable to an adequate supply of the constituent and to its solubility in the very soft waters that are characteristic of the lakes. Not all the boron that enters the lakes necessarily remains in solution, however. Boron, though a mobile component in many hydrochemical environments, is not immune to incorporation within organic and clay-mineral phases (Rankama and Sahama, 1950, p. 488; Oborin and Zalkind, 1964; Harder, 1964).
The ultimate source of boron in the lakes is conjectural; however, the volcanic environment of the closed basins and the characteristic association of boron (as well as chloride and fluoride) with volcanic activity suggests such a source. An extensive part of the Western United States— which includes the study area—may be peculiar in this regard. White (1957, p. 1679) stated that “Data on the boron content of glassy volcanic rocks * * * and of thermal waters * * * suggest that California, western Nevada, and southern Oregon constitute a high-boron province.”
TRACE ELEMENTS
Only meager data are available on the trace-element content of closed lakes and their tributaries in south-central Oregon. Spectrographic analyses for 17 elements were performed on samples collected June 12, 1962, from
Summer and Goose Lakes and from the Chewaucan River, and collected January 12, June 12, and September 17, 1962, from Lake Abert (table 29).
Among the 17 trace elements tested for in the lake-water samples, only five—aluminum, iron, molybdenum, nickel, and vanadium—occurred in amounts consistently greater than the lower limits of quantitative determination. Other elements whose concentrations were measurable in one or two of the five lake-water samples are cobalt, manganese, lead, and titanium. Largest measured amounts, in parts per billion, of the nine constituents are tabulated below:
[>, greater than; =, about, or nearly equal to]
Element Parts per
billion
Aluminum (Al)___________________________ 270
Cobalt (Co)______________-______________ 19
Iron (Fe)_________________________________ >96
Lead (Pb)__________________________________ 23
Manganese (Mn)__________________________ =4.8
Molybdenum (Mo)___________________________ >48
Nickel (Ni)________________________________ 17
Titanium (Ti)___________________________ =1.9
Vanadium (V)_____________:______________ 77
Maximum concentrations for all but one of the nine constituents occurred in Lake Abert (the exception, 19 ppb of cobalt, occurred in Summer Lake).
Six elements were abundant enough to be determined quantitatively in the lone sample from the Chewaucan River (table 29), although the relative abundances may not be at all representative because the dissolved-solids content (193 ppm) is about 1.6 times the estimated average. Nonetheless, the amounts may suggest the general order of magnitude in stream inflow compared with that in the lakes themselves.
A comparison of the relative amounts (expressed as percentages of the total dissolved-solids content) of several trace elements in the samples collected in June 1962 from the Chewaucan River and from Goose, Summer, and Abert Lakes is given in table 19. The analyses, which are arranged in order of increasing dissolved-solids content, show few consistent trends of trace-element enrichment or depletion. (The same is true for the analyses of
Table 19.—Relative amounts of six trace elements in samples from Chewaucan River near Valley Falls and Goose, Summer, and Abert Lakes, June 12, 1962
[Element data shown as percentage of dissolved solids multiplied by 105. All data recalculated from table 29. <, less than]
Chewaucan Goose Summer Lake
Constituent River near Lake Lake Abert
Valley Falls
Aluminum (Al)_____ 460 87 270 16
Cobalt (Co)_______ 78 <11 30 <1.0
Iron (Fe)_________ 3,300 40 24 18
Molybdenum (Mo) 21 50 70 2.6
Nickel (Ni)_______ 150 13 3.0 3.5
Vanadium (V)______ 300 110 2.6 16
Dissolved-solids content, in parts per
million_____________ 193 1,270 6,270 49,100HYDROLOGY AND GEOCHEMISTRY, CLOSED-BASIN LAKES, OREGON
B53
three Lake Abert samples of different dissolved-solids content. See table 8.) The lack of clear and consistent trends may reflect the basin-to-basin differences in the chemical character of inflow or the different rates and forms of biologic and non-biologic reactions within the lakes. (See, also, discussion on reliability of trace-element data, p. B56.)
On the assumption that data in tables 8 and 19 are at least semiquantitatively representative of the trace-element content of the lakes and stream, several general characteristics are indicated by the percentages. Five of the elements are substantially depleted as the dissolved-solids content increases from 193 ppm (Chewaucan River near Valley Falls) to 1,270 ppm (Goose Lake); only molybdenum is enriched. Relative (percentage) amounts of all six trace elements are significantly less in saline Lake Abert than in either Chewaucan River near Valley Falls or comparatively dilute Goose Lake. Depletion mechanisms doubtless include (1) biologic uptake as trace nutrients (cobalt, iron, molybdenum, and vanadium, among others) (Hutchinson, 1957, p. 824, 826; Oborn, 1960, p. Ill; Rankama and Sahama, 1950, p. 598-600), (2) precipitation in a reducing environment (cobalt, iron, and vanadium, among others), and (3) diagenetic mineral alteration (aluminum).
LOCATION SYSTEM
The location method used in this report is based on the rectangular system of land division. For example, in the symbol 33/21-2M, the two numbers preceding the dash indicate the township and range south and east, respectively, of the Williamette base line and meridian (T. 33 S., R. 21 E.) in Oregon. The last number after the dash indicates the section (sec. 2), and the letter “M” denotes the specific 40-acre tract among the 16 tracts within section 2, using letter designations, as shown in figure 37.
Location symbols for the California part of Goose Lake basin indicate the township and range north and east of the Mount Diablo base line and meridian. They are distinguished from locations in Oregon by inclusion of the letters “N” (north) and “E” (east) as parts of the symbol (for example, 45N/14E-16J).
RECORDS OF LAKE-SURFACE ALTITUDE
Data for studied lakes are presented as follows: Lake Abert (table 20), Summer Lake (table 21), Goose Lake (table 22), Silver Lake (table 23), Malheur and Harney Lakes before 1900 (table 24), Malheur Lake after 1900 (table 25), Harney Lake after 1900 (table 26), and Crump, Hart, and Bluejoint Lakes (table 27).
D C B A
E F G H
M L K J
N P Q R
SECTION 2
Figure 37.—Rectangular system of land division, showing letter designations of the sixteen 40-acre tracts within a given 1-square-mile section. For example, the location 33/21-2M referred to in text (patterned tract shown above) is the NWJSWf sec. 2, T. 33 S., R. 21 E., in Oregon.
CHEMICAL ANALYSES METHODS OF ANALYSIS
Analytical determinations performed during the period of intensive reconnaissance study, April 1961-June 1963, are, for the most part, those used for standard water analysis by the U.S. Geological Survey (Rainwater and Thatcher, 1960). Colorimetric determinations employed a Beckman model “B” spectrophotometer, and flame-photometric determinations utilized a similar instrument with a flame attachment. The ranges of values determined in the Geological Survey’s Portland, Oreg., laboratory for the constituents and properties of closed-basin waters of south-central Oregon during the study are as follows:
[Data reported in parts per million except as indicated]
Constituent or property Smallest value Largest value
Silica (Si02) 9.1 207
Calcium (Ca)__ 1.4 59
Magnesium (Mg) .3 44
Sodium (Na) _____ 3.2 28,400
Potassium (K) 1.6 752
Bicarbonate (HC03) _ __ 35 5,930
Carbonate (C03)_ __ _ __ 0 9,460
Sulfate (S04) _ .2 990
Chloride (Cl) _ .0 25,000
Fluoride (F) __ _ _ _ _ .0 13
Bromide (Br)__ _ _ _ 1 101
Nitrate (N03)__ _ ______ .0 1.2
Orthophosphate (P04)_ .12 101
Boron (B) .51 85
Specific conductance
(micromhos at 25° C) 56 85,100
pH 6.8 9.7
Density at 20° C — 1.058B54
CLOSED-BASIN INVESTIGATIONS
The actual methods used and the refinements made necessary by the unusual character of the more saline samples are summarized below. Reference is made throughout to the detailed analytical procedures described by Rainwater and Thatcher.
Silica.—Silica content was determined colorimetrically (Rainwater and Thatcher, 1960, p. 259) on all samples containing less than 100 ppm of Si02, whereas gravimetric procedures (Rainwater and Thatcher, 1960, p. 261) were used for most concentrations exceeding that value (Abert and Summer Lakes). All samples were filtered through
0. 45.micron Millipor pads as soon as possible after collection (usually within 1 week), and were stored in polyethylene bottles to prevent silica pickup from the conventional borosilicate glass containers. When dilutions were necessary prior to colorimetric determination, they were made immediately before the analysis was made because a significant amount of silica pickup was noted in the dilutions stored in conventional volumetric flasks, owing to the caustic high-pH nature of the more saline waters. (Of course, the error introduced by such pickup would be magnified even more by the dilution factor.)
Refinements employed as part of the gravimetric determination were as follows:
1. 100-ml sample aliquots were used, even though the
resulting dissolved material totaled as much as 8 grams. The abundance of salts necessitated extreme care, but did not seem to alter the analytical accuracy of the determination.
2. Silica was determined colorimetrically from the residual
filtrate (produced by means of step 8 of Rainwater and Thatcher’s procedure), and the measured quantity was a significant part (3 to 10 percent) of the total silica. However, because of low silica concentrations in the presence of abundant orthophosphate in the filtrates (measured P04 concentrations in Lake Abert were as much as 101 ppm during the period of study), an arithmetic correction was necessary to allow for enhancement of the molybdate-blue color from orthophosphate. Using cells with a 10-mm optical depth, the amount of enhancement was found to be about 0.13 absorbency units per milligram of orthophosphate in the 10-ml aliquot. Thus, if the initial 100-ml portion contained 8.0 mg of P04 (equivalent to 80 mg/1 in the original water sample) and the filtrate were diluted to 250 ml, a 10-ml aliquot thereof would contain 0.32 mg of PO4, and the absorbency-correction factor would be —0.042 units. A correction of this type is unusual, necessary only when the concentrations of orthophosphate are large; most natural waters contain less than 1 ppm of orthophosphate and do not require gravimetric techniques for the determination of silica.
Calcium and magnesium.—Determinations for calcium and magnesium were made by using normal complexometric procedures (Rainwater and Thatcher, 1960, p. 127, 174, 197) for samples containing less than 10,000 ppm of dissolved solids. However, the reliability of values, especially for calcium, at salinities exceeding about 2,000 ppm is questionable. Above 10,000 ppm, the complexometric determination for calcium plus magnesium (results of which are expressed as hardness, in terms of parts per million of calcium carbonate) seems to give good semi-quantitative results. This method, therefore, was used on several samples in preference to the expensive, time-consuming permanganimetric and pyrophosphate methods (Rainwater and Thatcher, 1960, p. 129, 198). In fact, the more costly methods were used only on sample If (table 28).
Sodium and potassium.—Sodium and potassium were measured using conventional flame-photometric techniques (Rainwater and Thatcher, 1960, p. 55). Samples were diluted when necessary to bring the amounts of sodium and potassium below 50 and 10 ppm, respectively. As with the determination of silica, dilutions were made shortly before analysis to prevent sodium pickup from the sodium borosilicate glass of volumetric flasks. Because the emission intensity of the potassium flame is enhanced by accompanying sodium, the amount of sodium in standards, blanks, and samples or dilutions thereof was augmented, as necessary, to obtain a constant background. This was accomplished by adding as much as 5.00 ml of a 2,000-ppm sodium standard to 25.00 ml of the sample, dilution, standard, or blank. The final aliquot volume was then brought to 30.00 ml by the addition of distilled water.
Each sample or dilution thereof was analyzed in combination with standards containing slightly smaller and greater sodium or potassium concentrations. This technique of close bracketing helped to ensure the accuracy of results, and it permits calculations to be made from simple proportionalities.
Carbonate, bicarbonate, and pH.—Carbonate, bicarbonate, and pH were determined in the laboratory by use of the conventional potentiometric method (Rainwater and Thatcher, 1960, p. 94) and a Beckman pH meter (model H-2). For samples containing more than about 75 epm of carbon species, a sulfuric-acid standard of 10 times the usual (0.01639A) strength was employed as the titrant.
Sulfate.—Sulfate was determined by one of two methods, depending on the amount present in, and the chemical character of, the sample. For dilute waters the visual thorin method (Rainwater and Thatcher, 1960, p. 279) was employed. However, modifications of the procedure listed by Rainwater and Thatcher (1960, p. 280) were used by the U.S. Geological Survey’s Portland laboratory for the routine determination of sulfate. The followingHYDROLOGY AND GEOCHEMISTRY, CLOSED-BASIN LAKES, OREGON
B55
modifications were made (keyed to the procedural steps listed by Rainwater and Thatcher):
Step 1.—The cation ion-exchange resin was first washed with distilled water before rinsing with sample.
Step 2.—(Identical with the second procedural step of Rainwater and Thatcher, 1960, p. 283.)
Step 3.—A volume of deionized sample (10-ml maximum) containing less than 5 mg of sulfate was pipeted into a porcelain evaporating dish. Excessive amounts of chloride interfere with a visual determination: a maximum of 1,000 ppm chloride is tolerated, but below that amount, the chloride-to-sulfate ratio must be less than 20:1. Where the chloride content of reconnaissance samples did not meet the qualifications, sulfate was determined gravimetrically (Rainwater and Thatcher, 1960, p. 283).
Steps 4, 5, and 6.—Ethyl alcohol, rather than dioxane, was used as the organic medium. It was combined and stored with the thorin indicator solution (1.3 ml of 0.2 percent thorin to 1.0 liter of alcohol), rather than added separately at the time of determination. Forty milliliters of the alcoholic indicator solution was added to the deionized sample of step 3 above, and the pH was then adjusted to about 4.5, using a pH meter and 0.1 or 0.5 normal sodium hydroxide, or 2-percent hydrochloric acid.
Step 7.—A blank was carried through the entire procedure and titrated with samples. Its titration volume (generally less than 0.06 ml) was subtracted from the sample titration volumes for more accurate low-concentration results.
Chloride.—Concentrations of chloride were measured using the Mohr volumetric method (Rainwater and Thatcher, 1960, p. 141). For samples containing only a small amount of chloride (less than 10 ppm), the determinations were made on 100-ml aliquots evaporated to about 50 ml. Where chloride concentrations exceeded about 1,500 ppm, dilutions were made to maintain the amount of chloride in a 25-ml aliquot at 40 mg or less.
Fluoride.—Determination of fluoride was by the colorimetric Zirconium-Eriochrome Cyanine R method (Rainwater and Thatcher, 1960, p. 163). Preliminary dilutions were made, when necessary, to keep the dis-solved-solids content of samples below about 8,000 ppm.
Bromide.—The concentrations of bromide plus iodide and of iodide alone were initially determined by the conventional oxidation method (Rainwater and Thatcher, 1960, p. 121). The determination of bromide plus iodide is considered accurate, but later determinations of iodide alone (in 1964), by the more accurate photometric method, showed that actual concentrations of iodide were far less than those indicated by the oxidation technique. (For example, results of the oxidation method gave an iodide concentration of about 3 ppm in Lake Abert, whereas the photometric method indicated only 0.2-0.5 ppm.) Therefore, bromide values listed in table 28 represent the concentration of bromide plus iodide (oxidation method), corrected for the small amounts of iodide that later were shown to be typical.
Nitrate.—The nitrate content in dilute samples was
measured colorimetrically by use of the phenoldisulfonic acid method (Rainwater and Thatcher, 1960, p. 216), whereas the more conventional reduction technique (p. 219) was used for those samples containing more than 50 ppm of chloride.
Orthophosphate.—Orthophosphate was determined
colorimetrically by use of the phosphomolybdate method (Rainwater and Thatcher, 1960, p. 245). However, several minor refinements of the basic procedure described by Rainwater and Thatcher (1960, p. 247) were employed to increase the analytical accuracy. To prevent deterioration, all reagents were prepared with extreme care immediately prior to analysis, and were then carefully added to the sample aliquots by means of pipets. (Use of volumetric pipets aided accuracy by reducing the small variations in reagent volume which cause sizable differences in sample absorbency.) The digestion period between the addition of the molybdate and stannous chloride reagents was kept at 7.0 minutes. Finally, the determination of sample absorbency was made exactly 20 minutes after the stannous chloride reagent had been added. Exact timing was vital, for the absorbency “plateau” that follows color development but precedes color deterioration was characteristically attained during the period between about 15 and 25 minutes after the stannous chloride had been added.
Boron.—Concentrations of boron were determined by the colorimetric dianthrimide method (Rainwater and Thatcher, 1960, p. 113).
Specific conductance.—Measurements of specific conductance were made using the Wheatstone bridge method (Rainwater and Thatcher, 1960, p. 275) and a U.S. Geological Survey instrument. The conductivity cell had a constant of 0.345.
Trace elements.—Trace element contents of six samples (table 29) were determined by means of the spectro-chemical separation-concentration technique perfected by William D. Silvey of the U.S. Geological Survey (Silvey and Brennan, 1962). The analyses were performed by Silvey and his coworkers in the U.S. Geological Survey laboratory at Sacramento, Calif.
ACCURACY OF ANALYTICAL RESULTS
The overall accuracy of chemical analyses performed as part of this study is good; however, several minor points deserve mention. Chloride concentrations smaller than 1 ppm are of only semiquantitative accuracy when determined by the conventional Mohr volumetric method, due to the very small titration volumes. The difficulty of accurate determination is compounded by natural water color, which is intensified during evaporation of a 100-ml aliquot to 50 ml prior to titration.
After the reconnaissance analytical work had been completed, analysts discovered that even the very small chloride concentrations (much less than the 50-ppmB56
CLOSED-BASIN INVESTIGATIONS
maximum indicated on p. B55) interfered with the accurate determination of nitrate by the phenoldisulfonic method. Thus, the apparent nitrate values shown in table 28 for samples 14b, 31d, 31e, and perhaps for several earlier samples as well, may be somewhat less than the amounts of nitrate actually present.
The chemical changes that occurred within samples during the interim between the time they were collected and analyzed may also have caused some inaccurate analytical results. Several of the potentially most important chemical changes are discussed briefly below.
The determination of large silica concentrations in high-pH waters, such as those of Abert and Summer Lakes, is subject to some error because the solubility of amorphous silica is strongly dependent on temperature (Morey and others, 1964, fig. 1). Thus, any turbid sample collected at a relatively low temperature (35°-50° F) and later filtered at laboratory temperature may have gained a significant amount of dissolved silica in the interim. The effect was minimized by filtering the samples in the laboratory as soon after collection as possible. (Field filtration was difficult or impossible with available equipment, owing to the very fine grained particles in turbid suspension.)
All considerations of pH and of carbonate to bicarbonate ratios, based on data shown in table 28, are necessarily tentative because the analytical determinations were performed in the laboratory as much as 1 month after sample collection. Both parameters are subject to rapid change, owing to their variability with temperature change and with partial pressure of C02, biologic activity, and reactions between the water sample and accompanying sediment. Similar caution should be applied to the interpretation of the laboratory-determined values given for calcium in saline samples with high pH and abundant carbonate because of possible loss or gain of alkaline-earth carbonates in solution due to solubility considerations.
Nitrate and orthophosphate were determined on samples that were neither treated with preservative at the sampling site nor refrigerated between collection and analysis. Thus, because of biologic activity in the samples after collection, the amounts of nitrate and orthophosphate reported for streams and dilute lake waters may not accurately reflect the amounts actually present at the time of sampling. Orthophosphate concentrations exceeding about 1 ppm in the more saline lakes probably are accurate.
Finally, the limitations imposed on the significance of spectrographic trace-element analyses (table 29) by inadequate sample-treatment techniques may be important. For example, large amounts of aluminum reported in the Summer Lake analysis may be correlated with
(1) the pronounced turbidity of the sample when collected;
(2) the long interval between collection and filtration (about 30 hours); and (3) the presence of a near-colloidal suspension in the sample even after vacuum filtration
through 0.45-micron pore-size pads. In other words, the reported aluminum content, and perhaps that of other trace elements as well, may represent components of the fine dispersion, rather than comprising true solutes. Alternatively, elements originally present in the sample as true solutes may have subsequently been sorbed by the clay minerals in suspension, or depleted or enriched owing to biologic activity or decay within the sample under conditions of varying temperature, C02 pressure, and pH during the period between collection and filtration. Thus, analytical results in table 29 may not represent the exact trace-element content of the aqueous phase of a sampled water. However, they are considered to give at least a semiquantitative indication of the trace-element characteristics.
TABLES OF CHEMICAL ANALYSES
A compilation of chemical analyses is presented in tables 28-31, as follows: Table 28 gives results of analyses for the lakes and their tributary streams. Table 29 shows the trace-element content of samples from Abert, Summer, and Goose Lakes and from the Chewaucan River. Tables 30 and 31 contain representative chemical analyses of water from springs and wells in the several topographically enclosed basins. Sample-collection sites are numbered in each table (lakes and streams, 1-41; springs, 101-111; wells, 112-136), and the locations of many are shown on plates 1 and 2. In addition, location of each site is indicated by the number and letter code of the rectangular system of land division (p. B53).
REFERENCES
Allison, I. S., 1940, Study of Pleistocene lakes of south-central Oregon: Carnegie Inst. Washington Yearbook 39, p. 299-300.
------ 1945, Pumice beds at Summer Lake, Oregon: Geol. Soc.
America Bull., v. 56, no. 8, p. 789-807.
------ 1946, Pluvial lakes and pumice, in [Symposium] Early man
in Oregon: Sci. Monthly, v. 62, no. 1, p. 63-65.
------ 1954, Pluvial lake levels of south-central Oregon [abs.]:
Geol. Soc. America Bull., v. 65, no. 12, p. 1331.
Allison, I. S., and Mason, R. S., 1947, Sodium salts of Lake County, Oregon: Oregon Dept. Geology and Mineral Industries GMI Short Paper 17, 12 p.
Antevs, Ernst, 1938, Rainfall and tree growth in the Great Basin: Am. Geographical Soc. Spec. Pub. 21, 85 p.
Berner, R. A., 1966, Chemical diagenesis of some modern carbonate sediments: Am. Jour. Sci., v. 264, p. 1-36.
Brown, S. G., 1957, Occurrence of ground water near Ana Springs, Summer Lake basin, Lake County, Oregon: U.S. Geol. Survey open-file report, 26 p.
California Department of Water Resources, 1965, Ground water quality, in Hydrologic data, 1963—Northeastern California: California Dept. Water Resources Bull. 130-63, v. 2, app. E, p. 147-195.
■------ 1966, Goose Lake water quality control policy: Central
Valley Regional Water Quality Control Board Rept. 5 [basic data, prelim, ed.].
Free, E. E., 1914, The topographic features of the desert basins of the United States, with reference to the possible occurrence of potash: U.S. Dept. Agriculture Bull. 54, 65 p.HYDROLOGY AND GEOCHEMISTRY, CLOSED-BASIN LAKES, OREGON
B57
Gladstone, Samuel, 1942, An introduction to electrochemistry: Princeton, N.J., D. Van Nostrand Co., 557 p.
Hampton, E. R., 1964, Geologic factors that control the occurrence and availability of ground water in the Fort Rock Basin, Lake County, Oregon: U.S. Geol. Survey Prof. Paper 383-B, 29 p.
Harbeck, G. E., Jr., 1955, The effect of salinity on evaporation: U.S. Geol. Survey Prof. Paper 272-A, 6 p.
------- 1962, A practical field technique for measuring reservoir
evaporation utilizing mass-transfer theory: U.S. Geol. Survey Prof. Paper 272-E, p. 101-105.
Harbeck, G. E., Jr., and others, 1951, Utility of selected Western lakes and reservoirs for water-loss studies: U.S. Geol. Survey Circ. 103, 31 p.
Harder, Hermann, 1964, To what extent is boron a marine index element?: Geochem. Soc., Geochemistry Internat. No. 1, 1964, p. 105-112 [in English, 1964, translated from Fortschr. Geol. Rheinld. u. Westf. 10, p. 239-253 (1963)].
Harding, S. T., 1965, Recent variations in the water supply of the western Great Basin: California Univ. Water Resources Center Archives, Archives Ser. Rept. 16, 226 p.
Hubbs, C. L., and Miller, R. R., 1948, The zoological evidence— Correlation between fish distribution and hydrographic history in the desert basins of Western United States, pt. 2 of The Great Basin, with emphasis on glacial and postglacial times: Utah Univ. Bull., v. 38, no. 20, p. 18-166.
Hutchinson, G. E., 1957, A treatise on limnology: New York, N.Y., John Wiley & Sons, v. 1, 1015 p.
Jones, B. F., Rettig, S. L., and Eugster, H. P., 1967, Silica in alkaline brines: Science, v. 158, no. 3806, p. 1310-1314.
Jones, B. F., and Van Denburgh, A. S., 1966, Geochemical influences on the chemical character of closed lakes, in Symposium of Garda, October 9-15, 1966, Hydrology of lakes and reservoirs Proceedings: Internat. Assoc. Sci. Hydrology Pub. 70, p. 435-446.
Keen, F. P., 1937, Climatic cycles in eastern Oregon indicated by tree rings: Monthly Weather Rev., v. 65, no. 5, p. 183-188.
Kohler, M. A., Nordenson, T. J., and Baker, D. R., 1959, Evaporation maps for the United States: U.S. Weather Bur. Tech. Paper 37, 12 p.
Langbein, W. B., 1961, Salinity and hydrology of closed lakes: U.S. Geol. Survey Prof. Paper 412, 20 p. [1962].
Livingstone, D. A., 1963, Chemical composition of rivers and lakes, in Fleischer, Michael, tech, ed., Data of geochemistry, 6th edition: U.S. Geol. Survey Prof. Paper 440-G, 64 p.
Morey, G. W., Fournier, R. O., and Rowe, J. J., 1964, The solubility of amorphous silica at 25°C: Jour. Geophys. Research, v. 69, no. 10, p. 1995-2002.
Oborin, A. A., and Zalkind, I. E., 1964, Geochemistry of boron in supergene processes: Geochem. Soc., Geochemistry Internat. No. 1, 1964, p. 63-72.
Oborn, E. T., 1960, A survey of pertinent biochemical literature: U.S. Geol. Survey Water-Supply Paper 1459-F, p. 111-190.
Piper, A. M., Robinson, T. W., and Park, C. F., Jr., 1939, Geology and ground-water resources of the Harney Basin, Oregon, with a statement on Precipitation and tree growth, by L. T.
Jessup: U.S. Geol. Survey Water-Supply Paper 841, 189 p. [1940].
Rainwater, F. H., and Thatcher, L. L., 1960, Methods for collection and analysis of water samples: U.S. Geol. Survey Water-Supply Paper 1454, 301 p.
Rankama, K. K, and Sahama, T. G., 1950, Geochemistry: Chicago, 111., Univ. Chicago Press, 912 p.
Russell, I. C., 1884, A geological reconnaissance in southern Oregon: U.S. Geol. Survey 4th Ann. Rept., p. 435-462.
Silvey, W. D., and Brennan, Robert, 1962, Concentration method for the spectrochemical determination of seventeen minor elements in natural waters: Anal. Chemistry, v. 34, p. 784-786.
