Vous êtes sur la page 1sur 29

Journal

of
Hydrology
ELSEVI ER
[21
Journal of Hydrology 172 (1995) 31-59
Solute and isotopic geochemistry and ground water flow in
the central Wasatch Range, Utah
Al a n L. Ma y o a' *, Ma r k D. Lo u c k s b
aBrigham Young University, Department of Geology, P.O. Box 24646, Provo, UT 84602-4646, USA
bMontgomery Watson American Inc., 261 East Cranberry Hill Drive, Draper, UT84020, USA
Received 17 March 1994; revision accepted 8 March 1995
Abstract
Gr ound water flow systems in the rugged central Was at ch Range, Ut ah, were investigated by
solute and isotopic methods. Six types of gr ound water systems were identified on the basis of
rock type and structure. The systems are broadl y grouped i nt o four categories: (1) granitic
systems; (2) non-granitic systems: (a) unconsolidated alluvial systems, (b) consolidated
sedimentary rock systems, and (c) fault controlled systems; (3) thermal systems; (4) mine
drainage systems. These ground water systems have distinctive solute and isotopic
chemistries. Based on an analysis of 3H and 14C data, all of the ground waters have a
c omponent of post-1952 recharge water, and mos t of the gr ound waters are compos ed almost
entirely of post-1952 recharge water. A few of the gr ound water systems cont ai n some water
which is hundreds t o perhaps t housands of years old. The/~lSo and/~2H dat a pl ot on a local
meteoric water line, and none of the dat a exhibit a positive 61So shift. The absence of a positive
6[ So shift suggests maxi mum aquifer temperatures are approximately 100C.
Granitic terrains are domi nat ed by fracture controlled, local gr ound water flow systems
which respond rapidly t o recharge events. The granitic gr ound water systems have estimated
maxi mum circulation depths of about 160 m below land surface. All the alluvial, mos t of the
consolidated sedimentary bedrock, and some of the fault controlled gr ound water systems are
shallow circulating, local flow systems with short travel times. Gr ound water circulation of
mos t of the local flow systems is within 200 m of land surface and is bedding plane controlled.
Some of the consolidated bedrock and many of the fault controlled ground water systems are of
the intermediate type and have circulation depths t o 500 m. Four thermal gr ound water systems
have been identified. The thermal systems are heated by the geothermal gradient and have not
circulated deeper t han 2- 2. 5 kin. The thermal systems are of the intermediate and regional type.
* Corresponding author.
0022-1694/95/$09.50 1995 - Elsevier Science B.V. All fights reserved
SSDI 0022-1694(95)02748-3
32 A.L. Mayo, M.D. Loucks / Journal of Hydrology 172 (1995) 31-59
T h e solute chemistry o f m o s t o f the non-granitic g r o u n d waters m a y b e attributed to the
dissolution o f c a r b o n a t e minerals a n d m i n o r a m o u n t s o f g y p s u m . Contributions o f external
CO2(s) to s o m e o f the t h e r m a l systems f r o m either m e t a m o r p h i c o r silicate hydrolysis processes
is suspected.
1. Introduction
The Wasatch Range, Ut ah, lies along the boundary between the Middle Rocky
Mount ai n and Basin and Range physiographic provinces. Al t hough the numerous
springs in the range provide the base flow for i mport ant Ut ah streams and rivers,
previous investigations (Milligan et al., 1966; Mundorff, 1971; Fuhri man and Merritt,
1976; Kohler, 1979; Cole, 1983; Homes et al., 1986; Ut ah Depart ment of Nat ural
Resources, 1987; Roark et al., 1991; Mayo et al., 1992) have provided little insight
into the overall character of the ground water flow systems. The purpose of this
investigation is to identify and characterize ground water flow systems in the central
Wasatch Range. This investigation is limited to t hat port i on of the central Wasatch
Range which is bounded by Interstate 80 (I-80) to the north, Provo Canyon to
the south, Salt Lake and Ut ah Valleys to the west, and Heber Valley to the east
(Fig. 1).
2. Methods o f study
Field investigations included locating spring and mine discharge sites, estimating or
measuring discharge rates, evaluating the structural and stratigraphic relationships at
each site, and collecting wat er and rock samples for laboratory analysis. Wat er
quality samples were collected from 88 cool and warm wat er sites. Temperature,
pH, Eh, and bicarbonate were determined in the field. Solute water samples were
collected and preserved according to standard procedures (US Environmental
Protection Agency (EPA), 1981) for l aborat ory analysis.
Stable isotopic compositions are reported as the per mil (%o) difference of the
sample relative to St andard Mean Oceanic Wat er (SMOW), for 62H and 6180, Pee
Dee Format i on Belemnite (PDB) for 613C, and the Canyon Diablo meteorite (CD)
for 634S. 14C values are reported r.elative to per cent modern carbon (pmc) and 3H
values are reported in tritium units (TU).
Saturation indices were calculated using the comput er code WATEQF (Plummet et
al., 1976). Gr ound water flow systems are described in terms of concepts of local,
intermediate and regional flow as defined by Tot h (1963).
3. Geologic setting
The rugged central Wasatch Range rises more t han 2000 m above the adjacent
A.L. Mayo, M.D. Loucks / Journal of Hydrology 172 (1995) 31-59 33
valleys and is domi nat ed by pronounced topographic relief. The range is geologically
complex, composed predominantly of faulted and folded Paleozoic carbonate rocks
and Tertiary granitic intrusion (Fig. 2). Precambrian basement rocks contain as much
as 3050 m of thickly bedded quartzite and conglomerate with interbedded calcareous
shale and slate (Davis, 1983). Gently to tightly folded Paleozoic strata, which are
highly dissected by faults, crop out over most of the region. Predomi nant lithologies
are limestone and dolomite. Calcareous shale and quartzitic sandstone occur less
' 44
V
20O
39 41
V V V 4 2 ~ 3
,....40
3~7 62 (
r
II
STUDY
AREA
UTAH
Salt Lake
Valley
Ut ah
Valley
18 '
81
8 0 1
60
17
V
8 2
8 3 7 6
30
V
4766
V6 5 4
67 52
5 3 ~ '
1 4 9
V
50
V
I I
57 2 62 4
V5 6 VV
22 ~ V
)13 V ~ 25
) 1 4
I I (
61~
71'
Heber
Valley
Explanation
Sample Locations
0 Gnmitic Sys~am
UncomolJda~d
Alluvial Systems
V Consolklated Sed~nzntal'y
Rock Systems
Fault C o n t r o H ~ l
S y s t e m s
[ ] The/'mal Systems
Mine Drainage Sysh.ms
~ R o a d s ,
~ d g e of Mountain
Front
0 1 2 3 4 5
Scale in Kilometers
Fig. 1. Index map of t he study area in t he central Was at ch Range, Ut ah. The study area is bounded by
Parley' s Canyon t o t he nort h, Provo Canyon t o t he south, Salt Lake and Ut ah Valleys to t he west and
Parley and Hebcr Valleys t o t he east.
34 A.L. Mayo, M.D. Loucks / Journal of Hydrology 172 (1995) 31-59
often (Bromfield et al., 1970). Mes ozoi c strata are compos ed o f sequences o f
l i mestone, dol omi t e, sandst one and shale. Tertiary rocks include quartz monzoni t e t o
diorite stocks and tuffaceous sandst ones and congl omerates o f the Tibble Fork
Format i on. The Tibble Fork Format i on is as much as 760 m thick and locally
cont ai ns l i mestone lenses ( Davi s, 1983). Pleistocene glacial morai nes are l ocated in
many o f the hi gh mount ai n valleys and are predomi nantl y compos ed o f carbonate,
quartzitic or granitic clasts ( Davi s, 1983).
Hne$
u m
F i l l .
rm.
~cks
~m.
