97
GEOCHEMISTRY OF SPRING WATER FROM THE BLACKFOOT RESERVOIR REGION, SOUTHEASTERN IDAHO: APPLICATION TO GEOTHERMAL POTENTIAL by kny Uutsinpiller A thesis submitted to the faculty of The . University of Utah in partial fulfillment of the requirements for the degree of Master of Science in Geology Department of Geology and Geophysics The Un;versi ty of Utah December 1979 4952 ·'50'

GEOCHEMISTRY OF SPRING WATER FROM THE BLACKFOOTgeology.isu.edu/Geothermal/References/Thesis... · the Blackfoot Reservoir region of southeastern Icaho and detection ~f. a geothermal

  • Upload
    others

  • View
    2

  • Download
    0

Embed Size (px)

Citation preview

  • GEOCHEMISTRY OF SPRING WATER FROM THE BLACKFOOT

    RESERVOIR REGION, SOUTHEASTERN IDAHO:

    APPLICATION TO GEOTHERMAL POTENTIAL

    by

    kny Uutsinpiller

    A thesis submitted to the faculty of The . University of Utah in partial fulfillment of the requirements

    for the degree of

    Master of Science

    in

    Geology

    Department of Geology and Geophysics

    The Un;versi ty of Utah

    December 1979

    4952 ·'50'

  • I...1 5 1 5'

    I-J

  • THE UNIVERSITY OF UTAH GRADUATE SCHOOL

    FINAL READING APPROVAL

    To the Graduate Council of The University of Utah:

    I have read the thesis of Amy Hilts; npi lJ er In Its final form and have found that (I) its format, citations, and bibliographic style are consistent and acceptable; (2) its illustrative materials including figures, tables, and charts are in place; and (3) the final manuscript is satisfactory to the Supervisory Committee and is ready for submission to the Graduate School.

    Parry Member. Supervisory Commillee

    Approved for the Major

    7 .)

    Stanley H./ L Chairman; Dcan

    Approved for the Graduate Council

    James L. Cl ayton {?Dean of The Graduate Sehool

  • The Blackfoot Reservoir region in southeastern Idaho is

    recognized as a potential geothermal area because of the presence of

    several young rhyolite domes (50,000 years old), Quaternary basalt

    flows, and warm springs. North- to northwest-trending high-angle

    normal faults of Tertiary to Holocene age appear to be the dominant

    structural control of spring activity in the Blackfoot Reservoir

    region. Surface spring-water temperatures average 14°C except for a

    group of springs west of the Reservoir Mountains which average 33°C.

    Chemica1 geothermometers, which indicate temperatures of last

    water-rock equilibrium, applied to fifty water samples give

    temperatures less than 75°C except for eight springs along the Corral

    Creek drainage. The springs along Corral Creek have Na-K-Ca

    temperatures that average 354°C, which are a direct result of 1arge

    potassium concentrations in the water. A correction for carbon

    dioxide applied to the Na-K-Ca geothermometer lowers the estimated

    temperatures of the anomalous springs close to the measured surface

    temperatures. Mixing model calculations suggest that hot water with a

    maximum temperature of approximately 67°C may be mixing with cooler,

    more dilute water in the springs from the Corral Creek drainage.

    Stability relations of low-temperature phases in the systems

    K20-A1203-Si02-H20-C02 and Na20-A1203-Si02-H20-C02 indicate that the

    large concentrations of potassium in the eight anomalous springs are

  • t derived from equilibrium reactions with the potassium-bearing minerals

    muscovite and microcline. Other springs in the Blackfoot Reservoir

    region do not appear to obtain their sodium and potassium contents

    from equilibrium reactions with feldspars. Carbon dioxide and

    hydrogen sulfide gasses may be derived through the oxidation of

    organic matter by the reduction of sulfate. Concentrations of major

    and minor elements, and gasses found in springs of the Blackfoot

    Reservoir region are due to water-rock reactions at temperatures less

    than 100°C. Meteoric water circulating along faults may reach

    te~peratures up to 100°C within a few kilometers of the surface.

    Travertine deposited by the springs is co~posed primarily of calcite

    and aragonite, with minor amounts of gypsum, aoatite, and phosphate.

    Based on spring geochemistry. a geothermal reservoir of less than

    100°C may exist at shallow (less than 2 km) depths in th~ Blackfoot

    Reservoir region.

    v

  • t

    { { 'f

    > ,,

    I i

    TABLE OF CONTENTS

    Page

    AB STRACT •...••••...•••••••.•.•...•••••.....•••.•.•••.••••.••...••. 1 v

    LIST OF FIGURES •.•.••••••••••.••••..•.•..••••••••••••••••.•..•.••.vi i i

    LIST OF TABLES • . • •• • • • • • • • • •• • • . • • • •• • • • • • • • • • • • . • • • • • • . • • • . • • • •• x

    ACKNOHLEDGMENTS •••••••••••••••••••••.•••••.•.••••.•••••••••.•...•. xi

    INTRODUCTION ••••••.••••••••••••.••.•.••••••••••••••••••••••••..••. 1

    GEOLOG I C BACKGROUND 2

    Physiography ...•.•••.•.••.•••••..•.•.••••••••.•.•......•••... 2 Stratigraphy and Igneous Geology •....••.•••••.••.....•.•...•• 2 Structur,: ..••.••.••••••..••••••..•.•.•••••••••.••.•..•.•..... 8 Se i smi city •.••••••••••••••••••..•.....•.•.•..•...••••.•••.••• 10 Gravity and f1agnet ics .•...•••••..•.•.••.•.••••..•.•••••••.... 13 Heat Flow ••.••.•••.•••.••..••.•.••...•••..•••••.••••.•••...•. 14

    PREVIOUS SPRING WATER STUDIES •••••••••••••.••.•••••.••••.••.•.•.•. 22

    i4ETHODS ••...•.•.......••••••.•••.••••••.•..••••••••••••..••••..... 24

    Collection and Analysis 24 Water ••••••••••.••••••••.•••••••••••••••••.....••....•.• 24 Traverti ne ••••••.•.•....••••.•••••••••.•••••••••••••.••• 28

    Thermodynamic Modeling of Ion Species in Solution •••••••...•• 29 Geothermo:i1eters ••••••••••••••••••.••••••••••••••••••••••••••• 30

    Bas i c assumptions •••••••••••••••..•••••••••••••••••••••. 30 SiOz geothermometer •••••••.•••.••.•.•••••••••••••••••••• 32 ~a-K-Ca geothermometer ••••••.•••.••..•.•.••.••••.•••.••. 33 Na-K-Ca-COz geotherrnometer ••••..•.••••.•••.••••••••••.•• 34 Na-K-Ca-Mg geotherrnometer .••......••••..•.•.••••.•.••••• 35 Mixing models ••.••.•••••••.•..••.•.•.••••.....•...•••.•. 36

    WATER CiiEiJ11 STRY 38

    Chern i ca 1 Cons t itl!ents ..••..•••.••••.•.••..••••••.•.••.•.....• 38 ;'la j 0 r e1ernent s .•.••••...•.••.................•.•........ 38 Mi nor e1ernent s .•..•..•.•.......••.••••••...•••••••..•.•. 47

    Discussion of Chemical Te:'1peratures 42 5i02 and Na-K-Ca temperatures ..•..•.•..•.••.•••.•....... 43 rlixing models •••..•...•.•..••....•....•....••.•.•....... 52

    Chemical £q:.dlibria ..•.•...••...................••.••...•...• 55

    http:�.�...��...................��.��...�http:���..�...�.�..��....�....�....��.�.�http:�..�.�..�.��.���.�http:�..�..�.�.......��.������...�������..�.�http:�.����...�.��.................�.�http:��..���.����.�.��..������.�.��.�http:��.��.�������.�..��.�.�.����.....�...���.�http:��......����..�.�.����.�http:����..�.����.���.����������http:������.���.��..�.�.��.����.���.��http:�������.���.��.�.��������������������http:������.�.�....����.���������.�������������http:����������.��������.�����������������.....��....�http:��...�.�.......������.���.������.�..������������..����http:�������������.��.�����.����.��.�.�.�http:��.��.���.���.��..��.�.��...���..�����.����.���...�http:�...�����..�.�.��.�.����..�.�������http:�.������������������..�.....�.�.�..�...����.���http:��.��.������..������..�.�.���������.��.�..�.�http:�....��.�����.��.....�.�http:�.���.�.��.�����..�.�.��������.�.�......���http:������.������������.��.�.������������������������..��http:���������������������.�����.�.����.���������.�...�http:�.�.����������.����..�.�..����������������.�..�.��http:�...����...�������.�.�...�����.....���.�.���.����.��...��

  • TABLE OF CONTENTS (CONTINUED)

    Page

    GASSES •••.••.••••••••.••••••••••••••••••.••..••••.••.••••..••••..• 62

    Source of CO 2 and H2S ••••••••••••••••••••.••.••.••••••..•••.• 62 Source of Sulfate •.•••.••..•.••••••........•..•.••........... 68

    TRAVERTINE ••••••••••••••••••••••••••••••••.•...•••.••.•..•........ 70 -

    Description ••.•••••••••••••••.••...........••••.•..••.••.•... 70 Composition ••.••••••••••••.•••••..•..••.••.•••.•••.•......... 72

    HEAT FLOW, GEOTHERNS, AND WARM SPRINGS .••••••••••••••••••••••.••.• 75

    SUMMARY AND DISCUSSION ••.•••••••••.••••••••••••••••.••••••.••••.•• 78

    CONCLUSIONS •••••••••••••••••••••••••••••.••••••••••.•••••••••••••• 82

    REFERENCES 83

  • LIST OF FIGURES

    Figure 1. Sketch map of major physiographic features of the Snake River Plain and adjacent regions, adapted from King (1969} 3

    2. Index map of southeastern Idaho, adapted from Mansf i e1d (1 927 ) . . . . . . . • . . . . . . . . . . . . • . . . • . . •• • . . • • . . . . . .. 4

    3. Generalized map of seismicity in southeastern Idaho and western Wyoming based on a compilation of University of Utah epicenters, ca. 1975-1977 11

    4. Generalized geologic map of the Blackfoot Reservoir region, southeastern Idaho, modified from Dion (1974) .... 18

    4a. Simplified structural map of faults in the Blackfoot Reservoir region, southeastern Idaho, adapted from Armstrong and Cressman (1963). Overlay to figure 4 18

    4b. Location map of water and travertine samples collected during Ju1Yt 1978, in the Blackfoot Reservoir region, southeastern Idaho. Overlay to figure 4 18

    5. Spring and well numbering system used by the U.S. Geological Survey in Idaho, adapted from Mitchell (1975).27

    6. Histograms of sodium, potassium, chloride, silica, calcium, magnesium, and bicarbonate concentrations in spring and well water from the B1ackfoor Reservoir region, southeastern Idaho 43