Snyder, C. T., Hardman, George, and Zdenek, F. F., 1964, Pleistocene lakes in the Great Basin: U.S. Geol. Survey Misc. Geol. Inv. Map 1-416.
Stott, W. J., 1952, Investigation of saline deposits in southern Oregon: Bonneville Power Adm. and Portland Univ., 60 p.
Sverdrup, H. U., Johnson, M. W., and Fleming, R. H., 1942, The oceans, their physics, chemistry, and general biology: New York, N.Y., Prentice-Hall, Inc., 1087 p.
Trauger, F. D., 1950, Basic ground-water data in Lake County, Oregon: U.S. Geol. Survey open-file report, 287 p.
Turekian, K. K., and Wedepohl, K. II., 1961, Distribution of the elements in some major units of the earth’s crust: Geol. Soc. America Bull., v. 72, no. 2, p. 175-192.
U.S. Geological Survey, 1916, 1917, 1918, 1921, Surface water supply of the United States, 1913, 1914, 1915, 1917, pt. 10, The Great Basin: U.S. Geol. Survey Water-Supply Papers 360, 390, 410, 460, respectively.
U.S. Supreme Court, 1932, The United States of America v. The State of Oregon: Joint Abs. Record, October term, 1932, v. 1 and 2, 1166 p.
Van Winkle, Walton, 1914, Quality of the surface waters of Oregon: U.S. Geol. Survey Water-Supply Paper 363, 137 p.
Walker, G. W., 1963. Reconnaissance geologic map of the eastern half of the Klamath Falls (AMS) quadrangle, Lake and Klamath Counties, Oregon: U.S. Geol. Survey Mineral Inv. Field Studies Map MF-260.
Waring, G. A., 1908, Geology and water resources of a portion of south-central Oregon: U.S. Geol. Survey Water-Supply Paper 220, 86 p.
Welch, P. S., 1952, Limnology, 2d edition: New York, N.Y., McGraw-Hill Book Co., 538 p.
Whistler, J. T., and Lewis, J. H., 1916, Warner Valley and White River projects, irrigation and drainage: U.S. Reclamation Service, 123 p.
White, D. E., 1957, Magmatic, connate, and metamorphic waters: Geol. Soc. America Bull., v. 68, no. 12, p. 1659-1682.
White, D. E., and Waring, G. A., 1963, Volcanic emanations, in Fleischer, Michael, tech, ed., Data of geochemistry, 6th edition: U.S. Geol. Survey Prof. Paper 440-K, 29 p.
Whitehead, H. C., and Feth, J. H., 1961, Recent chemical analyses of waters from several closed-basin lakes and their tributaries in the Western United States: Geol. Soc. America Bull., v. 72, no. 9, p. 1421-1426.TABLES 20-31B60
CLOSED-BASIN INVESTIGATIONS
Table 20.—Altitude of water surface in Lake Abert, 1843-1963 [All altitudes through June 1951 determined on the basis listed; those since June 1951 determined at gage on east shore of lake]
Year Date Altitude 1 2 (ft) Basis
1843 . Dec. 20 24,256 Sketch by Capt. J. C. Fremont, discoverer.
1861-62 Presumed high Period of generally heavy runoff.
1865 . Aug. 21 Large body of water (diary of George Conn, with military party).
1881 Probably high Goose Lake filled to outlet.
1890 Probably high Generally very wet period following dry years.
1904 Probably high Wet year; heavy runoff.
1905 . June 6 Ivan Landes set copper bolt in bedrock on southwest shore at an altitude of 4,261.23 ft; lake must have been several feet lower.
1906 October 4,251 Areal extent of lake, as shown by notes of Eastern Oregon Railroad survey, east shore.
1906 Autumn 4,251 Approximate estimate from pho-
tograph by Waring (1908, pi. X). Water about 5 ft below level of July 3, 1963.
1915 Oct. 4— 4,254.2 Leveling from Landes’ copper bolt.
1917 __ Apr. 29 4,254.2 Do.
May 15 4,254.6 Do.
July 8 4,254.5 Do.
Sept. 25 4,253.5 Do.
1918 Apr. 28 4,253.5 Do.
June 15 4,252.9 Do.
1919 May 18 4,251.1 Leveling from Landes’ copper bolt; may be too low because of southwest wind.
1920 ... Aug. 7 4,249.1 Levels to water in river at south end of lake, allowing 0.2 ft drop to lake.
1921 ... Mar. 26 4,250.2 Leveling from Landes’ copper bolt.
Aug. 24 <4,249.5 Leveling from Landes’ copper bolt; unable to reach lake because of mud.
Year Date Altitude 1 (ft) Basis
1924 Aug. 14 «4,244 Dry, (except for small pools fed by springs and Poison Creek) for first time in memory of old inhabitants (Lake County Examiner, news report).
1926, 1930, 1931, 1933, 1934, 1937. Lake practically dry, in late summer or autumn of each year, according to local residents; confirmed by study in 1964 of water budget for previous years.
1931 October «4,244 Observations by Antevs (1938, p. 19).
1938 A wet year; lake probably rose moderately and has since contained water.
1939 July 15__ .. 4,248.2(±0.2) Photograph, by S. T. Harding, of boulder later used to hold
1941 June 4,247.5(±1.0) gage. Lake outline on map of Fremont National Forest, 1954, and comparison with soundings of 1959.
1947 Aug. 2_ 4,247.5(dbl.O) Aerial photograph by U.S. Forest Service of extreme south end of lake.
1950 July 21 4,248.02 Leveling from bench mark, east shore.
Nov. 11 4,246.60 Do.
1951 Mar. 31 4,249.47 Do.
June 5 4,250.52 Do.
June 23 4,250.13 Leveling from bench mark, east shore. Gage set on boulders on east shore, 10 miles northeast of Valley Falls, used for all later water levels; zero at 4,247.78 ft.
Year Date Altitude 1 (ft) Year Date Altitude 1 (ft) Year Date Altitude 1 (ft) Year Date Altitude 1 (ft)
1951.- Aug. 10 ._ 4,249.10 1956.. Mar. 8. .. _. 4,253.88 1961 — June 21 __ 4,253.41 1962.. Sept. 10 .. 4,250.78
1952.. Jan.12 _. 4,248.80 Apr. 4 __ 4,254.73 June 23— _ .. 4,253.37 Sept. 20 .. 4,250.65
May 3 .. 4,251.15 Apr. 22 .. 4,255.30 June 27 .. 4,253.31 Sept. 30 ._ 4,250.52
June 4 4,252.34 May 31 .. 4,257.28 Aug. 18 _. .. 4,252.72 Oct. 10 ._ -. 4,250.60
June 25 __ 4,252.78 June 2 _. 4,257.36 Sept. 7 _. 4,252.04 Oct. 20 4,251.11
Oct. 16. 4,251.86 June 20. .. 4,258.6 Sept. 26 _. 4,251.9 Oct. 31 ._ 4,251.19
1953.. Feb. 26 4,252.74 July 2 .. 4,258.2 Sept. 29 ... 4 4,251.81 Nov. 10 .. 4,251.19
Mar. 9 4,252.7 July 17 _. 4,258.02 Oct. 10. .. 4,251.69 Nov. 20 4,251.26
Apr. 19. 4,252.95 July 28 .. 4,257.84 Oct. 20. . .. __ 4,251.69 Nov. 30 _. 4,251.30
4,252.97 . 4,257.68 Oct. 31 __ 4,251.67 4,251.56
May 8 _ 4,254.03 Aug. 13 ._ 4,257.66 Nov. 10 ... 4,251.65 Dec. 20 .. 4,251.76
May 25 __ 4,253.67 Oct. 15— .. 4,257.28 Nov. 20 ... 4,251.62 Dec. 31 .. 4,251.87
May 30 __ 4,253.84 Nov. 30 ... 4,257.18 Nov. 30 ... 4,251.69 1963- Jan.10 __ ._ __ 4,252.02
July 14 . __ 4,254.58 1957.. Mar. 2 ... 4,257.58 Dec. 10 _ ... 4,251.77 Jan.20 ._ 4,251.95
July 20 __ 4,254.49 Mar. 3 _. 4,258.28 Dec. 20 ... 4,251.76 Jan. 31._ — __ 4,252.02
July 31 __ 4,254.38 Apr. 22 ... 4,258.53 Dec. 31 ... 4,251.86 Feb. 10 ._ 4,252.56
Sept. 27 4,253.56 May 9 _ .. 4,258.66 1962.. Jan.10 ... 4,251.87 Feb. 20 ._ 4,252.80
Oct. 8 — .. 4,253.53 June 15 ... 4,258.98 Jan. 20.. . . ... 4,251.92 Feb. 28 __ 4,252.93
1954.. Feb. 27 4,253.98 July 26 ._ 4,258.47 Jan. 31 ... 4,251.93 Mar. 10 __ 4,253.12
Mar. 14 __ __ 4,254.26 Aug. 22 ... 4,258.00 Feb. 10 _. 4,252.07 Mar. 20 4,253.09
Apr. 17 4,255.08 Oct. 22 ... 4,256.78 Feb. 20 ... 4,252.23 Mar. 31. ._ 4,253.13
Apr. 28 __ 4,255.13 1958__ May 5 ._ 4,259.12 Feb. 28 ... 4,252.08 Apr. 10 ._ 4,253.28
May 23 4,255.64 May 30. _ _ . ... 4,259.88 Mar. 10 ... 4,252.24 Apr. 20. __ 4,253.63
June 16. _ __ __ 4,255.63 June 22 ... 34,260.48 Mar. 20 ... 4,252.22 Apr. 30 ._ 4,253.90
4,255.43 July 9 ... 4,260.43 Mar. 31 4,252.23 __ 4,254.19
July 11 __ 4,255.43 July 25 ... 4,260.23 Apr. 10 ... 4,252.27 May 20 .. 4,254.55
Sept. 15 __ 4,254.38 Nov. 7 ... 4,259.13 Apr. 20. ... 4,252.31 May 31. ._ 4,254.83
Sept. 22 __ 4,254.48 1959.. Jan. 30. _ ... 4,259.2 Apr. 30 ... 4,252.21 June 7 .. 4,254.89
Nov. 22 __ 4,254.02 Apr. 8 . ... 4,259.12 May 10_ _ ___ 4,252.16 July 3 ... 4,254.88
1955.. Jan.17 4,253.97 May 7 ... 4,258.92 May 20 ... 4,252.17 Aug. 6. _ _ _. .. 4,254.47
Mar. 8 ... 4,254.03 May 19_ ... 4,258.98 May 31 _ 4,252.45 Aug. 14 ... 4,254.36
May 1__ __ 4,254.08 May 20 ... 4,258.82 June 10. ... 4,252.40 Oct. 21 -.- 4,253.58
May 10 ... 4,253.93 Oct. 9 ... 4,257.0 June 20 — ... 4,252.23 Nov. 12 __ 4,253.58
July 9. ... 4,253.26 Nov. 11 ... 4,256.6 June 30 ... 4,252.03 Dec. 6 ... 4,253.72
Aug. 5 ... 4,252.78 Nov. 13 ... 4,256.65 July 10 ... 4,251.83 Dec. 26.. _. __ 4,253.63
Sept. 16 ... 4,253.08 I960.. June 3__ _ _ ... 4,256.25 July 20 ... 4,251.65
Oct. 6__ ... 4,252.78 July 28 ... 4,255.3 July 31 ... 4,251.45
1956.. Jan.22 ... 4,253.28 Aug. 10. ... 4,254.8 Aug. 10 _ ... 4,251.26
Feb. 12 ... 4,253.84 1961.. Feb. 15 ... 4,254.32 Aug. 20 ... 4,251.15
Feb. 16 ... 4,253.84 Apr. 25. ... 4,253.93 Aug. 31 ... 4,250.94
1 Altitudes are based on the datum of 1929, later adjustments of the level net the beaches that were exposed in 1915.
being disregarded. Many of the lake-level readings were made by personnel of U.S. 3 Highest stage ever observed.
Soil Conservation Service and Oregon State Engineer. 4 Continuous records available from Sept. 29, 1961, to June 7, 1963.
2 Approximate minimum level in 1843 inferred from sketch showing lake coveringHYDROLOGY AND GEOCHEMISTRY, CLOSED-BASIN LAKES, OREGON
B61
Table 21.—Altitude of water surface in Summer Lake, 1905-63
[Observations made by U.S. Geological Survey personnel, except as indicated. Altitudes based on supplementary adjustment of level network of 1947. Levels during period January 1905-May 1959 based on copper bolt in rock at southwest corner of lake, 14 miles south of present Summer Lake Post Offiaa (saa. 25. T. 32 S., R. 16 E.). Permanent reference marks set September 2, 1959, by leveling
at Hunters Point, on west shore of lake 10 miles south of Summer Lake Post Office; later records of lake altitude obtained at this site. Altitude entries followed by the letter “a” (4,145.2a) were approximated; those followed by the letter “e” (4,144.5ft) were estimated]
Date Altitude (ft) Remarks
1906
Jan. 9_______________ 4,151.0 Gage and reference mark set.
Feb. 25, 26__________ 4,151.4 Highest level ever observed.
Mar. 21, Apr. 5________ 4,151.4 Do.
1912
Nov. 10.............. 4,151.3
1916
Oct. 3_________________ 4,150.12 Beach line and top of alkali, alt 4,151.2 ft.
1918
Apr. 29.............. 4,150.48
1920
Mar. 26................ 4,149.64
1921
Mar. 9................. 4,149.42
1929-34
Summers______________ __________ Lake low but not dry (Antevs, 1938, p. 19).
1931
Summer_______________ __________ Lake dry; note on Fremont National Forest
Map, 1932.
October______________ __________ Lake at lowest known level; 2 mi wide and
reached a point 2 or 3 mi short of normal south end (Antevs, 1938, p. 19).
1937
Sept. 22_____________ __________ Lake dry except for small sump.
1941
May 1-June 10________ __________ Field party of Cadastral Engineer, Bureau of
Land Management, sounded depths over lake bottom. Greatest depth of water was less than 2.5 ft; water level probably was between altitudes of 4,146 and 4,147 ft.
1944
Sept. 1............... 4,146.2 Allison (1945, p. 791).
1946
Aug. 8_______________ 4,145.2a Deduced from aerial photograph, U.S. Forest
Service, and from comparison with level notes for Sept. 30, 1961.
1948
(No date).___________ __________ Nearby farmers in 1950 reported lake was
dry about 1948.
1949
September-...................... “Practically dry” (Harbeck and others,
1951, p. 29).
1960
July 23______________ 4,145.5a Level party could not reach water because of
mud.
Summer minimum_______ 4,144.5e Area about 10 acres (A. B. Claggett, Oregon
State Game Comm., oral commun., 1961).
Date Altitude (ft) Remarks
1961
June 6--------------- 4,147.20 Recent high-water mark at an altitude of
4,148.1 ft.
1962
April---------------- 4,149a Highest in many years (A. B. Claggett, oral
commun., 1961).
June 30.............. 4,147.43
1963
Apr. 24.............. 4,147.57
Sept. 9______________ 4,145.8a Deduced from aerial photograph.
Oct. 8_______________ 4,144.91
1964
Apr. 28.............. 4,147.93
Sept. 23_____________ 4,146.03
1966
Mar. 10_............. 4,146.7
1966
Aug. 10.............. 4,147.71
1967
Apr. 23.__........... 4,148.73
Oct. 22______________ 4,147.25
1969
Jan. 30______________ 4,148.2 By hand level.
May 4................ 4,148.17
21_____________ 4,148.06 Depth 3.0 ft, mi east of Hunters Point.
Sept. 2.............. 4,146.05
Nov. 11.............. 4,146.08
1960
July 27.............. 4,145.62
1961
Apr. 25______________ 4,146.33 White crust at an altitude of 4,148.97 ft.
June 21-.............- 4,145.86
Aug. 14-------------- --------- Nearly dry; lake covered about 10 acres
(A. B. Claggett, oral commun., 1962).
Sept. 30_____________ 4,144.86 West edge of lake about % mi east of Hunters
Point.
1962
Jan. ll-__........... 4,146.38
Apr. 16______________ 4,147.00
May 28--------------- 4,146.39 High watermark for 1962 at an altitude of
4,147.0 ft.
July 26.............. 4,145.56
Aug. 18______________ 4,145a
Sept. 18_____________ _________ Nearly dry; south edge of lake about 2 mi
northeast of Hunters Point.
Oct. 30______________ 4,146.57
1963
July 2_______________ 4,147.97
Aug. 20...........4,147.23B62
CLOSED-BASIN INVESTIGATIONS
Table 22.—Altitude of water surface in Goose Lake, 1832-1963
(Data for 1832-86 from Harding (1965), and page numbers listed in parentheses gages or on leveling from bench mark H-92 at New Pine Creek (alt 4,720.499
for those years refer to his report. Data for 1936-41 from S. T. Harding (written ft, datum of 1929). Gage used in period 1947-49 was unstable; hence, results
commun., 1962). Records for 1945-63 based chiefly on occasional readings of may be in error a small fraction of a foot]
1832, October.—4,701-4,706 ft (p. 24).
1846— About 4,695 ft (p. 26).
1849, late summer.—4,694-4,695 ft (p. 26).
1854-—About 4,698 ft (p. 26).
1868.—Slight overflow (4,716 ft; p. 32).
1872— About 4,711 ft (p. 27).
1874-—About 4,713 ft (p. 27).
1877.—About 4,708 ft (p. 27).
1879, 1880.—About 4,711 ft (p. 27).
1881. —Brief overflow during windstorm (p. 27).
1882. —About 4,714 ft (p. 27).
1886— About 4,713.5 ft (p. 27).
About 1895.—About 4,715 ft (C. T. Watkins, Alturas, oral commun., 1956).
1904, July.—About 4,715 ft (from survey made by U.S. Bur. Reclamation).
1914, August.—About 4,713 ft.
1914-34-—Precipitation was below average in every water year except 1925 and 1927. Lake was dry at times in each calendar year, 1926 and 1929-34 (Antevs, 1938, p. 18-19).
1936, October.—About 4,694 ft, based on news account stating that lake then contained more water than in any October during previous 5 years.
1937. —Lake reported practically dry late in year.
1938-41.—Lake level was about 4,700 ft in 1938, 4,698 ft in 1939, and 4,699 ft in 1941.
Date Altitude Date Altitude Date Altitude
1945 1949—Con. 1958
Oct. 30 - 4,699.71 June 6— 4,698.1 Sept. 17 . 4,706.1
Nov. 30 . 4,699.32 July 7— 4,697.5 Oct. 16 _ 4,705.8
1946 Aug. 9 __ 4,697.4 Nov. 11 _ 4,705.8
Mar. 22 _ 4,700.74 Nov. 14 __ 4,697.4 Dec. 9 . 4,706.0
Apr. 3 . 4,700.86 Dec. 6_ 4,697.4 1959
May 9. __ _ 4,700.90 17 4,697.5 Jan. 28 . 4,706.3±
June 4__ . 4,700.98 1960 Mar. 17 . 4,706.1 ±
July « _ 4,700.19 Mar. 4__ 4,697.3 Apr. 23 _ 4,705.8
16 . 4,700.17 Apr. 7— 4,697.5 May 6 _ 4,705.73
Aug. 9. __ - 4,699.87 May 8 _ __ 4,697.6 19 . 4,705.7
Sept. 5 _ 4,699.32 June 5— __ 4,697.7 June 24 . 4,705.2±
Oct. 8 — _ 4,698.93 July 6— 4,697.5 July 13 _ 4,704.8
Nov. 16 _ 4,698.89 Aug. 7— __ 4,697.4 21 _ 4,704.7
1947 Nov. 13 __ __ <4,697 Aug. 21 _ 4,704.2
Mar. 15 4,699.3 1951 25 _ 4,704.0
Apr. 14. _ _ 4,699.4 Apr. 2— 4,697.8 Sept. 16_ __ _ 4,703.7
May 5 _ _ 4,699.1 June 21 — __ 4,698.3 Nov. 13 - 4,703.23
June 4- _ 4,698.8 1952 1960
July 11 _ 4,698.4 June 24 — __ 4,700.96 July 28 . 4,703.02
Aug. 9. _ _ 4,697.7 1953 1961
Sept. 9 _ 4,697.6 Apr. 21 — __ 4,701.47 Apr. 26 - 4,702.15
Oct. 11 _ 4,697.3 Oct. 7— __ 4,701.19 June 22 _ 4,701.51
Nov. 19 _ 4,697.4 8 .. 4,701.15 Sept. 7 _ 4,699.98
Dec. 11 _ 4,697.4 1954 28 _ 4,700.0
1948 Apr. 27— __ 4,702.87 1962
Apr. 19 — _ 4,698.1 Sept. 23— 4,700.85 Jan. 13 _ 4,699.81
May 10 _ _ 4,698.2 1966 Apr. 19 . 4,700.57
June 7 _ 4,698.4 Oct. 4 4,699.47 May 30-_ _ 4,700.50
July 10 _ 4,698.4 1966 June 11 _ 4,700.31
Aug. 5 - 4,697.7 June 6— .. 4,704.96 Aug. 17 _ 4,699.4 ±
Sept. 3 _ 4,697.4 July 9 -- 4,704.45 Sept. 3. __ _ 4,698.57
Oct. 6 _ 4,697.5 Aug. 11 -- 4,703.98 Nov. 14 _ 4,698.94
Nov. 4 . 4,697.4 1967 1963
1949 Apr. 22. 4,705.54 July 3 _ 4,700.94
May 6 _ _. . 4,698.0 Oct. 24— 4,703.94 HYDROLOGY AND GEOCHEMISTRY, CLOSED-BASIN LAKES, OREGON
B63
Table 23.—Altitude of water surface in Silver Lake, 1861-1963
[Datum used is that of 1929, with supplementary adjustment of 1947 (U.S. Coast and Geodetic Survey). Altitudes for 1905-17 were originally published in Water-Supply Papers 360, 410, and 460 (U.S. Geol. Survey, 1916, 1918, 1921), but were baspH on local datum 130.2 ft too high. Corrected values are shown in this table. Records for period 1905-21 were obtained in sec. 11, T. 29 S., R. 15 E., from staff gage or by leveling; those for period 1950-63 were obtained by leveling
from bench mark on north shore, or from staff gage at inlet, in sec. 21, T. 28 S., R. 15 E., 5}4 miles east of Silver Lake (town). Many of the readings were made by personnel of U.S. Soil Conservation Service and Oregon State Engineer. Additional readings for 1905. 1915, and 1956-60 are in file* of the TT.S. Geological Survey, Portland, Oreg.]
Date Altitude (ft) Date Altitude (ft) Date Altitude (ft) Date Altitude (ft) Date Altitude (ft) Date Altitude (ft)
1861-1908 0) 1915 1956—Con. 1968—Con. 1959—Con. 1960
190 A 4,303.90 Floodmark 4,308.9 Feb. 13_. 4,305.83 Jan. 29 4,307.20 Feb. 10 4,304.24
4,311.8 Oct. 3. 4,302.87 July 28 __ 4,307.88 27-- 4,305.93 Feb. 10 4,307.21 18 4,304.11
4,310.36 1916 Aug. 13 4,307.50 Mar. 14. _ 4,306.33 27 4,307.21 Mar. 16 4,304.11
1905 4,303.67 30 4,307.20 31._ 4,306.45 Mar. 11 . 4,307.14 Apr. 3 4,304.03
Jan. 17.. 4,310.54 1917 1 2 1967 Apr. 21._ 4,306.72 27 4,307.12 May 5 4,303.81
Feb. 27.. 4,310.8 4,300.78 Apr. 17 4,307.34 30.. 4,307.21 Apr. 16 4,306.85 31 4,303.59
Mar. 20. - 4,310.8 1921 2 29 4,307.34 May 16. 4,307.96 May 4 4,306.62 June 17 4,303.27
4,310.8 May 10_ 4,302.65 May 15 4,307.61 29.. 4,308.20 7 4,306.70 July 27 4,302.59
4,310.7 15.. 4,302.54 31 4,307.89 June 16. _ 4,308.95 20 4,306.51 1961
4,310.5 4,299.85 June 7 4,307.74 26.. 4,309.05 June 1 . -. 4,306.37 Apr. 25 4,300.85
1906 1951 July 24 4,306.84 July 11.. 4,308.95 11 4,306.19 June 20 4,300.32
A pr 4 308 8 4,299.21 4,306.46 28.. 4,308.71 29 4,305.86 Became
26_ _ 4\309.6 1953 30 4,306.11 Aug. 19.. 4,308.25 July 15 4,305.46 dry.
May 6 _ 4,309.9 Apr. 24 4,300.87 Sept. 11 4,305.91 Sept. 1 _ _ 4,307.94 29 4,305.26 1962 Dry.
1908 4,300.79 30 4,305.80 Sept. 17_. 4,307.74 Aug. 17 — 4,304.86 1963
4,304.8 1954 Oct. 10 4,305.80 30.. 4,307.63 30 4,304.80 July 4. 4,300.32
1912 Apr. 28. 4,303.58 28 4,305.73 Oct. 15.. 4,307.48 Sept. 3 4,304.56 September Dry.
Nov. 11 _ 4,307.09 Floodmark 4,304.8 Nov. 12 4,305.68 Nov. 7_ _ 4,307.31 29 4,304.42
1913 Sept. 23. 4,303.32 27 4,305.63 19. _ 4,307.31 Oct. 1 4,304.36
Feb. 5__ 4,306.9 1955 Dec. 11 4,305.59 Dec. 2__ 4,307.21 15 4,304.31
1914 Mar. 10. 4,302.79 31 4,305.68 17-. 4,307.21 30 4,304.22
Aug. 2__ 4,306.47 Oct. 7. 4,300.1 1958 1969 Nov. 11. 4,304.21
Oct. 9_. 4,305.35 1956 Jan. 13 4,305.76 Jan. 7._ 4,307.15 27 4,304.11
Nov. 2__ — 4,305.28 May 31 4,307.9 27 4,305.68 15__ 4,307.18 Dec. 29 4,304.05
1 For the period 1861-1903 the descriptive references by Antevs (1938, pages cited below) give fragmentary data on lake extent or altitude. In 1861 (p. 12) Paulina Marsh was practically a lake, extending almost to present site of Silver Lake (town). In 1879 (p. 14) Silver Lake was 4 feet deep. In 1882 (p. 14) water had risen 6 feet since 1879, had a depth of 10 feet, and was confluent with Thorn Lake. From 1885 to 1889 (p. 15) water subsided, and lakebed became dry in 1889, exposing sagebrush presumed to have grown in an earlier dry period in the 1840’s
(p. 12). Heavy runoff in 1890 filled Silver Lake (p. 15), so that it overflow'ed into
Thorn Lake and Christmas Lake Valley. In spring of 1904 (p. 16) Silver Lake filled, and overflow into Thorn Lake continued at least as late as September 17.