~m.
ian
F a u l t
u l t
~Jon
0 I 2 3 4 5
Scale in Kilometers
Fi g . 2. Ge ne r af i z e ~l g e o l o g i c ma p o f t he c e n t r a l Wa s a t c h Ra n g e , Ut a h .
0
0 ~
0
"1:
"d
. g
c ~
z
e
r~
r~
ad
I m
A.L. Mayo, M.D.~ucks/~nal ~ol ogyl ~(1995) 31-59 35
VVVVV V VVV VV V V VVV
V V V V V V VVV V
e e e e e e e N N . e e ~NNNNNeNNNN ee NNNNN~NNN
V V V V V V V V V V O0 V O V V O V O V V V V O0 O V V V V V V V V
V V V V V V V V V V V V V V V V V V V V VV VVVV
V V V V V V V V V V V V V V V V V V V V V V V V V V
. . . . . . . . . . . o o ~ o o o ~ o ~ o oo o ~ o o o ~ o o o
V V V V V V V V V V VV V V V V V V V V V V V V V V V V
V V V VV
~ ~ ~ ~ ~ ~ oo~ ~ ~ . . . . . . . . .
VV ~V
V
. . . . . . . . . . . . o o o o o o o o o
VV V V
~ - ~ ~
. . . . . . . . . . . . 0 0 0 0 ~ 0 0 0 0 ~ 0 ~0 ~ o o Z o 0 0 0 0 0
~ ~ o o o ~ ~ ~ N ~ N ~ . . . . . = . .
36 A.L . Mayo,
~d
0
e~
e~
M.D. L oucks / J ournal of Hydrol ogy 172 (1995) 31- 59
i
U
~ _ _ _ . _ _ . ~ . ~ m ~ m m m . . . . ~
. . . . . . . ~ _ - ~
V V V V VVV V V
V V V V V V V V V V VV V V V
v ' ~ v ' v ' v ' ~ v ' ~ v ' o ~ v ' ~ v ' v ' ~ o o v ' ~ v ' O~ v '
VVVVVVVV VVVV VV VVVV V VVVVV
VVVVVV V V V V V V V V V V V V V V V V VVVV
e ~ e e ~ e NN~ e e ~ e e N~ N~ N~ e~ ~ N~ N~ N
V V V V V V V V VVVVVV VVVVV V VVVV
v v V v
V V V v V v V
V V V V V V V
~ - ~
A.L. Mayo, M.D. Loucks / Journal of Hydrology 172 (1995) 31-59 37
v V v
V V V V V V
V V V V V V V V O0 0 0 0 0 0 V O0 V V V O V V V O V V V V V V O0
V VV V V V V V V V V V V V V V V V V V V VV
V V V V V V V V V V V V V V V V V V V V V V V V V V
V V VVVV V VVVV V V V V V V V V V V V V V
V V V V V V V V V
o ~ - ~ ~ ~ o ~ o ~ ~ ~ o o ~ o o ~ o o o o ~ ~ -
V V V V
~ ~ . .
38 A. L. May o, M. D. Loucks / Journal o f Hydrol ogy 172 (1995) 31- 59
0
. o
0
0
0
8
o
o
t ~
I I I I I I
1
I
q ~ q q
I I I I 7
I I I I I I
I I I I
I I I I I
I
o o o o o o o d o o ~d
I I I I I I
A. L. Mayo, M. D. Loucks / Journal of Hydrology 172 (1995) 31-59
4. Ground water flow systems defined
39
Six types of ground water systems have been identified in this study on the basis of
rock type, geologic structure, discharge temperature, and discharge location. The six
system types are broadly grouped into four categories: (1) granitic systems; (2) non-
granitic systems: (a) alluvial systems, (b) consolidated sedimentary bedrock systems,
and (c) fault controlled systems; (3) thermal systems; (4) mine drainages. Waters from
the six types of ground wat er systems may be distinguished from each other on the
Table 3
Isotopic compos i t i ons and ages of gr ound waters in the central Was at c h range, Ut ah
62H (%0) 6180 (%0) 613C (~oo) 6348 (~oo) 14C (pmc) 3H (TU) Age (years)
Granit~ systems
3 - 1 2 9 - 1 6, 8
6 - 1 2 6 - 1 6. 0 - 1 4. 3
7 ~124 - 1 5. 7
8 - 1 2 7 - 1 6. 2
Mean - 1 2 7 - 1 6. 2 - 1 4. 3
SD 2.1 0.5
Unconsolidated alluvial systems
15 - 1 3 2 - 1 6. 4 - 6. 3
Consolidated semimentary rock systems
22 - 1 2 2 - 1 4. 9
23 - 1 2 3 - 1 5. 2
24 - 1 2 7 - 1 6. 2
27 - 1 35 - 1 6. 9
28 - 1 2 8 - 1 6. 3
30 - 1 31 - 1 6. 8
31 - 1 3 2 - 17. 1
32 - 1 2 8 - 1 6. 6
36 - 1 4 0 - 1 7. 2
41 - 1 2 8 - 1 6. 4
42 - 1 3 4 - 1 7. 0
44 - 1 2 3 - 1 5. 9
45 - 1 2 2 - 1 5. 0
50 - 1 2 7 - 1 6. 4
Mean - 1 2 9 - 1 6. 3
SD 5.4 0.8
Fault controlled systems
54 - 1 31 - 1 7. 2
55 - 1 2 7 - 1 6. 4
59 - 1 3 0 - 16. 1
62 - 1 3 7 - 1 7. 4
66 - 1 2 9 - 1 6. 8
67 - 1 2 2 - 1 5. 3
68
Mean - 1 2 9 - 1 6. 5
SD 4.9 0.8
- 10. 1
- 9. 8
- 1 0. 3
- 8 . 2
- 9 . 8
- 9 . 6
0.8
- 1 0. 0
6.2
6.2
74.7 59.3 Mode m
74.7 59.3
43.7 1 7 . 3 Mi xed- Moder n
71.0 39.9 Mode m
63.4 33.4 Mode m
85.6 52.5 Mode m
57.3 1 4 . 1 Mi xe d- Mode m
74.9 29.4
70.4 33.9
10.9 14.1
- 7 . 5 8.7 47.3 22.4
- 1 0. 9 54.5 38.1
- 8 . 4 - 3. 8 57.9
- 9 . 2 2.5 53.2 30.3
1.5 8.8 5.4 11.1
Mode m
Mi xe d- Mode m
Mode m
40 A.L. Mayo, M.D. Loucks / Journal of Hydrology 172 (1995) 31-59
bas i s o f t ot al di s s ol oved s ol i ds ( TDS) , Mg 2+, SO 2 - a n d CI - c onc e nt r a t i ons , Ca : Mg
r at i os , a n d cal ci t e a n d dol omi t e s a t ur a t i on i ndi c es ( Tab l es 1 a nd 2).
5. Isotope geochemistry
5. 1 . 6 1 8 0 and 62H
Gr o u n d wa t e r 6180 a n d 62H val ues r ange f r o m -17. 7/oo t o - 1 3. 6%o a n d f r o m
- 1 40%o t o - 1 1 5%o, r es pec t i vel y ( Tab l e 3). Two gener al ob s e r va t i ons ma y be ma d e
r e ga r di ng t he 6180 a n d 62H da t a : (1) t he d a t a pl ot a l ong a l oc al me t e or i c wa t e r l i ne
( MWL ) ( Fi g. 3); (2) i s ot opi c al l y l i ght g r o u n d wat er t ends t o oc c ur we s t o f t he
Wa s a t c h Ra n g e t o p o g r a p h i c di vi de, wher eas i s ot opi c al l y he a vy g r o u n d wat er t ends
t o oc c ur eas t o f t he di vi de ( Fi g. 4).