    7. A portion of the (Ca+Mg)-Na-K triangular diagranl showing mole percentages of the major cations in spring and well water from the Blackfoot Reservoir region, southeastern Idaho 45

    8. A portion of the HC0 3-C1-S04 triangular diagram showing mole percentages of the major anions in spring and well water from the Blackfoot Reservoir region, southeastern Idaho ......•................................ 46

    9. Mixing models for springs in the Corral Creek drainage, Blackfoot Reservoir region, southeastern Idaho 53

    http:Idaho......�http:1927).......�............�...�..���..��

  • LIST OF FIGURES (CONTINUED)

    Figure 10. Stability of phases in the system K20-A1 203-Si0 2H20-C0 2 at 25°C, Ptotal= 1 atm, = 0.26 atm 58PC02

    11. Stability of phases in the system Na20-A1203-SiOzH20-C02 at 25°C, Ptotal= 1 atm, PCOz= 0.26 atm 60

    12. Stability of phases in the system CaO-MgO-CO z-H20 at 25°C and 1 atm pressure 61

    13. Correlation between concentrations of bicarbonate and sulfate in spring and well water from the Blackfoot Reservoir region, southeastern Idaho 65

    14. Comparison of water compositions from the Blackfoot Reservoir region with the calcite equilibriumcomposition defined by: CaC0 3 + H+.= Ca 2+ + HC0 3 - •.••••• 71

    15. Generalized conductive temperature profiles for major heat-flow provinces in the western United States, modified from Lachenbruch and Sass (1978) 76

    ix

  • LIST OF TABLES

    Table 1. K-Ar dates for rhyolites in the Blackfoot Reservoir region, southeastern Idaho (S. Evans, University of Utah, personal commun q 1979) 7

    2. Locations of spring- and well-water samples collected from the Blackfoot Reservoir region during July, 1978.... 25

    3. Concentrations of chemical species in spring and well water from the Blackfoot Reservoir region, southeastern Idaho 39

    4. Chemical geothermometers indicating temperatures of last water-rock equilibria for spring and well waters in the Blackfoot Reservoir region, southeastern Idaho .... 49

    5. Mixing model results for springs in the Corral Creek drainage, Blackfoot Reservoir region, southeastern Idaho. 52

    6. Reactions used to establish stability relations among mi nera1s 56

    7. Composition of travertine from the Blackfoot Reservoir region, southeastern Idaho 73

  • ACKNOWLEDGMENTS

    Dr. William T. Parry supervised the project. Dr. William P. Nash

    provided guidance and reviewed the manuscript. Dr. Walter J. Arabasz

    reviewed the manuscript. M. Cleary assisted in water sampling.

    J. Ballantyne revised WATEQF for use on the UNIVAC 1108. The University

    of Utah Research Institute Earth Science Laboratory permitted use of

    the rcp for water analyses. Mr. Maurice Magee, of Scintrex, Inc.,

    performed the uranium analyses. The project was supported by U.S.

    Geological Survey Research Grant l4-08-0001-G-545.

  • INTRODUCTIDrl

    Geothermal potential has been recognized in the region south and

    west of the Blackfoot Reservoir in southeastern Idaho because of the

    presence of several very young rhyol ite dornes, associated Quaternary

    basalt flows, warm springs, and anomalously high chemical temperatures

    in several soring waters. Evaluation of the geothermal potential of

    an area logically proceeds from analysis of regional geologic data to

    careful evaluation of volcanic rock age and petrology, spring

    geochemistry, and thermal gradients (Ward and others, 1973). The

    purpose of this report is evaluation of spring-Hater geochemistry in

    the Blackfoot Reservoir region of southeastern Icaho and detection ~f

    a geothermal reservoir in the subsurface. High apparent che~ical

    temperatures and the presence of abundant gasses are suggestive of a

    high-temperature co~ponent in the cooler springs, which may be better

    cefined with detailed study of the spring-\-1ater gecche:nistry.

    Water samples from forty-five springs an1 four wells Here

    collected and analyzed for major and minor constituents, and the

    mineralogy of six travertine samoles \'1-35 determined. ~Iater chemistry

    is interpreted in terms of mineral equi~ibria. Temperatures of last

    water-rock equilibria are estimated from tne silica c~ntent of the

    water ~~i the rla-K-Ca geothermorneter with suitable corrections fJr

    carbon dic,;dde. :"1agr.::sium, and f71ixing ;.lith cooler s'Jfface \laters •

    . Mecha~i~~s 3re oropcsed to explain the oresence of gasses and the war~

    teT.perct~res of t~e springs.

  • 1 GEOLOGIC BACKGROUND

    Physiography

    The Soda Springs-Blackfoot Reservoir region lies in a transition

    region between the Basin and Range and the Middle Rocky Mountain

    physiographic provinces, approxi~ately 75 k~ southeast of the Snake

    River Plain, Idaho (fig. 1). The boundary between the two provinces

    has been variously placed by different workers (Fenneman, 1917, D. 82;

    1931, footnote 2, p. 170; ~1ansfield, 1927, p. 11). In general, the

    eastern province is characterized by mountains formed by folding and

    thrust faulting with a high proportion of mountains to valleys. The

    western province is typified by block-faulted ~ountain ranges

    separated by wide, deeply-filled basins anc valleys. A com~or.ly

    accepted boundary is placed along the western edge of the Bear River

    Range and northward along the eastern edges of the Blackfoot and

    Willow Creek Lava Fields, although arguments ~ay be made for placement

    along the western edge of the Aspen Range.

    Stratiqraohy and Iq~eou$ Geology

    Published reports of the geology include those of ~ansfield

    (1927,1929), Armstrong and Oriel (1955), Armstrong (1969), and a

    summary by f·1abey an'd Oriel (1970). In the '.lountain ranges of the

    western, or Basin and Range, portion of the area, tilted and faulted

    Paleozoic and lavler 1vlesozoic (Triassic) sedi:nentary units are expos2d

  • .'.f $'II! o o oo o

    (0 NV

    iddle Rocky

    P~. 9 Mts.

    Utah

    Salt Lake City

    , t-

    EXPLANATION

    ~ Terrestrial volcanic rocks Quaternary (Snoke River Plain) and Tertiary age

    Wt;l'1 Thick deposits In Intermontane depressions Quaternary and Tertiary age

    o Undifferentiated Paleozoic and Mesozoic rocks

    Figure 1. Sketch map of major physiographic features of the Snake River Plain and adjacent regions. adapted from King (1969).

  • 4

    t i j

    :i 1. !

    i

    43°00' I

    I ~ I ~ Ul I U'l ....... 1 ~

    ~ ~ U'l I

    ~ ~ ~ Ig'

    \) Soda ~

    "Z-. 'Z G'l (~0")J ·Springs"t ~ I; ~ :-z.• G'l I ~ ~ I ~ \) ")J I 42°30'0 C4.Ic z ~ o'getown '2 I ~ ~

    V'l IV'l

    z ")J I~ Vl l>Q:: Iz G) Q:: tTl I

    C4.I I~ .... Q::

    Q::

    ~ ~

    Q) Idoh 42°00'Utah

    0 5 10 15 milea I I I I 0 8 16 24 kilometers

    ••• Area of fioure 4

    Figure 2. Index map of southeastern Idaho, adapted from Mansfield (1927).

  • 5 -

  • o

    basalt flows, but Middle Cone contains inclusions of d still older

    basalt. China Hat and Middle Cone have been dated at less than 0.1

    m.y. by K-Ar methods (Armstrong and others, 1975) and at about 0.05

    m.y. by K-Ar, thermoluminescence and hydration rind methods (Leeman

    and Gettings, 1977). Two islands in the Blackfoot Reservoir are

    composed of slightly older rhyolite which has been dated by Phillips

    Petroleum Company. Rhyolitic tuff from Sheep Island, in sections 11

    and 14, T. 6 S., R. 41E., gives a K-Ar aqe of 1.3 ~ 0.2 m.y.

    Rhyolitic vitrophyre from the smaller island immediately south of

    Sheep Island in section 14, T. 6 S., R. 41E., gives a si~ilar age of

    1.5 ! 0.6 m.y. (D. Hayr.lOnd, personal cO~'TIun. to t4. P. Nash, 1979).

    Recalculation of these dates using decay constants AS and \_ from

    table 1, gives ages of 1.4 ± 0.2 m.y. and 1.5 : 0.6 m.y.,

    respectively. Several exposures of rhyolite occur in the northern

    part of the study area. In the Cranes Flat 15-~inute quadrangle,

    three stubby flows or flat domes (Mansfield, 1927) of rhyolite cover a

    tota~ area of approximately 3 square miles. Two of the exposures have

    been dated by K-Ar methods at the University of Utah (see table 1).

    Rhyolite from the S~~NW~ sec. 33, T. 4 S., R. 42 E. (sample BFR-3A), .J,.

    gives an age of 1.59 ~ 0.06 m.y. and rhyolite from the NE~~E~ sec. 25,

    T. 4 S., R. 41 E. (sample 8FR-5A), gives an age of 1.41 ± 0.15 m.y.

    (S. Evans, personal cornmun., 1979). A small exposure of rhyolite tuff

    at the north end of the Chesterfield Range in sec. 20, T. 5 S., R. 40

    E., has been dated by Phillips Petroleu~ Com~Jny at 10.2 ! 0.5 ~.y.

    f (D. Haymond, personal commun. to \!. P. :lash, 1979) and recalculated using the new decay ccnstants to an age of 10.7 ! 0.5 m.y. 1:

  • 7

    Table 1. K-Ar dates for rhyolites in the Blackfoot Reservoir region, southeastern Idaho (S. Evans, University of Utah, personal commun., 1979).

    Samol e BFR-3A BFR-5A

    Location S\~!4NW;4 sec. 33, NP.d~E~ sec. 25, 1. 4 S., R. 42 E. T.4S.,R.41E.

    Rock type rhyolite rhyolite

    :-1aterial dated sanidine sanidine

    Wei oht (gm) 0.79201 0.92494

    ;; K 9. 15 7.81

    Moles/gm Ar~~d (xl0 11 ) 2.469 1.913

    :~ Ar 40 56 87-'--' rad

    Age (m.y.) 1.59 =: 0.06 1.41 :: 0.15

    Constants

    A = 0.581 x 10-10/yrE:

    K40/Ktota1 = 1.167 x 10-4 mole/mole

  • 8

    Quaterna~ travertine is widespread and delineates areas of past

    and present spring activity. The exposures vary in size up to about

    12 square kilometers and occur primarily along the edges of the lava

    fields adjacent to upper Paleozoic and lower Mesozoic calcareous

    rocks. The character of the deposits is discussed in a later section.