2 In September or October 1917 the bed of Silver Lake became dry and was nearly or completely dry in each summer or autumn from 1917 to 1950. In summer of 1951, Paulina Marsh overflowed and several hundred acres of Silver Lake bed was flooded to shallow depth. Paulina Marsh again overflowed into Silver Lake in 1952, and the lake contained water thereafter until August 11, 1961. It was dry in 1962.B64
CLOSED-BASIN INVESTIGATIONS
Table 24.—Hydrologic references to Malheur and Harney Lakes, 1826-99 [Data from U.S. Supreme Court (1932). All page numbers in parentheses refer to that report]
1826, November 1.—Peter Skene Ogden observed that “A small ridge * * * divides the fresh water [extension of Mud Lake] from the salt lakes [Harney Lake]. The lakes have no intercourse” (p. 1014-1059). Lack of overflow on that date does not necessarily mean that runoff in 1826 was below average; in fact, overflow to Harney Lake has never been observed as late as November 1.
1827, June 8.—Ogden again visited the area and noted that the water of Harney Lake was “very high. The waters of Sylvailles [Silvies] River and the lakes [Malheur and Mud] discharge into it” (p. 1014-1059).
1831, July.—John Work reported Harney Lake to be unusually high, possibly as compared to Ogden’s report of 1826 (p. 1014-1059).
1853.—Malheur Lake was high, based on observation of depth at fords (p. 631).
1860.—Lt. Joseph Dixon prepared map showing the Silvies River flowing into Harney Lake (p. 1014-1059).
1873, December.—Malheur Lake about 2J feet deep near midlake (p. 370).
1878. —Water near top of Sand Reef; none flowing to Harney Lake (p. 841).
1879. —Water 2 or 3 feet deep east of Sand Reef; none flowing to Harney Lake (p. 832).
1881, about May 1.—Water practically to top of reef, according to witness M. H. Brenton, who, with his foot, kicked out a small trench, allowing water to start flowing to Harney Lake. Others described the flow they saw as “a little bit of a stream” to “a stream 60 to 80 feet wide” (p. 836, 841).
Early descriptions of the Sand Reef as a barrier to the flow do not correspond to observations after 1916, when reef no longer was an effective barrier.
1882-83.—Water observed flowing from Mud Lake to Harney Lake (p. 836, 837); flow probably not continuous.
1887-89.—Dry years; Malheur Lake very low in 1889 (p. 400, 401, 561, 1085).
1894-—Malheur Lake very high (p. 843, 844, 926).
1895.—Malheur Lake very high; attempt made to open larger channel at Sand Reef to lower Malheur Lake level (p. 771-774, 850).
1899.—All lakes high; water flowing to Harney Lake at Sand Reef (p. 815-818, 846-848).HYDROLOGY AND GEOCHEMISTRY, CLOSED-BASIN LAKES, OREGON
B65
Table 25.—Altitude of water surface at staff gages in Malheur Lake, 1932-62
Altitude: Altitudes reported in feet above mean sea level. Lake dries at about 4,089 ft. Records are based on reports of the Bureau of Sport Fisheries and Wildlife (Fish and Wildlife Service, U.S. Dept, of the Interior) and (prior to 1922) of the Oregon State Land Board. Several additional readings in 1943, 1944, 1948, 1951, 1952, and 1953 are in files of the U.S. Geological Survey, Portland, Oreg. For records of the period 1903-31, refer to Piper, Robinson, and Park (1939, p. 129).
Gage at Narrows Bridge: Gage referred to U.S. Geological Survey bench mark
WA-314 (alt 4,095.87 ft). Doubtful connection between Malheur Lake and Narrows gage below about 4,092 ft; none below 4,091.5 ft. Previously published records below that level are not representative of level of Malheur Lake.
Gage near Voltage: Gage in channel of the Donner und Blitzen River 2% miles north of Voltage, datum at mean sea level (leveling by Bur. Sport Fisheries and Wildlife). Water level at gage is somewhat higher than lake level at times because of river flow, wind, and obstructions, such as tule growth. For records of the period 1921-31, refer to Piper, Robinson, and Park (1939, p. 130).
Date Altitude Date Altitude Date Altitude Date Altitude Date Altitude Date Altitude
(ft) (ft) (ft) (ft) (f« (ft)
Gage at Narrows Bridge
1943
Feb. 28 (>)
Apr. 9 4,094.12
23 .. 4,094.42
May 5 4,094.71
25 4,094.79
30 .. 4,094.43
June 14 .. 4,094.51
30 __ 4,094.29
July 20 .. 4,093.95
Aug. 7 __ 4,093.55
27 __ 4,092.97
Sept. 24 __ 4,092.75
Oct. 15 __ 4,092.71
Nov. 3 __ 4,092.83
1944
Mar. 3 4,093.11
Apr. 7 __ 4,093.00
29 4,092.79
May 22 __ 4,092.35
June 13 .. 4,092.26
27 __ 4,092.35
July 11 4,092.15
28 __ 4,091.57
1945
June 1 __ 4,093.20
17 .. 4,093.57
1945—Con.
July 2________ 4,093.51
23 ...... 4,093.21
Aug. 10_______ 4,092.85
28_____ 4,092.47
Sept.13________ 4,092.17
30_....... 4,091.93
1946
Jan. 28_______ 4,092.86
Mar. 11_______ 4,092.85
31________ 4,092.91
Apr. 20_______ 4,092.91
May 24________ 4,093.00
June 10_______ 4,092.83
25 ....... 4,092.73
July 11_______ 4,092.39
24 _______ 4,092.01
1947............. (2)
1948
Aug. 3________ 4,092.39
21________ 4,091.93
Sept. 3_______ 4,091.61
1949
Apr. 4........ 4,091.86
12_______ 4,092.23
27________ 4,091.93
May 12....... 4,091.81
1949—Con.
May 24. 4,092.08
June 4. 4,091.97
29_. 4,091.55
1950. — (*)
1951
May 14.. 4,091.71
30.. 4,092.43
June 6_. 4,092.51
23.. 4,092.13
July 9_. 4,091.53
1952
Apr. 18_. 4,092.62
23_. 4,094.17
May 2_. 4,094.58
12.. 4,095.14
26.. 4,095.25
June 4.. 4,095.11
17 4,094.88
30.. 4,094.86
July 18.. 4,094.52
31_. 4,094.28
Aug. 13_. 4,093.92
30.. 4,093.66
Sept. 15.. 4,093.52
29.. 4,093.26
Oct. 17_. 4,093.06
1962—Con.
Nov. 1_ 4,092.99
28. 4,092.96
1963
Jan. 23. 4,092.84
Feb. 13. 4,092.90
26. 4,092.82
Mar. 15. 4,092.76
Apr. 4. 4,092.58
28. 4,092.46
May 18. 4,092.52
28. 4,092.85
June 9. 4,093.34
26. 4,093.66
July 15.. 4,093.45
29. 4,093.18
Aug. 17. 4,092.76
30.. 4,092.46
Sept. 8.. 4,092.30
30. 4,091.90
Oct. 28.. 4,091.66
Dec. 3_. 4,091.89
1964
Feb. 9. 4,092.22
24.. 4,092.24
Mar. 25. 4,092.22
Apr. 18. 4,092.28
1964—Con.
May 11_ 4,091.87
June 25. 4,091.57
1955. (")
1966
May 9. 4,092.45
16. 4,093.49
29. 4,093.62
June 10. 4,093.67
27. 4,093.37
July 5. 4,093.43
25. 4,093.09
Aug. 4. 4,092.97
28.. 4,092.33
1967
Mar. 3. 4,093.07
15. 4,093.77
Apr. 18. 4,094.25
30. 4,094.45
May 16.. 4,094.57
31_. 4,094.90
June 2. 4,094.93
15_. 4,094.79
30.. 4,094.67
July 10.. 4,094.45
28.. 4,093.97
Aug. 14. 4,093.69
26.. 4,093.43
1957—Con.
Sept. 16. 4,092.90
30. 4,092.49
Nov. 20. 4,092.63
Dec. 24. 4,092.79
1958
Jan. 15. 4,092.85
Feb. 3_ 4,092.97
16. 4,093.55
Mar. 2_ 4,093.90
16. 4,094.20
Apr. 5. 4,094.20
25. 4,094.28
May 2. 4,094.48
16. 4,094.72
21. 4,094.72
30. 4,094.52
June 12. 4,094.48
27. 4,094.27
July 10. 4,093.98
29. 4,093.51
Aug. 5. 4,093.12
27. 4,092.42
1959
Mar. 3. 4,092.08
Apr. 28. 4,092.58
1960- -62. — 3 sodium
(calculated)
Specific
conductance pH Density
(micromhos at 20° C
at 25° C)
BASIN—Continued
33/18-26E) about 32 river miles upstream from mouth
1.3 ............... 0.3 ....................... 80 22 ............................ .........
.4 ........... .......... .3 ______________________ 51 18 __________ _____________ _____________
.5 ...................... .3 _________ ____________ 75 27 33 ............. ..........
1.0 ......................... ......................... ............ 75e 25 ............ 85 6.9
12 .3 ____________ _________ 81 31 30 104 7.4
.0 6.4 .......... .5 ______________________ 58 20 30 57 7.2
.0 ...................................... .......... ............ 66 23 29 69 7.4
.2 ___________ __________ _______________ __________ ____________ 65e ________ ______________ 65 _________
about 16 river miles upstream from mouth
0.8
70e
77
Falls (35/21-21P) about 1 river mile upstream from mouth
2.0 105e 44 127 402 285 ... 7.3
17 9 5 0.4 0.6 0.24 255 185e 108 47 7.9
2 8 140e ... 202 ...
125e 180 ...
1.0 .1 .15 94 42 30 124 7.7
1 n 90e _ _. 111 ...
138
5 5 105e 139 ...
137
185e ... 276 ...
7.0 .4 .8 .32 193 72 50 293 7.6
247 ...
3 2 .12 _ 140e ... 237 ...
1 0 .25 85e ._. 105 ...
75e ... 83 ...
northeast of Valley Falls (35/21-1C)
2.0 0.1 ........... ........................ ......... 73 24 34 84 6.8
1.8 _______________________ ____________ 0.19 _________ 70e ...................... 82 .........
BASIN
of Summer Lake Post Office (32/16-2NEJ4)
1,600 5.4 7.6 ......... 18 34
3,360 ..................... ............. 41 .........
2,720 .................... .............. 36 _________
630 .................................. 8.8 .........
840 .................................. 11 _________
1,350 5.4 __________ ___________ 18 ________
375 ...................... ............ 5.2 _________
7,200 15,000e 12,500e 2,900e 3,900e 8 6 98 10,600 20,800 ... 17,700 ... 4,860 6,320 ... 9.6 9.5 .. 1.004 1.011 1.008
6,270 l,800e 8 97 9,500 2,990 ... 9.6 _.
of Summer Lake Post Office (32/16-25L)
1,840
8,400e 10
11,900 9.7 1.005e
of Summer Lake Post Office (31/16-35NEJ4)
1,930
8,400e
97
1.005e
shore, but exact site unknown
3,040 ............ .......... .................................... 16,800e ............... 98 .......................... .......... 1.015
5,350 __________ ______________ _________ ___________ ____________ 27,000e .................. 98 .................................. 1.020e
Summer Lake Post Office (30/17-6)
0.2 154 30
0.2 .1 0.5 161 26 74 213 8.4
24 223 6.5
2.6 164 24 75 ... 7.7
22 .... 209 8.5
.3 158 22 76 218 8.8B72
CLOSED-BASIN INVESTIGATIONS
Table 28.—Chemical and physical character of
Discharge Water
Analysis Collection (cfs) temperature Silica Calcium Magnesium Sodium Potassium Bicarbonate Carbonate Sulfate
No. date or lake level (ft) C°F) (Si02) (Ca) (Mg) (Na) (K) (HCOa) (COs) (SO.)
GOOSE LAKE 21. Goose Lake, east shore at Oregon-
21a July 9, 1958 4,706.5e --- 160 400 0
21b Jan. 30, 1959 4,706.3 452 24
21c Apr. 26, 1961 4,702.15 58 18 5.1 358 31 628 79 62
21d Sept. 7, 1961 4,699.98 64 74 666
21e Jan.13,1962 4,699.81 33 66 592
21 f Apr. 19, 1962 4,700.56 66 50 411
21g June 12, 1962 4,700.31 66 58 13 4.0 450 36 746 94 83
21h Sept. 2, 1962 ... 4,698.57 70 80 970
22. Goose Lake, west shore near Oregon-California State line
22a May 12, 1912. 22b Aug. 13, 1955 22c Aug. 23, 1956. 22d Sept. 4, 1957. 22e July 9, 1958..
23a Sept. 16, 1953 23b May 5, 1954.. 23c Sept. 4, 1957.. 23d July 9, 1958..
4,714e 50 18 2.0 350 34 638 98 45
4,700e 15 5 450 550 195 80
4,703.8e ... 10 5 255 455 25
82 90 14 3.4 264 27 484 49 46
4,706.5e ... 190 440 50 ..
23. Goose Lake, south end near Davis Creek, California (approx.
4,701e 72 68 12 4.9 469 44 704 146 76
4,703e 63 64 13 4.7 292 32 566 30 51
71 86 14 3.4 264 27 494 47 46
4,706.5e ... 210 495 55 ..
24. East Branch Thomas Creek?, 1
24a June 26, 1958.
3e
80 46 24 4.4 141 8.4 166 0 138
25. Drews Creek, 11 miles
25a June 26, 1958........ 15e 73 29 10 5.1 7.3 2.8 76 0 1.0
26. Dry Creek, 14 miles
26a June 26, 1958....... 12e 75 36 16 2.8 8.5 4.7 86 0 2.9
27. Davis Creek, 1 mile northeast of
27a June 26, 1958....... 25 58 32 5.6 0.5 3.4 1.8 30 0 0.0
28. Willow Creek, 12 miles north of town
28a June 26, 1958.
5e
77 42 11 5.0 7.5 3.2 78 0 1.9
29. Lassen Creek, 11 miles north of town
29a June 26, 1958....... 15e 64 35 4.0 1.7 4.0 2.1 28 0 2.9
30. Pine Creek, 1 mile southeast of
30a June 26, 1958 52 56 25 5.2 1.0 4.5 1.2 28 0 3.8
SILVER LAKE
31. Silver Lake, northeast shore (28/16-31 or 32, except 28/15-22
31a Feb. 1912 ... 4,308± ... 57 42 25 M9 319 23 9.2
31b Jan. 29, 1959 ... 4,307.2 37 ._ 232 0 ...
31c May 18, 1959 4,306.50 48 40 29 268 0
31d Apr. 25, 1961 ... 4,300.85 55 22 59 44 137 28 700 28 .4
31e June 20, 1961 ... 4,300.32 ... 19 16 18 234 32 354 161 1.6
32. Silver Creek, 1.5 miles southwest of
32a Aug. 20, 1956........ 23 70 32 6.4 2.1 3.7 1.1 38 0 0.3
32b Apr. 25, 1961........ 11 39 .......... 5.5 1.6 3.2 ________ 35 0 .........
See footnotes at end of table.HYDROLOGY AND GEOCHEMISTRY, CLOSED-BASIN LAKES, OREGON
B73
lakes and their tributary streams—Continued
Chloride Fluoride
(Cl) (F)
Bromide Nitrate
(Br) (NO*)
Orthophos- Boron
phate (PO4) (B)
Dissolved Hardness as solids CaC03 (calculated)
Specific
Percent conductance
sodium (micromhos
at 25° C)
pH Density
at 20° C
BASIN
California State line (41/20-23P)
45 ..................-............-................
70 ___________ __________- -........- ----------
117 0.8 ....................... 5.3
224 ___________ ___________ -........... 9.2
600e ____________ ____________ 860 8.5
660e 55 ___________ 1,010 8.8
1,040 66 88 1,580 9.0
1,900e ........................ 2,820 -----------
8.6 1,600e
l,100e
152 335 .9 1 6.9 12 3.8 1,270 2,700e ___ 49 91
2,590
1,770
1,980
4,070
9.1
(41/19-20 or 21 in Oregon; 48N/13E-20 or 21 in California)
100
160
85
83
55
0.9
1.5 1.1
6.8
2.3
2.3
1,010 1,280e 790e 823 680e
53 89 .........................
58 90 1,990 ___________
46 ____________ 1,190 8.5
49 87 1,230 8.9
................ 1,000 8.8
sites: 23a, 45N/13E-14L; 23b, 45N/13E-3; 23c and 23d, 46N/13E-34)
162 .9 1.2
98 .4 1.9
84 .9 7.8
65 ........... .......................-
4.8 1,340 50 91 2,110 9.2
2.1 868 52 87 1,310 8.8
2.6 824 49 87 1,240 8.9
740e 1,100 8.9
mile west of Lakeview (39/20-9N?)
74 1.2
1.8
3.8 525 78 78 801 7.9
southwest of Lakeview (40/19-18P)
1.5 0.0
0.5
0.0 94 46 24 120 7.7
southwest of Lakeview (41/19-7B)
0.8 0.3
0.9
0.1 115 52 24 136 7.9
town of Davis Creek, Calif. (45N/14E-16J)
1.5 0.2
0.0 .......0.0 60 16 29 54 7.7
of Davis Creek, Calif. (47N/14E-22Q)
1.0 0.5
0.5
0.0 111 48 24 128 8.0
of Davis Creek, Calif. (47N/14E-27L)
1.0 0.4
0.1
0.1 65 17 30 53 7.5
New Pine Creek, Calif. (48N/15E-32L)
1.0 0.3
0.2
0.0 56 17 35 54 7.6
BASIN
for May 18, 1959, sample; exact site unknown for 1912 sample)
3.3 0.2 366 208 34 ..
1.0 _ 300e 141 361 7.9
4.0 _ 370e 162 427 7.5
21 1.0 .9 __ 686 330 45 1,090 8.5
40 1.4 1.2 698 114 77 1,140 9.7
Silver Lake Post Office (28/14-28SWM)
0.2 ........................ 1.0 ................................... 66 25 24 67 6.9
______________ 20 ......... 56 7.6B74
CLOSED-BASIN INVESTIGATIONS
Table 28.—Chemical and -physical character of
Discharge Water
Analysis Collection (cfs) temperature Silica Calcium Magnesium Sodium Potassium Bicarbonate Carbonate Sulfate
No. date or lake (°F) (SiO*) (Ca) (Mg) (Na) (K) (HCOa) (CO*) (SCh)
level (ft)
HARNEY
33. Malheur Lake, south
33a Mar. 8, 1912___........ 4,091e .......... 14 27 20 117 27 439 0 37
33b Nov. 29, 1961._........ 4,090e .......... 26 38 12 31 4.0 177 0 51
34. Harney Lake
34a Mar. 10, 1912_______ 4,086.5 .......... 31 ...................... 8,820 335 5,760 1,590 1,930
34b Jan. 4, 1962 6..... 4,081e .......... 73 ...................... 13,200 488 2,730 4,720 2,590
35. Donner und Blitzen River, 2.5 miles
35a Aug. 21, 1956.
43
24 8.3 3.3 9.9 1.4 63 0
36. Silvies River, 6 miles northwest
36a Aug. 22, 1956_.......... 9.2 67 23 15 5.5 11
3.6 100 0
WARNER
37. Crump Lake, west shore about
37a Sept. 7, 1912............................... 3.0? 13
37b Apr. 27, 1961________ 4,472.48 62 39 34
37c Sept. 8, 1961........ 4,470.80 48 .....................
37d Apr. 19, 1962......... 4,474.9 60 23 .........
5.6 3 18 .......... 95 0 8.6
14 54 6.3 246 6 26
____ 90 ....................... .........................
____ 16 ............................................................
38. Hart Lake, northwest shore about
38a Sept. 5, 1912________ 4,471.0 ....... 19 22 9.8 *49 ......... 211 0 16
38b Apr. 27, 1961________ 4,465.48 59 35 52 30 192 16 518 43 94
39. Bluejoint Lake, west shore about
39a Sept. 6, 1912.
20 21 23 1,370 82 1,820 511 206
40. Deep Creek, 5.5 miles
40a Mar. 31, 1951....... 235 __________________________________ __________ ____________________46 0 9.2
40b Aug. 21, 1956_______ 20 65 31 7.9 3.8 7.1 2.8 62 0 ---------
41. Honey Creek, 0.7 mile northwest
41a Aug. 21, 1956.
1.2 72 22 19 8.1 12 3.4 126 0
1 One of 17 samples collected at 15 sites during lake traverses on May 20, 1959. 1914, p. 122). Composites 12a and 12b, respectively, had largest and smallest
(See pi. 2.) dissolved-solids contents reported for the period Aug. 11, 1911-Aug. 14, 1912.
2 Analysis of a composite of daily samples for the 10-day period (Van Winkle, 3 Calculated amount of sodium plus potassium, expressed as sodium. ValuesHYDROLOGY AND GEOCHEMISTRY, CLOSED-BASIN LAKES, OREGON
B75
lakes and their tributary streams—Continued
Chloride Fluoride Bromide
(Cl) (F) (Br)
Nitrate Orthophos- Boron
(NOa) phate (POO (B)
Dissolved Hardness as Percent solids CaC03 sodium
(calculated)
Specific
conductance pH Density
(micromhos at 20° C
at 25° C)
BASIN
shore near wildlife refuge
22 .................... 2.4 1.6
4.5 0.3 __________ 3.3 .42
484 150 58 ........................
258 142 31 409 7.0
(exact site unknown)
6,800 .................-....... 2.8
11,700 13 47 ..........
10 .......... 22,400 28
32 83 34,300 4
98 .......................... 1.021
98 47,700 9.8 1.026
southeast of Frenchglen (32/32 J^-8SW^)
1.8
0.1
87e 34 37 111 7.9
of Burns (22/30-8SWM)
2.2 ...................... 0.3 ....................... 120e 60 27 170 7.8
BASIN
6 miles north of Adel (38/24-22)
4.2 ....................... 1.7 .......... 0.0
20 0.8 ...................... ........... 1.0 --------------------
30 ....................... -....................---------------
5.0 ............................................................
101 56 41 _________ _______________
322 144 44 488 8.4
500e ......................... 752 ...........
120e .......... .............. 162 ..........
3 miles northeast of Plush (36/24-14)
9 .......... ........... 0.8 0.8 0.0 230 96 53 .......................
62 2.1 ............................................. 781 252 60 1,190 8.8
25 miles north-northeast of Plush
504 ...................... 0.8 2.0 ......... 3,630 147 92
west of Adel (39/23-15NWM)
1.0 ....................... ............. ............ ......................... 28 ....................... 79 6.9
1.2 ....................... 0.2 ...................... 90e 35 28 105 7.1
of Plush (36/24-29NEM)
2.2
0.2
135e 81 23 204 7.9
for samples collected in 1912 from Silver, Crump, and Hart Lakes recalculated * 6 Calculated.
from Van Winkle’s data (1914, p. 113, 117). 6 Lake depth at point of collection was only 2 inches, of which 1 inch was ice.
4 Mean values calculated from analyses of 37 composite samples.B76
CLOSED-BASIN INVESTIGATIONS
Table 29.—Trace-element content of samples from Abert, Summer, and Goose Lakes and from the Chewaucan River
[Spectrographic analyses by U.S. Geol. Survey, Sacramento, Calif. Results reported in parts per billion. Symbols: >, more than amount indicated; <, less than amount indicated; £, about equal to but slightly more than amount indicated; about equal to but slightly less than amount indicated]
Analysis No.1 Collection date in 1962 Appearance when collected._ Time lapse until filtration 2_ Lake Abert Chewaucan River Summer Lake Goose Lake
If Jan.12 Turbid; light tan. 2.0 li 8a June 12 Sept. 17 Clear Clear 1.1 0.8 141 June 12 Slightly turbid; light amber. 1.2 16f June 12 Very turbid; light tan. 1.2 21g June 12 Turbid; tan. 1.4
Aluminum (Al) 270 76 77 8.6 168 11
Beryllium (Be). - <1.0 <1.9 <1.9 <.57 <2.0 <.57
<.48 <1.0 <•9 <.29 <1.0 <.29
Cadmium (Cd) - <2.4 <4.8 <4.7 <1.4 <5.0 <1.4
Chromium (Cr) <2.4 <4.8 <4.7 <1.4 <5.0 <1.4
Cobalt (Co) <2.4 <4.8 8.3 1.5 19 <1.4
Copper (Cu)_ _ <2.4 <4.8 <4.7 <1.4 <5.0 <1.4
Gallium (Ga) <9.6 <19 <19 <5.7 <20 <5.7
<.48 <1.0 <■9 <.29 <1.0 <.29
Iron (Fe) >96 90 84 64 15 5.1
<2.4 <4.8 23 <1.4 <5.0 <1.4
Manganese (Mn) £2.4 <4.8 <4.7 <1.4 <5.0 <1.4
Molybdenum (Mo) >48 13 31 .4 44 6.3
Nickel (Ni) £.48 17 17 2.8 1.9 1.7
Titanium (Ti) <1.0 <1.9 <1.9 <.57 <2.0 <.57
Vanadium (V) >48 77 £47 5.7 1.6 14
Zinc (Zn) <19 <19 <19 <5.7 <20 <5.7
1 Number of the companion major-constituent analysis in table 28. pore-size pads. Sample from Summer Lake remained opalescent even after filtration;
2 Time, in days, between sample collection and filtration through 0.45-micron others were clear.
B78
CLOSED-BASIN INVESTIGATIONS
Table 30.—Chemical and physical
[All analyses by U.S. Geological Survey. Results reported in parts per million, except as indicated. Estimated values throughout
Site Location * Water
No Collection Discharge 1 2 temperature Silica Calcium Magnesium Sodium Potassium
(on pi. 1) date (gpm) (° F) (SiO,) (Ca) (Mg) (Na) (K)
Lake Abert
102a 33/22-16F May 20, 1959, _ 66 46 159
102b 33/22-16F 200e 66 46 4.5 2.5 160 7.6
103 33/22-20A 200e 71 33
104 34/21-6 J 3e 68 51 8.0 6.9 270 16
105 35/21-1C- June 12, 1962 1.5e 56 34 23 9.3 23 3.5
106 35/21-21D 150e 66 65 9.5 7.5 290 14
Summer Lake
107 30/17-5A Oct. 17, 1955 500e 60 37 3.2 1.9 43 3.6
108 33/17-12A. Oct. 8, 1948 21 116 96 1.4 .4 399c 6.8
Goose Lake
109 39/20-27B 49 157 66 15 0.4 152c 2.2
110 39/20-27L. May 22, 1948 __ 50 185 140 8.5 1.4 268c 8.8
Warner
111 38/24-27N Sept. 16, 1948 5 104 125 18 2.0 175c 8.7
1 Spring locations are designated according to the rectangular system of land 2 449 gpm equals 1 cfs.
division. (See p. B53.)HYDROLOGY AND GEOCHEMISTRY, CLOSED-BASIN LAKES, OREGON
B79
character of sampled spring waters
table are indicated by “e” following the value. Calculated sodium and sulfate concentrations are indicated by “c” following the value]
Bicar- Ortho- Dissolved Specific
bonate3 Carbonate8 Sulfate Chloride Fluoride Nitrate phos- Boron solids Hardness Percent conductance r*H8
(HCOa) (COs) (SOO (Cl) (F) (NO,) phate (B) (calculated)3 as CaCO, sodium (micromhos
(PO4) at 25° C)
basin
198 8 29 - 285e 44 483 8.5
216 12 104 22 781 8.6
224 8 32 104 0.7 - 0.21 475 22 92 777 8.5
256 16 141 580e 17 ... 973 8.7
298 0 40 260 .6 .14 800 48 90 1,360 8.1
69 0 24 50 .0 .15 201 96 33 345 7.0
282 0 47 295 .3 .13 867 54 80 1,450 8.1
basin
109 3 6 __ 0.2 162 16 82 206 8.4
374 30 in 285 2.2 .1 1.0 1,120e 5 98e 1,760 8.5
basin
84 0 152 99 3.1 0.2 7.0 531 39 89 813 7.7
208 0 223 146 6.9 .3 9.9 905 27 89 1,320 7.3
basin
4130 116 150 1.9 1.5 7.3 662e 53 86e 935 8.7
3 See headnote in table 28.
4 Carbonate plus bicarbonate, expressed as bicarbonate.B80
CLOSED-BASIN INVESTIGATIONS
Table 31.—Chemical and physical
[All analyses made by U.S. Geological Survey. Well data largely from Trauger (1950, table 1). Results reported in parts
Site No. Location 1 Depth Water-bearing material Collection Water temperature Silica Calcium Magnesium Sodium
(on pl.l) (ft) date (°F) (SiO 2) (Ca) (Mg) (Na)
Lake Abert
112 33/18-24L 114 Gravel Oct. 7, 1948-.-. 54 14 3.2 21c
113 33/18-24 (?)P 165 Apr. 4, 1961. __ 44 5.0 2.9 69
114 36/21-6B 80 Sand, black Jan. 12, 1962.. 58 22 13 31
Summer Lake
115 30/16-lF-.......... 339 Sand, fine, red..............