5. 1. 1. Loc al me t e or i c wat er l i ne
St abl e i s ot opi c c o mp o s i t i o n s o f c ent r al Wa s a t c h Ra n g e g r o u n d wat er s have b een
pl ot t e d ( Fi g. 3) wi t h r es pec t t o t he c ont i ne nt a l M WL def i ned as
Table 3 (Continued)
82H (%0) 6180 (%0) 6t 3C (%0) 634S (%0) t4C (pmc) 3H (TU) Age (years)
Thermal systems
69 -132 -17. 3 -2. 5 13.6 5.6
70 -136 -17. 5 -3. 4 10.4 10.6 7.8
72 -140 -17. 4 -10. 7 4.6
73 -132 -17. 2 -2. 7 15.0 5.4
74 -115 -13. 6
Mean -131 -16. 6 -4. 8 10.9 7.2 7.8
SD 9.5 1.7 3.9 4.6 2.9
Mine drainage systems
75 -134 -17. 2
76 -134 -17. 0 -10. 2 -1. 1 96,2 36.8
78 -126 -15. 5 -18. 0 -0. 2 36.5
81 -128 -16. 3 1.6
82 -125 -16. 3
85 -138 -17. 3
87 -128 -16. 2 -9. 3 -2. 1 47.4 40.8
88 -139 -17. 7 -8. 5 6.4 46.2 20.9
Mean -132 -16. 7 -11. 5 0.9 63.3 33.8
SD 5.5 0.7 4.4 3.4 28.5 8.8
Rock 13 C
Cambrian Ophir Fm. -2. 7
Mississippian Fitchfield Fm. 2.4
Pennsylvanian Oqurrih Fm. 2.4
Triassic Thaynes Fm. -0. 9
NE
Modern
Modern
Mixed-Modern
Mixed-Modern
Age: NE (not estimated); Modern ( < 40 years); Mixed-Modern (mixed < 40 years with older water).
A.L. Mayo, M.D. Loucks / Journal of Hydrology 172 (1995) 31-59 41
O
O
-1-
- - 1 0 0
- 1 1 0
- 1 2 0
- 1 3 0
- 14- 0
- 1, 50
- 1 8
I I I
0 Gr a n i t i c S y s t e ms
A l l u v i a l S y s t e ms
V C o n s o l i d a t e d B e d r o c k S y s t e ms / .
F a u l t S y s t e ms / -
[] T h e r ma l S y s t e ms
Mine Discharges /
~ 0 J r ~ '''~ _ _ _
z ~ ~ ~ 6 / ~ ' ~ ~ ~ ~ 0 V V V
6 0
0 [ 3 V
~ V V
[3 v
[] V
, , , I , , , I , , , I ,
- 1 7 - 1 6 - 1 5
~lao (o/oo)
Fi g. 3. Scat t er pl ot of 62H vs. 6180 relative t o t he MWL.
\ ,
"
818o c ~w, arl
c ~ t ~ lm~tval
i s -1 ~,
S c ~ i| m ~
Fi g. 4. Spat i al di s t r i but i on of 82H and 6~80. Mean i s ot opi c c ompos i t i ons are: wes t er n sl ope (62H, - 126.2%o,
t518 O, - 16.0%o), eas t er n sl ope (62H, - 133.0%o, 61SO, - 16.99~), Amer i c an For k Canyon (62 H, - 128.6%o,
di '8 O, - 16.4%~).
42 A.L. Mayo, M.D. Loucks / Journal of Hydrology 172 (1995) 31-59
62H = s(6180) + d(%o), where s is the slope and d is the deuterium excess (Merlivant
and Jouzel, 1983). Values of s = 8 and d = 10 have on the global scale been shown to
approximate coastal meteoric water (Craig, 1961; Dansgaard, 1964) and are used to
define the MWL.
Central Wasatch Range dat a plot subparaUel to the MWL along a local MWL
defined as
62H = 86180 + 1.9%o
(1)
Local MWLs (i.e. isotopic compositions t hat do not have d values of ten) are common
in the western interior of the USA (White and Chuma, 1987; Mayo and Mitchell,
1988) and elsewhere (Bath, 1983; Gut, 1983; Moser et al., 1983). Paleoclimatic con-
ditions have generally been attributed to these 62H depletions. However, elevated
tritium concentrations (Tables 2 and 3) and the wide distribution of stable isotopic
compositions of the consolidated bedrock and fault controlled ground waters along
the MWL (Fig. 3) suggest t hat (1) paleoclimatic effects are minimal in ground waters,
and (2) there is minimal significance in the differences between the mean stable
isotopic compositions of the various ground water system types.
Continental fractionation, the combined effects of inland rainout and the precipi-
tation of isotopically light re-evaporated fresh water has resulted in variations in d
values less t han ten in many continental regions (Dansgaard, 1964), and is the likely
cause of the observed d value in the central Wasatch Range.
5.1.2. Isotopic distribution in the central Wasat ch Range
Isotopically heavy ground water (mean 6180 = -16.0%o, 62H = -126.2%0) tends to
occur on the western side of the topographic divide of the central Wasatch Range,
whereas isotopically light ground water (mean 61aO =-16. 9%o, 62H -133%o)
tends to occur on the eastern side of the topographic divide (Fig. 4). Western and
eastern side ground waters are statistically different based on a t-test at a 95%
confidence interval.
A third geographic group of isotopic dat a were collected from the northern end of
American Fork Canyon (Fig. 4). These sites are enclosed by topographic divides to
the west and east and have mean 6180 and 62H values of -16.6%o and -128.6%o,
respectively.
The 6180 empirical temperature scale of Dansgaard (1964) has been used to
estimate the atmospheric temperature at the time of precipitation formation:
6180 + 13.6%o
C = (2)
0.695
Because Eq. (2) was not developed for the mountainous region of the western USA, it
cannot be used to calculate absolute precipitation temperatures. Instead, it is useful
for approximating differences in precipitation temperatures. Relative recharge
temperatures calculated from Eq. (2) indicate t hat the estimated mean recharge
temperature of western slope ground water is 1.5C higher t han the estimated
mean recharge temperature of eastern slope slope ground waters. Gr ound water in
A.L. Mayo, M.D. Loucks / Journal of Hydrology 172 (1995) 31-59 43
the American For k area has an estimated mean recharge temperature only 0.6C
higher than the estimated mean for eastern slope ground waters.
Orographic effects, including storm morphology, local topography, and micro-
climate, appear to be largely responsible for the spatial distribution of mos t observed
isotopic values. The isotopes 180 and 2H are preferentially ' rained out ' of the air
masses as storms pass over western slope areas. As the 180 and 2H depleted air masses
then pass over the topographic divide and descend into lower elevation canyons on
the eastern slope area, the resulting isotopic composition of precipitation over the
eastern slope becomes more negative with respect to 61So and 62H.
Not all dat a are consistent with the propos ed orographic effect. It is possible that
the isotopic compositions of a few ground waters have resulted from (1) recharge
occurring during different paleoclimatic conditions, or (2) seasonal recharge effects.
The calculated recharge temperatures for Samples 22, 36, 62, and 74 are a few degrees
lower or higher than the calculated mean recharge temperatures of water from their
respective t opographi c locations. These differences may suggest the presence of some
paleo ground water which recharged during a cooler climatic period (Samples 36 and
62) and ground water which was recharged during a warmer climatic period (Samples
22 and 74). The anomal ous calculated recharge temperatures cannot be attributed to
seasonal isotopic variations in modern meteoric water because ground water recharge
on the western and eastern sides of the range is subject to similar spring snowmelt and
summer t hunderst orm conditions.
5.2. 6 1 3 C
Mos t st udy area ground water discharges have log PCO2(g ) and 613C values (Tables
2 and 3) which are consistent with the dissolution of carbonat e minerals in the
presence of soil zone CO2(g ). However, some of the thermal ground waters have
elevated log PCO2(g) values and 813C values that are more positive than woul d be
anticipated. Calculated PCO2(g) values of thermal ground waters range from
logPCO2(g) of - 0. 33 to - 3. 1 2 atm, and 613C values range from - 2 . 5 to -18.0%o
13
(Table 3). Fact ors contributing to anomalous PCO2(g) and 6 C values are described
elsewhere in this paper.