    Structure

    The structural evolution of southeastern Idaho has been discussed

    by Mansfield (1927), Armstrong and Cressman (1963), and Ar~strong and

    Oriel (1965). Westward-dipping thrust faults have long been

    recognized and were originally described by Richards and Mansfield

    (1912) as a single large folded thrust fault which they namea the

    Bannock overthrust. Later work by the U.S. Geological Survey as part

    of a study of the \~estern (U.S.) phosphate field resuited in the

    recognition of several discontinuous, separate thrust fcults. Current

    interpretation of the Bannock thrust zone by Armstrong and Cressman

    (1963) stems from the recognition of a westward-dipping imbricate

    thrust zone extending from southwestern nontana to north-central Utah,

    and the presence of tear faults and younger block faults as

    progressive stages in the tectonic development of the region.

    Tnrusting occurred fro~ Late Jurassic through Cretaceous ti~e,

    preceded and accompanied by the develop~ent of large north-trending

    folds. Individual thrust faults are younger to the east, as are the

    rocks exposed in the overriding ~Iestern plates (Armstrong and

    Cress~an, 1963, Armstrong and Oriel, 1965). In the Soda Springs area,

    two major thrust faults are documented. The Paris fault is 2xposed

  • 9

    near Cavanaugh Siding approximately 17 km southeast of Soda Springs

    where Cambrian Brigham Quartzite and Triassic Thaynes Limestone are

    juxtaposed (Mabey and Oriel, 1970). The trace of the Paris fault is

    covered to the north. The Meade fault places Mississippian limestone

    over Jurassic and Cretaceous strata and is exposed to the east of

    Georgetown where it bends to the north and splays into several slices

    (!~abey and Oriel, 1970).

    High-angle faults are abundant and more conspicuous than thrust

    faults in this portion of southeastern Idaho. Two principal sets of

    high-angle faults have been recognized. The east- to northeast

    trending faults are thought to be tear faults originating in response

    to differential eastward movement among the thrust plates and

    therefore contemporaneous with thrusting (Armstrong and Cressman,

    1963). Several miles of horizontal movement has been taken up along

    tear faults such as the Blackfoot fault in T.l 5., Rs. 43 and 44 u. and a set of faults south of Tenmile Pass in the Soda Springs Hills.

    The other set of high angle faults trend north to northwest and are a

    result of block faulting which began in the Tertiary and has continued

    to the present (Armstrong and Oriel, 1965). Gem Valley, Bear River

    Valley, and Slug Creek Valley are recognized as grabens. Steep

    northwest-trending normal faults cutting basalt and o~e of the

    rhyolite domes in the Blackfoot Lava Field, and the presence of fault

    scarps on modern alluvial fans, attest to the recency of block

    faulting in the Blackfoot Reservoir-Soda Springs area. Travertine

    deposits located along the east side of the Bear River Valley and

    similar deposits elsewhere in the region ~re associated with the

  • 10

    north- to northwest-trending block faults and not with thrust fault3.

    Sei smi city

    Southeastern Idaho is located within the Intermountain seismic

    belt (ISB), a northerly-trending zone of earthquake activity which, in

    this region, roughly parallels the boundary betl-/een the Basin and

    Range and Middle Rocky Montain physiographic provinces (fig. 3).

    Smith and Sbar (1974) and Smith (1978) discuss the ISB in relation tG

    the tectonics of the western United States, and Bones (1978) has

    studied the ISB in southeastern Idaho. 80th have judged seismic

    activity in southeastern Idaho to be characterized by earthquake

    s\'Iarms, \·,hich are composed of r.1any events \-Ih;ch increase in frequency

    to a maximum and then gradually decrease, and which lack a single

    outstanding mainshock. Ward (1972) reviews the association of

    swarm-type earthquake activity with major geothermal ~reas in Iceland,

    the United States, Ne\~ Zeal and, Japan, and other countri es. \Shere

    detailed data are available, there is a close spatial relationship

    between microearthquakes and geothermal areas. ihe microearthquakes

    rarely had Richter magnitudes greater than 4.5, probably because of

    the small dimensions of most geothermal areas and the weakening of the

    crust due to hydrothermal alteration and the presence of fluids.

    Focal depths ranged from near surface to 6 km. Ward (1972) and Smit~

    and Sbar (1974) suggest :hat detailed seismic studies may be useful ~n

    lJeothermal areas to delineate active fault zones \"hich l7Iay control the

    movement of fluids in the geothermal system. A microearthquake suney

    in the Scda Springs region during May 31 to June 22, 1977 (Bones,

  • 11

    Figure 3. Generalized map of seismicity in southeastern Idaho and western Wyoming based on a compilation of University of Utah epicenters, ca. 1975-1977 (see Bones, 1973, and Smith and others, 1976). Known and suspected active faults from Witkind (1975).

  • v.

  • 13

    1978) indicated a pattern of diffuse seismicity with little clustering

    along the maDped active faults in the region. Focal depths averaged

    about 7 km. Earlier microearthquake surveys in southeastern Idaho

    have also shoYn shallow fecal depths, generally less than 15 km (Smith

    and Sbar, 1974). Even though the pattern of diffuse seismicity does

    not correlate with the surface expression of active faults in the

    region, fault plane solutions indicate general east-west crustal

    extension along normal faults with small co~ponents of strike-skip

    movement (Smith and Sbar, 1971, Bones, 1973). One exception is a

    fault plane solution derived from six events southeast of the

    Blackfoot Reservoir which shows approxi~ate1y equal amounts of

    strike-slip and reverse faulting related to a local cluster of

    slightly deeper earthquakes (Bones, 1978).

    Gravitv and ~39netics

    Gravity and aeromagnetic data in the Soda Springs region have

    been assembled and interpreted in an informative paper by Mabey and

    Oriel (1970), from which the following qenera1 izations are made.

    North\.,est-trendi ng gravi ty anomalies present ; n Gem Vall ey and the

    Bear River Valley are probably a result of low-density Tertiary and

    Quaternary sedimentary rocks filling the valleys. A compound gravity

    low of a more complex nature is found in the Blackfoot Lava Field

    south of Henry. Steep gravity gradients along the margins of the

    ano~a1y s~ggest the possibility of high-anqle faults. The anomaly

    could te produced by several thousand feet of low-density material.

    However, the irregular shape of the anomaly, the abundance of volcanic

  • 14

    craters and cones, and the presence of rhyolite do~es sug~(S~ that the

    gravity anomaly could be caused by a volcanic collapse structure or

    possibly a buried granitic intrusive body.

    Magnetic anomalies are complex over areas of exposed Cenozoic

    volcanic rocks, and ~uch less complex when volcanic rocks are absent.

    The complexities can be interpreted as variations in the thickness end

    distribution of basalt flows and related basaltic intrusive bodies.

    Several negative magnetic anomalies east of the Blackfoot Reservoir

    are attributed to an older reversely-magnetized basalt flow,

    indicating an age of greater than 0.7 m.y. In the valley of Corral

    Creek, magnetic and gravity data indicate that the valley contains

    low-density fill and little or no basalt except for an area of thicker

    basalt in the northern part of the valley. The high magnetic

    intensity near China Hat may reflect thick basalt flows and intrusives

    related to the numerous vents in the immediate vicinity. This high

    magnetic intensity may also be attributed in part to an underlying

    granitic intrusive mass, as suggested by gravity interpretations.

    Leeman and Gettings (1977) have modeled a laccolithic silicic

    magma body of 330 km 3 to fit the observed gravity data. The body is

    estimated to extend from 0.5 km to a maximum of 6 km below the surface

    and to have a center of mass at 0.9 km. This remarkably large and

    shallow ~luton would, if present, constitute a generous thermal

    source.

    Heat Flow

    Brott and others (1976) have studied the regional heat flow of

  • 15

    the Snake River Plain to identify zones of high heat flolv related to

    recent volcanism and to evaluate the Qeother~al potential of the

    Plain. Most heat-flow measurements were made in the western Snake

    River Plain, which is not complicated by the effects of the Snake

    Plain aquifer that underlies most of the eastern Snake River Plain.

    Ground water in the Snake Plain aquifer flows from the Island Park

    caldera near the Idaho-Wyoming-Montana border to the southwest beneath

    the Snake River Plain and discharges primarily at Thousand Springs in

    south-central Idaho. The water table is deepest near the caldera

    (approximately 1.5 km) and becomes shallower to the southwest

    (approximately 1.0 k~) (Mundorff and others, 1964). The rapid flow of

    water through the aquifer transfers heat laterally which results in

    low suriace he~t flow. High heat flow may exist beneath the eastern

    Snake River Plain; however, no wells were ~easured by Brott and others

    (1975) that penetrated the aquifer. In a subsequent paper (Brott and

    others, 1978), an eastward increase in heat-flow values is proposed on

    the basis of systematic increase in elevation eastward, and

    west-to-east time-progression of silicic volcanism across the Snake

    River Plain.

    Heat flow values are generally higher along the northern and

    southern margins of the western Snake River Plain, compared to values

    within the Plain (Brott and others, 1976, 1978; Blackwell, 1978).

    Heat-flow values along the southern margin vary from 2.2 to 4.1 HFU

    (Heat flow units) (Brott and others, 1978). High heat flow may also

    extend along the southern margin of the eastern Snake River Plain.

    The Rexburg-St. Anthony area, which is approximately 120 km north of

  • 16

    ."" "j

    "

    '

    ir

    ,

    SOGa ~prlngs, is at the eastern boundary of tne Snake Plain aquifer

    and has average heat-flow values of 5.0 HFU (Brott and others, 1976).

    Based on a time-progressive thermal model. Brott and others (1978)

    predict a regional heat flow of 2.5 to 3.0 HFU in the area fro~ Idaho

    Falls to the Idaho-Wyoming border. The corresponding thermal

    gradients would range from 85° to 100°C/km. The high heat flow

    predicted for the eastern Snake River Plain and its margins imply that

    there should be many active geothermal systems where heat f1o\1 is

    sufficient for hydrothermal convection to become an important

    heat-transfer mechanism.

    There is a large ga~ in heat-flow data in southeastern Idaho,

    which includes the Blackfoot Reservoir region. Although the region is

    approximately 50 to 75 km from the southern edge of the Snake River

    Plain, upper Cenozoic volcanic rocks that are present in the Blackfoot

    and Gem Valley Lava Fields join northward with similar rocks of the

    Snake River Plain. Evidence that volcanic rocks of the 8lackfoot

    Reservoir region were extruded from local sources indicates that heat

    flow may be high in this area. Further heat-flow measurements may be

    complicated by aquifers in the basalt, uhere the major water-bearing

    zones occur in layers of scoriaceous ~ateria1 and cinders, and at the

    contacts between flows. The aquifer beneath the Blackfoot Lava Field

    is recharged by leakage fro~ the Blackfoot Reservoir and by

    precipitation and irrigation water (Dion, 1974). Ground water flo~s

    to the south past the town of Soda Springs into the Bear River and

    Soda Point Reservoir. Wells drilled in the basalt aquifer have yields

    that vary from 300 gpm (gallons per minute) to 3500 gpm (Dian, 1974).