116 30/16-lR___________ 209 Lava above red and black cinders.
117 30/16-12A__.... 248 Lava_...........................
118 30/16-12J__________ 375 Lava, red, pumiceous_________
119 30/17-7F___________ 414 Lava, red, spongy, with pumice. _
120 33/18-15B...... 134 .............................
Oct. 15, 1955. _ 69 35 12 1.4 49
Oct. 15, 1955. _ 67 40 5.6 1.4 58
Oct. 15, 1955.. 60 42 3.2 1.9 82
Oct. 15, 1955. _ 66 39 3.6 .7 98
Oct. 15, 1955._ 66 38 2.8 2.6 68
Oct. 10, 1948_ _ 92 106 16 4.7 194c
Goose Lake
121 38/19-24B.......... 343
122 38/19-36A.......... 205
123 38/20-33Q......... 400
124 39/19-15H........... 90
125 39/20-4C............ 38
126 39/20-16A._....... 365
127 39/20-29L.......... 210
128 40/19-5D........... 100
129 40/19-22M........ 3,000
130 41/20-13Q.......... 17
Gravel, sand. Gravel_______
Gravel, fine.
Gravel, medium to fine.
Sand (?)_............
Gravel_______________
____do_______________
July 2, 1948... 54 44 18 1.8 7.3c
July 2, 1948 53 46 10 4.2 13c
June 30, 1948._ 182 95 12 2.6 209c
Aug. 2, 1948. __ 54 62 20 4.7 7.7c
June 30, 1948._ 212 145 15 4.4 236c
July 10, 1948.. 79 77 1.1 .8 45c
July 13, 1948_ _ 64 4.7 4.6 40c
Aug. 2, 1948.__ _ 57 48 9.2 47c
Nov. 2, 1950... " 66 64 5.0 12 546
June 25, 1948. _ 65 52 8.2 3.1 183c
Silver Lake
131 27/14-35M 61 Sand
132 28/14-2 IQ 240 Cinders and clay, red.
133 28/14-22L 34 Gravel, sand
Feb. 18, 1950. _ Dec. 12, 1958. _ Feb. 18, 1950..
50 51 17
43 14
48 53 13
Warner
134 33/26-20L 600
135 36/24-17J 128
136 39/24-34L 156
Aug. 13, 1948............
Feb.13,1950.. .........
Aug. 9, 1948. 50
67 42 46 402c
45 13 8.7 21
65 5.9 5.5 46
1 Well locations are designated according to the rectangular system of land division.
2 See headnote in table 28.
3 Also includes 0.42 ppm of orthophosphate (PO4).HYDROLOGY AND GEOCHEMISTRY, CLOSED-BASIN LAKES, OREGON
B81
character of sampled well waters
per million, except as indicated. Calculated sodium and sulfate concentrations are indicated by “c” following the value]
Potassium (K) Bicarbonate 2 (HCOs) Carbonate2 (CO.) Sulfate (SO.) Chloride (Cl) Fluoride (F) Nitrate (NO*) Boron (B) Dissolved solids 2 (calculated) Hardness as CaCOa Percent sodium Specific conductance (micromhos at 25° C) pH 2
basin
8.8 96 0 15 7.0 0.3 1.3 0.03 172 48 44 202 8.0
6.4 215 0 .4 1.5 .7 .2 237 24 82 326 7.8
7.4 199 0 8.0 10 .2 .51 »249 110 36 354 7.2
basin
6.0 41 19 19c 44 1.2 208 36 71 310 9.2
4.8 114 3 20c 22 .2 .. 212 20 83 277 8.4
9.1 180 2 34c 24 .0 288 16 87 384 8.4
5.8 228 3 11c 19 .2 293 12 92 433 8.5
4.8 157 4 17c 12 .1 227 18 86 299 8.6
13 132 0 227 104 1.9 2.0 0.8 734 60 85 1,050 7.7
basin 4
4.4 80 0 4.3 1.5 0.0 0.1 0.1 120 52 22 140 8.2
5.4 90 0 3.7 1.8 .0 .2 .1 128 42 37 144 7.8
5.8 72 0 259 119 4.7 .1 6.7 749 40 90 1,100 8.1
5.1 5104 ... 4.5 1.5 .1 .2 .01 157 69 18 176 8.8
6.6 107 0 289 132 4.6 .2 7.6 893 56 89 1,240 7.9
2.0 121 0 1.4 3.2 .6 .3 .52 192 6 92 196 8.2
3.0 138 0 1.0 2.8 1.0 .9 .02 190 31 72 234 8.2
15 5200 ___ 12 16 .4 98 .01 401 158 37 553 8.7
36 1,520 0 1.1 39 .4 1.5 6.6 1,460 62 92 2,150 7.9
6.0 338 0 47 80 2.4 .5 4.4 553 33 91 874 7.7
basin
5.6 184 0 1.0 1.0 0.2 3.2 0.01 207 112 27 285 7.5
2.8 107 0 4.4 1 .5 .1 .5 137 66 24 177 8.2
4.8 94 0 9.0 8.0 .2 8.8 .00 166 73 27 210 7.4
basin
24 748 0 226 246 0.5 1.2 3.5 1,420 294 73 2,190 ...
7.0 122 0 8.5 5.0 .2 1.9 .01 170 68 37 225 8.1
5.6 156 0 .7 10 .8 .3 .04 217 37 69 266 ...
4 For additional published analyses of well waters in California, refer to California 4 5 Carbonate plus bicarbonate, expressed as bicarbonate.
Department of Water Resources (1965, p. 157; 1966, table 44).AINDEX
A Page
Acknowledgments______________________________ B3
Agriculture___________________________________ 5
Alkaline-earth content, surface waters___ 49
Aluminum_____________________________________ 52
Ana River, chemical character________________ 30
earthen dam, effect on Summer Lake. _ 25, 28
inflow to Summer Lake.................. 26
Ana Springs_____________________________ 24, 28
chemical character..................... 30
discharge_____*_____________________ 25
Anabaena, Lake Abert_________________________ 20
Analyses, chemical, accuracy of results__ 55
chemical, Ana River_____________________ 70
Bluejoint Lake______________________ 74
Crump Lake__________________________ 74
Chewaucan River_____________________ 70
Goose Lake__________________________ 72
ground water, Goose Lake basin.. 78
Silver Lake basin_______________ 80
Harney Lake_________________________ 74
Hart Lake___________________________ 74
Lake Abert_______________________68, 76
Malheur Lake________________________ 74
methods used_______________________ 53
Silver Lake_________________________ 72
Summer Lake..................... 70, 76
surface waters, Goose Lake basin.. 72
Harney Basin_________________40, 74
tables described................... 56
trace elements______________________ 76
Anderson Lake________________________________ 38
Anions, bicarbonate abundance in streams . 44
net long-term change, Goose Lake____ 36
Malheur Lake______________________ 41
Summer Lake_________________________ 30
Artemia salina, Lake Abert___________________ 20
B
Basins, drainage, general features____________ 3
Beachlines above present-day shore, Blue-
joint Lake...............___ 38
Goose Lake______________________________ 33
Lake Abert.__________________________ 11,12
Silver Lake_____________________________ 37
Summer Lake__________________________12, 24
Bicarbonate, accuracy of analytical results.. 56
analysis method used____________________ 54
geochemical trends and processes____ 4.4
relation to carbonate and pH in lakes.. 45 Biologic controls on chemical quality of
surface waters__________46,45,50,51
Lake Abert____________________________ 20
Biologic nutrients, Goose Lake_______________ 36
Lake Abert______________________________ 18
Summer Lake_____________________________ 29
Bluejoint Lake.............................. 37
altitude, lake-surface________________ 65
beachlines above present-day shore__ 38
chemical analyses_______________________ 74
chemical character_____________________ 5
chemical relation to other lakes in basin. 41
lake-level fluctuations__________ 38, 65, 67
outflow_________________________________ 6
relation to surface runoff_______________ 5
solute depletion----------------------- 42
[Italic page numbers indicate major references]
Page
Bluejoint Lake—Continued
summary of data______________________ B5
water supply______________________________ 38
Boron, analysis method used__________________ 55
geochemical trends and processes_____ 52
Bridge Creek, supply to Silver Lake____________ 38
Brines, interstitial, playa deposits_________ 28
Bromide, analysis method used________________ 55
geochemical trends and processes_____ 50
relation to chloride in closed-basin lakes. 50
sources___________________________________ 51
Buck Creek, supply to Silver Lake______________ 38
C
Calcium, accuracy of analytical results___ 56
analysis method used____________________ 54
geochemical trends and processes_____ 49
Campbell Lake__________________________________ 38
Carbonate, accuracy of analytical results_ 56
analysis method used____________________ 54
chemical reaction in lakes___________44, 49
geochemical trends and processes_____ 44
relation to bicarbonate and pH in the
lakes____________________________ 45
Carbonate plus bicarbonate, relation to dis-solved-solids concentration in
lakes__________________________ 44
Cations in lake waters____________________ 7, 8
Chemical analyses. See Analyses, chemical. Chemical character of water, Bluejoint Lake. 5, 41
Chewaucan River_________________________ 21
Crump Lake______________________________ 5, 41
effect of man’s activities_______________ 8
evidence for leakage from Silver Lake.. 40
Goose Lake______________________________34, 35
inflow______________________________ 36
Harney Lake_____________________________ 5, 41
Hart Lake_______________________________ 5, 41
Lake Abert___________________________17,18
inflow________________________ 20, 23, 24
lakes________________________________ 5, 6
reaction with lakebed minerals___ 45
48, 49,50
Malheur Lake____________________________ 5, 41
precipitation on Lake Abert basin____23, 24
Silver Lake_____________________________40, 41
Summer Lake_____________________________27, 28
inflow______________________________ 30
Chewaucan Marsh, drainage, effect on lake
inflow__________________________ 8
effect on solutes in Chewaucan River. _ 21
evapotranspiration losses_______________ 13
Chewaucan River, chemical character_______ 21
drainage area___________________________ 13
relation between discharge and inflow to
Lake Abert_____________________13, 15
solute contributions to Lake Abert___21, 47
trace-element concentrations____________21, 52
Chloride, analysis method used_______________ 55
geochemical trends and processes_____ 46
relation to bromide in closed-basin
lakes__________________________ 51
relation to dissolved-solids content in
closed-basin lakes_____________ 47
sources_________________________________21, 47
Cladophora, Lake Abert_______________________ 20
Page
Clay minerals, chemical reaction with potassium_________________*___________________ B43
effect on pH of lake waters______________ 50
Climate, study area___________________________ 4
at and near Lake Abert, 1961-62__________ 14
Closed-basin lakes, reconnaissance___________ 37
Cobalt_______________________________________ 52
Cottonwood Creek, supply to Goose Lake.. 32
Cottonwood Reservoir, storage capacity____ 8, 33
Crump Lake___________________________________ 38
altitude, lake-surface___________________ 65
average annual inflow________________ 5, 39
chemical analyses________________________ 74
chemical character________________________ 5
chemical relation to other lakes in basin. 41
lake-level fluctuations___________ 38, 65, 67
leakage into peat deposits____________ 6, 40
relative chloride content________________ 47
solute depletion____________________5, 6, 42
summary of data___________________________ 5
water loss______________________________ 40
water supply_____________________________ 38
D
Deep Creek, average annual discharge_____ 39
supply to Crump Lake_____________________ 38
Density, saline lake waters__________________ 43
Desiccation, lake, effect on dissolved-solids
content_________________________ 42
Diatoms, Lake Abert__________________________ 20
Dissolved-solids concentration, geochemical
trends and processes____________ 42
principal constituents in the lakes__ 5
relation to carbonate plus bicarbonate. 44 relation to chloride content in closed-
basin lakes_____________________ 46
relation to lake levels___________________ 7
Goose Lake_______________________3.4, 49
Lake Abert___________________________ 17
Silver Lake__________________________ 40
Summer Lake__________________________ 27
relation to orthophosphate concentration____________________________________ 51
relation to potassium content____________ 43
relation to sodium content_______________ 43
seasonal changes__________________________ 7
sources, Goose Lake______________________ 36
Lake Abert___________________________ 20
Summer Lake__________________________ 30
summary of data___________________________ 5
Donner und Blitzen River, average annual
discharge_______________________ 39
supply to Malheur Lake___________________ 38
Drews Creek, supply to Goose Lake------------ 32
Drews Reservoir, storage capacity_____________ 8
E
Evaporation, Goose Lake------------------- 5, 33
lake-surface, summary of data for lakes. 5
mass-transfer method for estimating,
Lake Abert______________________ 13
seasonal changes in rates----------------- 7
Silver Lake_____________________________ 39
Summer Lake______________________________ 26
B83B84
INDEX
Page
Evaporite salts, eolian removal______________ B44
Evapotranspiration, dissipation of lake
inflow___________________________ 6
losses from Chewaucan Marsh______________ 13
F
Faulting, Lake Abert basin_____________________ 9
Silver Lake basin________________________ 39
Summer Lake basin________________________ 24
Fertilizers, effect on lake waters_____________ 8
Flagstaff Lake________________________________ 38
Fluctuations in lake levels____________________ 6
Bluejoint Lake______________________ 38, 65, 67
Crump Lake__________________________ 38, 65, 67
Goose Lake_______________________________31, 62
Harney Lake_________________________ 37, 64, 66
Hart Lake____________________________38, 67
Lake Abert___________________________ 9, 60
Malheur Lake________________________ 37, 64, 65
Silver Lake_____________________________ 37, 63
Summer Lake_____________________________ 24, 61
Fluoride, analysis method used________________ 55
geochemical trends and processes..:______ 50
G
Geochemistry, Goose Lake_____________________ 31*
Lake Abert_______________________________ 17
Silver Lake__________________________ 1*0
Summer Lake______________________________ 27
trends and processes, major constituents £2
minor constituents__________________ 1*7
trace elements_______________________ 52
Geologic setting, study area___________________ 3
Goose Lake, absence of leakage________________ 33
altitude, lake-surface___________________ 62
area and volume______________________ 31
beachlines above present-day shore___ 33
carbonate, variations relative to bicarbonate__________________________________ 35
chemical analyses of streams feeding_ 72
chemical character_______________________34, 35
chloride content, long-term net increase 36
dissolved solids, sources________________ 36
tonnage depletion___________________ 34
effect of irrigation on inflow___________ 32
evaporation._____________________________ 5, 33
geochemistry____________________________ Sit
inflow, average annual________________ 5, 33
lake-level fluctuations_______________31, 62
lake-surface precipitation_______________ 32
lakebed configuration____________________ 31
orthophosphate content, seasonal variations _________________________________ 36
overflow_________________________________ 6, 33
relative abundance of major anions___ 36
silica content, seasonal variations__ 36
solute depletion_________________________34, 42
springs and seeps________________________ 32
sulfate content, long-term net increase. 36
summary of data-------------------------- 5
trace elements___________________________34, 52
water loss--------------------------- 33
water supply.--------------------------- 32
Goose Lake basin, chemical analyses, ground
water___________________________ 78
springs and streams______________________ 32
Ground water, chemical analyses of spring
and well waters_________________ 78
contributions to the lakes________________ 5
discharge of Ana Springs________________ 24
springflow contribution, Goose Lake__ 32
Harney Lake_________________________ 38
Lake Abert______________________ 12,22
Summer Lake_________________________24, 30
Growing season, study area_____________________ 5
H
Harney Basin__________________________________ 37
chemical analyses, ground water_________40, 41
streams-----------------------------40, 74
Page
Harney Lake________________________________ B27
absence of leakage_______________________ 40
altitude, lake-surface___________ 37, 64, 66
average annual inflow________________ 5, 39
chemical analyses________________________ 74
lake-level fluctuations__________ 37, 64, 66
outflow___________________________________ 6
overflow_______________________________37, 40
relation to surface runoff__________________ 5
spring inflow____________________________ 38
summary of data___________________________ 5
water loss_________________________________ 40
water supply__________________________ 37, 38
Hart Lake__________________________________ 37
altitude, lake-surface____________________ 67
area and volume........................ 38
average annual inflow____________________ 5
chemical analyses....................... 74
chemical relation to other lakes in basin 41
lake-level fluctuations_________________38, 67
leakage into peat deposits----------- 6, 1*0
overflow________________________________ 5, 40
relation to surface runoff________________ 5
summary of data___________________________ 5
water loss_______________________________ 40
water supply_____________________________ 38
Honey Creek, average annual discharge_____ 39
supply to Hart Lake______________________ 38
Humidity, diurnal range______________________ 5
I
Inflow, annual, Bluejoint Lake_______________ 39
annual, Crump Lake________________________ 5
effect on dissolved solids in lakes.. 42
Goose Lake___________________________ 33
Harney Lake....................... 5,39
Hart Lake_____________________________ 5
Lake Abert......................... 13
straight-line equation__________ 15
Malheur Lake_______________________ 39
Silver Lake________________________38, 39
Summer Lake___________________________ 26
sources, lakes in Warner basin___________ 38
Iodide, analytical methods used______________ 55
Iron________________________________________ 52
Irrigation, depletion of streamflow, Chewaucan River____________________________________ 13
diversion, effect on lakes________________ 8
effect on inflow, closed-basin lakes_ 38
Goose Lake_________________________ 33
Lake Abert___________________________ 13
Summer Lake___________________________25, 26
pumping from Hart Lake___________________ 38
L
Lake Abert, absence of leakage_______________ 15
altitude, lake-surface_____________9,11, 60
Anabaena_________________________________ 20
area and volume___________________________ 9
Artemia salina___________________________ 20
average annual inflow____________________ 13
expressed as straight-line equation. 15
beachlines above present-day shore___11, 12
biologic controls on chemical quality_ 20
carbonate, variation relative to bicarbonate_______________________________ 18,45
chemical analyses.................... 68,76
chemical character___________________17,18
Cladophora.............................. 20
climatic data, 1961-62_______________ 11*
diatoms____________________________________ 20
dissolved solids_________________________17, 18
sources______________________________20, 23
eolian salt removal______________________ 23
evaporation______________________________13, 14
geochemistry_____________________________ 17
inflow from Chewaucan River________________ 9, 13
lake-level fluctuations.............— 9,60
Page
Lake Abert—Continued
lake-surface precipitation_____________B12, 23
lakebed configuration___________________ 9, 10
mass-transfer coefficient_______________ 14
orthophosphate content, seasonal variations____________________________________18, 51
outflow__________________________________ 6
overflow________________________________ 15
plankton________________________________ 20
relation between inflow and upstream
discharge of Chewaucan River. 13, 15
relative abundance of major anions--- 20
salts received from overland flow_______ 24
silica content, seasonal variations__18, 48
solutes, depletion during desiccation_ 18
inflow from Chewaucan River______21, 47
springs and seeps_______________________12, 22
storage in playa deposits___________ 23
summary of data__________________________ 5
trace elements________________17,19, 21, 52
tree-growth records__________________ 9,11
water budget_________________________ 15
water supply ________________________ 12
water loss______________________________ 13
Lake Abert basin, chemical analyses of
ground water____________________ 78
chemical analyses of stream water____ 70
chemical character of precipitation__ 23
faulting_________________________________ 9
Lake-bottom muds, bacterial sulfate reduction _______________________________________ 46
Lake levels, fluctuations, Bluejoint Lake_38, 67
fluctuations, Crump Lake________________38, 67
Goose Lake_______________________31,62
Harney Lake__________________ 37, 64, 66
Hart Lake_______________________ 38, 67
Lake Abert_______________________ 9,60
Malheur Lake_________________ 37, 64, 65
Mud Lake_____________________________ 37
Silver Lake_________________________37, 63
Summer Lake_________________________21*, 61
relation to dissolved-solids concentration _____________________________________ 7
Goose Lake___________________________ 40
Lake Abert___________________________ 17
Summer Lake__________________________ 27
Lake-surface altitude, records_______________ 53
Lakebed altitudes_____________________________ 5
Lakebed configuration, Goose Lake____________ 31
Lake Abert____________________________ 9,10
Lakes, general features_______________________ 3
Lead_________________________________________ 52
Leakage, Bluejoint Lake______________________ 40
Crump Lake___________________________ 6,1*0
Goose Lake______________________________ 33
Harney Lake_____________________________ 40
Lake Abert______________________________ 15
Malheur Lake_________________________ 6,1*0
Silver Lake___________________________ 39
Summer Lake___________________________ 26
Livestock_____________________________________ 5
Location of study area________________________ 3
Location system, rectangular system of land
division. ..................... 53
Lower Campbell Lake________________________ 38
M
Magnesium, analysis method used_____________ 54
depletion in lakes______________________ 50
geochemical trends and processes___________ 1*9
Malheur Lake______________________________ 37
altitude, lake-surface____________ 37,64,65
average annual inflow____________________ 39
chemical analyses_______________________ 74
chemical character_____________________ 1*1
lake-level fluctuations__________ 37, 64, 65
leakage into peat deposits___________ 6,1*0
overflow___________________________5,6,57,40
solute depletion____________________ 41
summary of data__________________________ 5INDEX
B85
Page
Malheur Lake—Continued
water loss_____________________________ B40
water supply____________________________ 38
Manganese____________________________________ 52
Marsh, Chewaucan, Lake Abert basin________ 9,13
floor of Warner basin___________________ 37
Malheur Lake basin______________________ 40
Paulina, Silver Lake basin______________ 38
Summer Lake basin_______________________ 26
Mass-transfer coefficient, Lake Abert________ 14
Mass-transfer equation, evaporation from
Lake Abert computed_____________ 13
Medford Weather Station, pan evaporation
data____________________________ 39
Molybdenum___________________________________ 52
Mud Lake.................................. 37
Muds, lake-bottom, bacterial sulfate reduction ________________________________________ 46
Mugwump Lake_________________________________ 38
N, O
Nickel....................................... 52
Nitrate, analysis method used________________ 55
concentrations in Goose and Silver
Lakes___________________________ 51
Organic matter, anaerobic decomposition___45, 46
Orthophosphate, accuracy of analytical results........................................ 56
analysis method used____________________ 55
fluctuations in relative concentration.. 51
Goose Lake__________________________ 36
Lake Abert__________________________ 18
Summer Lake_________________________ 28
geochemical trends and processes_____ 51
Outflow from lakes________________________ 5, 6
Overflow, altitude of lakes______•________ 5
Crump Lake, supply to Hart Lake______38, 40
Goose Lake______________________________ 33
Hart Lake_______________________________ 40
lake, depletion of dissolved solids__ 8, 42
Lake Abert............................ 15
Malheur Lake, supply to Harney Lake. 38
Pleistocene lake---------------------15, 37
Silver Lake----------------------------- 40
Summer Lake_____________________________ 26
to Bluejoint Lake_______________________ 38
P
Pan evaporation data, Medford Weather
Station_________________________ 39
Peat deposits, lateral seepage from lakes_ 6, UO
Pelican Lake_________________________________ 38
pH, accuracy of anlytical results____________ 56
affected by clay minerals--------------- 50
analysis method used____________________ 54
geochemical trends and processes_____ US
relation to carbonate versus bicarbonate_____________________________________ 45
relation to depletion of calcium________ 49
relation to photosynthesis______________ 50
relation to solubility of silica________ 49
Phosphate. See Orthophosphate.