5.3. 6S4 S
Al t hough the sulfur isotopic dat a are limited, the dat a fall into five categories which
are related to the aquifer system types (Table 3):
(1) mine drainages with 634S values of 2.0%o or less.
(2) Mine discharges with 634S values of +4. 6% and +6.49/00 (Sites 72 and 88,
respectively). Site 72 has been classified as a thermal system; however, it discharges
from a mine.
(3) Ground water discharging from the Weber Quartzite and from a faul t -carbonat e
rock system with 834S values of +6.2%0 and +8.7%O (Sites 36 and 66, respectively).
(4) A fault-controlled ground water discharge with a 634S of -3.8%0.
(5) Thermal ground waters with ~345 values of 10.0%o or more.
44 A.L. Mayo, M.D. Loucks / Journal of Hydrology 172 (1995) 31-59
The 634S dat a suggest t wo different SO 2- sources: (1) dissolution of gypsum or
anhydrite; (2) oxidation of sulfide minerals. Positive 6345 values greater than about
+40/00 are attributed to sulfate mineral dissolution. The anticipated range of ~34S
values in Paleozoic gypsum and anhydrite in the central Wasat ch Range is +10 to
+20O/~ (Holser and Kaplan, 1966). At typical central Wasatch Range non-thermal
aquifer temperatures, isotopic fractionation accompanying gypsum dissolution may
be represented as
345 >
CaSO4(s) enrichment Ca2+ + SO 2_ (3)
(634S ~ +15/oo) (~34S ~ +9%o)
where 634S = +15%o has been arbitrarily assigned.
Isotopic compositions less than about +4%o appear to result from the sulfide
mineral oxidization. Typical 634S values of magmatic pyrite are about 0%o (Faure,
1986). A 634S of -2.2%0 has been reported for pyrite in the Park City District (Thode
et al., 1961), and Mayo and Kl auk (1991) found a me a n ~34S of +1.3%o in ground
water from crystalline rock aquifers in nort h-cent ral Utah. At low temperatures,
isotopic fractionation accompanying sulfide mineral oxidation is minimal and the
resulting SO42- will have a 634S value similar to the mineral phase.
The odor of H2S(g) at many thermal and at some other ground water discharge
locations suggests sulfate reduction. The isotopic compositions of these waters will be
more positive than those of the non-reduced water deeper in the aquifer (Kaplan and
Rittenburg, 1964).
5.4. Ground water age
Gr ound water 3H levels range from 7.8 to 59.3 TU (Table 3). These dat a suggest
that all ground waters sampled for 3H contain some modern recharge water and
that some ground waters are composed mostly of modern recharge water (i.e.
post-1952).
Carbon-14 activities of ground waters range from 5.4 to 96.2 pmc (Table 3). It was
not possible to calculate absolute radi ocarbon ages for ground waters in the central
Wasat ch Range because: (1) many samples with a low 14C pmc contained an appreci-
able component of modern recharge as evidenced by elevated 3H levels; (2) some
samples, particularly the thermal water sample, have 613C values indicative of an
external source of ' dead' carbon (i.e. 613C is less than approximately +4%o), which
cannot be accounted for by 14C ground-water age estimating equations; (3) many
samples have 14C pmc contents of some 60 pmc or more, which are indicative of
pos t q952 recharge water (Fontes and Garnier, 1979).
A qualitative analysis of each radi ocarbon sample is listed in Table 3; in this table,
' modern' means the ground water received most of its recharge since 1952 and
' mi xed-modern' means the ground water contains some modern recharge and
some pre-1952 recharge. It is likely that some of the the pre-1952 recharge occurred
hundreds to possibly thousands of years ago.
A2L. Mayo, M.D. Loucks / Journal of Hydrology 172 (1995) 31-59
6. Grani t i c ground wat er s y s t e ms
45
Gr ani t i c gr ound wat er fl ows ar e of t he l ocal t ype whi c h ar e pr e domi na nt l y f r ac t ur e
c ont r ol l ed. Di s c har ges oc c ur f r om s eepi ng j oi nt s and as s mal l di s c har ge s pri ngs f r om
f r ac t ur es or f r om t he r es i dual gr ani t e over b ur den. The me a n di s c har ge r at e is 4. 41 s - l ;
however , di s c har ge r at es as hi gh as 11.3 1 s -1 oc c ur ( Tab l e 1).
Thes e s ys t ems ar e gr eat l y affect ed b y s eas onal r ec har ge f r om s nowmel t and
t hunder s t or ms , and s ome s pr i ngs dr y up dur i ng t he s umme r mont hs . A c or r el at i on
o f 1980-1983 t hunde r s t or m event s wi t h s umme r and a u t u mn di s c har ge dat a f or Site 6
s ugges t s t hat a b o u t a 1 mo n t h l ag b et ween ma j or r ec har ge event s is t he t i me r equi r ed
f or t he c hange in pr es s ur e t o r eac h t he di s c har ge l oc at i on (Fig. 5).
Mos t di s c har ge t emper at ur es ar e l ow ( bel ow 8C). A ma x i mu m dept h o f c i r c ul at i on
is es t i mat ed t o be a b out 160 m, b as ed on di fferences bet ween me a n annual ai r
" o,
o
30
i
!
~ 20
Q,
0
.~ o
J F MA MJ J A S OND [ J F MA MJ J A S OND [ J F MA MJ J A S OND[ J F MA MJ J A S OND
1980 1981 1982 1983
Fig. 5. Bar graphs of average monthly ground water flow rates from Site 6 (inside the Later Day Saints
(LDS) storage vault) in Little Cottonwood Canyon and monthly precipitation data collected in the Little
Cottonwood Canyon about 0.6 km from the vault.
46 A.L. Mayo, M.D. Loucks / Journal of Hydrology 172 (1995) 31-59
temperatures and discharge temperatures, and a geothermal gradient of 40C km -1
(Kohler, 1979; Cole, 1983). 3H and 14C contents in Sample 6 indicate the presence of
appreciable quantities of post-1952 recharge water.
Gr ound waters discharging from the granitic terrains are of the calcium-
bi carbonat e type, and have low TDS concentrations (Table 1). The overall solute
chemistry of the granitic terrain waters is consistent with granitic terrain ground
waters found elsewhere (White et al., 1963; Fet h et al., 1964). The 613C value of
-14.3%o and logPCO2(g) of - 2. 01 atm of Sample 6 (Table 3) suggest that soil zone
gas is the principal cont ri but or of the dissolved carbon.
Elevated F- concentrations occur in ground waters at Sites 8, 9, and 10 (Table 1),
which issue from the Little Cot t onwood Stock near the Paleozoic sedimentary rock
contact. Dissolution of hydrothermally deposited veins which contain fluorite may be
responsible for the elevated F - levels. Fluorite has been identified in the nearby
American For k Mining District (Calkins and Butler, 1943).
7. Non-granitic ground water systems
The non-granitic ground water systems include unconsolidated alluvial systems,
consolidated sedimentary rock systems, and fault controlled discharge systems.
Gr ound waters in glacial moraines are classified as alluvial systems (Fig. 2). Alluvial
systems have a cont i nuous hydraulic connection with the atmosphere. Discharge
rates are commonl y less than 5 1 s -! (Table 1), and these rates decline noticeably
during the dry season. Sites 11 and 15 have unusually high discharge rates and have
corresponding large recession characteristics during dry summer months.
Consolidated sedimentary rock ground water systems occur in Paleozoic and
Mesozoic age strata and the Tertiary age Tibble For k Format i on (Fig. 2). Gr ound
water discharges typically occur from bedding planes and minor fractures. Mos t
springs discharge from carbonat e and coarse clastic sequences which directly overlie
shales or other rock units having low hydraulic conductivities, or where the carbonate
and clastic sequences are highly fractured. The mean discharge rate is 13.4 1 s -1 and
discharge rates tend to be seasonally stable.