  • 17

    This is equivalent to about 19 and 220 liters per second,

    respectively. The depth to the top of the aquifer is generally less

    than 100 meters (Dian, 1974); therefore, wells or holes used to obtain

    heat-flow measure~ents should be greater than 100 meters deep in the

    Blackfoot Lava Field.

  • 18

    Figure 4. Generalized geologic map of the Blackfoot Reservoir region, southeastern Idaho, modified from Dion (1974).

    Figure 4a. Simplified structural map of faults in the Blackfoot Reservoir region, southeastern Idaho, adapted from Armstrong and Cressman (1963). Overlay to figure 4.

    Figure 4b. Location map of water and travertine samples collected during July, 1978, in the Blackfoot Reservoir region, southeastern Idaho (see table 2). Overlay to figure 4.

  • ~..--:1~~r:d••.." Spt'lnV

    FIQure 4b_

    21

    fXPLANATION

    ot wll wotlr loapll

    oTrovertlACI sampl'

    EXPLANATION

    ~-.- ...... ..,... Block foull hotchurls on ~O~ thrown side

    :. ~-r-r TronlYl"l fault hotJ:burel on do_· thrown 51delotrow indiCltl. dlrlctloA 01 relative Inovement

    -....--?T yTtlrUit fel ult querlld where doubtful i 10WltlU. in upper plale and poinlinll down dill

    fioure 40. Figure 4.

  • PREVIOUS SPRING WATER STUDIES

    Springs in southeastern Idaho have been described by Mansfield

    (1927, 1929) as normal springs (non-saline and non-mineral de~ositing)

    \~hich grade into mineralized and thermal springs of four types:

    calcareous, iron-bearing, sulphureted, and saline. Springs in the

    Soda Springs-Blackfoot Reservoir area are various combinations of the

    first three types, all of which are carbonated. Most of the sulfur

    springs and vents are concentrated along the eastern edge of the Bear

    River Valley. Saline springs occur to the east in the Craw Cfeek and

    Freedom 7~-minute quadrangles. No chemical\analyses of springs or

    spring dep0sits were reported by Mansfield except from A~burn Hot

    Springs in western Wyoming and from a group of hot springs in the Sear

    La~e Valley.

    Dion (1974) studied the hydrologic budget of the Blackfoot

    Reservoir, Soda Creek, and the gear River, to dete~ine the amount of

    leakage from the reservoir under present conditions and the effect of

    raisi~g the water level of the reservoir. The Blackfoot River basin

    (Snake River drainage) is separated from the Bear River basin (Great

    Basin drainage) by a barely perceptible divide across the Blackfeot

    Lava Field between the reservoir and Soda Springs. Dion deter~ined

    that about 10 cubic feet of water per second was leaking i~to the Sear

    River basin fro~ the Blackfoot Reservoir and that the actual yielas of

    Soda Creek and the Bear River were greater th2n the expected yields

  • 23

    due to contributions from numerous carbonated springs. He speculated

    that the source of the carbonated springs was deeply circulating

    meteoric water from the nearest higher drainage area, which is the

    upper Blackfoot River basin.

    Young and Mitchell (1973), and Mitchell (1976) have identified

    geothermal potential in the Blackfoot Reservoir area based on the

    geologic setting and anamalous subsurface water temperatures predicted

    by the Na-K-Ca geochernical thermometer. Later work consisted of

    sampling and analyzing eleven springs in the region: Woodall Springs,

    Hooper Springs, Soda Springs, Sulfur Springs, four Corral Creek

    springs, two springs along the Blackfoot River below the dam, and a

    spring near Wilson Ridge. Mitchell (1976) concludes that there is a

    possibility of deep geothermal resources or shallow low-temperature

    ~eothermal energy in the region.

  • ~aj~r dete~ined a Pe~kin-Elmer

    Model 603 atomic absorption spectrophoto~eter. Sodium, potassi~rn,

    METHODS

    Collection and Analysis

    Water

    Fifty water samples were collected from an area of approximately

    900 km 2 surrounding the Blackfoot Reservoir in southeastern Idaho.

    Sample locations are given in table 2 and are shown in figure 4b.

    Springs and wells are numbered according to the system used by the

    U.S. Geological Survey in Idaho (fig. 5). Sampling proc2dures of

    Brown, and others (1970) and Presser and Barnes (1974) were followed.

    The water was filtered through a·0.45-~icrometer filter and four

    500-ml samples were collected from each site in polyethylene bottles

    with polyseal caps. Two of the samples were acidified to below pH 2

    with hydrochloric and nitric acid. The third sample was diluted 10:1

    to prevent polymerization of silica, although analysis later showed

    that the concentrations of silica were so low that the dilutions were

    unnecessary. Undiluted samples were actually used for silica analysis

    to eliminate any error introduced by dilution. Temperature, pH, and

    bicarbonate were measured in the field. The high CO 2 content of the

    water and the rapid re-equilibration of aqueous carbonate species made

    it necessary to perform the bicarbonate titration as quickly as

    possible at the sample site. Even so, the calculated bicarbonate

    contents are probably sl.ightly lower and the p~ measurements slightly

    higher than those of the actual springs.

    dissolved cations \vere using

  • 25

    Table 2. Locations of spring- and well-water samples collected from the Blackfoot Reservoir region, southeastern Idaho, during July, 1973.

    Sample

    Woodall 1

    Location

    7S 42E 34baa5

    Map Designation

    A ~'Jooda 11 2 75 42E 27dbcS B ~~ooda 11 3 7S 42E 27acaS C Woodall 4 75 42E 23badS o Hooper Spring 85 41E 36dddS Formation Spring 85 42E 27cbbS

    EF

    Poison Creek 1 45 41E 32bbb5 G Poison Creek 2 45 41E 32cca5 H Big Spring 85 41E 33aacS Corra1 Creek 1 65 41E 19baclS

    IJ

    Corra1 Creek 2 65 41 E 19bac2S J Cerra1 Creek 3 65 41E 19bac3S J Corral Creek 4 6S 41E 19bac4 Carra1 Creek 6 6S 41 E 19bac65

    JJ

    Sulfur Spring 1 9S 42E 14aadS K Sulfur Spring 2 9S 42E 13bca5 SP 100 9S 42E 6badS SP 101 9S 42E 5bcd SP 102 9S 41 E 1cbcS

    LMNo

    SP 103 6S 40E 12dbbS P SP 104 65 40E 12bddS SP 105 6S 41E 6bac1S SP 106 6S 41E 6bac25

    QR S

    SP 107 6S 41E 6dbc5 SP 108 7S 42E 16abc SP 109 5S 41E 30cddS SP 110 7S 40E 1aadS

    TUV \~

    SP 111 7S 40E 12adbS SP 112 7S 40E 23abaS SP 113 75 40E 23baa5

    XYZ

    SP 114 SP 115 SP 116 SP 117 SP 113 SP 119 SP 120 SP 121

    5S 40E 15bdcS 55 40E 15abdS 5S 40E "' bbdS 5S 40E 14bcdS 5S 40E 25aabS 85 41E 23aaaS 8S 42E 15cccS 8S 40E 26dcc5

    AA BB CC DO EE FF GG HH

    "

  • 26

    Table 2. Locations of spring- and well-water samoles collected from the Blackfoot Reservoir region t southeastern Idaho, during July, 1978 (continued).

    Sample Location Map Designation

    Chubb Spring 5S 42E llcacS II Soda Spring 9S 41£ l2addS JJ Lone Tree Spring 6S 41£ laddS KK Hopkin's Landing 6S 41£ 22add LL Henry 1 6S 42£ 10cbaS ~lM Henry 2 6S 42£ 9dbdS NN Henry 3 6S 42£ 9bccS 00 Henry 4 6S 42E 8addS pp Henry 5 6S 42£ 8dbaS QQRh Spring 1 4S 42£ 30cbbS RR Rh Spring 2 4S 41£ 25adbS SS

  • Ff t---+--~.~ ~ t::l ~ EE

    -lii:t-- ~/N

    IW 801S£

    IS 1£ 2£

    R. 41 E.

    6 5

    7

    13

    21 22 23 24

    30 29 28 27 26 25

    3/ 32 33 34 35 36

    T.

    6

    s.

    IBAS: LIN I 40141£ 42£ 43£

    S£CTlON 19

    I b a I I ,,·:-;1 I

    ---1- .L - t - - -~ I - -C I d I I

    I I - ~- -/9- -1-

    I I I II ---~---l--- ~---

    II I I I I I

    6S-4IE-19bool

    Figure 5. Spring- and well-numbering system used by the U.S. Geological '\

    Survey in Idaho. Example is well 6S 41E 19baal, which is located in the NE~NE~NW~ sec. 19, T. 6 S., R. 41 E., and is the first well sampled in that tract. Spring locations are followed by IISII and travertine locations are followed by liP. Adapted from Mitchell (1976).

  • 1 1 t ! i

    calcium, silica, magnesium, and strontium were determined by direct

    methods. Sulfate was determined indirectly by adding a known

    concentration of barium-chloride solution, precipitating the sulfate

    as barium sulfate, and measuring the amount of bariu~ left in

    solution. Aluminum and iron were determined using the HGA 2000

    graphite furnace. Chloride was measured with a specific ion electrode

    and by Mohr titration. An average value is reported, since the two

    methods were generally in agreement. Fluoride was also ~easured with

    a specific ion electrode. A phosphomo1ybdate method was used to

    dete~ine orthophosphate calorimetrically. Total orthophosphate is

    reported as P04 • Boron, lithiu~, barium, ~anganese, and strontium

    contents \iere determined with an ARl inductively coupled plasma

    spectrophotometer. Strontium values from both the AA and rcp methods

    are reported. At concentrations above 2 mg/l the results are similar.

    However, in samples which contain lesser amounts of strontiu~ the rcp

    results are roughly twice those of ato~ic absorption. Seven spring

    samples were analysed for uranium using a Scintrex UA-3 uranium

    anaiyser.

    Travertine

    Sixteen travertine sa~p1es were collected from active springs and

    from tufa mounds associated with extinct springs. After initial

    petrographic examination, six representative specimens were chosen for

    more detailed study. X-ray diffraction techniques were used to

    determine the mineralogy of the whole rock and of the acid-insoluble

    residu~. Several specimens were examined using the electron

  • 29

    microprobe and the scanning electron microscope for se~i-quantit3tive

    chemical analysis.

    Ther~odynamic Modelinq of Ion Species in Solution

    Activities, fugacities, activity ratios, and activity products of

    solution species were calculated using HATEQF (Plurnrner and others,

    1976), a Fortran IV version of \lATEQ (Truesdell and Jones, 1973). The

    computer program calculates the equilibrium distribution of inorganic

    aqueous species and complexes for major and minor elements based on

    chemical analyses. Activity coefficients are first calculated usinq

    the extended Debye-Huckel equation for dilute aqueous solutions.