Photosynthesis, effect on pH_________________ 50
Physiographic setting, study area_____________ 3
Plankton, Lake Abert_________________________ 20
Playa deposits, solute content____________23,1*2
Pleistocene lake, Harney Basin_______________ 37
highest known levels_____________________ 5
Lake Abert basin________________________ 12
present-day lakes as remnants___________ 6
Silver Lake basin_______________________ 37
Summer Lake basin______________________ 24
Warner basin___________________________ 38
Poison Creek, chemical character of water 24 source of dissolved solids to Abert Lake 23
Potassium, analysis method used______________ 54
depletion in lakes__________________41, 43
geochemical trends and processes____ US
Precipitation, Goose Lake____________________ 32
Lake Abert____________________________ 12
Page
Preci pi tation—C ontinued
range in average annual________________ B5
summary for lakes________________— 5
Summer Lake____________________________ 2U
source of dissolved solids for Lake
Abert________________________20, 23
Purpose of report____________________________ 3
R
Reconnaissance, closed-basin lakes__________ 37
References__________________________________ 56
Runoff, surface, entering the lakes___________ 5
S
Salts, accumulation in lakes______________ 8, U2
depletion during desiccation___________ 42
Goose Lake_________________________ 34
Lake Abert._______________________ 18
Summer Lake________________________ 28
overland flow into Lake Abert__________ 24
wind-transported, Lake Abert basin__ 23
Sand Reef, temporary dam____________________ 37
Scope of report_______________________________ 3
Seepage, lateral, into peat deposits______ 6, U0
Seeps, shoreline__________________________ 6, 12
Silica, accuracy of analytical results______ 56
analysis method used------------------- 54
downstream depletion in Chewaucan
River.......................... 21
fluctuations in relative concentration,
Goose Lake_____________________ 36
Lake Abert.......................18, 48
Summer Lake________________________28, 30
geochemical trends and processes_____ U7
relation to dissolved-solids concentration____________________________________48, 49
Silicate minerals, breakdown______________ 45, 48
Silver Creek________________________________38, 39
Silver Lake------------------------------- 37
altitude, lake-surface_________________37, 63
beachlines above present-day shore___ 37
chemical analyses------------------------ 72
chemical character----------------------- UO
depletion of inflow______________________ 9
dissolved-solids concentrations________40, 41
evaporation rate______________________ 5, 39
geochemistry_____________________________ UO
lake-level fluctuations________________37, 63
leakage________________________________39, 40
overflow_________________________________ 6
relative abundance of major ions_______40, 41
summary of data__________________________ 5
water budget_____________________________ UO
water loss_______________________________ 39
water supply_____________________________ 38
Silver Lake basin, chemical analyses of
ground water_____________________ 80
chemical analyses of stream water---- 72
faulting_________________________________ 39
water supply to lakes___________________ 38
Silvies River, average annual discharge___ 39
supply to Malheur Lake___________________ 38
Sodium, analysis method used__________________ 54
geochemical trends and processes_____ US
Solute losses, Bluejoint Lake_________________ 42
Crump Lake_______________________________ 42
Goose Lake_______________________ 34, 36, 42
Lake Abert_______________________________ 18
Malheur Lake_____________________________ 41
mechanisms causing_______________________ U%
Silver Lake_____________________________ 41
Summer Lake______________________________ 28
Solutes, contributed by inflow________________ 42
wind-transported, Bluejoint Lake_____ 42
Goose Lake___________________________ 34
Summer Lake.......................... 28
Specific conductance, method used for
measurements_____________________ 55
relation to dissolved-solids content-21, U3
Page
Springs, chemical analyses__________________ B28, 78
inflow to Summer Lake____________________24, 30
peripheral, Goose Lake____________________ 32
Harney Lake___________________________ 38
Lake Abert---------------------------12, 22
Summer Lake___________________________ 26
source of dissolved solids......... 46, 47, 52
supply to lakes___________________________ 6
Stone Corral Lake_____________________________ 38
Streamflow, chemical analyses__________________ 70
Chewaucan River, inflow to Lake Abert 15 depletion by irrigation, Goose Lake
basin___________________________ 33
Lake Abert basin______________________ 13
Summer Lake basin_____________________ 26
effects of man’s activities--------------- 9
Silver Creek, inflow to Silver Lake-- 38
Streams, abundance of bicarbonate-------------- UU
calcium content__________________________ 49
chloride content_________________________ 47
feeding Goose Lake_______________________ 32
feeding Lake Abert_____________________ 13
feeding smaller lakes in study area__ 38
feeding Summer Lake______________________ 26
fluoride content_______________________ 50
magnesium content_________________________ 49
nitrate content___________________________ 51
orthophosphate content____________________ 51
silica content____________________________ 47
sulfate content__________________________ 46
Sulfate, analysis method used_________________ 54
geochemical trends and processes_____ US
Summary of data for lakes---------------------- 5
Summer Lake___________________________________ 2U
average annual inflow____________________ 26
beachlines above present-day shore___12, 24
carbonate, variation relative to bicarbonate___________________________________ 30
chemical analyses________________________70, 76
chemical character_______________________27, 28
chloride, long-term net increase_________ 30
dissolved solids, sources________________ 30
variations in concentration__________ 27
evaporation rate_________________________ 26
geochemistry_____________________________ 27
inflow from Ana River________________9, 2U, 26
inflow from springs______________________ 5, 24
lake-level fluctuations__________________24, 61
lake-surface precipitation_______________ 24
orthophosphate content, seasonal variations____________________________________28, 51
outflow___________________________________ 6
overflow_________________________________ 26
relative abundance of major ions_____ 27
silica content, seasonal variations__28, 49
solutes, depletion during desiccation_ 28
storage in playa deposits............23, 28
springs and seeps________________________ 24
sulfate, long-term net increase__________ 30
summary of data___________________________ 5
trace elements___________________________27, 52
water budget_____________________________ 26
water supply_____________________________ 2U
Summer Lake basin, chemical analyses of
ground water____________________ 78
faulting_________________________________ 24
Surface waters, alkaline-earth content____ 49
anionic character________________________ 44
biologic controls on chemical quality— 46,
U8, 50, 51
Swamp Lake____________________________________ 38
T
Temperatures, air, seasonal variations... U
Lake Abert, relation solute content_ 18
The Narrows, channel from Malheur Lake.. 37
Thomas Creek, supply to Goose Lake_______ 32
Thorn Lake, overflow from Silver Lake____ 9, 37
Titanium____________________________________ 52B86
INDEX
Page
Trace elements, accuracy of analytical re-
sults________________________ B56
analysis method used------------------- 55
chemical analyses______________________ 76
Chewaucan River_____________________21, 52
geochemical trends and processes____ 52
Goose Lake__________________________34, 52
Lake Abert__________________________19, 52
Summer Lake_________________________27, 52
Tree-growth records, Lake Abert_________ 9, 11
relation to lake levels_________________ 6
Twentymile Creek, average annual discharge_____________________________________ 39
supply to Crump Lake___________________ 38
V, W
Vanadium_________________________________ 52
Page
Vegetation along beach line, Lake Abert_ B9, 12
Volcanism, ultimate source of chloride in
lakes_____________________________ 47
Warm Spring Creek, supply to Harney Lake 38
Warner basin____________—.--------------37, 38
chemical analyses of ground water-- 78
chemical contrasts among lakes------------- 41
Water budget, Goose Lake________________________ 33
Lake Abert_________________________________ 15
Silver Lake______'----------------- 40
Summer Lake________________________________ 26
Water-level fluctuations, Bluejoint Lake 38, 65, 67
Crump Lake______________________ 38, 65, 67.
Goose Lake_________________________________31, 62
Harney Lake_____________________ 37, 64, 66
Hart Lake__________________________________38, 67
Lake Abert_________________________________ 9, 60
Malheur Lake____________________ 37, 64, 65
Page
Water-level fluctuations—Continued
Silver Lake_______________________ B37, 63
Summer Lake_______________________24, 61
Water Loss, Goose Lake____________________ S3
Lake Abert____i_______________________ is
minor lakes__________________________ 1+0
Silver Lake_______________________ 39
Summer Lake___________________________ 26
Water supply, Goose Lake__________________ 32
Lake Abert____________________________ 12
minor lakes___________________________ 38
Summer Lake___________________________ 2k
Wind-transported salts as means of solute
depletion____________________ 42
Goose Lake____________________________ 34
Lake Abert___________________________ 23
Summer Lake___________________________ 28
☆ U.S. GOVERNMENT PRINTING OFFICE: 1971 0-411-034
UNITED STATES DEPARTMENT OF THE INTERIOR GEOLOGICAL SURVEY
R.20E. 120° 20'
R.21E.
R.22 E.
PROFESSIONAL PAPER 502-B PLATE 2
120°05'
T.31S.
42° 50' ->
T.32S.
o (\ \
/
36
(
V
\V-''N U \
y7^ \ \> 4\ \\ 4V dv
v\
■r
Ov X -4
*> /'
/
/ \
r \
A
>
/
A
"7*
t,
rjs
T.33S.
T.
Jki
40’
"**N \ Cogian
...y* J/< Buttes
—•' V .J
-iJE-
f C
4-7 A >V...^V
' _ y2U . ;
r i r
A--
...4
T.34S. -
=====4
iiNorrolws
35'
•'7>-
V /”
■■ a-
o.
•A
.r?s
/ \
\
\
_z_
T"
■?*"
-U=-
i x # \ i1 \ !.\ T /
■■ v\ V . Cn\ V \ i ”
sx 'W^EUCHRES butte/{/ V*U<, x i V ./ / )
V H. ^ .y \— A .y / \
1A / ^ ■ })— / .1
\
\
N
<
yj
A
V
■r
V
\ • \y '
n|Vv ^ v
\A} -A
I
1/ HZ
■^ -L H.
" \
\
-t
...-/A / **-1' ,
V7
/
I »>
CX
"•V.
> --/y
\&v JF vTT
JU6 MTN
z- \s\
\ ’
\ V
\ ^
\
\
\
\
) \
________ )
\
A,
■y-
. A'*
"A
•M
XX ■
,«v
w
4^-'* X,.
>
s,»
Tp?/ . // ///
\p-—
-IT
S BM
tvy 4
\
y
.z/
y \ SH ---------------N
A
'1 . - -.. x,
V""36
y i,_.—...31 “
"5SLa>->
v-
rr
/
EXPLANATION
1/
Irrigated and marshy areas adjacent to lake
Fine-grained playa surface
,106
a2 r
Surface water Spring
Sampling sites
Numbers correspond to sampling sites
listed in tables 28-31
• 114 Well
A
Sample site during lake survey of May 20,1959
Lake level U,258.9 feet above mean sea level. First figure to right of symbol is water temperature (degrees Fahrenheit); second figure is chloride content, in parts per million. All samples collected at mid-depth, ex- |\ cept as noted. Maximum lake depth at time of sample collection, about 15 feet
T.35 S.
42°30' -
Clover
. 1 /
^ I /3* ^ v
- r 'f-’ - y\
T.36S
\
» , • \ 1 yj
r h/ 7—-^"
? ' r r ./rssr-
l^AVV
-r- h f K /■
^ - .’> /
7' t "'/><
11\ K ' i i K
(i i v... \h
V j
A 1 / &
■ T
A \
\ \
1.■?
)
■W toAe >
'^=jJ
i
'-TCHMXfl
w#
T«:
y Twifry
' lakes/
/ y.
JUMPER
T.31S.
42° 50'
T.32S.
45'
T.33S.
\
\ ..yx — V.. ^
\
\
i 1— - ^—..
S"'^. Clover At
\56
LUCE\
RES'
A
T.34S.
35'
T.35 S.
42°30'
T.36S.
R.20 E.
120°20'
15' R.21E.
10’
R.22 E.
120°05'
Base compiled from maps by
U.S. Forest Service and
Oregon State Highway Department
*6
SCALE 1:62 500 1 2
4 MILES
1 .5 0
HZHZHZHZHLE
4 KILOMETERS
DATUM IS MEAN SEA LEVEL
APPROXIMATE MEAN DECLINATION, 1971
MAP OF LAKE ABERT AND VICINITY, SOUTH-CENTRAL OREGON, SHOWING SAMPLE-COLLECTION SITES FOR CHEMICAL ANALYSIS, AND REPRESENTATIVE WATER-QUALITY DATA FROM THE LAKE SURVEY IN 1959. erop-- £1
l£) fP - "S>
Solute Balance at Abert and Summer Lakes, South-Central Oregon
GEOLOGICAL SURVEY PROFESSIONAL PAPER 502-CSolute Balance at Abert and Summer Lakes, South-Central Oregon
By A. S. VAN DENBURGH
CLOSED - BASIN INVESTIGATIONS
GEOLOGICAL SURVEY PROFESSIONAL PAPER 502 -C
A description of the quantity and chemical character of incoming, outgoing, and stored solutes and the mechanisms of solute accumulation and depletion in a saline environment
UNITED STATES GOVERNMENT PRINTING OFFICE, WASHINGTON : 1975UNITED STATES DEPARTMENT OF THE INTERIOR
GEOLOGICAL SURVEY V. E. McKelvey, Director
Library of Congress Cataloging in Publication Data Van Denburgh, A. S.
Solute balance at Abert and Summer Lakes, south-central Oregon.
(Closed-basin investigations)
(Geological Survey Professional Paper 502-C)
Bibliography: p.
Supt. of Docs. No.: I 19.16:502-C
1. Saline waters—Oregon —Abert Lake. 2. Saline waters—Oregon—Summer Lake.
I. Title. II. Series. III. Series: United States Geological Survey Professional Paper 502-C. QE75.P9 No. 502-C [GB1625.07] 557.3'08s 1553'.72*09795931 75-619062
For sale by the Superintendent of Documents, U.S. Government Printing Office
Washington, D.C. 20402 Stock Number 024-001-02689-9CONTENTS
Pap-
Abstract ..................................................... Cl
Introduction................................................... 2
Location and general features of the area................... 2
Purpose and scope of investigation.......................... 2
Acknowledgments............................................. 2
Location system............................................. 4
Hydrologic characteristics of the two lakes and their basins. 4
Climate..................................................... 4
Inflow...................................................... 5
Lake Abert............................................. 5
Summer Lake............................................ 7
Lake extent and fluctuation................................. 8
Parameters that characterize the lakes...................... 9
Lake Abert............................................ 10
Summer Lake........................................... 10
Page
Solute balance............................................ CIO
Incoming and outgoing solutes.......................... 12
Lake Abert.......................................... 12
Summer Lake........................................ 18
Summary' of solute gain and loss................... 18
Stored solutes......................................... 19
Solutes dissolved in the lakes..................... 19
Lake-bottom and peripheral solutes................. 24
Lake Abert..................................... 24
Summer Lake.................................... 26
Summary of stored solutes.......................... 27
Contrasting solute balances at the two lakes........... 27
Literature cited........................................... 29
ILLUSTRATIONS
Figure
3-7. Graphs showing:
3. Wind movement and representative seasonal patterns of precipitation, streamflow, air temperature,
and evaporation in and adjacent to the Abert and Summer Lake basins.......................................... 5
4. Fluctuations of Lake Abert in water years 1916-65, compared with precipitation at Valley Falls
and discharge of the Chewaucan River......................................................................... 6
5. Correlation between coefficient of area variation and geometric-shape factor for Lake Abert,
Summer Lake, and several other closed lakes.................................................................. 11
6. Estimated discharge and dissolved-solids content of Chewaucan River near Valley Falls, water
years 1961-65................................................................................................. 13
7. Changes in the chemical character of Chewaucan River near Valley Falls with increasing
dissolved-solids content...................................................................................... 16
8. Map showing source and chemical character of sampled spring and seep flow peripheral to
Lake Abert........................................................................................................... 17
9, 10. Bar graphs showing relative abundance of major anions in:
9. Playa-surface salts adjacent to Abert and Summer Lakes......................................................... 17
10. Abert and Summer Lakes, 1912 and 1961........................................................................ 19
11-14. Graphs showing:
11. Relation of dissolved-solids tonnage to fluctuations of volume and surface area of
Lake Abert, 1937-69.......................................................................................... 21
12. Variations in lake-surface altitude and dissolved-solids tonnage at Summer Lake,
1959-69...................................................................................................... 22
13. Relation of increasing dissolved-solids tonnage in Summer Lake during 1963-68 to fluctuations of lake
volume and surface area, and time since last prior desiccation............................................... 23
14. Hypothetical long-term cycles of dissolved-solute tonnage and lake volume for a closed lake.................... 24
1. Map showing physiographic and hydrologic setting and the chemical character of brines in shallow lake-bottom
and peripheral sediment, Lake Abert and Summer Lake, south-central Oregon................................... In pocket
1. Index map showing location of Abert and Summer Lakes................................................................ C2
2. Aerial photographs of Abert and Summer Lakes........................................................................ 3
illIV
CONTENTS
TABLES
Page
Table 1. Hydrologic budgets for Summer Lake before 1900 and after 1926........................................................................ C8
2. Area and volume of Abert and Summer Lakes......................................................................................... 8
3. Parameters that describe the physical and hydrologic characteristics of Abert and Summer Lakes................................... 10
4. Chemical analyses of representative surface and ground waters in the Abert and Summer Lake basins................................ 14
5. Changes in dissolved-solids tonnage during passage of Chewaucan River through Chewaucan Marsh in water years 1961 and
1965, and in an average year............................................................................................. 14
6. Solute contributions to Abert and Summer Lakes and their peripheral areas by streams, springs, and seeps......................... 15
7. Dissolved-solids concentrations and tonnages in Lake Abert at various times since 1911........................................... 20
8. Dissolved-solids concentrations, tonnages, and chloride : sodium ratios in Summer Lake at various times since 1911............... 20
9. Analytical data for sediment and interstitial brine at representative lake-bottom and playa sampling sites....................... 24
10. Relative abundance of major ions in representative samples of lake water and interstitial brine from lake-bottom and
peripheral-playa sediments............................................................................................... 25CLOSED-BASIN INVESTIGATIONS
SOLUTE BALANCE AT ABERT AND SUMMER LAKES, SOUTH-CENTRAL OREGON
By A. S. Van Denburgh
ABSTRACT
Neighboring Abert and Summer Lakes occupy the broad flat floors of deep, topographically enclosed basins in the high desert of south-central Oregon. The two saline lakes and their tributary areas provide contrasting hydrologic environments well suited to an investigation of the various aspects of solute economy. Exclusive of direct precipitation, Lake Abert receives about 90 percent (62,000 acre-ft) of its average annual water supply from the snow-fed Chewaucan River. Summer Lake, in contrast, derives about 74 percent (67,000 acre-ft per year) of its inflow from Ana Springs via Ana River. Prior to impoundment of the river and inundation of the springs, the annual contribution probably averaged about
100.000 acre-feet (83 percent of the total). The Chewaucan River at its mouth carries an average of about 120 ppm (parts per million) of dissolved solids, whereas the flow of Ana Springs contains 160 ppm.
The total solute loads contributed to Abert and Summer Lakes from all sources average about 13,000 and 24,000 tons per year, of which ground-water increments constitute approximately 25 and 95 percent. Prior to 1900 the annual increment to Summer Lake was about 26,000 tons. Constituents that dominate in the lakes (sodium plus equivalent chloride and carbonate-bicarbonate) have been contributed by inflow in the following approximate amounts during an average year: Lake Abert, 5,600 tons; Summer Lake since 1926, 14,000 tons; Summer Lake prior to 1900,
15.000 tons. The chloride:sodium ratio has been about 0.2 for each of the increments, on the basis of equivalents per million.
During the 50 years beginning in 1916, the area and maximum depth of Lake Abert have averaged almost 50 square miles and 6'A feet. The lake dried several times during the drought of the 1920’s and 1930’s, whereas in 1958 it covered 64 square miles and had a maximum depth of 1654 feet. By 1969 the lake contained almost 15 million tons of dissolved solids and had a chloride:sodium ratio of 0.57. The solute concentration generally is
30.000 to 60,000 ppm except at unusually high lake levels or near-dryness.
At Summer Lake prior to 1900 the average maximum depth was about
6H feet, and the water body characteristically covered about 60 square miles. The solute load totaled approximately 5 million tons and was chemically similar to the load in Lake Abert, except that the chloride:sodium ratio was only 0.29. Since 1926 Summer Lake has dried frequently and has had an average maximum depth of only about IH feet and a surface area of about 40 square miles. In 1969, after containing water continuously for almost 6 years, the lake held 1.4 million tons of solutes, which included a somewhat greater proportion of chloride than prior to 1926 (chloride:sodium ratio characteristically was about 0.35). The dissolved-solids content of Summer Lake generally is between 5,000 and 50,000 ppm.
The solute tonnage in broad, shallow closed lakes with wide marginal playas fluctuates in response to long-term changes in lake area and volume. At Abert and Summer Lakes during a generally receding phase, major solute losses occur in only the final stages of recession, when much of the tonnage becomes entrapped within the peripheral playa sediments. The depletion can continue at a lesser rate with the aid of the wind during a characteristically low-level period. Much of the lost solute tonnage is recovered during or soon after a return to high level, and additional solutes accumulate if the lake remains at high level.
Fine-grained lacustrine deposits that underlie the lakes and adjacent playas harbor large solute tonnages. Beneath the lakes, interstitial brines constitute as much as 80-85 percent of the muds, by volume. Measured salinities at 2-5 feet depth are 40,000-50,000 ppm beneath Summer Lake, and they exceed 100,000 ppm beneath much of Lake Abert. The interstitial solutes are dominated by sodium among the cations and are richer in carbonate plus bicarbonate and poorer in chloride than the overlying lake waters. Characteristic chloride:sodium ratios are 0.25-0.30 for brines beneath Summer Lake and about 0.50 for those at Lake Abert.
Shallow playa sediments adjacent to the lakes contain somewhat smaller quantities of brine (50-80 percent) than the bottom sediments. At Lake Abert measured salinities also are less (80,000-100,000 ppm), whereas at Summer Lake they are greater (80,000-140,000 ppm), reflecting differences in peripheral ground-water input. The most pronounced contrast between playa and lake-bottom solutes at both lakes is in the relative abundance of anions. Deposits at the playa surface are far richer in carbonate and bicarbonate than are those in either the lake or the lake-bottom interstitial fluids. The enrichment apparently is associated with evaporation of upward-moving fluids: the first-formed salts of sodium carbonate and bicarbonate presumably tend to shield the residual chloride-enriched brine from continued evaporation and resultant crystallization. Wind removal of the powdery components of carbonate- and bicarbonate-rich salts from the playa surface causes long-term enrichment of chloride in residual brines; the relative abundance of chloride beneath the playas is far greater than in or beneath the lakes (characteristic chloride:sodium ratios beneath the playas are about 0.8 at Lake Abert and 0.5 at Summer Lake).
At Lake Abert, solute quantities in the top 5 feet of playa and lake-bottom deposits average about 550 tons per acre (the greatest amounts per acre underlie the lake), and total 25-35 million tons, of which about one-third is chloride. Comparable values at Summer Lake are about 300 tons per acre (w ith greatest per acre amounts beneath the eastern playa) and 15-20 million tons overall, of which only 20-25 percent is chloride.
ClC2
CLOSED-BASIN INVESTIGATIONS
The relation between incoming and stored solutes at the two lakes contrasts sharply. Total salts and chloride stored in Lake Abert and within the top 5 feet of bottom and playa sediments are, respectively, 8,000 and 16,000 times the average present-day income, whereas comparable values at Summer Lake are only about 1,000 and 2,000 (the higher values for chloride reflect its enrichment relative to carbonate and bicarbonate). This contrast between the two lakes contradicts the situation that would be expected because of (1) the different solute quantities that may have been retained in the two basins after the final shrinking of Pleistocene Lake Chewaucan (Summer Lake should have inherited a tonnage greater than that of neighboring Lake Abert), and (2) apparent differences in the long-term uniformity of water supply and, therefore, lake stability in the two basins (Summer Lake should have been more stable, because of the dominance of ground-water inflow). Even with allowance for possible differences in lake-basin sedimentation rates, the solute balances contrast, suggesting strongly that Summer Lake has not always had an income as abundant as it did immediately prior to man’s alteration of the natural hydrologic situation. A major prehistoric increase in the amount of interbasin ground-water flow feeding Ana Springs, because of faulting, is possible, though other evidence discourages this. Alternatively, because the area-to-depth ratio at shallow levels is much greater at Summer Lake than at Lake Abert, the effects of evaporation and wind may have been magnified at the former during long prehistoric dry periods, when inflow, regardless of its source, was diminished. Large deposits of windblown sand east of Summer Lake argue in favor of extended periods of near-desiccation, with attendant depletion of stored playa salts by deflation.
INTRODUCTION
LOCATION AND GENERAL FEATURES OF THE AREA
Adjacent Abert and Summer Lakes lie in the high desert country of south-central Oregon, within the northwest limits of the Great Basin (pi. 1; fig. 1). They occupy deep, topographically enclosed basins, and lose water only by evaporation. The two saline lakes and their tributary areas cover 1,250 square miles, of which 860 square miles belongs to Lake Abert. Altitudes within the combined basin range from about 4,145 feet at Summer Lake and 4,250 feet at Lake Abert to almost 8,400 feet at one place on the western drainage divide. The lowest point of possible overflow from the combined basin probably exceeds 4,600 feet altitude. The lakes themselves are separated by a divide with a low point at about 4,385 feet.
During the late Pleistocene a large lake filled both basins to a maximum level of about 4,520 feet, covering almost 500 square miles with a volume of water that exceeded 21 cubic miles (70 million acre-ft). The fluctuations of that lake are recorded by prominent shoreline features at many places throughout the two basins.
The area surrounding the lakes is, for the most part, one of intricately faulted Tertiary volcanic flows and pyroclastic debris, mapped on a reconnaissance basis by George W. Walker (1963). Dip-slip movement along the faults has formed numerous horsts, grabens, and tilted fault blocks of spectacular proportion, such as those described by Donath (1962). Two of the most prominent structural features are Winter Ridge, a 3,000-foot scarp
0 50 100 KILOMETRES
Figure 1.—Map of Oregon showing location of Lake Abert (A), Summer Lake (S), and the area included on plate 1.
that overlooks Summer Lake on the west (pi. 1; fig. 2), and Abert Rim, an even more abrupt escarpment that rises, at places almost vertically, to a height as much as 2,450 feet above Lake Abert within 1 mile of its eastern shore. The two lakes occupy depressions that formed during the faulting and subsequently filled with clastic and lacustrine sediments.
PURPOSE AND SCOPE OF INVESTIGATION
Because of their physical and hydrologic characteristics, Abert and Summer Lakes offer an excellent combined site for a study of various aspects of solute balance within a closed-basin environment. This investigation has furnished information on: The amounts and chemical character of incoming, outgoing, and stored solutes; the mechanisms and effectiveness of accumulation and depletion; the stability of the stored-solute resources; the manner in which these closed-lake systems react to and record changes in climate and activities of man; and the factors that may account for a contrast between solute balances at the two lakes.
A reconnaissance of the two lakes and their inflow has been made by Phillips and Van Denburgh (1971). The present study utilizes information presented in the reconnaissance report, along with new data collected since that time and previous information not included in the preliminary study.
ACKNOWLEDGMENTS
I am particularly indebted to Blair F. Jones, of the U.S. Geological Survey, who has been an invaluable source of information, advice, and great encouragement through-Figure 2.—Aerial photographs of Abert and Summer Lakes. A, High-altitude vertical photograph of the two lakes and adjacent areas, October 24, 1972. North at top; picture width about 52 miles. Estimated lake levels: Abert, 4,258.5 feet; Summer, about 4,145 feet. (From ERTS-I satellite photograph 81093181615G000; multispectral scanner, infrared wavelength range.) B, Oblique view of Lake Abert, viewed to the north-northeast. In foreground,
Chewaucan River meanders north to lake. Abert Rim dominates skyline east of lake. Photograph taken in August 1963; lake level 4,254.5 feet. C, D, Oblique views of Summer Lake, looking south. Marshy wildlife-management area in foreground. Broad mudflats east of lake contrast with Winter Ridge and Slide Mountain to the west and south. Photographs taken in July 1964; lake level 4,147.2 feet.
n
ua
SOLUTE BALANCE AT ABERT AND SUMMER LAKES, OREGONC4
CLOSED-BASIN INVESTIGATIONS
out the investigation, and to Glenn E. Tyler, retired State Watermaster, District 12, who has collected many of the critical water samples and streamflow data, and who, along with his family, has provided a hospitality that makes the high desert country of south-central Oregon so special. I am also grateful to Walter B. Langbein, Kenneth N. Phillips, Shirley L. Rettig, George I. Smith, and Alfred H. Truesdell of the U.S. Geological Survey, and Boyd Claggett of the Summer Lake State Game Management Area, who have provided valuable help during the study, as well as to Meyer Rubin of the Survey, who provided carbon-14 age determinations on four sediment samples.