Discharge locations of many of the consolidated sedimentary ground water systems
are controlled by thrust and extension faults. Such ground water systems have highly
variable discharge rates, ranging from 0.3 1 s - t at Sites 55, 58 and 65 to 240.7 1 s -z at
Site 54 (Table 1). There does not appear to be a correlation between discharge rates
and fault types. At many discharge locations travertine deposits occur. Travertine
thickness ranges from mi nor encrustation on rocks surrounding the discharge area to
a terrace covering about 200 m 2 at Cascade Spring (Site 54).
7.1. Chemical character
Non-granitic ground water systems are predominantly of the cal ci um-bi carbonat e
and cal ci um-magnes i um-bi carbonat e types (Table 1). A few ground waters have
elevated sodium concentrations. Each of the types of non-granitic ground water
47
2500
system types tends to have slightly different solute chemistries. The alluvial ground
waters tend to have lower Mg 2+ and SO 2- contents than do the consolidated bedrock
ground waters. The fault controlled ground waters tend to have the highest SO 2- and
Mg 2+ concentrations of all the non-granitic ground water systems (Table 1).
Cumulative frequency distribution plots of TDS exhibit similar straight-line trends
for each of the non-granitic type systems, suggesting similar salinity controls (Fig. 6).
Solute compositions are consistent with the dissolution of magnesium-bearing lime-
stone containing varying amounts of gypsum. Gypsum and dolomite dissolution are
reflected in the SO ] - and Mg 2+ concentrations. Ground waters from Sites 36 and 66
have 634S values (+6.2%0 and +8.7%0, respectively; Table 3) which are consistent with
gypsum dissolution. Fault controlled discharge (Site 68) has a 634S of -3.8%0, which
may be the result of sulfide mineral oxidation; however, sulfate reduction may also be
a factor.
The cumulative frequency distribution of Ca 2+ : Mg 2+ molar ratios shows that
none of the systems recharge in pure dolomite and that the consolidated bedrock
and the fault controlled systems have a greater contact with dolomite than do the
alluvial systems (Fig. 7). A ratio of unity indicates the dissolution of a pure dolomite,
whereas higher ratios are indicative of greater calcite contributions. The Ca 2+ : Mg 2+
molar ratios of some waters have been influenced by the dissolution of gypsum.
However, when potential Ca 2+ contributions from gypsum dissolution are
subtracted from measured Ca 2+ values the molar ratios remain above unity.
Most of the alluvial and consolidated bedrock aquifer waters are undersaturated
2250
2000
1750
1500
1000'
750
500
250
0
I I I I I
7
. J
~3n
E
6 O
r-~
A. L. Mayo, M. D. Loucks / J ournal of Hydrology 172 (1995) 31-59
Gr ani t i c Syst ems
Unconsol i dat ed Al l uvi al Syst ems /
Consol i dat ed Sedi ment ar y Rock Syst eys
Fault Cont r ol l ed Syst ems
Thermal Syst ems
S
Mi ne Syst ems
I I I I I
10 .30 5 0 7 0 9 0 99
Pe r c e n t a g e of s a mpl e s whose val ue is equal to or l ess t han t he val ue shown
Fi g. 6. Cumul at i ve f requency di st ri but i on o f TDS i n ground wat er sampl es f rom t he cent ral Was at c h
Range, Ut ah.
48
15
A.L. Mayo, M.D. Loucks / Journal of Hydrology 172 (1995) 31-59
I I I I I
14 o Gr ani t i c Syst ems
13 Unconsol i dat ed Al l uvi al Syst ems /
12 v Consol i dat ed Sedi ment ar y Rock Syst ems /
Faul t Cont r ol l ed Syst ems ]
11 o
Thermal Syst ems / / '
o /
~o 1 0 M i n e S y s t e m s / / /
ne 9
a
_B
1
o
1 1 0 30 50 70 9 0 9 9
Per cent age of sampl es whose val ue is equal t o or l e s s t han t he val ue s h o w n
Fig. 7. Cumulative frequency distribution of Ca:Mg molar ratios of ground waters from the central
Wasatch Range, Utah.
wi t h respect t o bot h cal ci t e and dol omi t e. Ab o u t one- t hi r d o f t he f aul t cont r ol l ed
gr ound wat ers are sat urat ed t o super sat ur at ed wi t h respect t o cal ci t e ( Fi g. 8),
al t hough car bonat e aqui f er gr ound wat ers are of t en cal ci t e under sat ur at ed
( Ha r mo n et al ., 1975). The sat ur at i on and super sat ur at i on o f some f aul t cont r ol l ed
gr ound wat ers is at t r i but ed t o t he r api d ascensi on o f sl i ght l y war mer gr ound wat er
al ong f aul t s. The sl i ght el evat i on o f di scharge t emperat ures is pr obabl y t he r esul t of
gr ound wat er ci r cul at i on whi c h is deeper t han t hat o f ot her non- gr ani t i c system
gr ound waters.
7.2. Gr ound wat er ci rcul at i on
All the alluvial, most of the consolidated bedrock, and some of the fault controlled
ground water systems are shallow circulating local type systems. The discharge rates
of the local flow systems respond to annual recharge events, and their 3H and 14C
contents indicate appreciable recharge since 1952. Circulation depths of most local
flow systems are less t han 200 m. The met hod for estimating maximum circulation
depths is described below.
Some of the consolidated bedrock and many of the fault controlled ground water
systems have elevated discharge temperatures (i.e. around 9C or more) and are
of the intermediate type. Although slightly elevated discharge temperatures indi-
cate deeper circulation t han local flow systems, in most instances it is unlikely
t hat circulation depths exceed 500 m. Estimates of maxi mum circulation depths
are based on differences between mean annual air temperatures and estimated
A.L. Mayo, M.D. Loucks / Journal of Hydrology 172 (1995) 31-59 49
I I I I I
0 Supersaturated ~ . ~ v
- 2 = / / = - ' - =
/
--3 / 0 Granitic Systems
Unconsolidated Alluvial Systems
v Consolidated Sad. Rock Systems
--4 Fault Controlled Systems
o Thermal Systems
= _ Mine Systems
--5 I I I I I
10 30 50 70 90 99
Percentage of sampl es whose value is equal to or l ess than the value shown
Fig. 8. Cumulative frequency di st ri but i on of calcite s at urat i on of ground waters from t he central Was at c h
Range, Ut ah.
maxi mum aquifer temperatures, and assume a geothermal gradient of 40C
km -] (Kohler, 1979; Cole, 1983). Most, i f not all, of the intermediate systems
have a component of recharge which occurred before 1952; however, some of
these systems may have some ground water which recharged hundreds to
perhaps a few t housand years ago.
8. Thermal ground water systems
In this investigation, ground water systems are considered thermal if they have a
discharge temperature higher than 15C. The 15C criterion is consistent with White' s
(1957) thermal definition at mos t locations. Based on the 15C criterion, four thermal
systems are identified, each of which has distinctive chemistry (Table 1). The solute
chemistries may be classified as:
(1) Mi dway thermal system (Sites 70, 71 and 73): high TDS, Ca2 +- Mge+- HCO3
type water having elevated concentrations of Na +, K +, SO 2-, and C1-.
(2) Mayflower Mine thermal system (Site 72): high TDS, Ca2+-SO 2- type water
having a low concentration of K +.
(3) Mount ai n Meadows thermal system (Site 69): high TDS, Ca2+-Mg2+-SO2-(?)
type water.
(4) Parley' s Canyon thermal system (Site 74): low TDS, Mg2+-Ca2~--HCO~ - type
water having relatively low concentrations of Na +, K +, SO ] - and CI-.