    Calculations are then made of the concentrations of weak acids and ion

    pairs by a series of mass action and mass balance equations. The

    calculated concentrations reduce the amount of free ions in solution

    and thus change the ionic strength and activity coefficients. New

    values for the activity coefficients and concentrations of species are

    re-entered in the equations, and the process is repeated until the

    sums of all species agree with the analytical values v/ithin 0.5

    percent. The ratios of the ion activity products to the sOlubility

    products of minerals are compared. If the ratio is greate~ than one,

    the water is supersaturated with respect to that mineral. When the

    ratio is less than one, the wate~ is unde~saturated. Neither Eh nor

    dissolved oxygen were measured in the BlackfoJt Reservoir springs,

    therefore WATEQF was unable to calculate pE, oxidation-reduction

    reactions, or the partial pressures of 02 and CH 4 in the water.

  • 30

    ...• K!

    Geothermometers

    '

    Basic assumptions

    Temperatures of last water-rock equilibria were calculated for

    each water analysis using the basic tools provided by Fournier and

    Rowe (1966), Fournier and Truesdell (1973), Paces (1975), and Fournier

    and Potter (1979). The assumptions made in using chemical

    constituents of hot spring waters to estimate the subsurface

    temperature can be outlined, after Fournier and others (1974), as

    follows:

    1) Chemical reactions which are temperature dependent

    occur in the subsurface. The silica geothe~ometer

    assumes that the amount of silica in the surface

    spring water is controlled by the solubility of

    quartz or chalcedony at depth. The Na-K-Ca

    geother~ometer assumes that reaction betHeen

    Na-fe1dspar and K-fe1dspar controls the con

    centrations of these cations in solution and

    also takes into account the influence of ca1cium,

    which competes with sodium and potassium in the

    silicate reactions. The solubilities of silica and

    feldspars vary as a function of temperature and

    to a lesser degree, of pressure.

    2) The mineral phases are sufficiently abun1ant in

    the subsurface so that supply is not the limiting

    factor.

  • 3) Equilibrium between rocks and water is approached

    at the reservoir temperature. This assumption

    is true primarily for high subsurface temperatures

    and long reservoir residence times.

    4) The water does not re-equilibrate or change

    composition as it flov/s to the surface. This

    depends upon factors such as the upward travel

    time of the water, armoring of the channel rocks with

    CaC03 or 5i02, dilution of the high-temperature

    water with water from another source, and the

    mechanism of cooling. Deeper water entering a

    shallow reservoir may be diluted with chemically

    different, cooler water and the spring water

    emerging at the surface will be a mixture of the

    two types. Hot water on its way to the surface

    cools by different mechanisms. Deep water that

    is at a higher temperature than the surface

    boiling temperature will c~ol adiabatically (by

    boiling) as it rises through l~wer pressure

    regions. As the steam separates, non-volatile

    components will be concentrated in the fluid

    phase. Lower temperature w2ter cools conductive1y.

    The amount of heat lost to the wa11rock through

    conductive cooling depends upon the flow rate.

    If the channel is sufficiently non-reactive

    and the flow rate rapid, the water composition

  • 32

    should re~aln necrly tne sa~e. The task is to

    distinguish between springs which cool adiabatically,

    conductively, or by a combination of both mechanisms.

    SiD? geothelinOmeter

    Althouqh amorphous silica is the precipitated silica phase under

    surface conditions in hot spring pools, the a~ount of silica in

    solution is controlled by the solubility of quartz or other silica

    phases found at depth. The solubilities of amorphous silica,

    cristobalite, chalcedony, and quartz at vario~s te~peratures have been

    studied by many workers (Fournier and Rowe, 1977, Morey and others,

    1962, 1964, Siever, 1962). Studies of active geothermal systems show

    that silica concentrations are relatively constant over long time

    periods, which implies that steady-state conditions have Dersisted at

    depth (Fournier and Rowe, 1965). Evidence also suggests that little

    ~e-2Gu;librat;on occurs during the transport of water to the surface

    due to armoring of the channels with amorphous silica, rapid co01in9,

    and the sluggish deposition of quartz and other crystalline silica

    phases. Thus, the silica content of a surface hot spring can be used

    to estimate the temperature of last equilibriu~ with a silica phase at

    depth, provided the cooling takes place by conduction or that a

    correction is ~ade for steam separating from the solution during

    ascent. Adiabatic cooling by ste3m separation was not considered for

    springs in the Blackfoot Reservoir area because of l~w to moderate

    surface temperatures (well below 100°C) and the absenl.e of stea~

    fumaroles.

  • Silica temperatures were calculated fro~:

    tOC (quartz, conductive cooling) = 5.205 1315

    - log SiO z (mg/l) - 273.15

    tOC (chalcedony, conductive cooling) = 4.655 1015. 1

    - log SiO z(mg/l) "73 15

    - L •

    (Truesdell, 1976).

    The quartz geothermometer works best at temperatures from 150°

    225°C. At 10\ver ternperatures, the chalcedony geothermo:-:leter '":lay give

    more accurate results (Fournier, 1977).

    Na-K-Ca oeother~o~eter "

    The exchange reaction between sodium and potassium feldspars is

    the basis of a quantitative geothermometer from which several

    empirical curves relating atomic Na/K ratios in natural waters to

    temperature have been constructed (White, 1965; Ellis, 1970; Mercado,

    1970; and Fournier and Truesdell, 1970). Increasing evidence that

    calcium-rich waters did not yield reasonable Na/K temperatures,

    especially in low to moderate temperature systems, led to the

    development of the Na-K-Ca geother'llometer. Eve" though the amount of

    calcium in solution is controlled by the solubility of a caicium

    fr

    carbo~ate at certain conditions of te~perar.ure, pH, and PCQ2, Fournier

    and Truesdell (1973) have empirically shown that aqueous Na-K-Ca

    relationships can be explained in terms of silicate reactions. The ,

    reactions are shown by tnree configurations:

    (x + 2y) K+ + solid = x Na+ + y Ca2 + + solid

    (2y - x) K+ + x ~a+ + solid = y Ca2 + + solid

  • 34

    (x - 2y) K+ + Y Ca 2+ + solid = x rJa+ + solid.

    If one potassium ion is used in writing the reactions, then a

    generalized equilibrium constant, K*, can be written for all

    configurations:

    log K* = log Na + B 10gfCa K Na

    where S depends on the stoichiometry of the reaction. By plottin~

    previously known hot spring compositions and temoeratures on a graph

    of log K* versus 10 3 /T (absolute), a best fit curve was obtained when

    s = 4/3 or 1/3, depending upon the temperature. Evaporative

    concentrations do not affect the a1k~li ratios; however, there is

    evidence that significant water-rocx reaction occurs during ascent

    which alters the ratios and results in deceptive tla-K-Ca ternperatL!res

    (Fournier and Truesdell, 1973). tla-K-Ca geotherl1Cl71eter temperatures

    were calculated from:

    1647 tOC = log (Na/K) + slog ()lCa/Na) + 2.24 - 273.15

    ...,here 8 = 4/3 for (;JCd/Na) > 1 and t < lOO°C,

    a '" 1/3 for (fu/Na) < 1 or t4/3 > leQoC,

    and Na, K, Ca are in mol es/l iter

    (Fournier and Truesdell, 1973)

    Na-K-Ca-C~2 geothermometer

    Pace$ (1975) ~as e~pirical1y derived a correctic~ factor (I) for

    the Na-K-Ca ~eother~o~eter to be apolied at ground water te~Deratures

    of below 75 u C with the partial pressure of CO~ above 10-4 atmoscheres.

  • 35

    At these temperatures, an inf:~x of 2n 2c~ci~ying age~t such as ;~

    may cause a steady-state situation in the silicate reactions to exist

    rather than an equilibrium state. I, the disequilibrium index, is the

    deviation of the water from its equilibriu~ co~position and is shown

    to be related to PC02 by a least squares method yielding the equation

    I = -1.36 - 0.253 log PC02 . Subsurface temperatures are calculated using a modified form of the

    Na-K-Ca geothermometer:

    o 1647 t C = log (Na/K) + 4/3 log ()IC'a/Na) _ 1+ 2.24 - 273.15

    where ~a. K, Ca are 1n moles/liter

    (Paces. 1975).

    Na-K-Ca-Mq geother~o'11eter

    Observations of active geothermal syste~s and exoerimental

    evidence by Ellis (1971) show that ~agnesium concentrations are

    generally low when temperatures are high (greater than a~out 175°C).

    Cooler waters contain more r:1agnesium and may yield abnormally high

    ~a-K-Ca temperatures. Fournier and Potter (1979) have devised an

    empirical method for correcting Na-K-Ca te~peratures for low to

    moderate amounts of magnesium. The correction depends on both the

    estimated Na-K-Ca temperature and the equivalent percent magnesium

    (O'lg/Ulg+K+Ca) x 100). Fournier and Potter recom::lend using the

    correction only when Na-K-Ca temperatures are above 70°C and when R,

    the equivalent percent magnesium, is less than 50. Waters that have

    R > sa do not fit the empirical curve as well, perna os reflecting

  • 36

    equilibration at lower temperatures or a nonequilibrium situation.

    The correction factor can be obtained graohically or by the equation:

    ttMg = 10.66 - 4.7415(R) + 325.87 (log R)2 - 1.032 X 105

    (log R)2/T

    -1.968 x 107 (10g R)2/T2 + 1.605 X 107 (log R)3/T2,

    where T = Na-K-Ca temperature (K),

    and R = (Mg/(Mg + K + Ca)) x 100, with Mg, K, Ca in equivalent units.

    iltMg is then subtracted from the esti~ated fla-K-Ca temperature.

    i1ixing Models

    Fournier and Truesdell (1974) have proposed a method to deteJ.nine

    the temperature of the hot water co~ponent and the fraction of cold

    water present in \'iarm spri ngs Vlhi ch are a mi xture of deep hot water

    and cold water of a shallower origin. The model used here is one of

    conservation of enthalpy and silica; that is, any steam separated

    during the ascent condenses in the cold water, and no solution or

    deposition of silica occurs after the hot water leaves the reservoir.

    Two equations aie written; the first relates entha1pies to the

    fraction of cold water present

    (H cold) (X) + (H hot)(l-X) = H spring

    where X is the fraction of cold water, and H is enthalpy, and the

    second equation relates silica contents ~o the fraction of cold Hater

    (Si cold)("X) + (Si hot) O-X) = Si spring

    ,. where Si is the silica content in ~g/l. The silica content cf the hot

    \·/ater is related to the te'l1perature by the solubility of quartz.

  • ;)/

    Selected vJlues of enthalpy and silica for different hot \'/ater

    temperatures are given in Fournier and Truesdell (1974). A graphical

    approach can be used to solve the equations. The two equations are

    plotted on a graph of II temperature of hot \~ater" versus II fract i on of

    co~d water," by assuming a series of values for (H hot) and (5i hot)

    and knowing the temperatures and silica contents of the warm springs

    and the cold water component. The point of intersection of the curves

    gives the estimated hot water temperature and fraction of CQld water.