LOCATION SYSTEM
The numbering system for hydrologic sites in this report indicates location on the basis of the rectangular subdivision of public lands, referenced to the Willamette base line and meridian. Each number consists of three units: the first is the township south of the base line; the second unit, separated from the first by a slant, is the range east of the meridian; the third unit, separated from the second by a dash, designates the square-mile section. The section number is followed by letters that indicate the quarter section, quarter-quarter section, and so on; the letters a, b, c, and d designate the northeast, northwest, southwest, and southeast quarters, respectively. For example, site 33/21-2cdb is in NW14SE14SWW sec. 2, T. 33 S„ R. 21 E.
All locations are based on quadrangle maps of the Oregon State Highway Department (scale, about 1 inch to the mile) and U.S. Geological Survey (about 2/i inches to the mile).
HYDROLOGIC CHARACTERISTICS OF THE TWO LAKES AND THEIR BASINS
CLIMATE
Climatic conditions in the high desert country of south-central Oregon range from arid or semiarid on the valley floors to semihumid in the timbered mountain areas. Average yearly precipitation ranges from 12 inches or less in the dry northeastern parts of the two basins to more than 25 inches in the lofty, well-timbered southwestern parts. In the lowlands about 90 percent of the average annual precipitation is fairly evenly distributed throughout the 9-month period of October-June (fig. 3). From November through March most of the precipitation occurs as snowfall, whereas thundershowers are principal contributors during the remaining 7 months of the year.
Long-term records of annual precipitation at Valley Falls are shown in figure 4. The cumulative departure curve in that figure shows that rain and snowfall were characteristically less than average prior to 1940 and generally greater than average during 1940-65.
Precipitation records are also available for two other stations on the valley floors of the Abert and Summer Lake basins. These Weather Bureau data are summarized below for the short period of concurrent record, 1958-64:
Average annual
Station Location1 Altitude precipitation,
name (ft) 1958-64
_______________________________________________________________(in.)
Valley Falls....... S6/21-6aba 4,320 13.8
Paisley............ 33/18-24dbd 4,360 11.6
Summer Lake........ 30/16-23acd 4,190 13.8
'See p. C4 for explanation of location system.
On the basis of this information and data for sites outside the two basins, the estimated lakewide average may be about 11 inches per year at both lakes. For Lake Abert, this would represent about 90 percent of the quantity at nearby Valley Falls.
Potential evaporation on the floors of the Abert and Summer Lake basins is about 40 inches per year, two-thirds of which occurs from May through September (fig. 3). The actual amount of water lost from Abert and Summer Lakes themselves is uncertain, because evaporation varies from place to place depending on differences in air temperature, relative humidity, amount of wind, and the physical, thermal, and chemical character of the water body.
Water losses from a Class-A land pan have been measured at the Summer Lake National Weather Service station during part of each year since 1961. The nearest long-term record of year-round pan evaporation is that for the Medford Experiment Station, 110 miles southwest of Summer Lake. Concurrent measurements at these two stations are available for 36 months (mostly May-Oct.) during 1961-66. The coefficient of correlation between the two sets of monthly data is a favorable 0.89. On the basis of these concurrent measurements and the year-round records at Medford, the land-pan evaporation rate at Summer Lake station for the long term is an estimated 57 inches per year. This is equivalent to about 41 inches from a freshwater body, assuming a pan coefficien t of abou 10.73 (Kohler and others, 1959, pi. 3). The long-term freshwater rate for the lake itself may be slightly higher because the weather station is on the cooler, western side of the basin in the shadow of Winter Ridge (pi. 1).
Evaporation at Lake Abert is less certain. During May-October 1962, K. N. Phillips and his coworkers measured a lake-surface water loss equivalent to 32.5 inches from a freshwater body (Phillips and Van Den-burgh, 1971, p. B14). At the Summer Lake station during the same period, pan evaporation was 44.7 inches, which converts to an almost identical freshwater value of 32.6 inches, using the coefficient 0.73. Comparison of monthly lake and pan evaporation can be risky, owing to the effects of inflow and heat storage within the lake (Nordenson,SOLUTE BALANCE AT ABERT AND SUMMER LAKES, OREGON
C5
PRECIPITATION
STREAMFLOW
AIR TEMPERATURE
EVAPORATION
WIND
□
s
0 10 20
PERCENT OF TIME
Figure 3.—Wind movement and representative seasonal patterns of precipitation, streamflow, air temperature, and evaporation in and adjacent to the Abert and Summer Lake basins (from records of the National Weather Service and the U.S. Geological Survey). Precipitation at Valley Falls for water years 1916-65. Streamflow for Che-waucan River near Paisley, water years 1913-21, 1925-65. Temperature at Valley Falls for calendar years 1931-60. Evaporation fromClass-A land pan at Summer Lake station for water years 1961-65 (mostdata for May-Oct. measured; most data for Nov.-April estimated on basis of
records at Medford Experiment Station, 115 miles west-southwest of Summer Lake). Average wind directions and speeds at Lakeview airport, 26 miles south of Lake Abert, based on observations during period January 1959-May 1962 (Howell, 1965). Length of each direction bar shows percentage of time that winds exceeding 3 mph (miles per hour) were from the direction indicated. Winds of 0 to 3 mph, which occurred 38 percent of the time, are not shown. Type of bar indicates wind speed: single-line bar, 4-17 mph; double-line bar, 18-35 mph.
1963, p. 279). At Lake Abert, however, the risk is minimized because inflow during 1962 was well below normal, and the lake was shallow (mean depth averaged only 5 feet during the period). Thus, the comparison suggests that the freshwater evaporation rate at Lake Abert is about 41 inches per year (slightly greater than the 40.3 inches per year estimated earlier by Phillips and Van Denburgh (1971, p. B15) ). The actual rate of annual evaporation from the lake would be less than 41 inches, however, because of the effect of salinity.
INFLOW
Abert and Summer Lakes are fed by inflow of contrasting hydrologic character. The principal source of flow into Lake Abert is the Chewaucan River, whereas Summer Lake is fed principally by Ana Springs, which rise about 6 miles north of the lake (pi. 1). Both lakes are sustained partly by direct precipitation. Also, small intermittent streams drain areas adjacent to the lakes, and the discharge of peripheral springs and seeps reaches each lake.
LAKE ABERT
The Chewaucan River drains about 490 square miles, much of which is timbered and mountainous. The principal source of flow is snowmelt runoff, and most of the discharge therefore occurs during the spring and early summer. At the gaging station near Paisley (pi. 1) in an average year, 74 percent of the total annual runoff occurs during March-June (fig. 3). The station (drainage area, 275 sq mi), is about 32 river miles upstream from the mouth, but the seasonal distribution of flow into Lake
Abert probably is not much different, although the average yearly volume is considerably less.
The annual flow of the Chewaucan River near Paisley has averaged 96,000 acre-feet during the period 1916-65 (maximum value was 250,000 acre-feet during the 1956 water year, minimum value was an estimated 18,000 acre-feet during the 1924 water year). The record from 1916 through 1965 is summarized in figure 4. The accompanying cumulative departure curve shows that flow was generally below normal during the period 1918-37 but was much greater than normal from 1951 to 1958. The pronounced recovery from a period of lower than average streamflow, which began in 1951, apparently lagged 11 years behind a similar recovery from generally below normal precipitation (fig. 4). The lag may reflect ground-water replenishment upstream from Paisley following the pre-1940 drought.
The average discharge of the Chewaucan River at its mouth for the period 1916-65 is an estimated 62,000 acre-feet—65 percent of the upstream value. The relation between runoff at the two sites is a refinement of that developed by Phillips (Phillips and Van Denburgh, 1971, p. B15), using an extrapolation of climatic and lake-level data obtained during the fairly well documented period 1951-65.
Agricultural activity in Chewaucan Marsh apparently has not as yet (1970) had a pronounced effect on the amount of runoff reaching Lake Abert (Phillips and Van Denburgh, 1971, p. B13). This is, at least in part, because much potential inflow was lost through natural evapo-transpiration in the swampy flatlands prior to drainage ofC6
CLOSED-BASIN INVESTIGATIONS
20
15
w
LU
I
u
z
10
5
0
0
(0
LU
I
o
z
-10
-20
-30
-40
400
iu 300
cc o . <
D
_l
o
>
200
100
Ui
LU
LU
CC
o
<
-200
-400
-600
250
LU
CC
o
<
200
150
C/D
□
z
<
(/)
D
0
1 h
100
50
63
60
(/)
LU
i
LU
CC
<
(/)
50 ?
40
0
<
LU
CC
<
1920 1930
1940 1950 1960
Figure 4.—Fluctuations in volume and surface area of Lake Abert in water years 1916-65, compared with precipitation at Valley Palls and discharge of the Chewaucan River. In bottom graph, shaded bars indicate measured annual discharge near Paisley; heavy-line
bars indicate estimated total annual inflow to Lake Abert. Horizontal lines in top, middle, and bottom graphs indicate average annual precipitation (12.2 in.), lake volume on September 30 (111, 000 acre-ft), and annual discharge near Paisley (96,000 acre-ft), respectively.SOLUTE BALANCE AT ABERT AND SUMMER LAKES, OREGON
Cl
the marsh for agricultural use in the early 1900’s. These natural losses may have been comparable in magnitude to present losses due to irrigation.
Inflow to Lake Abert from other streams may average about 2,500 acre-feet per year, on the basis of altitude-runoff relationships developed by D. O. Moore (US. Geological Survey, oral commun., 1970) for nearby Honey and Silver Creeks (U.S. Geological Survey gaging stations 10378500 and 10390000). Data for the Chewaucan River were not used in developing the relationships because that stream’s headwaters area, above Paisley, is in a much wetter altitude-runoff environment than are Lake Abert and vicinity.
The total annual discharge of all seeps and springs along the periphery of Lake Abert may be about 5,000 acre-feet, though the amount of ground-water flow actually reaching the lake may average only 2,000 to 2,500 acre-feet per year, owing to evaporation. Some of the flow doubtless is derived from recharge within the Lake Abert basin, but most of it may come from an adjacent basin or basins (Phillips and Van Denburgh, 1971, p. B13). No large submerged springs are known (none was reported in periods when the lake was dry).
In summary, then, the combined flow supplied to Lake Abert and its peripheral areas in excess of direct lake-surface precipitation probably averaged about 70,000 acre-feet per year (62,000 acre-ft from the Chewaucan River, 5,000 acre-ft from seeps and springs, and 2,500 acre-ft from ephemeral streams).
SUMMER LAKE
Ana Springs, the principal source of discharge into Summer Lake, originally provided a relatively consistent year-round supply that flowed directly to the lake via the Ana River. On the basis of an estimated average discharge of 130 to 150 ftVs (cubic feet per second) from the several orifices, the annual contribution under natural conditions prior to 1900 amounted to about 100,000 acre-feet. A dam was completed immediately downstream from the springs in 1926, and the impounded water inundated the orifice area to a depth ranging from 16 to 46 feet (Phillips and Van Denburgh, 1971, p. B25). The resulting hydrostatic head has reduced the average discharge rate to about 92 ftVs (or 67,000 acre-ft per year).
At present, additional springflow apparently feeds the river below the gage. Several measurements, about 2 miles downstream from the gage (listed below), suggest that the pickup in that reach averages about 10,000 acre-feet per year. Some of this increment may represent ground water that would issue from the main Ana Spring orifices under natural (pre-dam) conditions. Limited and seemingly conflicting data prior to construction of the dam (table in right column) suggest that inflow below Ana Springs may not have been much less than now, though the approximate quantities are uncertain. For the purpose of
Discharge (fts/s)
Date Near Ana Springs Near lake Increase (+) or
(30/17-6d) (30/17-9ccd) decrease (-)
After construction of Ana Reservoir1
5-16-68........ 27 38 +11
7- 10-68............. 54 62 +8
11- 27-68............ 93 114 +21
12- 5-69............. 94 93 -1
2- 19-70............ 119 118 -1
3- 24-70............. 94 106 +12
8- 6-70.............. 34 42 +8
Before construction of reservoir
7-17-04........ 165 179 +14
3-28-05........ 148 s150 +2
'Data from Glenn K. Tyler, Watermaster (written communs., 1968-70).
^Data from Henshaw and Dean, 1915, p. 756-757; accuracy of near-lake measurements uncertain because of backwater (this does not apply to measurements of 1968-70 when lake was about 5 ft lower).
^Measured Jan. 8. 1905; assumed comparable with later upstream measurement.
solute tonnage computations, the pre-dam increment is assumed to have been about half the present-day quantity, or 5,000 acre-feet per year.
The amount of flow actually reaching Summer Lake from Ana Springs and other springs tributary to the Ana River is depleted by diversions for irrigation of agricultural and waterfowl resting areas. A crude estimate of present-day net depletion due to these diversions is about 20,000 acre-feet per year (under natural conditions, only an estimated 10,000 acre-ft was lost through evapo-transpiration in marshy areas north of the lake). Thus, net annual flow reaching the lake from the spring systems is now about 57,000 acre-feet—only a little more than half of the pre-1900 quantity. Furthermore, most of the present inflow occurs during the period October-April, rather than throughout the year at a nearly constant rate as it did prior to impoundment and diversion.
Many small streams, in part springfed, drain the rugged, timbered areas west and south of Summer Lake (pi. 1), and far smaller quantities of runoff occasionally reach the lake from the east during brief thundershowers. The streams may contribute about 10,000 acre-feet per year, on the basis of altitude-runoff relationships. Much of this flow is dissipated by evapotranspiration—both natural and man-caused—before reaching the lake. Under natural conditions prior to 1900, the annual depletion may have averaged only about a third of the 10,000 acre-foot total, whereas, under present-day conditions, irrigation and natural losses may dissipate about two-thirds of the total.
Springs and seeps in addition to those that feed the Ana River contribute a small amount of flow to the lake and its peripheral areas. Best known of the group is Summer Lake Hot Spring (33/17-12aac), which flowed at a rate of about 34 acre-feet per year in 1948. Many springs on the basin floor were visited in 1948-49, after a 10-year period of above-normal precipitation (fig. 4). For visited springs not contributing to the river, the combined flow was aboutC8
CLOSED-BASIN INVESTIGATIONS
2,200 acre-feet per year (Trauger, 1950, p. 218-224). A reasonable estimate of total spring discharge on the basin floor, in addition to that feeding the river, may be 3,000-4,000 acre-feet per year, of which only a small part overcomes evapotranspiration and reaches the lake. Much of the springflow west and south of Summer Lake doubtless represents the reappearance of recharge from streams draining the Winter Ridge-Slide Mountain area.
At some time in the past Summer Lake received overflow from the Lake Abert basin via the channelway north of Paisley (Phillips and Van Denburgh, 1971, p. B15), but such overflow probably has not occurred since Pleistocene time.
Interbasin ground-water flow toward Summer Lake fed by percolation from the Chewaucan River north of Paisley is considered to be small. Ground-water gradients and lithologic logs for wells on the alluvial-lacustral fan north of the river indicate that the total northward flow may average only about 2,000 acre-feet per year, and well-water chemistry suggests that most of the solutes are not derived directly from the river.
Table 1 summarizes the water budget of Summer Lake for conditions before and after man’s alteration of the natural hydrologic environment. The data indicate that present-day net inflow averages only about 60 percent of the quantity under natural conditions.
Table 1Hydrologic budgets for Summer Lake before 1900 and after 1926
[Acre-feet per year, except as indicated1]
No. Item Pre-1900 Post-1926
1. Flow of Ana Springs........................... 100,000 67,000
2. Inflow to Ana River below springs............... 5,000 10,000
3. Depletion of Nos. 1 and 2 by evapotranspiration
before reaching Summer Lake............... 10,000 20,000
4. Mountain-front streamflow west and south
of lake...................................... 10,000 10,000
5. Depletion of No. 4 by evapotranspiration before
reaching Summer Lake.......................... 3,000 7,000
6. Net inflow to Summer Lake (1+2 + 4- 3-5,
rounded).................................... 100,000 60,000
7. Lake area corresponding to No. 6 (acres)2... 38,000 22,000
8. Lake level corresponding to No. 7 (feet above
mean sea level)............................... 4,151 4,146
'Probable accuracy: Item 1, ± 15,000 acre-feet before 1900, ± 5,000 acre-feet after 1926; items 2, 3, and 6, ± 5,000 acre-feet; items 4 and 5, ± 3,000 acre-feet; item 7, * 2,000 acres; item 8, pre-1900, ±1 foot, post-1926, ± H foot.
2 Assumes average net evaporation of 2.6 feet per year before 1900 and 2.7 feet per year after 1926 (table 3).
LAKE EXTENT AND FLUCTUATION
Abert and Summer Lakes occupy broad, flat, elongate basins (pi. 1). At maximum depths of only 5 feet, the 2 water bodies cover about 44 and 57 square miles, respectively. Data relating lake volume, areal extent, and lake-surface altitude are presented for the two lakes in table 2. Data for Lake Abert are based largely on a reconnaissance
Table 2.—Area and volume of Abert and Summer Lakes
Lake level
(ft above Area Volume
mean sea (sq (thousands
level) mi) of acre-ft)
Lake Abert'
4,244 ± Dry 0.0
4,245 12.2 4.0
4,246 21.4 14.8
4,247 30.5 31.4
4,248 38.8 53.5
4,249 43.9 80.0
4,250 47.6 110
4,251 50.6 141
4,252 53.0 174
4,253 55.1 209
4,254 56.8 245
4,255 58.3 282
4,256 59.7 320
4,257 60.9 358
4,258 62.0 397
4,259 63.0 437
4,260 63.7 478
4,262 65 560
Summer Lake
4,144.4 ± Dry 0.0
4,145 23.5 4.5
4,146 37.5 23.9
4,147 45.5 50.4
4,148 51.0 81.0
4,149 55.2 116
4,150 58.7 152
4,152 64.0 230
4,155 70 360
'Data are slightly modified from those of Phillips and Van Denburgh (1971, table 4) on the basis of reevaluated bathymetry below 4,248 feet.
survey made in 1959 and are considered to be reasonably accurate. The data for Summer Lake are approximations based on shoreline configurations from three sets of aerial photographs, information from Geological Survey 1:24,000- and 1:250,000-scale topographic maps, and personal observation. (Data from aerial photographs are as follows: 27 sq mi at 4,145.2 ft on Aug. 8, 1946; 35 sq mi at 4,145.75 ft on Sept. 9, 1953; and 49H sq mi at 4,147.7 ft on July 23-24, 1963.)
The long-term variations of lake level and volume reflect changing climatic conditions. The record of volume fluctuations at Lake Abert since 1915 is shown in figure 4. (Most values for the poorly documented period from 1915 to 1951 are based on lake-level estimates, using precipitation at Valley Falls and runoff in the Chewaucan River near Paisley (Phillips and Van Denburgh, 1971, p. B15). The validity of these estimates is supported by close agreement with the few recorded levels.) Comparison with curves showing cumulative departures from average precipitation and runoff in the basin illustrate how sensitive the lake is as an index of climatic and hydrologic variation. Only a few years of deficient precipitation and runoff are necessary to cause lake-basin desiccation. OnceSOLUTE BALANCE AT ABERT AND SUMMER LAKES, OREGON
C9
the lake is dry, an excess of local precipitation alone is not sufficient to cause lake-volume recovery, as shown during 1940-48 (fig. 4). Only after return to above-normal runoff rates does the lake respond by filling to former levels, as it did during 1951-58.
The average amount of water in Lake Abert during the present century may be less than under comparable climatic conditions in the past, owing to agricultural activity in Chewaucan Marsh. Presumably, the difference is relatively small, however (p. C5).
In contrast to conditions at Lake Abert, the extent and fluctuations of Summer Lake have been markedly altered by the activities of man. Prior to impoundment of the Ana River, Summer Lake received most of its water supply from the Ana Spring system, and the inflow occurred at a fairly constant rate. Therefore, seasonal and long-term fluctuations of lake level doubtless were much more subdued than those for Lake Abert because of the secondary importance of highly variable surface runoff. Since construction of the reservoir, however, the submergence of orifices and the seasonal diversions for irrigation have greatly reduced the amount and month-to-month consistency of flow into the lake. Present inflow averages only about 60 percent of the pre-reservoir quantity (table 1), and most of it occurs from October through April. Hence, seasonal variations in lake level are more pronounced, and the long-term average is appreciably lower than pre-reservoir values. In fact, Summer Lake has dried almost completely during the summers of many years since 1926.
The changes brought about by impoundment and diversion of the Ana River are shown more exactly by comparing hypothetical “equilibrium” lake levels under the natural and altered conditions. At “equilibrium” level the lake area is such that evaporation exactly balances the incoming water volume. Thus, the “equilibrium” lake area, in acres, equals the average annual inflow, in acre-feet, divided by average net evaporation (total evaporation less precipitation), in feet per year. Table 1 shows that Summer Lake is an average of about 5 feet shallower now than under natural conditions. (The budgets in table 1 are refined and more detailed versions of those presented by Phillips and Van Denburgh (1971, p. B26); the changes result from additional information concerning several of the budget items.)
PARAMETERS THAT CHARACTERIZE THE LAKES
Langbein (1961, p. 6, 13, 14) presented several interrelated parameters that depict the geometric and hydro-logic characteristics of a closed lake and its tributary basin. These characteristics in turn strongly influence the chemical character of the lake. Among the important parameters are
1. The geometric lake-bottom shape factor, which equals
the square root of the average lake-surface area, \&T^, divided by the mean lake depth, D. A high value indicates a flat saucer-shaped lake, whereas a low value indicates a deep lake.
2. The coefficient of lake-area variation, U, which indi-
cates the variability of lake area with time, relative to the mean lake area. A high value indicates an unstable lake whose area varies greatly with time; a low value indicates a stable lake in which the variations are small relative to the mean.
3. The response time, k, which indicates how rapidly a
lake reacts to changes in climate. As Langbein says (1961, p. 6): “A lake with a low value of k, near 1 year, is a playa lake. It fills and dries up in a year. It responds to the current year’s rainfall and virtually not at all to that of preceding years. A lake with a high value of k, on the other hand, reacts slowly, and may be at a high level during a period of low rainfall and vice versa.”
4. The long-term stability factor, P, which indicates the
probability of desiccation or, at the other extreme, of freshening by overflow. Low values indicate near-desiccation or near-overflow, whereas a high value indicates greater stability.
The coefficient of lake-area variation can be calculated directly from detailed records of lake level, after converting the levels to areas. The data are applied to the general formula for a coefficient of variation (J = S jA, where
sm,2:
(ia,:
s 2
(1)
In the absence of detailed lake-level data, Langbein’s equation 8 (1961, p. 6) is used. The equation is
0.26re /e(At/AL)k
D J
\ 2+1/*
(2)
where n is the exponent in the proportionality relationship between lake area and volume, AL oc Vn; D is mean depth; E is net annual evaporation (total evaporation minus precipitation); A T is the tributary area exclusive of the lake; A ^ is lake area; and k is the response time.
The response time is defined as:
k =
V"-V' E(A"-A') ’
(3)
where V" and V' are lake volumes at a high and low stage, A" and A' are the equivalent lake areas, and E is net annual evaporation.
The long-term stability factor, P, is defined in terms ofCIO
CLOSED-BASIN INVESTIGATIONS
mean lake depth, D, in feet;tributary area, Ap, in square feet; and basin volume at altitude of overflow, Vq , in cubic feet, as follows:
LAKE ABERT
Lake Abert invites a rather detailed analysis of characterizing parameters because of the comparatively simple hydrology and the relative abundance of data. The period of most complete record at the lake spans water years 1952-65. Unfortunately, however, the average lake level during that period was considerably higher than the average for the much longer period of record, 1916-65 (fig. 4). Nonetheless, Langbein’s general parameters are evaluated for both periods. Values for several of the descriptive parameters are summarized in table 3. The geometric lake-bottom shape factors and coefficients of lake-area variation are compared with those of other closed lakes in figure 5.
Table 3.—Parameters that describe the physical and hydrologic characteristics of Abert and Summer Lakes
Lake Abert Summer Lake
Parameter 1952-65 1916-65 Post-1926 Pre-1900
Tributary area excluding lake, in square miles (Aj,) '730 •740 350 320
At average lake level: Lake-surface altitude, in feet .... 4,254.7 4,250.5 4,146 4,151
Lake area, in square miles (/i/J2 58 49 38 62
Lake volume, in acre-feet 126,000 24,000 190,000
Mean lake depth, in feet (D) 7.3 4.0 1.0 4.8
At high lake level: Altitude, in feet .... 4,260 4,260 4,150 4,154
Area, in acres (A') .... 41,000 41,000 38,000 44,000
Volume, in acre-feet (V,f) .... 478,(KK) 478,000 152,000 315,000
At low lake level. Altitude, in feet .... 4,248 4,246 4,145 4,147
Area, in acres (A') .... 25,(KK) 14,000 15,000 29,000
Volume, in acre-feet (V') .... 54,000 15,000 4,500 50,000
Net evaporation, in feet per year (E) 2.3 2.5 2.7 2.6
Geometric lake-bottom shape factor (y/A^/D) 1.0 1.8 6.2 1.6
Exponent in area-volume equation (n) .20 ’.32 .28 .21
Coefficient of lake-area variation (U; eq 2) 4.085 .23 .36 .074
Response time, in years (L eq S) 11.5 6.9 2.4 6.8
Long-term stability factor (P; eq 4)5 5.6 3.5 0.99 4.6
‘Lake Abert basin covers about 860 square miles, but effective drainage area, including lake, is only an estimated 790 square miles.
2Average lake area (At) for each period is close but not identical to area at average lake level because area-altitude curves are not linear (for example, average area of Lake Abert during 1952-65 was about 57 sq mi).
’Exponent is about 0.45 for lake levels below 4,248 feet and 0.20 above that level; weighted average for period is 0.32.
’Calculated using lake-area estimates at 2-month intervals (eq 1); agrees well with value calculated using equation 2 (0.090).
’Basin volume at altitude of overflow, Vq , is based on drainage divide between the two lake basins, 2.2 miles north of Paisley (alt about 4,385 ft; see pi. 1). Basin volumes at that altitude are Abert, 12 X 106 acre-feet (5.2 X 10" cu ft); Summer, 25 X 106 acre-feet (10.9 X 10" cu ft).
SUMMER LAKE
Summer Lake is similar to Lake Abert in that the period with most adequate data is not representative of long-term conditions. Almost all information on the lake and its inflow has been obtained since Ana River Reservoir was
completed in about 1926, yet hydrologic conditions at the lake since that year have been far different than during the preceding period. Nonetheless, the lake provides an excellent site to test the applicability of Langbein’s generalized parameters in a situation where about 90 percent of the inflow, exclusive of lake-surface precipitation, is provided by ground water. The calculations of several hydro-logic parameters for Summer Lake use the records since 1926, as well as the largely estimated data for natural conditions prior to settlement of the area by man.