50 A.L. Mayo, M.D. Loucks / Journal of Hydrology 172 (1995) 31-59
The Midway thermal system is the largest of the four thermal regimes. This system
has deposited travertine layers as thick as 75 m over a 10 km 2 area. Locally, travertine
mounds rise as much as 20 m above land surface. The aerial extent and thickness of
the travertine deposits suggest t hat the Midway thermal system has been active for an
extended period of time. The system currently discharges into small pools t hat occupy
craters (known locally as ' hot pots' ) or from springs issuing from openings in the
travertine. Little or no flow discharges from most of the craters, and discharge rates
from openings in the travertine terrace do not exceed 11.3 1 s -1. A slight odor of
H2S(s ) can be detected at some of the discharge sites.
The Mayflower Mine (Site 72) is located about 9 km nort h of the Midway thermal
system. The temperature of the flowing discharge at the surface is 21.2C (Table 1);
however, temperatures reaching 60C have been reported at the 3000 ft level (Kohler,
1979).
At Mount ai n Meadows (Sulfur Spring; Site 69) approximately 14.2 1 s -1 discharge
from a seepage front of 200-300 m length. The thermal system discharges from the
organically rich Manni ng Canyon Shale, and the distinct odor of H2S(g) lingers in the
air. The seepage front appears to be associated with the Deer Creek Thrust Faul t and
numerous extension faults which are located in the immediate area.
The Parley' s Canyon thermal system (Site 74) is located on the nort h side of
Interstate-80 in Parley' s Canyon. The thermal system discharges as a single
spring t hat has a current discharge of only 0.3 1 s -1. A travertine cone esti-
mated to be 8 m high is found at this site. Gr ound water flow appears to be
controlled by fracturing of Triassic limestones and sandstones within a small,
tightly folded anticline.
9. Mi ne discharges
The mine discharges have a wide range of discharge rates and solute and isotopic
compositions (Table 1). Discharge rates range from 7.1 1 s -l in the Par k- Heber
Tunnel (Site 84) to about 88 1 s -1 in bot h the Spiro Tunnel (Site 88) and the Steam-
boat Tunnel (Site 75). Discharge rates appear to be a function of bot h tunnel length
and the structure and/ or stratigraphy intercepted.
Mine drain waters issuing from sulfide mineralized zones are neutral (pH 6.0-7.7),
and only a few have acid drainage (pH 3-4). Causes of these acid mine drainages and
the geochemical evolution of mine drainage waters have been discussed in detail by
Mayo et al. (1992).
10. Di scussi on
10.1. Travertine deposition
Some spring discharge waters, currently undersaturated with respect to calcite,
have previously deposited travertine, and some discharge waters saturated or
supersaturated with respect to calcite do not have associated travertine deposits.
A.L. Mayo, M.D. Loucks / Journal of Hydrology 172 (1995) 31-59 51
For example, ground water issuing from Cascade Spring (Site 54) is not
currently depositing travertine and is currently calcite undersaturated
( l ogS It =- 0 . 1 5 , where S ! is saturation index), yet the discharge area has the
mos t extensive travertine deposit outside of the Mi dway thermal area. A large
travertine terrace is located approximately 150 feet downgradient of the spring
discharge area. Travertine deposition resulted from the exsolution of CO2(g) as
the water rushed over a steep embankment between the spring area and the
travertine terrace.
The cause of calcite undersaturation at Cascade Spring is problematic. Three
mechanisms could potentially explain why the discharge waters are currently under-
saturated: (1) travertine deposition may only occur during unusually warm summer
periods, when surface water temperatures are elevated by summer heating; (2) the
faults along which the ground wat er ascends may have become partially sealed since
the time of active travertine deposition; (3) the PCO2(g) value of the discharge water
may have decreased since the time of travertine deposition. The summer heating
hypothesis was tested using using the comput er code PHREEQE (Parkhurst et al.,
1980). Possible variations in calculated discharge Sic were modeled for incremental
temperature rises. Model ed calcite saturation is not attained until the water tempera-
ture more than doubles (e.g. attains 21C). Because the surface water temperature
only rises by 1-2C over the short distance from the discharge area to the travertine
area, this mechanism is not reasonable.
Direct evidence is not available for the fault sealing hypothesis. Faul t sealing
requires that the ground wat er temperature becomes elevated at dept h and that
mineral deposition occurs along the flow pat h as the result of temperature-induced
calcite supersaturation. Faul t sealing woul d decrease the flow rate and woul d result in
lower present-day discharge temperatures, which woul d in turn decrease the calcu-
lated Si c. However, it is difficult to reconcile the large flow rate of Cascade Spring
(240 1 s -1) with fault sealing.
A decreased PCO2(g) value of Cascade Spring recharge water or the elimination
of an external source of CO2(g) since the time of travertine deposition are possible
mechanisms to explain calcite undersaturation. Decreased PCO2(g) in the recharge
wat er could be related to vegetation changes owing to climatic changes. The loss of
an external source of CO2(g) could be due to the sealing of deeper faults along
which t he gas migrated upward t oward the Cascade Spring ground water system.
The 613C value (-10.0%o) of Cascade Spring discharge water (Table 3) does not
suggest an external source of CO2(g); however, fault-related external gas is sus-
pected at the nearby Mi dway Thermal area, which also has extensive travertine
deposits.
10.2. Circulation and chemi cal evolution o f t hermal wat ers
10.2.1. Circulation and heat source o f t hermal wat er
Silica and Na - K- C a geochemical thermometers (Fournier and Rowe, 1966;
Fourni er and Potter, 1979) were used to estimate subsurface temperatures of the
thermal ground waters (Table 4). Estimated or measured ground water temperatures
52 A.L. Mayo, M.D. Loucks / Journal of Hydrology 172 (1995) 31-59
may be used to estimate circulation depths, assuming the temperatures are controlled
by the geothermal gradient. The Mg 2+ corrected Na - Ca - K geothermometer
( Foumi er and Potter, 1979) is potentially applicable to thermal ground water
aquifers in the central Wasatch Range because the Na - Ca - K t hermomet er results
are higher t han 70C.
The low geot hermomet er temperature calculated for the Parley' s Canyon thermal
system (Site 74) appears to be an underestimation of the subsurface temperature
(Table 4). This underestimation may be the result of a slow circulation rate, as
evidenced by the small discharge volume, which allows the system partially to cool
by conduction. Alternately, the low calculated temperature may be the result of r oc k -
water re-equilibration near the surface. In either case, the geothermometers would
indicate a mi ni mum subsurface temperature.
The Midway, Mayflower mine, and Mount ai n Meadows thermal systems have
mean Mg 2+ corrected maxi mum aquifer temperature estimates (Table 4) of 70-
86C, 49C, and 102C, respectively (Sites 70, 71, 73, and 69, and 72). The estimated
Mi dway aquifer temperature shows good agreement with Na - K- Ca thermometers
used by Kohl er (1979). Because of probable mixing of near-surface cold ground water
with the thermal water, Kohl er (1979) used a silica mixing model (Fournier and
Rowe, 1966) to estimate a 125C maxi mum temperature for the Midway thermal
system. Mixing is consistent with the tritium and 14C contents of these ground
waters; however, their unstable isotopic compositions may only be indicative of
modem ground water (Table 3). Kohl er (1979) may have slightly overestimated the
aquifer temperature because silica geothermometers were used. Quartz and
chalcedony geothermometers often overestimate subsurface temperatures of aquifers
where there is a continual acid supply owing to: (1) organic decay (i.e. organic-rich
shales); (2) oxidation of sulfides (i.e. mineralized zones); (3) the influx of H2S(g) (e.g.
Mount ai n Meadows thermal system); (4) an excess CO2(g) (e.g. Midway thermal
system).