  • H.lHER CHGlI STR \'

    Chemical Constituents

    Table 3 lists the surface te~peratures, pH, and concentration of

    major and minor elements present in water samples from the Soda

    Springs - Blackfoot Reservoir area. All of the elements are in mg/l

    except uranium, which is in pg/l. Most of the waters are neutral to

    very slightly acidic. Approximately 85% of the scmples are between pH

    6.2 and 7.0.

    Major ele::lents

    Figure 6 shows the distribution of major element concentrations

    and also indicates those samples which show consistently anomalo~s

    ccncentrations of several elements. Major cations present are sodium,

    potassium, calcium, silica, and ~agnesium. Silica concentrations are

    generally lJw, averaging 11 mg/l 5i. HooDer Springs contains 33 mg/l

    5i, the largest amount measured in the region. The relatively high

    silica at Rh Spr. 1 and 2 was expected because these springs flow

    direct1y from glassy rhyolitic rocks. Hole percents of the other

    major cations are plotted on a portion of the (Ca+Mg)-Na-K triangular

    diagram in figure 7. Most of the springs contain very little

    potassium relati>le to othel~ cations, are high in calcium and

    magnesiu~, and contain varyin1 a~our.ts af sodium. The most noticeable

    exceptions are SP 104, SP 114, SP 117, and Corral Creek 1,2, 3,'~,

  • " 39 .~

    ,f

    Table 3. Concentrations of chemical species in spring and well water from the Blackfoot Reservoir region, southeastern Idaho (reported in mg/l; except uranium, which is in ~g/l).

  • ,",~,~.",""i!o

  • .;.p, .i .""~'''J;,\~,_.~"." ,~~~:~~",."

    Ca Si Mg Mn 5r Sr Ila Al Fe Il L1(1 F P04 504

    I!COJu

    5alllllle T: C vI! Na K (ICP) (ICP)(ICP) ocr) ( lCP)

    1,2 3fJO SID 17 1011 19 a8 9 24 ~2 1'15 10 53

    J7 1.24 0.00 890 35flOO.O,? 4.5 5.3 0.25 0.04 0.78 0.53 0.646.74 95 241 635 10 2~HSP 104 22 7.5 0.14 0.05 229

    52 5

  • ,_ ._~. _._ -,-,.~~.------""._",~_ '.fA. @:t ~.,' ---'"- ~un TS Zt" . • ~ __ __ ...-.< _~ ... ""....... '.... 4-.' -,., --~".·ot_" ~"-"",'.,.,_ ~l_' •

    Sample T:C pH Na K Ca SI Mg ·Mn (ICP)

    Sr (lCP)

    Sr 8a (ICP)

    Al Fe 8 (lCP)

    L1 (lCP)

    Cl F .P04 S04 He03 u

    SP 117 28 6.42 148 207 615 II 268

  • 43

    3S

    30

    25

    20 ~ 20U Z LLI ::::I 0

    15 LLI

    • 15cr l&.()~ .... ...U ()~Z ....It)LLI 10 ..::::I 10 It)... ~ 0

    --I

    ..IW .... .... ...a: ....~ ....l&. ....~ .....5 ~:::: ~ ~

    "" 0 50 100 150 200 60 100 150 200 250

    Na,ppm K,ppm

    30

    25

    20 20 ~ U Z W

    15 ::::I 15 0

    u ~ cr

    w Z l&. W

    10 10 0 ::::I

    Wa: l&.

    e

    o 150 o 50 100 SiCa, ppm

    Figure 6. Histograms of sodium, potassiunl, chloride, silica, calcium, magnesium, and bicarbonate concentrations in spring and well water from the Blackfoot Reservoir region, southeastern Idaho.

  • •., 4 j. f

    20 >u z IlJ " => 10a L&J a: u. 5

    0

    44

    .... ~ ..:'.. Q.::: Ct).. ...

    I~ l:I- "llCl. \.)

    '0 Ct) \.) 0 \.) Ct)

    \> r-/I 100 200 300 400 ~OO 600 700 800 900 1000

    20

    15

    >U Z L&J :=I '0 0 L&J a: IJ.

    5

    0 200 2~0 300

    15

    .......... ~

    't ....

    '~......

    Co, ppm

    >U 10 Z \LI :=I a L&J a: 5 u.

    4000350030002000I~1000500

    't I)... Cl. Ct)

    Fig~re 6 (continued). Histograms of sodium, potassium, chloride, silica, calcium, magnesium, and bicarbonate concentrations in spring and well water from the Blackfoot Reservoir region, southeastern Idaho.

  • ,, Co + Mg

    f

    • . • SP 103

    ccr·~t .SP 104 .SP 1/4 -SP 1/7

    .SP/oo

    • HOPKINS LANDING

    Rh Spr. I .R" $pr. 2

    No K

    Figure 7. A portion of the (Ca + Mg)-Na-K triangular diagram showing mole percentages of the major cations is spring and well waters from the Blackfoot Reservoir region, southeastern Idaho.

  • 1 46

    CI 50 4

    Figure 8. A portion of the HC0 3 -Cl-S04 triangular diagram showing mole percentages of the major anions in spring and well waters from the Blackfoot Reservoir region, southeastern Idaho.

  • 47

    and 6, whic~ contain unusually large conce~trations 8f potassi~m.

    Major anions present are bicarbonate, sulfate, and chlorine. The

    eight springs that contain anomalous amounts of potassium also contain

    the largest amounts of sulfate and bicarbonate. Mole percents of the

    anions are plotted on a portion of the HC03-Cl-S04 triangular diagram

    in figure 8. Bicarbonate is the dominant anion in springs in this

    region. SP 100 contains anomalous concentrations of fluorine and

    phosphate, which is part of the reason for its unusual position within

    the triangle, as these elements are not considered when calculating

    the mole percentages. This spring is located less than one mile from

    the Monsanto Company·s elemental phosphorous plant and is probably

    contaminated with water draininq from the tailings and wastes. The

    contaminated spring water then enters Soda Creek, which flows through

    the town of Soda Springs.

    Minor Elements

    Minor elements that make up less than one weight percent of the

    total dissolved solids include strontiu~, barium, aluminum, iron,

    boron, lithium, fluorine, and phosphate. The largest concentrations

    of strontium, boron, lithium, and fluorine are found in waters from

    the Corral Creek springs, SP 104, SP 114, S~ 117. and Soda Springs.

    These elements are among the few whose concentration in natural

    ther~a1 waters is not governed by water-rock equilibria (Ellis and

    !1ahon, 1964). C1, B. Li, Sr, and F are not readily accomodated in

    secondary silicate structures and so tend to accumulate in the liquid

    Dhase, once liberated from their source mineral. Ellis and ~ahon

  • 48

    (1964, 1967) reported that chloride and boron are very soluble

    elements which tend to be concentrated on the surfaces of minerals and

    thus released easily from the rock. Lithium is more apt to be

    incorporated into the structure of minerals and thus its concentration

    in waters increases with temperature. The concentrations of these

    elements can be accounted for by water-rock interactions without

    calling upon addition of magmatic fluid. Their presence in certain

    springs in southeastern Idaho is probably due to water-rock reactionsf f enhanced by high bicarbonate content and/or increased temperature.

    Iron and aluminum are present in barely detectable concentrations in

    most samples. The largest iron concentrations occ~r in springs \~ith

    noticeably dark brown to red, iron-stained travertine deposits.

    Discussion of Chemic~l Te~peratures

    5i02 and Na-K-Ca te~peratures

    Springs sampled in southeastern Idaho vary in surface temperature

    from 8°C to 40°C t with an average of 16°C. The war~est waters occur

    at a group of springs and wells drilled by FMC Corporation in 1966 to

    1970 which are located on the west side of the Reservoir r10untains in

    the Corral Creek drainage. The relatively low temperatures (as

    compared to 100°C for boiling) and lack of evidence of steam fumaroles

    in the area indicate primarily conductive cooling.

    Quartz and chalcedony saturation temperatures assuming no stea~

    loss during cooling were calculated from the measured silica content

    of the springs and are tabulated in table 4. Estimated temperatures

    vary from 43°C to 119uC for quartz saturation, and from SoC to 8goC

  • 49

    Table 4. Chemical geothermometers indicating temperatures of last water-rock equilibria for spring and well waters in the Blackfoot Reservoir region, southeastern Idaho.

    Measured Quartz Chalcedony Temp. (cond.) (cond. ) Na-K-Ca Na-K-Ca-CO Na - K-Ca -:~g

    tOC tOC tee tOC 2 tOCt"'C

    ilooda 11 1 12 43 8 -19 -59-.,-:)1"'Jooda i 1 2 14 53 19 -16

    \'Jooda: 1 3 17 58 23 -6 -52 Woodall 4 8 53 19 -2 -42 Hooper Spr. 10 119 89 58 -15 Formation Spr. 12 43 3 -19 -57 Poison Ck. 1 19 53 19 16 -23 Poison Ck. 2 14 58 23 18 -32 6ig Spr. 12 77 43 -6 -46 Carra1 Ck. 1 40 77 43 366 34 114 Corral Ck. 2 38 77 43 365 34 106 Corral Ck. 3 32 74 40 367 35 107 Carra 1 Ck. 4 28 82 49 368 32 111 Corral Ck. 6 25 62 28 347 23 101 Su1 f:Jr Spr. 1 21 95 63 '2 -116 Sulfur Spr. 2 11 58 23 -15 -58 S? 100 10 93 61 74 11 31 SP 101 21 103 71 49 6 SP 102 12 91 58 43 -18 SP 103 12 66 31 74 5 70 S? 104 22 66 31 370 36 108 SP 105 8 43 8 -22 -50 SP lD6 8 43 8 -20 -56 SP leo7 8 48 14 -24 -53 S? 103 13 53 19 -6 -.16 SP 109 10 58 23 -8 -49 SP 11 Q 2 48 14 -27 -61 SP 111 26 43 14 20 -23 SP 112 9 80 46 6 -30 S1> ~ 13 9 71 37 -9 -J5 SP 114 18 48 14 324 32 33 SP 115 11 53 19 3 -41 SP 116 25 58 23 31 -24 SP 117 28 71 37 325 33 81 S? 11B 3 62 22 -1 -43 S? 119 10 107 76 50 -13 S? 120 10 43 14 -11 -51 S; 121 16 77 43 34 -25 Chubb Sor. 13 43 14 10 -25

    .~; Soda SOl'. 30 S~ 49 28 -27 Lone Tt'ee Spr. 24 62 28 47 -14 Hcokir,s Ldg. 9 80 46 25 -24

    :; ...~enry 1 15 ..;) 19 1 -40 Henry 2 16 58 23 -1 -42 Henry 3 18 53 19 7 -39 ~enry c: 20 62 28 16 -34 Henry 5 24 71 37

  • 1

    r Wi:

    J

    J

    ..

    50

    for chalcedony saturation. The calculated ternperattires are only

    moderately high for the springs with the highest surface temperatures.