The response time and long-term stability factor for the period since 1926 are the lowest recorded in table 3; they, along with the highest recorded lake-bottom shape factor, indicate a very unstable lake, which Summer Lake is known to have been.
The low coefficient of lake-area fluctuation prior to 1900 reflects the relative seasonal and year-to-year uniformity of an unregulated inflow composed principally of ground water (Langbein, 1961, p. 15). The lake-bottom shape factor was higher than normal in combination with such a low coefficient of area variation (fig. 5). This too shows the influence of a presumably steady inflow prior to 1900. The calculated response time of only about 7 years is misleading, because equation 3 reflects only indirectly the steadying influence of ground-water increments.
SOLUTE BALANCE
Evaluation of the solute balance for any closed lake requires a knowledge of the amount and nature of solutes brought into the lake by various agencies, solutes removed from the lake and its immediate surroundings, and solutes accumulated and stored either within the lake water or as components of the lake-bottom and periph-eral-playa deposits. Unfortunately, vagueness concerning the quantitative importance of various facets of a solute budget and the complexity of interrelations between those facets make anything more than a semi-quantitative guess difficult even under fairly well documented present-day conditions. The task of a reliable backward extrapolation of present data over a period of several thousand years is even more formidable.
In studying the salt economy of Abert and Summer Lakes, each basin is divided into two units. The unit of principal concern includes the lake-water body, the upper few feet of saturated sedimentary deposits underlying the lake, and the near-surface parts of adjacent mudflats, which also harbor lake salts. The secondary unit comprises the remaining tributary areas and the atmosphere above. An evaluation of the salt budget, then, involves a quantitative and qualitative study of solutes entering, leaving, and remaining within the unit of principal concern. This in turn necessitates a similar study of solute movement within the secondary unit.SOLUTE BALANCE AT ABERT AND SUMMER LAKES, OREGON
Cll
/yCake area, in squareTmilesV
GEOMETRIC-SHAPE FACTOR ( ---------:--------------- )
V MEAN LAKE DEPTH, IN FEET '
Figure 5.—Correlation between coefficient of area variation and geometric-shape factor for Lake Abert, Summer Lake, and several other closed lakes. Trend of increasing lake-area variation with increasing area-to-depth ratio shows that broad, shallow lakes are more susceptible to expansion and contraction because of the greater relative importance of evaporation. Stabilizing effect of ground-water inflow is shown by the low coefficient for Summer Lake prior to
1900, when springs contributed more than 90 percent of the inflow exclusive of direct lake-surface precipitation. Data for Deep Springs Lake, Calif., from Jones (1965, p. 22; coefficient V is recalculated, using information from Jones’ fig. 13). Data for other lakes except Abert and Summer are from Langbein (1961, p. 7, 18); his data for Pyramid Lake have been recomputed on the basis of revised bathymetry (Harris, 1970) and the lake level as of 1970.
Two general categories of solutes exist within each basin. One category comprises solutes that are mobile (that is, tend to remain in solution) in a saline high-pH environment; it includes constituents dissolved in the
lakes, as well as those dissolved or dissolvable adjacent to and beneath the lakes. At the other extreme are solutes that are mobile in dilute inflow (table 6) but which do not accumulate in solution within an alkaline, saline lake and,C12
CLOSED-BASIN INVESTIGATIONS
therefore, do not play a role in the long-term solute economy of the lakes and their peripheries. These “immobile” solutes yield important organic and inorganic components of the lacustrine deposits beneath and adjacent to the lakes (Jones and Van Denburgh, 1966, p. 443-444).
The major immobile constituents are silica, calcium, and magnesium, whereas the mobile solutes are dominated by sodium, chloride, and carbonate-bicarbonate. Potassium and sulfate lie somewhere between the two extremes; they are depleted in the lakes, but not to the same drastic extent as silica or the alkaline earths.
In computing the balance between incoming, outgoing, and stored solutes, the mobile constituents are by far the most important and are easiest to deal with quantitatively. However, they are subject to recycling within and between the two major units of the basin, and their various increments are therefore difficult to separate into new and recycled portions.
INCOMING AND OUTGOING SOLUTES
LAKE ABERT
Among the several parameters of solute balance at Lake Abert, the contributions of surface flow are the best known quantitatively and qualitatively. (At Summer Lake, in contrast, the character of surface inflow exclusive of spring-fed Ana River is uncertain, but that flow contributes only a small percentage of the total incoming solute load.)
Chewaucan River is the largest tributary to Lake Abert. It drains about 57 percent of the basin and probably supplies almost 90 percent of the surface inflow. The chemical character of the river near its mouth is highly variable seasonally. The measured dissolved-solids content has ranged from about 70 to 260 ppm1 (parts per million) during the sampling period (April 1961 to September 1965). As shown in figure 6, peak values generally occur sometime during the winter or spring, whereas the smallest concentrations are characteristic of flow during late fall and early winter. Seasonal fluctuations of solute content are related in part to irrigation activity in Chewaucan Marsh (pi. 1). The largest concentrations of dissolved solids seem to coincide with periods when irrigation return flow from the marsh provides most of the stream water, whereas the dilute runoff represents flow passing directly through the marsh without diversion for irrigation.
'In calculated dissolved-solids concentrations, wh«ch are used almost exclusively in this report, bicarbonate values are conventionally multiplied by 0.492 to make the overall results comparable with those of dissolved-solids residue determinations. This convention is based on the fact that during the residue procedure, half of the bicarbonate is generally lost as COz (c arbon dioxide), and the other half is converted to carbonate (Hem, 1970, p. 218, 220). Under natural conditions, however, evaporation of a brine that is rich in sodium and bicarbonate generally produces a residue that contains appreciable amounts of bicarbonate as well as carbonate (Bradley and Eugster, 1969, p. 36-58). Thus, not as much COz is lost, and the quantity of natural residue is therefore greater than that determined either analytically or by standard computation. Water of hydration in the naturally occurring evaporite deposits makes the difference even greater. Nonetheless, all computations of solute concentration and tonnage in this report are based on the 0.492 convention for consistency of results.
Estimates of average monthly discharge for the Chewaucan River at its mouth during water years 1961-65 are shown in figure 6. The values are derived by relating estimated amounts of lake-surface precipitation and evaporation to changes in the volume of Lake Abert; the net increment is then ascribed to inflow, most of which is provided by the Chewaucan River. Dissolved-solids tonnages and concentrations based on the data in figure 6 are summarized below:
Water years Dissolved solids
(acre-ft per yr) Tons Parts per million1 per year
1961 12,000 2,700 165
1962 34,000 5,600 120
1963 170,000 28,000 120
1964 55,000 10,000 135
1965 230,000 36,000 115
1961-65 100,000 16,000 120
1916-65 62,000 10,000 120
'Rounded to nearest 5 ppm.
Silica, sodium, calcium, and bicarbonate are the principal dissolved constituents of the river, regardless of total dissolved-solids content. However, figure 7 shows that the relative abundances of sodium and chloride increase with respect to the other major constituents with increasing dissolved-solids content. At the average dissolved-solids content of 120 ppm, sodium and chloride comprise about 37 and 7 percent of the cations and anions, respectively, on the basis of equivalents per million (hereafter abbreviated “epm-percent”). Comparable percentages for Lake Abert, in contrast, are 98 and 56. Two representative chemical analyses of the Chewaucan River near Valley Falls are presented in table 4 (analyses 2 and 3). They show the chemical character at low and high dissolved-solids content (79 and 193 ppm).
Silica, the only abundant undissociated (nonionic) component of the streamflow, is generally present to the extent of about 30 ppm, regardless of dissolved-solids content; hence, it constitutes about 30 percent of the total salts at 100 ppm, but only 12 percent at 250 ppm. (Actually, most dissolved silicon is present as undissociated monomeric silicic acid (H4Si04) under normal hydrochemical conditions. By convention, however, the constituent is reported as silica (Si02).)
The present-day chemical character of the stream water changes appreciably as it passes through the Chewaucan Marsh. Table 5 shows estimated solute tonnages gained and lost within the marsh during the 1961 water year (about 20 percent of normal discharge at the river mouth), the 1965 water year (370 percent of normal discharge), and during a hypothetical average year. The data show that in dry years, such as 1961, incoming solutes—especially silica—are stranded in the marsh. Contrastingly, in years of greater than average runoff (such as 1965), more salts areSOLUTE BALANCE AT ABERT AND SUMMER LAKES, OREGON
C13
Figure 6.—Estimated discharge and dissolved-solids content of Chewaucan River near Valley Falls, water years 1961-65. Numbers above dissolved-solids curve indicate chloride concentration, in parts per million. Dots in discharge graph indicate monthly average streamflow of Chewaucan River measured near Paisley.
carried out of the marsh than are brought in (the silica increase is very small, however). In an average year, about a third of the silica is lost, but other components undergo moderate to substantial enrichment; as a result, the total incoming and outgoing tonnages are about the same. A significant point with regard to the possible solute contribution of agricultural activity is that about 90 percent of the chloride delivered to Lake Abert by the Chewaucan River is accrued within the marsh. Part of the increment doubtless was a component of (1) windblown alkali dust and (2) salts that were stranded during the contraction of pluvial Lake Chewaucan. However, an undetermined amount of chloride—perhaps as much as 5 percent or more of the total load—has been introduced by man in the
form of salt licks for the large cattle herd that grazes in the marsh each year. Thus, much of the chloride in the river at its mouth has been recycled, and other increments are attributable to the activities of man. As a result, the net amount of incoming chloride that is naturally “new” to the lake cannot be determined quantitatively. Similar statements apply to the other major constituents of Lake Abert as well.
Table 6 summarizes the total solute contribution to Lake Abert by the Chewaucan River, including the constituents that dominate in the lake (sodium plus equivalent chloride and carbonate-bicarbonate; see footnote 3 in table 6). The net effect of agricultural activity in the marsh on the chemical character of the river at its mouth, relativeTable 4.—Chemical analyses of representative surface and ground waters in the Abert and Summer Lake basins
[Analytical results are in parts per million, except pH]
Discharge 3 E £
No. Location1 Collection date or lake 2 £ _3 T3 C 2
level2 Water tempe: <°F) Silica (Si02) Calciu (Ca) c _ be M sS Sodiur (Na) C s> a, is .as ca C- Carboi (CO,) 0 ^ Dissol’ solids (calc.)3 xi ac
Lake Abert basin
1. 2. 3. 4. 5. Lake Abert Chewaucan River at mouth. do Seep N. of Lake Abert Spring SW. of Lake Abert. 34/21-24dca 35/21-2lcdb 35/21-21cdb 32/21-35bac 35/21-21bba .... Jan. 12, 1962 .... May 15. 1969 .... June 12, 1962 .... Aug. 12, 1963 .... June 9, 1962 4,251.81 800e 30e <0.0002e .33e 33 58 69 59 66 144 32 26 41 65 1.4 6.0 20 7.0 9.5 0.6 2.4 5.3 3.0 7.5 22,000 8.3 36 102 290 752 5,930 2.3 52 5.5 162 13 216 14 282 9,460 0 0 4 0 990 .4 12 26 47 19,300 1.8 7.0 40 295 <55,900 79 193 346 867 9.6 7.6 7.6 8.4 8.1
Summer Lake basin
6. Summer Lake 32 l6-2abd .... Apr. 25, 1961 4,146.32 68 110 2.5 .3 2,830 115 1,880 1,230 348 1,600 7,200 9.6
7. Ana River 3(1 17 tidd .... Apr. 25, 1961 92 58 36 5.0 2.3 39 3.6 91 9 5.8 12 158 8.8
•See p. C4 for description of location system.
2Discharge in cubic feet per second (ftVs); estimated values indicated by “e. ” Lake level in feet above mean sea level.
includes bicarbonate multiplied by 0.492. (See text footnote, p. C12.)
includes minor constituents not shown in tabulation. Density 1.045 g/cm5 at 20°C.
Table 5.—Changes in dissolved-solids tonnage during passage of Chewaucan River through Chewaucan Marsh in in an average year1 [All solute quantities are given in tons] water years 1961 and 1965 and
1961 water year 1965 water year Average year2
Upstream from marsh Downstream from marsh Percentage lost or gained Upstream from marsh Downstream from marsh Percentage lost or gained Upstream from marsh Downstream from marsh Percentage lost or gained
Discharge (acre-ft) 68,000 12,000 -82 212,000 230,000 +9 96,000 62,000 -35
Percentage of average annual discharge 71 19 210 370 100 100
Total dissolved solids (rounded) 6,000 2,700 -55 19,000 36,000 +47 8,500 10,000 + 15
Silica 2,800 490 -82 8,600 9,500 +9 3,900 2,500 -36
Calcium plus magnesium 740 360 -51 2,300 5,000 +54 1,000 1,500 +33
Sodium 440 440 0 1,400 4,400 +68 610 1,400 +57
Chloride 20 120 +83 60 790 +92 30 250 +88
Other components 2,000 1,300 -35 6,500 16,500 +61 2,900 4,600 +37
•Stations upstream and downstream from the marsh are near Paisley and Valley Falls, respectively (pi. 1).
2Based on runoff data measured or estimated for the period 1916-65, and chemical-quality data collected during the period 1959-65.
C14 CLOSED-BASIN INVESTIGATIONSTable 6.—Solute contributions to Abert and Summer Lakes and their peripheral areas by streams, springs, and seeps
[All values rounded to two significant figures or less]
Source
Acre-feet Silica (Si02) Calcium (Ca) plus magnesium (Mg) Sodium (Na) Chloride (Cl) Other constituents' 2 Total (rounded) Sodium plus equivalent chloride and bicarbonate2 s
per year Tons Tons Tons Tons Tons Tons Tons
ppm per ppm per ppm per ppm per ppm per ppm per ppm per
year year year year year year year
Lake Abert
Chewaucan River 62,000 30 2,500 17 1,500 16 1,400 3 250 54 4,600 120 10,000 38 3,200
Peripheral streams 2,500 20 70 10 30 15 50 2 7 35 120 80 270 35 120
Peripheral springs and seeps 5,000 45 310 10 70 140 950 85 580 170 1,200 450 3,100 340 2,300
Total 70,000 2,900 1,600 2,400 840 5,900 13,000 5,600
Summer Lake
Ana Springs:
(pre-1900) 100,000 37 5,000 8 1,100 39 5,300 13 1,800 63 8,600 160 22,000 92 13,000
(post-1926) 67,000 37 3,400 8 700 39 3,600 13 1,200 63 5,700 160 15,000 92 8,400
Peripheral springs and seeps:
(pre-1900) 8,000 40 440 10 110 60 650 15 160 90 980 220 2,400 140 1,500
(post-1926) 13,000 40 700 10 200 130 2,300 30 500 190 3,500 400 7,100 300 5,300
Peripheral streams 10,000 25 340 15 200 20 270 2 30 60 820 120 1,600 50 680
Pre-1900 total Post-1926 total 120,000 90,000 5,800 4,400 1,400 1,100 6,200 6,200 2,000 1,700 10,000 10,000 26,000 24,000 15.000 14.000
'Mostly bicarbonate.
2Bicarbonate is multiplied by 0.492. (See text footnote, p. G12.)
sDominant solutes in Abert and Summer Lakes. Concentrations include all sodium and chloride, plus an amount of bicarbonate equivalent to the “residual” sodium (that is, sodium in excess of chloride).
SOLUTE BALANCE AT ABERT AND SUMMER LAKES, OREGON C15C16
CLOSED-BASIN INVESTIGATIONS
z
o
TOTAL CATION OR ANION EQUIVALENTS PER MILLION
IT
UJ
Q.
Figure 7.—Changes in the relative amounts of major ions in Chewau-can River near Valley Falls with increasing dissolved-solids content. Solid lines for anions; dashed lines for cations. Chemical symbols as follows: HCOs, bicarbonate; Na, sodium; Ca, calcium; Mg, magnesium; Cl, chloride; S04, sulfate; K, potassium.
to natural conditions prior to settlement of the area, is unknown. The overall differences are assumed to be slight, however, and the estimates in table 6 probably characterize the river’s contribution before as well as after man’s development of the marsh, even though the estimates are based on postdevelopment data.
The amounts and chemical character of solutes contributed by small streams adjacent to Lake Abert are uncertain. Poison Creek, the only near-perennial stream aside from the Chewaucan River, carries the runoff from a 3-square-mile basin on Abert Rim. During periods of low flow, the stream characteristically contains about 75 ppm of dissolved solids. Overland flow and runoff in normally dry stream channels occasionally reaches Lake Abert from the arid peripheral areas to the north and west following summer thunderstorms and other heavy rains. Although the quantities of such flow can be appreciable during short periods of time, the contribution, in terms of average annual water volume, is small. Likewise, the solute contribution is small, and much of it represents increments derived from peripheral springflow that evaporated before reaching the lake, along with recycled salts, such as (1) components of windblown alkali dust, (2) salts isolated
from the pluvial lake during contraction, and (3) dissolved constituents of lake water previously blown out upon the flat northern playa and isolated above present lake levels. Excluding the solutes of peripheral springflow (which are accounted for below), the average dissolved-solids content of all inflow from small streams adjacent to the lake is assumed to be about 80 ppm. The estimated solute contribution is listed in table 6.
Ground-water contributions outside the Chewaucan River drainage are limited to seeps and springs adjacent to the lake. The chemical character of this flow is wide in range, as shown by the dissolved-solids values and chloride.sodium ratios in figure 8. (See representative analyses 4 and 5 in table 4; samples were collected from sites 3 and 9, fig. 8.) The figure also shows that, with one exception, the greatest solute content is characteristic of flow emerging closest to the lake. (The exception is seepage at site 8, which represents emergent underflow from the alluvial fan of Poison Creek and is fed in part by dilute percolating surface water.)
Although the peripheral ground-water flow actually reaching the lake may be only 2,000 to 2,500 acre-feet per year (p. C7), the total estimated spring and seep discharge, about 5,000 acre-feet per year, must be considered as a source of solutes. Table 6 summarizes the estimated solute contribution to the lake by peripheral ground water.
The net contribution of air-transported solutes at Lake Abert is the most difficult of the several elements of salt gain and loss to evaluate properly. The two media of transport are atmospheric moisture (ultimately precipitation) and windblown dust. Whereas precipitation is a salt contributor only, the transport of dust both contributes salts to and removes salts from the basin. The impression gained from observations at Summer Lake is that large salt tonnages are carried from the playa surface, especially during periods of low lake level. (For example, see Phillips and Van Denburgh, 1971, fig. 23.) Alkali dust clouds have been observed moving east from Summer Lake toward Lake Abert (R. W. Childreth, U.S. Geological Survey, oral commun., 1963) and moving both south and north from Summer Lake to areas outside its basin (Boyd Claggett, Oregon State Department of Game, oral commun., 1966). The actual quantities of salts entering and leaving the two basins, however, may be less impressive in magnitude than the billowing dust clouds are in appearance. Regardless of magnitude, though, the net result almost certainly is a loss of salts from the two tributary basins.
The chemical character of the dissolvable alkali dust from the peripheral salt flats has not been determined directly. However, it can be inferred on the basis of the character of efflorescent top salts that coat the dry playa surfaces at times. Dissolvable parts of the surface deposits adjacent to Abert and Summer Lakes are chiefly sodium carbonate and bicarbonate, as shown in figure 9. In fact,SOLUTE BALANCE AT ABERT AND SUMMER LAKES, OREGON
C17
120° 15'
0 5 KILOMETRES
Dissolved-solids
Sile Location Collection content Chloride: sodium
date (ppm)1 ratio*
1 32/21-26bbc......... 6- 9-62 285e 0.21e
2 -33dcb............ 9-28-68 355 .36
3 -35bac......... 8-12-63 346 .25
4 33/21-1 ldbb........ 8-14-63 550e .29
5 33/22- 16bdc........ 6- 9-62 475 .42
6 -20aab......... 6- 9-62 580e ,44e
7 34/21 -6da.......... 6- 9-62 800 .62
8 35/21-1 bac.......... 7-26-64 156 .87
9 -21bba......... 6- 9-62 867 .66
l"e" indicates estimated value. Includes bicarbonate multiplied by 0.492. (See text footnote, p. C12.)
zBased on equivalents per million; “e" indicates estimated value. Ratio for lake, 0.57.
Figure 8.—Source and chemical character of sampled spring and seep flow peripheral to Lake Abert.
LAKE ABERT
SUMMER LAKE
Carbonate
and
bicarbonate
Sulfate
Chloride
Playa Lake
salts water
Playa Lake
salts water
---
Figure 9.—Relative abundance of major anions in playa-surface salts adjacent to Abert and Summer Lakes, compared with the chemical character of the lake waters themselves. Relative abundances are based on equivalents per million (each bar represents 100 percent of the major anions). Playa data are averages for the following samples. Lake Abert: locations 33/22-9ccc (Sept.
4, 1944); 33/22-9db (May 20, 1959); 33/22-9cc (Sept. 17, 1962); 33/22-7ba (July 23, 1964); 33/21-9acd (Sept. 27, 1968); and 33/21-lObcc (Sept. 28, 1968). Summer Lake: locations 32/16-14, 32/17-26, 32/17-3 (all Aug. 30, 1944), 31/17-10bab (Sept. 30, 1968); and 32/16-2aad (Oct. 1, 1968). Analytical results for samples collected in 1944 are from Allison and Mason (1947, p. 3, 7). Sam-pl ing-site locations and other information were obtained from Ralph S. Mason (Oregon Department of Geology and Mineral Industries, oral commun., 1964).
the carbon species make up a far greater percentage of total salts in the surface deposits than in the adjacent lake water. Consequently, the net outgoing load of airborne salts may be removing carbonate-bicarbonate from the playa surface in preference to chloride.
The quantitative role of precipitation as a contributor of solutes is uncertain. In qualitative terms, however, the role is two-fold—incoming moisture brings soluble salts from outside the basins, and it also flushes local salts from the atmosphere above the basins. Therefore, the solute contributions of rain and snow are in part new to the basins and in part recycled, although the relative importance of the two increments is unknown.
Records of precipitation chemistry are meager. The five available analyses, for samples collected at or near the Valley Falls weather station, are tabulated on page C18. Considering the semiarid closed-lake environment, the samples are distinctive in their dilute character and near absence of sodium or chloride, and they support Gambell’s contention (1962, p. 94) that “excluding the immediate coastal areas, marine aerosols appear to constitute a small portion of the soluble material brought down in precipitation over the United States.” Regardless, the tonnage of solutes contributed directly to Lake Abert by lake-surface precipitation is insignificant in comparison with increments from other sources.C18
CLOSED-BASIN INVESTIGATIONS
[Results in parts per million, except as noted]
Date l .................... *1-12-62 5-1-64 6-9-64 7-30-64 to 11-12-64 to
8-1-64 12-11-64
Precipitation amount (inches
of water)................. 0.06 0.15 1.10 0.59 0.61
Type of precipitation........ Snow Rain Rain Rain Rain and snow
Sodium (Na).................. 1.2 3.8 0.8 1.1 1.2
Bicarbonate* * * * 5 (HCOs). 3 10 10 13 21
Sulfate (SO«)...................0 5.6 .0 2.0 3.0
Chloride (Cl)................ 1.0 .5 .0 .0 .0
Specific conductance (micromhos per cm
at 25°C).................... 9 35 23 37 52
Estimated dissolved
solids...................... 5 20 15 20 30
'All except first sample were collected at weather station and stored in closed poly-
ethylene or glass bottles as long as 1-2 months before analysis.
2Collected at east shore at Lake Abert, 10 miles northeast of Valley Falls. May have
been inadvertently contaminated by lake water during collection and melting.
5Samples contained no detectable carbonate when analyzed.
SUMMER LAKE
Among the natural sources of inflow to Summer Lake, the most plentiful contribution of solutes is provided by Ana Springs, via the Ana River. The spring flow contains about 160 ppm of dissolved solids—mostly silica, sodium, and bicarbonate (analysis 7, table 4). Sodium and chloride make up 76 and 15 epm-percent of the cations and anions, respectively, in the spring discharge, compared with 97 and 36 epm-percent in the lake.
Despite present-day depletion of the spring flow downstream from Ana Reservoir, the discharge at the reservoir should be considered in calculations of solute income because it gives a true indication of the solute contribution of the springs. Table 6 lists the present-day contribution, as well as that estimated for the more nearly natural conditions prior to 1900.
Additional natural ground-water flow in the Summer Lake basin includes springs tributary to the Ana River below the stream gage, seeps along the west and south shores, springs northeast of the lake, and Summer Lake Hot Spring to the southeast. Only the nearby seep and spring flows actually reach the lake, but all other ground-water sources contribute solutes that can be transported to the lake in surface runoff or dissolved when the lake is high. On the basis of scanty data regarding the chemical character of peripheral spring flow, the following estimates are made:
Average dissolved solids (ppm)
Spring inflow to Ana River below gage:
Pre-1900 (5,000 acre-ft per yr) .......................200
Post-1926 (10,000 acre-ft per yr) .....................450
Other peripheral springs and seeps (3,000 acre-ft per yr) .250
Combining these data, along with estimates for specific chemical constituents, the total solute contribution of peripheral springs and seeps is listed in table 6.
Water wells provide an additional source of solutes within the Summer Lake basin. However, their contribution has been restricted to the last few decades and
probably has not had more than a slight effect on the chemical character of the lake.
The percentage of total salts that has been recycled rather than newly introduced to the lake is unknown. Some of the increment emerging as springflow immediately adjacent to the western lakeshore may be derived from salts associated with the lake and its periphery; likewise, at least a small part of the salts in other locally derived ground-water discharge must originally have been a component of windblown alkali dust that was dissolved by percolating runoff. However, the Ana Springs system, which dominates the inflow to Summer Lake, is thought to be fed by ground water that originates outside the basin (Phillips and Van Denburgh, 1971, p. B39), and most of its solute contribution could therefore be considered new to the lake rather than recycled.
The solute contribution of small peripheral streams is as uncertain at Summer Lake as at Lake Abert. The dis-solved-solids content of drainage from the Winter Ridge-Slide Mountain area may be 100-140 ppm during near-low-flow conditions, on the basis of scanty data. Wet-season flow from that area doubtless is more dilute, but the occasional thundershower runoff from flat salty areas east of the lake may be more concentrated. For lack of definitive information, the average solute content of all peripheral streamflow (exclusive of Ana River) upstream from the influence of agriculture is assumed to be on the order of 120 ppm. Table 6 summarizes the contribution of specific solutes.
The contribution and depletion of solutes by atmospheric means is discussed on pages C16-C17.
In about 1916 commercial salt recovery by solar evaporation began after about 1,000 acres at the southeast end of Summer Lake had been dyked off (Allison and Mason, 1947, p. 1). The short-lived operation made a small profit but ceased when the salt-processing mill burned (Boyd Claggett, Summer Lake State Game Management Area, oral commun., 1966). According to R. S. Mason (Oregon Department of Geology and Mineral Industries, written commun., 1971), the total tonnage of salts removed was small.