Tabl e 4
Na - K- Ca , Mg corrected Na - K- Ca , and silica geot hermomet er results for t hermal ground waters in the
cent ral Wasatcli Range, Ut a h
Sample Na - K- C a Mg corrected Silica
no. geot hermomet er geot hermomet er geot hermomet er
temp. (C) temp. (C) temp. (C)
69 46 49 41
70 87 70 88
71 79 86 54
72 - 8 102 57
73 71 74 46
74 20 COOL 40
Mean 49.2 76.2 54.3
SD 37.2 19.7 17.9
A.L. Mayo. M.D. Loucks / Journal of Hydrology 172 (1995) 31-59 53
Excess silica tends to accumulate in relatively low temperature thermal ground
waters owing to kinetic factors, which result in a slow rate of quartz and chalcedony
precipitation relative to the rate of silica dissolution (Fournier, 1981). Quartz
supersaturation in all of the thermal ground waters (Table 4) suggests excessive silica
accumulation.
The lack of positive 6180 shifts also suggests subsurface temperatures below about
100C (Fig. 3). I f maxi mum aquifer temperatures are above about 100C > in the
Mi dway thermal systems, the absence of a 6180 shift requires wall rock depletion of
lSo in the aquifer. Depletion of 180 is caused by isotopic exchange between
circulating hydrot hermal fluids and surrounding cooling igneous rock (Shieh and
Taylor, 1969). However, significant ISo wall rock depletion has only been reported
within igneous rock bodies near the cont act zone with the count ry rock. Minimal 180
wall rock depletion occurs in carbonat e terrains (Shieh and Taylor, 1969). Because the
Mi dway region is underlain by carbonat e rocks (Bromfield et al., 1970) l So wall rock
depletion is unlikely.
Additionally, the nearest known igneous rock body suitable for lSo depletion is
located approximately 1-2 km to the nort h (Bromfield et al., 1970). It is unlikely that
the thermal ground waters circulate t hrough the igneous terrain and then discharge
from the carbonat e terrains. Geophysical dat a collected by Fox (1979) has been
examined by A. K. Benson (personal communication, 1994), who found little
evidence for an igneous body lying beneath the Midway thermal system,
Assuming the maxi mum aquifer temperature of the Midway thermal system is
85-100C and assuming an average temperature gradient of 40C k m- l (Cole, 1983),
the maxi mum probable circulation depth is 2-2. 5 km. Assuming a mi ni mum aquifer
temperature of 49C in the Mount ai n Meadows thermal system, the mi ni mum
circulation dept h is just over 1 km. The 49C estimate is based on the Mg 2+ corrected
geot hermomet er temperature estimate (Table 4). The maxi mum aquifer temperate of
the Mount ai n Meadows thermal systems is also estimated to be below 100C, and the
maxi mum probable circulation depth is 2-2. 5 km.
The measured 65C at the 3000 ft level (Kohler, 1979) represents t he mi ni mum
aquifer t emperat ure in the deeper portions of the Mayflower Mine. The Mg 2+
corrected geot hermomet er temperature of 102C is in good agreement with the lack
of a 6180 shift (Fig. 3). Assuming a maxi mum aquifer temperature of 100C, the
maxi mum probable circulation dept h in the Mayflower thermal system is 2-2. 5 km.
Most of the ground water flow models proposed for the Mi dway thermal systems
(Milligan et al., 1966) have suggested that meteoric waters have circulated to depth
along faults and within fractures in the Weber Quartzite. The fractures are related to
folding along the Park City anticline. Kohler (1979) suggested that the geophysical
dat a of Fox (1979) indicate t hat the area is not underlain by faulting at sufficient
depths to permit heating. Instead, Kohl er (1979) postulated that heating is due to a
cooling, shallow intrusive body. The hypothesis of Kohler (1979) is unlikely for
several reasons: (1) there is no evidence for recent volcanism in the region; (2) stable
isotopic dat a (613C, 61So, and 62H) are not consistent with volcanic gases of fluids; (3)
the gravity dat a of Fox (1979) suggest that a large fault may exist on the east side of
the Mi dway thermal system (A.K. Benson, personal communication, 1994).
54 A.L. Mayo, M.D. Loucks / Journal of Hydrology 172 (1995) 31-59
A deep fault structure probabl y intercepts ground water moving down-dip in the
carbonat e rocks which occur in the Wasat ch Range located immediately to the west.
The east dip of the rocks probabl y provides an avenue for deep circulation, and faults
provide an avenue for the rapid ascent of thermal ground water to the surface.
10. 2. 2. Che mi c al evol ut i on
Differences in mean of SO 2- and C1- concentrations of thermal waters (Table 1)
are the result of differences in gypsum, halite and sulfide mineral dissolution product s
in the various thermal ground waters. The large SO 2- content in the Mount ai n
Meadows ground water (Site 69) suggests that it intercepts more gypsum along the
flow pat h than do the other thermal systems. Non-bedded gypsum occurs with many
of the Paleozoic and Mesozoic limestones and more commonl y in the dolomites (J. K.
Rigby, personal communication, 1994).
The 634S values of the Mi dway thermal system (mean +12.79/oo) and Mount ai n
Meadows thermal system (+13.69/0o) are consistent with gypsum dissolution
with some sulfate reduction. The odor of H2S(g) at some of these discharge locations
suggests sulfate reduction processes. The 634S value of the Mayflower Mine thermal
system (+4.6%0) may be explained by a combination of gypsum dissolution and
sulfide mineral oxidation.
Thermal ground waters are supersaturated with respect to bot h calcite and
dolomite (Table 2; Fig. 8), and travertine deposits occur at many discharge
locations. Because calcite and dolomite solubility are inversely related to water
temperature, saturation or supersaturation often results from raised water
temperature and may involve the dissolution of additional aquifer and soil zone
minerals.
Elevated TDS concentrations of thermal ground waters are often attributed to
increased mineral solubility at higher temperatures, the common ion effect, and the
increased opport uni t y for contact with more soluble minerals associated with deeper
circulation depths and longer flow paths. The comput er code PHREEQE (Parkhurst
et al., 1980) was used to evaluate plausible geochemical evolutionary pathways at
elevated aquifer temperatures. This analysis demonstrates that at elevated aquifer
temperatures, calcite, dolomite, and gypsum dissolution, and the common ion effect
cannot account for the observed carbonat e mineral dissolution product s and the
carbonat e mineral supersaturation. The analysis assumed typical soil z o n e eCO2(g )
concentrations (log PCO2(g) = 10 -18 - - 10 -2s atm; Rightmire and Hanshaw, 1973;
Har mon et al., 1975). Specifically, the PHREEQE analysis demonstrates the follow-
ing: (1) the Mi dway thermal system should contain 4. 7-7. 1 mequiv 1-1 of carbonate
mineral dissolution product s (i.e. Ca 2+, Mg 2+ and HCO~) at the mean discharge
temperature of 34.3C; (2) the Mayflower mine thermal system should contain
6. 7-9. 5 mequiv 1-1 of carbonat e mineral dissolution product s at the discharge
temperature of 21.2C; (3) the Mount ai n Meadows thermal system should contain
7. 7-11. 7 mequiv 1-1 of carbonat e mineral dissolution product s at the discharge
temperature of 24.5C; (4) the 1-80 thermal system should contain 11. 2-
17. 1 mequiv 1-1 of carbonat e mineral dissolution product s at the
discharge temperature of 22.3C.
A.L. Mayo, M.D. Loucks / Journal of Hydrology 172 (1995) 31-59 55
Differences between the mean of the measured concentrations of carbonate mineral
dissolution products (i.e. Ca 2+, Mg 2+, HCO~-) for the four thermal systems and the
PHREEQE calculated concentrations are listed in Table 5. It is possible t hat the
PHREEQE overestimation of carbonate mineral dissolution products for the
Parley' s Canyon thermal sample (Site 74) is due to CO2(g) exsolving near the surface
before sampling. CO2(g) exsolution is suggested by the calculated log PCO2(g) value
(-3. 12 atm). Loss of PCO2(g) may be attributed to slow circulation near the surface
t hat would allow the escape of CO2(g). I f CO2(g) is lost, precipitation of carbonate
minerals is likely and would be accompanied by a decrease in the carbonate mineral
dissolution products. When the calculated log PCO2(g) (-3. 12 atm) was used in the
PHREEQE analysis, the sum of the calculated carbonate dissolution products was
6.7 mequlv 1-1 . The extremely small discharge rate (0.3 1 s -1) and the large travertine
cone at the Site supports the idea of escaping CO2(g).