    In some cases. the chalcedony te~peratures are sirni1ar to the measured

    surface temperatures. Two cold springs flo~ing through rhyolite to

    the northeast of the Blackfoot Reservoir (Rh Sor. 1 and 2) give high

    (100°C) calculated subsurface temperatures probably due to silica

    dissolved from volcanic glass. The variability and inconsistancies of

    the silica saturation temperatures may be explqined by returning to

    the basic assumptions listed previously. Equilibrium with quartz or

    chalcedony may not have been reached due to low temperatures or rapid

    movement of water through the reservoir. The Si02 content may be

    controlled by aluminosilicates rather than a pure silica phase. 110st

    of the warm springs have a rapid flow, which may mean little

    re-equilibration has occurred; however, mixing with cooler, more

    dilute waters may be partly responsible for the low silica content.

    Na-K-Ca temperatures are also tabulated in table 4. The

    estimates vary from -27°C to +74°C, except for eight springs which

    give anomalously high temperatures averaging 354°C. These sprin~s

    include Corral Creek 1,2, 3,4, and 6, SP 104, SP 114. and SP 117.

    They are roughly aligned in a north-northwesterly direction along the

    valley of Corral Creek. The anomalous temperatures are a resuit of

    iarge potassium concentrations which drive the log (ila/K) value used

    in the calculation toward zero or a slightly negative number. As can

    be seen from figure 6 the anomalous springs have potassium

    concentrations several times greater than any other springs in the

    area. Geother~omet2r temperatures below ooe are the result of

  • , I 51

    1

    deviations from tne basic assumptions. There is most likely an

    abundance of feldspar minerals in the rocks; however, at lower

    temperatures the amount of Ma, K, and Ca in the water may be

    controlled by reactions between minerals other than feldspars such as

    micas, clays, and carbonates.

    An important characteristic of the springs in the Blackfoot

    reservoir region, and one which is not taken into account in the

    silica or Na-K-Ca geothermometers, is the presence of large amounts of

    bicarbonate in the water. Concentrations range from less than 100

    mg/l to over 3500 mg/l, with the highest amounts found in the Corral

    Creek springs, SP 104, SP 114, SP 117, and Soda Springs (see fig. 6).

    It is evident upon comparison (table 4) that a correction for the C02

    content in the water drastically lowers the Na-K-Ca te~peratures. The

    Corral Creek springs, SPI04, SP114, and SPl17 show C02-corrected

    Na-K-Ca temperatures that average 32°C, very close to their ~easured

    surface temperatures (average 29°C). Most of the Blackfoot Reservoir

    area springs, other than those mentioned above, have N3-K-Ca-C0 2

    temperatures below O°C, suqqesting that the steady-state exists, not

    between feldspars, but between other Na, K, and Ca-bearing minerals.

    Re-equilibration between the water and wall-recks may also be

    occurring, as is evidenced by widespread travertine deposition in the

    region.

    A magnesium correction was applied to ten tta-K-Ca temperatures

    according to the recommendations of Fournier and Petter (1979). The

    correction lowered the eight anomalous Na-K-Ca te~Deratures to

    approximately 100°C, which ';s a more reasonable estimate. The

  • 52

    i1a-K-Ca-Mg temperatures (table 4) are still three to four times

    greater than the measured surface !e~Deratures of the springs, which

    ~ay or may not reflect high subsurface te~pera~ures.

    In using any sort of correction to the rJa-K-Ca geothermo'11eter,

    the basic assumptions which were discussed previously are an inherent

    part of the correction procedure. !f there is substantial deviation

    from the assumptions, then neither the corrected te~peratures nor the

    original Na-K-Ca te~peratures will be of significance.

    t'1ixing lTJodels

    The springs near Corral Creek exhibit variations in surface

    te~perature, silica content, and other constituents, thus indicating

    that they may be of a mixed origin. The cold water was assu~ed to be

    BOC and to contain 6 mgtl silica, which is similar to several of the

    coldest springs in the vicinity of Ca~ral Creek. Reasonable mixinq

    models were calculated for six springs (fig. 9, table 5). The

    estimated temperatur~ of the hot water comoonent ranges from 47°C to

    67°C and the fraction of cold water varies from 0.30 to 0.77.

    Table 5. Mixing model results for six springs in the Corral Creek drainage, Blackfoot Res2rvoir region, southeastern Idaho (from figure 9).

    Spring Temperature of Hot ~~ater,OC Fraction Cold Hater

    Corral Creek 53 ± 17 .30 +- . 13 +Corral Cre~k " c:: 56 13 .36 +- .22 ...Corral Creek 3 64 ± 15 .56 - .11

    Carral Creek 6 47 t 10 .58 + .10 +

    ~p 104 67 t 30 .77 - .091....S? : / 60 + 20 .66 + .11

  • 1 J.j

    300r

    250

    A 200

    150 r 100,

    U 0

    50 I-z

    01w z 0 02 0 U

    B0:: w I

  • 54

    41 ' 300r

    I

    ::r 0 150

    10C

    . I

    U 0

    I Z W Z 0 0... ~ 0 U

    0:: W I « 3

    5 I

    l.J... 0 w 0:: --...J ~ 0:: W 0... ~ W l

    50

    0

    300\

    250

    200

    150

    I 100 L

    50

    0 1

    3:)Cr

    250~ 200'

    I I

    150,

    loot

    E

    F

    50

    01 0

    I I I

    01 02 03 04 05

    FRACT!ON OF COLD

    I I

    06 07

    WATER

    !

    08 I

    09 1

    10

    Figure 9 (continued). Mixino models for springs in the Corral Creek drainage, Blackfoot Reservoir region, southeasterr. Idaho: 0, Corral Creek 5; E, SP 104; F, SP 117.

  • , I I

    I

    Chemical Eguilibria

    Activity diagrams are used to interpret so~e of the chemical

    characteristics of the spring waters in relation to the stability of

    minerals which may be in contact with the water. Carbon dioxide has

    been shown to be an important component of the springs in the area;

    therefore, reactions between minerals have been written involving C02

    and H2 0 as reactants and bicarbonate (HC03-) as a product. All

    thermodynamic data is from Robie and others (1978). The reactions

    used to construct the activity diagrams are listed in table 6.

    ihe relative stability fields of phases in the

    plotted in figure 10. PC02 is assigned a value of 0.26 at~osDheres,

    which is an average value for the 8lackfoot Reservoir springs. Each

    dot represents an individual spring. All of the springs have silica

    activities bet\teen those of quartz saturation and amorphous Si02

    saturation, as is typical of most ground waters (Garrels and Christ,

    1965; Stu~ and Morgan, 1970). Most of the waters plot in the

    stability field of kaolinite, a mineral lacking potassiu~. The

    cluster of points near the musovite, microcline, kaolinite junction

    represent Corral Creek 1,2, 3, 4, and 5, SPI04, SP114, and SP117.

    ihe high concentrations of potassiu~ in these springs apoarently are

    due to equilibriu~ reactions bet~een the potassium-bearing minerals

    ~uscQvite and feldspar, and kaolinite. A trend of increasing

    potassiu~ and silica concentr1tions in the spring-water 3ctiviti2S

    within the kaolinite stability field, ~ay be jue to irreversible

  • 5G

    Table 6. Reactions used to establish stabilityrelations among minerals.

    2 KAl 3Si3010(OH)2 + 2 CO 2 + 5 H20 = 3 A1 2SizOs(OH)4 + 2 K+ + 2 HC0 3muscovite kaolinite

    2 KA1Si 30s + 2 CO 2 + 11 H2O = A1 2Si 20s(OH)4 + 2 K+ + 2 HC0 3 + 4H4SiO~ microcline kaolinite

    3 KA1Si 309 + 2 CO 2 + 14 H2O = KA1 3Si 30 10 (OH)2 + 2 K+ + 2 HC0 3- + microcline muscovite 6 H4Si0 4

    KA13Si3010(OH)z + CO 2 + 10 H20 = 3 Al (Orl)3 + K+ + HC0 3 + 3 HuSi0 4muscovite gibbsite

    AlzSizOs(OH)4 + 5 H20 = 2 Al(OH)3 + 2 H4 Si04 kaolinite gibbsite

    NaA1Si30a + 3 H20 = NaA1Siz06'HZO + H4Si04 albite analcite

    2 NaA1Si206"H20 + 2 C02 + 5 H20 = A1zSi20s(OH)4 + 2 Na+ + 2 HCO~ + analcite kaolinite 2 H4Si0 4

    NaA1Si20o"H20 + CO 2 + 5 H20 = Al (OH)3 + Na+ + HCO - + 2 H4SiO~ analcite gibbsite

    A1 2SizOs(OH)4 + 5 H20 = 2 Al(OH)3 + 2 H4Si0 4 kaolinite gibbsite

  • 57

    Table 6. Reactions used to establish stability relations among minerals (continued).

    CaO-MgO-H70-C~

    Mg4(C03)3(OH)2·3H20 = 4 Mg(OH)2 + 3 CO 2 hydromagnesite brucite

    ++ ++4 CaC0 3 + 4 Mg + 4 H20 = Mg4(C03)3{OH)2·3H20 + 4 Ca + CO 2 calcite hydromagnesite

    2 CaMg(C0 3 )z + 2 14g++ + 4 H20 = Mg4(C03)3(OH)203H20 + 2 Ca++ + CO 2 dolomite hydromagnesite

    4 MgC0 3 + 4 H20 = Mg4(C0 3)3(OH)z·3HzO + CO 2 magnesite hydromagnesite

    ++ ++CaC0 3 + Mg + HzO = Mg(OH)2 + CO 2 + Ca calcite brucite

    2 CaC0 3 + Mg++ = CaMg(C0 3)2 + Ca++ calcite dolomite

    CaMg(C0 3)z + Mg++ = 2 MgC03 + Ca++ dolomite magnesite

  • 58

    at 25°e, Ptotal= 1 atm, and P eoz = 0.26 atm.

    .I

    0

    -I

    -2

    -3

    ....,I 0 u -4 :I:--• -~ C' 0

    , '"

    -9', 0

    rl k

    Figure 10.

    MICROCllN£

    I I I

    MUSCOVITE I I~ I~ I~ I~ IN0I(j;

    1(1) I§

    Iz 1& 0I~

    @ 1 ~==

    I~ • IN

    00

    ..I~

    0

    1:5 ~:,0 0•

    o • I • 0 • 0KAOLINITE 0• o •0

    0

    I I . , 00 I I 0 I I II II II

    ! I I! -5 -4 -3 -2

    log ( H4 Si04 )

    I I I I

    Stability of phases in the system K20-Al z03-SiO:-H:C-CO:

    0

  • 59

    reactions of the water with muscovite and microcline.