SUMMARY OF SOLUTE GAIN AND LOSS
The estimates of total solutes delivered to Abert and Summer Lakes and their immediate peripheries by streams, springs, and seeps during an average year are summarized in table 6. These sources contribute about 13,000 tons to Lake Abert in an average year, but less than half of that amount (about 6,000 tons) consists of constituents that dominate in the lake (sodium plus equivalent chloride and carbonate-bicarbonate). Most of the remainder includes such components as silica, calcium, and magnesium that are permanently removed from the solute cycle by organic processes and inorganic reactions.SOLUTE BALANCE AT ABERT AND SUMMER LAKES, OREGON
C19
Solutes contributed to Summer Lake have averaged about 25,000 tons per year, both before 1900 and since 1926. Sodium plus equivalent chloride and carbonate-bicarbonate total almost 60 percent of that quantity (about 15,000 tons). Thus, the annual contribution of these constituents to Summer Lake is about 2'A times the contribution to Lake Abert. The overall chloride:sodium ratio for inflow to both lakes has been about 0.2.
In addition to surface- and ground-water increments, salts also are contributed to the lakes and their peripheral playas as components of precipitation and windblown dust. However, the tonnages added in this manner are small compared with those delivered by surface and ground water, and most of the contribution probably represents recycled salts.
The only effective means of solute removal from the lakes and their peripheries is wind transport of powdery alkali dust from the dry playa surfaces. The quantity of salt removed by this means is unknown, but it presumably exceeds the amount brought in as a component of precipitation and dust.
Ground water in some parts of the Abert and Summer Lake basins may be subject to interbasin movement similar to that suggested for certain topographically enclosed basins in Nevada (for example, Eakin, 1966; Van Denburgh and Glancy, 1970, p. 13). In fact, most of the discharge of Ana Springs may originate in adjacent areas to the north, rather than in areas topographically tributary to Summer Lake. However, no salts are thought to be removed from the lakes by such mechanisms, because of the presence of peripheral springs at Abert and Summer Lakes and because of the low altitude of the two basin floors (they are the lowest within at least 50 miles).
STORED SOLUTES
Most of the annual increment of mobile solutes brought to each lake and its periphery by streams and ground water remains and accumulates within the lake and its immediate surroundings. The incoming solutes are stored in three environments—as dissolved constituents of the lake water itself, as interstitial components of the peripheral and lake-bottom sediments, and, part of the time, as thin salt deposits atop the peripheral playas. The first environment is easy to evaluate both quantitatively and qualitatively because the shallow lakes remain almost homogeneous except during periods of abundant inflow. In contrast, the peripheral and lake-bottom salts are not homogeneous, either areally or vertically, and an accurate evaluation of them is much more difficult.
SOLUTES DISSOLVED IN THE LAKES
Although both lakes are chemically similar, Lake Abert has by far the largest tonnage of dissolved solids. As of 1969 the lake contained almost 15 million tons of solutes, consisting mostly of sodium, carbonate, bicarbonate, and
chloride. In fact, those four constituents account for more than 96 percent of the total tonnage (table 4, analysis 1). Potassium and sulfate are of secondary importance; they make up an additional 3 percent of the salts. Several minor constituents, including silica, bromide, orthophosphate, and boron, are present in appreciable parts-per-million amounts, though together they make up less than 1 percent of the total tonnage (Phillips and Van Denburgh, 1971, p. B17). The most noticeable chemical characteristic of Abert and Summer Lakes is the near absence of calcium and magnesium. The two alkaline-earth ions are among the major constituents of most inflow, yet their combined concentrations in the lakes are less than 10 ppm. The dearth is ascribed to mineral reactions—most of the calcium apparently is tied up organically and inorganically as CaCOs, whereas magnesium is largely incorporated within the silicate fraction of lake-bottom sediments (Jones and Van Denburgh, 1966, p. 446).
The relative percentages of major constituents in the lake have not exhibited an appreciable net change since 1912 (fig. 10), on the basis of a comparison of recent analyses with one reported by Van Winkle (1914, p. 119).
The measured dissolved-solids content of Lake Abert has ranged from 18,700 to 95,000 ppm. Concentrations were much greater at near-dryness, but the low value probably represents almost the minimum overall concentration attained within the last century, and it occurred at the time of maximum observed lake extent, in mid-1958 (fig. 4). The broad shallow lake remains nearly homogenous throughout, except during periods of appreciable inflow, because of continual mixing by the wind.
The solute tonnage in Lake Abert has varied considerably during the 20th century, as indicated by the few reliable concurrent measurements or close estimates of dis-
LAKE ABERT
SUMMER LAKE
Carbonate
and
bicarbonate
Sulfate
Chloride
Feb. Apr.
Feb. Apr.
1912 1961
1912 1961
Figure 10.—Relative abundance of major anions in Abert and Summer Lakes in 1912 and 1961, on the basis of equivalents per million (each bar represents 100 percent of the major anions).C20
CLOSED-BASIN INVESTIGATIONS
solved-solids content and lake level listed in table 7. (The tonnage for February 1912 is only approximate, because the estimated lake level is based on a measurement on October 4, 1915; the altitude for 1915 has been extrapolated to 1912 using estimated inflow and net evaporation during the intervening 3.6-year period. Furthermore, the 1912 sample was collected near the south end of the lake and may not have represented a lakewide average because of dilution by stream or seep inflow; this would make the 14-million-ton value too low.)
Table 7.—Dissolved-solids concentrations and tonnages in Lake Abert at various times since 1911
Date Lake-surface altitude (ft above 4,200 ft) Dissolved solids
Parts per million' Millions of tons2
1912, Feb.5 56.5 29,600 14±
1939, July 21 48.2 95,000 8.2
1952, Aug 52.7 35,000 9.7
1955, Aug. 12 52.8 40,400 11.5
1958, July 9 60.5 18,700 12.8
1959, May 20 58.9 21,700 13.0
1961, Sept. 7 52.04 55,400 13.8
1962, Sept. 17 50.70 71,000 13.5
1963, Aug. 7 54.45 37,500 13.6
1964, July 23 53.94 40,000 13.7
1966, Aug. 4 56.2 31,400 14.4
1969, June 17 55.46 35,400 14.8
'Includes bicarbonate multiplied by 0.492. (See text footnote, p. C12.)
Tonnage = parts per million X density X volume, in acre-ft X 0.001360. Densities, in grams per millilitre at 20°C, are 1.025; 1.077; 1.027; 1.032; 1.014; 1.014; 1.044; 1.058; 1.026; 1.035; 1.026; and 1.029.
Take level is estimated and may be inaccurate by 1 foot or more; thus, the tonnage may also be inaccurate, by 2 million tons or more.
The chemical character of Summer Lake is similar in general to that of Lake Abert (table 4). However, chloride is considerably less abundant relative to carbonate-bicarbonate and sulfate in Summer Lake (fig. 10).
The distribution of major negative ions in Summer Lake underwent a significant net change between 1912 and the 1960’s—the amounts of chloride and sulfate increased relative to carbonate-bicarbonate, as shown in figure 10.
The dissolved-solids content of Summer Lake, in tons and in parts per million, varies even more than that of Lake Abert on a seasonal and long-term basis. Measured dissolved-solids contents have ranged from 1,800 to 81,800 ppm. Tonnage estimates for Summer Lake are based on relations similar to those for Lake Abert but are less accurate because of a less certain area-volume relation and a greater tendency for uneven distribution of solutes within the lake during some of each year. The most reliable approximate tonnages are listed in table 8.
Langbein (1961, p. 9-10) described the general pattern of variations in the dissolved-solute content of closed lakes. The amount of salts in solution at any particular moment depends on the lake-level history prior to that time and on
Table 8.—Dissolved-solids concentrations, tonnages, and chloride: sodium ratios in Summer Lake at various times since 1911
Date Lake-surface altitude (ft above 4,100 ft) Dissolved solids' Parts per Millions million2 of tons* Chloride- sodium ratio4
1912, Feb.5 51.5 16,800 5± 0.29
1944, Sept. 1 46.2 27,000 1.0 .32
1952, Aug 46.8 8,400 .5
1959, Jan. 30 48.2 8,400 1.0 .37
1961, June 21 45.86 15,000 .43 .37
1962, June 12 46.43 6,270 .30 .36
1963, Aug. 20 47.23 6,300 .49 .36
1964, July 21 47.15 9,200 .69 .35
1965, Sept. 9 48.5 7,200 .96 .35
1966, Aug. 6 47.42 12,500 1.08 .34
1967, Sept. 12 46.92 18,600 1.24 .35
1968, July 27 45.9 45,400 1.4 .33
1969, Sept. 6 45.26 51,500 .63 .33
'Samples collected during the period October-April may indicate tonnages lower than the lake wide average because of nonuniform dilution by inflow.
includes bicarbonate multiplied by 0.492. (See text footnote, p. C12.)
Tonnage = parts per million X density X volume, in acre-ft x 0.001360. Densities, in grams per millilitre at 20°C, are 1912 sample, 1.015; 1944, 1.020; 1952, 1.005; 1959, 1.005; 1961, 1.011; 1964, 1.009; 1966, 1.010; 1967, 1.014; 1968, 1.042; and 1969, 1.049.
4 Based on equivalents per million. Trend of net decreasing chloride.sodium ratio during 1959-69 is apparent rather than real, as indicated by additional data not listed in this table.
5Lake level is estimated on the basis of a measurement made Nov. 10, 1912 (4,151.3 ft) and, therefore, could be inaccurate by as much as 1 foot. Tonnage could likewise be inaccurate.
the level at that moment compared with the levels at dryness and at an unusually high stage. In this regard, a close relation exists between solutes in the lake and those stranded on flat peripheral playas, where such playas exist. At an unusually high lake level, most of the adjacent mudflats are inundated, and almost all of available salts are in solution. Hence, the dissolved-solids concentration within the lake may be relatively small, but the tonnage is large. In contrast, at near-dryness, the tonnage is small, but the concentration is large; at such a time most of the solutes are stored within the peripheral and lake-bottom sediments or are precipitated on the lake bottom.
Several of the general trends in solute content discussed by Langbein (1961) can be evaluated using observations at Abert and Summer Lakes. For example, Langbein (1961, p. 9) postulated a net loss of salts from solution throughout a period of generally receding lake stage (segment A-B of Langbein’s schematic fig. 8, which is shown in fig. 14 of the present report). Data for Lake Abert during 1958-62 suggest that such a trend may exist but not at high to medium lake levels. In June 1958 the lake reached the highest stage of modern times (16'A ft maximum depth). During the 4 subsequent dry years ending in September 1962, the level receded almost 10 feet, and the volume decreased from about 500,000 to about 120,000 acre-feet (fig. 11). Although solute concentrations during the 51-month interval increased from 18,000 to almost 80,000 ppm, the changes in tonnage were slight (table 7). In fact, the tonnage apparently increased from 12.8 million to 13.8 million between 1958 and 1961, even though the lake levelVOLUME, IN THOUSANDS OF ACRE-FEET VOLUME, IN THOUSANDS OF ACRE-FEET
SOLUTE BALANCE AT ABERT AND SUMMER LAKES, OREGON
C21
400
300
B. Relation of tonnage to volume and area
I
1958-69
64
60
50
4 6 8 10 12 14
DISSOLVED-SOLIDS LOAD, IN MILLIONS OF TONS
16
Figure 11.—Relation of dissolved-solids tonnage to fluctuations of volume and surface area of Lake Abert, 1937-69. Dots indicate chemical analyses used to compute tonnage (table 7). Yearly extremes of volume and area prior to 1951 estimated by K. N. Phillips (U.S. Geological Survey, written commun., 1964).
AREA, IN SQUARE MILES AREA, IN SQUARE MILESC22
CLOSED-BASIN INVESTIGATIONS
Figure 12.—Variations in lake-surlace altitude and dissolved-solids tonnage at Summer Lake, 1959-69. Numbers indicate dissolved-solids load, in
millions of tons.
declined 8'/i feet. Only during the final year of recession, in 1962, did a net depletion apparently occur, and even that loss was a mere 2 percent of the total tonnage (which may well be within the limits of error of the computations themselves). At Summer Lake, the period 1965-68 was one of generally receding lake level, yet the solute tonnage increased, rather than decreased, each year (fig. 12, table 8). The diminished tonnage in 1969 indicates that depletion occurs only at near-dryness (below about 4,146 ft).
On the basis of observations at Abert and Summer Lakes, a major loss of dissolved solids during the reduction in volume of a broad, shallow closed lake occurs for two reasons only—precipitation of evaporite salts and isolation of brine from the main body. In closed lakes, regardless of their chemical composition, no major loss of salts through mineral precipitation can occur in the dissolved-solids range from less than 10,000 ppm to about
100.000 ppm. When the increasing solute content reaches
10.000 ppm, virtually all the calcium carbonate or calcium sulfate or both have already been depleted, and no additional solubility limits are exceeded below about
100.000 ppm. Actually, depending on chemical character, the range can be even greater. At Lake Abert, for example, the “no precipitation” range may be from about 1,000 ppm to about 200,000 ppm or more, becauseof the low sulfate and high carbonate concentrations. Under present conditions (1970) precipitation of sodium carbonate and bicarbonate minerals, followed by sodium chloride, would not begin at Lake Abert until the volume had decreased to 35,000 acre-feet or less, which is equivalent to a lake stage about 3 feet lower than that attained in September 1962. In summary, mineral precipitation as a mechanism of dissolved-solids depletion is applicable only under certain well-defined hydrochemical conditions.
Langbein (1961, p. 9) suggested that during a period of low lake level, the dissolved-solute tonnage may undergo a significant net increase, as shown by segment B-C of his schematic figure 8. However, such a net increase would not be expected in broad, shallow water bodies that dry
every few years, because of solute depletion during the periods of low level. Under these conditions, some aspects of the solute cycle resemble those of an intermittent playa lake (Langbein, 1961, figs. 9B, C). At Summer Lake, for example, the solute tonnage was about 400,000 in June 1961, after a short period of dryness or near-dryness during the previous fall. The following summer, after a second brief dryness or near-dryness, the lake contained an even smaller load, about 300,000 tons, despite a seasonal lake-level maximum almost 1 foot higher than that of the previous year (fig. 12). Not until the lake had attained an even higher level following a third desiccation in September 1962 did the solute tonnage finally exceed that of 1961.
Wind can play an important role in the depletion of dissolved salts. When a broad shallow lake covers only part of its flat playa, wind can blow the residual water body out over peripheral mudflats. This mechanism, which has been observed at Summer Lake, isolates some of the brine as much as several feet above the lake level under calm conditions. Such action increases the amount of evaporation by increasing the temperature and surface area of lake water exposed to the atmosphere, and it also leaves the residual lake body depleted in solutes.
During a protracted period of desiccation and near-desiccation, wind also can cause a permanent removal of solutes from the basin floor (p. C16). Such a mechanism may have been an important factor in the major depletion that occurred at Goose Lake, 30 miles south of Lake Abert, in the drought-stricken 1930’s (Phillips and Van Denburgh, 1971, p. B34). Similarly, Summer Lake contained 5 million or more tons of solutes in 1912, before man’s activities began diminishing the quantity of inflow, yet the dissolved load is not known to have exceeded about 2 million tons since the onset of frequent desiccation following 1926. Even if the lake were to rise again to, and remain at, the high pre-1926 levels, the amount of salt regained from the lake-bottom sediments alone probably would fall short of the 5 million tons estimated for 1912 because of net depletion during the years since 1926. The effect of the salt-recovery operations between 1915 andSOLUTE BALANCE AT ABERT AND SUMMER LAKES, OREGON
C23
1920 (p. C18) on the total quantity of solutes dissolved in the lake is considered to be negligible because the tonnage removed apparently was small.
Langbein (1961, p. 9) suggested that during the third part of his long-term cycle, the dissolved-solute load should increase as the lake stage rises from a low to a high level over a period of years (limb C-D in his fig. 8). This phenomenon is illustrated clearly at Lake Abert (fig. 11). The figure shows that the lake rose about 12 feet and accumulated about 5 million tons of dissolved solids between July 1939 and July 1958. Only about 125,000 tons of lake constituents were contributed by inflow during that period. Thus, much of the salt increment was derived from the newly reinundated mudflats and doubtless represented solutes isolated from the lake during the 15-year period of desiccation or near-desiccation prior to 1938 (fig. 4). The principal mechanism of reaccumulation may be upward diffusion of solutes from the interstitial brines of lake-bottom sediment, as described by Lerman and Jones (1971).
Records at Summer Lake also show effective solute recovery during a rising stage (fig. 13). In the 6-year interval following desiccation or near-desiccation in the fall of 1962, the lake gained more than 1 million tons of solutes, of which less than 10 percent was contributed by Ana River and peripheral streams to the west and south. The greatest gains were in 1963 and 1965, the 2 years of net lake-level rise. However, the extraction of lake-bed salts continued at a relatively constant annual rate, even during the years of net falling level (fig. 13B).
The final stage in Langbein’s schematic salt cycle (1961, p. 9), “a stable high-level phase of a lake during which annual input of salts exceeds the losses * * can only be inferred at Abert and Summer Lakes, on the basis of a continual inflow of solutes, as well as the slow accrual of salts from the lake bottom and immediate periphery. Even the short period of generally moderate levels at Lake Abert during 1964-69 produced a net solute enrichment (fig. 1 IB).
The sequence of changes in dissolved-solute tonnage at shallow Abert and Summer Lakes that would accompany the idealized long-term fluctuation of lake volume used by Langbein (1961, fig. 8) can be compared with the saltcycle he proposed. Given Langbein’s pattern of lake-volume changes (top curve in fig. 14), the variations in dissolved-solids load observed or postulated at Abert and Summer Lakes are depicted in cycle 2 of figure 14. For comparison, the hypothetical pattern of variations suggested by Langbein (1961, p. 9) for closed lakes in general is shown as cycle 1. The greatest contrasts between the two cycles occur in segments A-B and B-C. Data in the present report indicate that a major loss of solutes occurs in certain lakes during only the final stages of a receding phase, and that the losses can continue during the subsequent period of
D1SSOLV E D-SOLI DS LOAD, IN MILLIONS OF TONS
TIME SINCE LAST DESICCATION (OCT. 1962), IN YEARS
Figure 13.—Relation of increasing dissolved-solids tonnage in Summer Lake during 1963-68 to (A) fluctuations of lake volume and surface area, and (B) time since last prior desiccation (Oct. 1962). Calculated tonnages numbered 1-6 are from table 8 (values for Aug. 20,1963, through July 27, 1968).
low level. Much of the loss is temporary, however, and salts are recovered during a return to high lake level (segment C-D in cycles 1 and 2). Additional salts are gained during any extended period of characteristically high stage (segment D-A in cycles 1 and 2), and completion of an ideal long-term cycle can even produce a net solute increase, as shown by segment A-A' in cycle 2.
The complete sequence depicted in cycle 2 (fig. 14) applies only to shallow closed lakes that occupy expansive flat playas, such as those on the floor of the Abert and Summer Lake basins. Water bodies in deeper, more nearly bowl shaped basins, such as Pyramid and Walker Lakes in Nevada (Langbein, 1961, p. 18), would have a different, more stable volume-tonnage cycle, given the same relative magnitude of long-term volume fluctuation. Likewise, inC24
CLOSED-BASIN INVESTIGATIONS
Eigure 14.—Hypothetical long-term cycles of dissolved-solute tonnage and lake volume for a closed lake, based on Langbein’s idealized sequence of lake-volume fluctuations (top graph; 1961, fig. 8). Cycle 1 in bottom graph is Langbein's schematic pattern of solute fluctuation; cycle 2 is the pattern observed or hypothesized for Aberl and Summer Lakes.
a basin where irregular lake-bottom features result in depressions and embayments that become isolated from the main water body at lower levels, the volume-tonnage cycle would be different from those at smooth-sided Abert and Summer Lakes.
Lake chemistry also influences the pattern of dissolved-solute fluctuation. For example, if a slightly saline lake at high level contains appreciable amounts of calcium and sulfate relative to sodium and carbonate-bicarbonate, the solute-tonnage variation during lake contraction will not resemble the upper part of segment A-B in cycle 2 (fig. 14) because of solute loss through precipitation of gypsum.
LAKE-BOTTOM AND PERIPHERAL SOLUTES
No sharp physical or chemical boundary exists between peripheral and lake-bottom solutes at Abert and Summer Lakes, because the areal extents of the two lakes are by no means stable. The two environments—one exposed to the atmosphere and the other covered by lake water—merge and could perhaps be considered as a single unit. However, the modes of occurrence and chemical character of the solutes, as well as the processes that affect them, are different in each environment.
LAKE ABERT
The lake and its broad northern playa are underlain by fine-grained sedimentary deposits that store large solute tonnages. Several short lake-bottom cores have been analyzed in detail. Representative results for one of the cores (table 9) indicate that the quantity of interstitial
Table 9.—Analytical data for sediment and interstitial brine at representative lake-bottom and playa sampling sites shown on plate 1
[Based on information from B. F. Jones, U.S. Geological Survey, written communs., 1966-71]
Depth interval (feet) Brine content (percent of volume) Dissolved-solids content of brine (thousands of parts per million) Chloride : sodium ratio1 Dissolved-solids content of mud (tons per acre-foot)
Lake Abert, bottom sediments, site 17
0.0-0.85 87 62 0.55 80
0.85-1.7 83 94 .50 110
1.7-2.4 285 125 .51 160
Lake Abert, playa sediments, site 23
0.0-0.003 (crust) 0.50 2300
0.003-0.1 47 173 .65 140
0.6-1.2 44 149 .61 100
1.2-1.7 53 143 .62 no
1.7-2.2 37 131 .67 70
2.2-2.8 56 125 .70 110
2.8-3.3 56 115 .72 100
3.3-3.8 76 102 .75 110
5.2-5.8 80 93 .76 110
5.8-6.3 80 95 .77 110
6.3-6.9 78 92 .78 100
Summer Lake, bottom sediments, site 5
0.5-1.0 73 43 0.34 46
1.0-1.5 69 37 .29 38
1.5-2.0 73 38 .27 38
2.0-2.5 72 40 .26 46
2.5-3.0 75 35 .30 42
3.0-3.5 74 33 .26 36
3.5-4.0 78 32 .28 36
4.0-4.5 83 33 .28 40
4.5-5.0 77 36 .28 42
5.0-5.5 78 39 .29 42
5.5-6.0 77 34 .30 40
6.0-6.5 82 34 .30 40
6.5-7.5 82 30 .30 34
Summer Lake, playa sediments, site 3
0.0-0.1 0.26
0.1-0.45 47 185 .47 140
0.45-0.8 51 163 .54 130
0.8-1.2 79 156 .54 190
1.2-2.1 66 156 .56 160
2.1-3.1 61 139 .55 130
3.2-3.5 70 137 .56 140
‘Based on equivalents pier million. Estimated.SOLUTE BALANCE AT ABERT AND SUMMER LAKES, OREGON
C25
brine is large (more than 80 percent of the sediment-brine total, by volume), and that brine salinity increases with depth. In fact, about 2 feet below lake bottom, the concentration is more than twice the long-term average salinity of the overlying lake. Other data (pi. 1) show similar characteristics. (Information in table 9 and on pi. 1 is based on analytical data provided by B. F. Jones (U.S. Geological Survey, written communs., 1966-71); chemical analyses by S. L. Rettig, U.S. Geological Survey. Shrock and Hun-zicker (1935, p. 14-17), Allison and Mason (1947, p. 1-7), and Stott (1952, p. 41-47) also have evaluated the chemical and physical characteristics of lake-bottom and peripheral deposits at the two lakes, but their information unfortunately cannot be used in appraising the stored-solute quantities because no data are available on the brine content of the sediments they sampled.)
Chemically, the brines resemble the overlying lake water, except that they are somewhat enriched in carbonate plus bicarbonate, relative to chloride (table 10). The enrichment may be due at least in part to the production of dissolved carbon species during anaerobic decay of organic material within the lake-bottom sediments (Jones and Van Denburgh, 1966, p. 444; Jones and others, 1969, p. 260).
On the playa north of Lake Abert, the environment and distribution of solutes are considerably different. The finegrained sedimentary deposits there are generally exposed to the atmosphere, rather than sealed off by overlying lake water, and entirely different chemical and physical processes operate. Immediately following inundation or “washing” of the playa surface by windblown lake water or rain, the remaining salts lie within the fine-grained deposits. During subsequent evaporation, some of the near-surface brine moves upward, producing an efflorescent salt accumulation on the playa surface that has both powdery and crusty components. Chemically, the surface salts are dominated by carbonate plus bicarbonate relative to the other anions (fig. 9). (As in the lake water, sodium is by far the most plentiful cation; it is accompanied by only
a small amount of potassium, and negligible calcium or magnesium.) The dominance of sodium carbonate and bicarbonate may result from differing solubilities. Bradley and Eugster (1969, p. 46-57) show that sodium carbonate and bicarbonate minerals precipitate much sooner than sodium chloride during evaporative concentration of brines such as the interstitial fluids at Lake Abert. The crustier components of the first-crystallized efflorescent salts may subsequently shield the underlying chloride-enriched (carbonate-bicarbonate-depleted) residual brine from appreciable evaporation. This situation is shown by the chemical character of efflorescences and shallow interstitial fluids at location 5, Summer Lake (p. C26).
In contrast to the paucity of chloride in efflorescent deposits, brines at 3- to 5-foot depth beneath the playa at Lake Abert are even richer in chloride than the lake itself and far richer than the lake-bottom brines. Table 9, which lists data for a representative site on the northern playa, shows this contrast, on the basis of chloride-to-sodium ratios. Plate 1 and table 10 also show the contrast. The apparent chloride enrichment may be due in large part to a net depletion of salts rich in sodium carbonate and bicarbonate, owing to wind transport of powdery playa-sur-face efflorescences (p. Cl 7). In fact, such a mechanism may help explain why the relative amount of chloride is far greater in, beneath, and adjacent to Lake Abert than in most waters feeding the lake, even with allowance for depletion of “immobile” solutes.
Brine salinity at 3- to 5-foot depth on the northern playa is greater than that of the lake, but less than that of the lake-bottom interstitial fluids. Brine quantities are characteristically less also.
Carbon-14 dates for sediment from the depth interval 4.0-5.0 feet at site 16 (pi. 1) suggest that the top 5 feet of sediment there has been deposited during about the last 5,000 years but that the carbonate and bicarbonate components of the interstitial brine are somewhat younger. The average age of leached sediment from the interval is 4,530±250 years (U.S. Geological Survey Radiocarbon
Table 10.—Relative abundance of major ions in representative samples of lake water and interstitial brine from lake-bottom and peripheral-playa sediments
Source Relative abundance, as a percentage of total cations and anions, on basis of equivalents per million
Sodium (Na) Potassium