    Stability relations of analcite, albite, kaolinite, and gibbsite

    in the syste~ Na20-A1203-Si02-H20-C02 at 25°C and 1 atmosphere are

    sho;m in figure 11. Pca is fixed at 0.26 at~ospheres. The 2

    activities of the springs fall well \~ithin the stability field of

    kaolinite, thus the amount of sodiu~ in the water does not aPDear to

    be controlled by the equilibrium reaction bet'tleen ;Ia-feldspar and

    kaolinite. The sodium could be the result of irreversible reactions

    of the water with tla-feldsoar or of reactions of other low-te~perature

    Na-bearing minerals such as s~ectites. The stabili~y field of

    ~ontmorillonite is not shJwn i~ this dia0ra~ because of variable

    co~position and lack of thermodyna~ic data.

    Fig,ure 12 sho~·ts the stable phases of ca1ciu:n and ;nagnesiuiTI

    caroonates at 25°C and 1 atmosphere. Spring-water activities fall

    near the ca1cite-do1o~ite-hydro~agnesitein~ersectior.. The abundance

    of carbonate rocks throughout the Blackfoot Reservoir reqion make it

    nigh1y probably th3t the calcium and magnesium in solution is derived

    from reactions with limestone, dolc;nite, and other carbonate rocks.

    As can be seen fro~ figure 6, the springs containing high amounts of

    bicarbonate also contain the most ca1ciu~ and ~agnesium. Bicarbonate

    acts as Gn acidifying agent which expedites the reaction rate between

    rock and \/ater (E 11; sand j'lahon, 1964).

  • • •

    60

    :3

    2

    ,.., -I I 0 U :I: -2

    + 0 Z

    -3 ~ 0

    'l -4

    ir

    J:.

    , .....h .-5, .. ...~~l: ..#1

    I : ••~l· -6 ~ . . ..,_1'

    KAOLINITE I

    I

    .. .,.

    ANALCITE

    lAJ...-en II) II)-(.:>

    Fig~re 11. Stability of phases in the system Na20-A1203-SiOz-H20-C02

    at 25°C, Ptota1= 1 atm, and = 0.26 atm. PC02

  • -----

    ......--,.,..."" •._".,,"' ... ,', ...:!':' .... ' -- ;r'~---~" -.~:: m,tj, VHim:':::::Jr:::""" ....,,! II·M llil ~ 1-."' .t ..

    5. \

    4

    3

    2

    +- +N N BRUCITEo 0\ U ~ - - 0

    0\ o

    -\

    -2

    -3"

    -4l I I

    I

    CALCITE

    ~~ I 4, <

    f~·

    DOLOMITE

    HYDROMAGNESI TE

    MAWES/TE

    I I I I I I I I I I,.

    -14 -12 -10 -8 -6 -4 -2 0 2 4 6

    log Peo 2

    Figure 12. Stabil j ty of phases in the system Ci:tO-~190-COrll::O at (J)25°C and 1 atlll pressure.

  • 1,$ I

    ·,.1 -.11'-"''''''

    ,.;.-.~::':' :;

    So~rce of CO~ era 4:S

    ~ll ~f the sa~pled sJrin~s in :~e Slc:kfoot ~eservoir area are

    characterized by lcr;e eGoun!s of carJon dioxice. There are ~lso a

    few areas of noticeable sulfur deoosition a~c H2S odor. Sulf~r

    Springs is located approximately four ~iles eas~ of Soda Sor;n~s on

    the western edge of the ~spen Ra~ge at the ~outh of Sulfur Canyon.

    Sulfur was produced here in the early nineteen-hundreds ~ut the plant

    was dis~antled before 192C. The sulf~r and assJc;ated s~a11 )YDSU~

    crysta1s occur as the ce~ent of a fault breccia ccnoosed of tuff,.

    limestone, and quartzite frag7.ents (Richards and Bridges, 1911).

    Hydrogen sulfide can also be detected in the Blackfoot Reservoir where

    it is associated with underwater springs and gas seeps.

    Sulfur-cemented rhyolite fraguents ~ere found on the small island

    south of Shee~ Island. Gas bubbles can be seen extending fro~ this

    island in a southeasterly direction, past the eas~ern ejge of Cinder

    Isl and to\/ard the southern end of the reservoi r. The eas:ern side of

    Cinder Island is a near-vertical cliff which appears to be a fault

    SCJrp. The northwest-southeast orientation of the islands and gas

    bubbles is si~ilar to that of other n~rmal faUlts to the south and

    west of the Blackfoot Reservoir.

    There are several ways to account for the oresence of hydrogen

    sulfide and carbon dioxide gasses. Previous authcrs (Mansfield, 1927,

    ~;...,.....Richards 3nd 8rid~es. 1911) have suggested a volcanic source for :....IC

    gasses due to the proxi~ity cf volcanic centers. The oossibility may

  • 1 63 not be entirely ruled out; however, it is far more likely that the

    gasses, in addition to the aqueous species discussed previously, are

    derived from reactions of water with the enclosing rocks.

    A simo1e model here proposed is the oxidation of organic ~atter

    (general formula CH 20) by the reduction of sulfate. The two half-cell

    reactions are (thermodynamic data from Stu~m and Morgan (1970)):

    A reduction process will tend to oxidize equi~olar concentrations of

    a~other redox process which has a lower pEo value. Thus, S04 2 - ~an

    oxidize CH20 to form CO 2 and H2S ~y the reaction:

    '1/8 S042 - + 1/4 CH20 + 1/4 H+ = 1/8 H2S(g) + 1/4 CO2 (g) +

    1/4 H2 0

    log K(25°C) = 6.45.

    To deter:nine \'Ihether the reaction is therllodynamically possible under

    actual c~nditi~ns, we can calculate the free-energy change for the

    reaction, ~G =2.303 RT log (Q/K). tG is the free-energy change for the reaction, K is the equilibrium constant at 25°C, and Q is the

    reaction quotient at the same temperature. t·le ~'/i11 assu~e that PC0 =

    f.

    2

    0.435 atm., the activity of H20 = 1, the activity of SOu = 5.82 X 10- 4

    and pH = 6.22 (from Sulfur Spr. 1, iJATEQF). Stur:m and ilorgan (1970, p. 233) sug~est that the concentration c~ orqanic matter in natural

    waters ranges from 0.1 to 10 mg/1. This corresponds to CH 20

    concentrations of about 10- 5 • 5 to 10- 3 • 5 ~o12s/l. In the Blackfoot

  • 64

    Reservoir area, the actual concentration is probably somewhere between

    the extremes, due to the abundance of organic-rich sedir.'lentary rocKs.

    Organic matter averages 2.1 weight percent of the Phosphoria For~ation

    (Gulbrandsen, 1967), whose thickness ranges fro~ approxi~ately 75 to

    150 meters in the Soda Springs area. Other black shale and li~estone

    units are scattered throughout the stratigraphic section. PHzS can be

    esti~ated between 10- 2 and 10- 8 atmosoheres for springs with a

    noticeable sulfur odor (Stu~m and Morgan, 1970, p. 364). For the

    reaction

    \~e can calculate a range of valuEs for Q from the above infor~ation: 11.l. 24

    Q = 10· . to 10 •

    10 51 6We know that K = • ; therefore,

    Q/K = 10- 33 to 10- 28

    and

    LG = 2.303 RT log (Q/K) = -51 to -38 Kca1/~01e.

    The oxidation of organic matter by su1fa:e is clearly favored

    thermcdynamical1y under the conditions given.

    Figure 12 shows a stoichiometric correlation between bicarbonate

    and su1 fate concentrati ons in \/ater fro~ the 81 ackfoot Reservoi r

    region. A least-squares linear regression technique apolied to the

    data yi~lds a linea~ equation of the fG~:

    Y = 2.8 X + 540

    with a correlation coefficient of 0.91. In the redox reaction above,

    the reduction of one mole of sulfate to one mole of hydrogen sulfi~e

  • 65

    10,OOO~--------------------/"""----

    / /

    C" 1-4/SP 104 ~/'

    SODA SPR. t/ SP 1/4C-:::< ··SP 1/7

    /- "",0

    /'

    • .. /' /'

    1000 I- ./ ........................ . .. _............... .-

    -------- - --- .. ·S? 100 .. ........

    0" 100 E Rh Spr.2 • Rh Spr. I

    1:

  • 66

    is accompanied by the oxidation of two moles of organic matter. The

    hydrogen sulfide that is produced, is probably oxidized under

    near-surface conditions to sulfate. The slope (~HC03/~S04) of 2.8 in

    the linear e~uation is correlative \Jith the stoichiometry of the redox

    reaction: one mole of H2S for two moles of CO z'

    Depending upon the oxidation potential, pH, and the presence of

    biological mediators, the oxidation of organic matter to CO 2 may also

    be accompanied by the reduction of oxygen, nitrogen, manganese, and

    iron species. The oxidation reaction most favored thermodynamically

    is:

    log K(25°C)=87.8

    (Stumm and t10rgan, 1970).

    To determine whether the reaction could occur in the \Vaters around the

    Blackfoot Reservoir, ~G must be determined. The concentr~ti0ns used

    in the calculation are PC02 ; 0.26 atm and (CH 20) = 10- 5 • 5 (~inimu~

    va 1ue) •

    A rough estimate of P02 for use in the calculation can be made

    at Sulfur Springs, where the amount of sulfate and pH is known, and an

    estimate (10- 2 to 10- 5 atm) can be made for PH S' Using the equatia~2

    1/8 S042- + 5/4 H+ + e- = 1/8 H2S + 1/2 H20

    \'/here

    pEO(25°C) = 5.25 (Stu~~ and Morgan, 1970),

    nE can be calculated,

    pE = pEa + log ([ox] /[re~J)

    = 5.25 + 1/3 leg [SO~:_] - 5/4 pH - li8 log PH2S

  • 67

    = -2.20 to -2.95.

    From the relation

    \-Jhere

    pEe (25°C) = 20.75 (Stum.-n and t1organ, 1970),

    the partial pressure of oxygen can be calculated from

    log PO = -4(pEO) + 2 log [H 2 0] + 4 pH + 4 pEz = -83.0 + 4 (6.22) + 4 (-2.20 to -2.95)

    = -67 to -70. iherefare, fer Sulfur Spring 1, Po is approximately 10-67 to 10-70

    2

    atmospheres.

    for the oxidation reaction

    CH 20 + 02(g) = H20 + CO 2(g),

    Q = 1064 ,

    Q/K = 10-24 ,

    and

    LG = 2.303 RT log (Q/K) = -33 Kca1/ncle.

    The oxidation reaction is favored ther~odynamica11y unGer these

    minimum conditions.

    S;~ilar calculations could be made for other redox precesses if

    the concentraticns and oxidation states of other elements and gasses

    were kno\'ln. Mitchell (1976) reports nitrate (NO~) UP to 0.4 m~/l and, ~.

    ammonid (NH3) up to 2 mg/1 in sprinq '..Jdters from this area, \,hich

    suggests that other reduction processes are also occurring.

    211ddingtorite, an ammc:1iuiTI f?ldspar, hcs been ieDcrted to occur in the

    ?hosohoria Formation in southeastern Idaho as an