Because carbonate mineral supersaturation cannot be attributed to normal mineral
dissolution processes, an external source of CO2(g) or saline fluids is suspected. These
processes are described below.
PCO2(g) levels in the Midway thermal system, which approach a calculated 1 at m at
some discharge locations, suggest an external source of CO2(g). An external CO2(g)
source is also suggested by the 613C values of HCO~- ( - 3. 4 to -2.5%0) in thermal
ground waters from both Midway and Mount ai n Meadows thermal systems (Table
3). However, the 613C (-10.7%o) and logPCO2(g) (-1. 84 atm) values from the
Mayflower Mine thermal water (Site 72) are within the normal ranges of carbon
derived from soil zone CO2(g) and the dissolution of soil zone and aquifer carbonate
minerals.
The two most likely external sources are thermal decarbonization of deep
sedimentary rocks and silicate hydrolysis. Thermal decarbonization of sedimentary
rocks during regional metamorphism has also been suggested as a source of ground
water CO2 (Craig, 1963; Mayo et al., 1985; Wexsteen et al., 1988). Excess H +
associated with external CO2(g) can liberate additional isotopically heavy HCO~
according to the following carbonate dissolution reactions:
CO2(g ) + H20 = H2CO3(aq ) (4)
H2CO3(aq ) ---~ H + + HCO~- (5)
Table 5
Summary of calculated and measured carbonate mineral dissolution products (Ca 2+ + Mg 2+ + HCO~) for
thermal ground water systems.
Thermal Measured PHREEQE Difference
system mean mean (mequiv 1-1)
(mequiv l- 1) (mequiv 1- l)
Midway 34.2 5.9 28.3
Mountain Meadows 24.7 9.7 15
Mayflower Mine 28 8.1 19.9
Parley's Canyon 9.3 14.2 -4.9
56 A.L. Mayo, M.D. Loucks / Journal of Hydrology 172 (1995) 31-59
CaCOa(s) + H + = Ca 2+ + HCO 3 (6)
CaMg(CO3)2(s ) + 2H + = Ca 2+ + Mg 2+ + 2HCO3 (7)
MgCOs(s) + H + = Mg 2+ + HCO3 (8)
Liberation of CO2(g) from siliceous-carbonate rocks occurs during high-temperature
(above 450C) metamorphic reactions (Turner, 1968; Winkler, 1976). In the Midway
and Mount ai n Meadows thermal areas, however, high metamorphic temperatures are
unlikely. In-situ temperatures of overthrust and block faulted siliceous-carbonate
rocks probably do not exceed 200-250C at reasonable overthrust and block fault
depths of 4- 6 km. Although the number of potential reaction pathways are greatly
reduced at these lower temperatures, several CO2(g)-producing reactions are possible.
For example, the following reactions would yield a PCO2(g) of 1 atm pressure at
temperatures of 180C and 190C, respectively:
5Dolomite + 8Quartz + H20 = Tremolite + 3Calcite + 7CO2 (9)
3Dolomite + 4Quartz + H20 -- Talc + 3Calcite + 3CO 2 (10)
The purpose here is not to suggest specific metamorphic reactions, rock-wat er ratios,
or H2 0 - C O 2 mole fractions which are unique to the postulated metamorphic system,
but rather to suggest t hat liberation of CO2(g) from a low-temperature metamorphic
source is possible. Determination of actual reaction pathways would require the
knowledge of the metamorphic mineral assemblages which lie beneath the thrust
and block faulted areas.
Silicate hydrolysis has been suggested as a source of excess CO 2 in clastic rock
diagenetic environments where temperatures reach 100C or higher (Hutcheon and
Abercrombie, 1990). CO 2 evolution by silicate hydrolysis would also have to occur at
considerable depth beneath the active thermal area.
Because the thermal ground waters have not been heated to temperatures in excess
of 100C as evidenced by the 6180 data, and the wat er - CO 2 ratio in the Midway
thermal system is high, external CO2(g ) would have had to diffuse upward, largely in
the absence of metamorphic water, along an extension fault(s) beneath the Midway
area. Such structures may be inferred from the gravity dat a of Fox (1979) as
interpreted by A. K. Benson (personal communication, 1994).
Other possible sources t hat could yield more positive 613C value (less t han about
-4%o) include production of excess acids, and magmatic and mantle CO2 sources.
The oxidation of pyrite, siderite, or similar iron-bearing minerals and the dissolution
of silicate minerals can form acids and liberate additional H +. Although these
minerals are found in many carbonate terrains within the study area, it is unlikely
t hat they are responsible for the elevated PCOE(g ) in the thermal ground waters. The
concentrations of dissolved silica are too low, and it is unlikely t hat sufficient O2(aq )
exists at depth in the thermal systems to liberate appreciable H + by sulfide mineral
oxidation.
Basin and range faulting is a potential avenue for the upward migration of mantle
A.L. Mayo, M.D. Loucks / Journal of Hydrology 172 (1995) 31-59 57
or ma gma t i c gas es and fluids. Such fluids c ont ai n CO2(g) ( Ne wt on and Shar p, 1975;
Lyon and Hul s t on, 1984). The 62H and 6180 dat a ( Tabl e 3) ar e not c ons i s t ent wi t h t he
s t abl e i s ot opi c c ompos i t i on o f ma g ma t i c fluids. Ma gma t i c wat er has a 62H of - 4 0 t o
- 80%0 a nd 6180 o f +6 t o + 9 % ( Tayl or , 1977; Trues del l and Hul s t on, 1980). Ho w-
ever, mant l e, der i ved CO2(g) (613C o f - 3 t o -10%o; Dei nes and Gol d, 1973; Ander s on,
1987) and t he di s s ol ut i on of aqui f er c a r b ona t e mi ner al s in t he c ent r al Wa s a t c h Range
(mean/~13C +0.3%0) woul d yield 613C val ues whi ch ar e c ons i s t ent wi t h t he ob s er ved
val ues. Mant l e CO2(g) is unl i kel y, however .
Vol c ani c CO2(g) is al s o an unl i kel y s our c e o f ext er nal H + in t he t her mal gr ound
wat er s bec aus e (1) t he mos t recent r egi onal vol c ani s m in t he t her mal ar eas c ul mi nat ed
in t he Ol i goc ene (Bromfi el d et al. , 1970), and (2) t he di s s ol ut i on of aqui f er c a r b ona t e
mi ner al s ( mean 613C +0.30%0) wi t h vol c ani c CO2(g ) (613C - 1 4 t o -28%o; Faur e, 1986)
woul d yi el d b i c a r b ona t e i s ot opi c al l y l i ght er t han f ound at t he Mi dwa y and Mo u n t a i n
Me a dows t her mal s ys t ems .
Acknowledgments
We woul d like t o t ha nk Dr . An t h o n y B. Mul l er and Vi c t or Heilweil of t he US
Geol ogi c al Sur vey f or t hei r t hought f ul revi ews and c omment s , and Li nda Ma y o f or
her c ar ef ul and pat i ent edi t i ng of t he manus c r i pt . Fi nanc i al s uppor t was pr ovi ded b y
t he Ut a h Geol ogi c al Sur vey and Br i gham Young Uni ver s i t y.
References
Anderson, R., 1987. Mantle and crust components in a carbonatite complex and the evolution of
carbonatite magma: REE and isotopic evidence from the Fen complex, southeast Norway. Chem.
Geol., 65: 147-166.
Bath, A.H., 1983. Stable isotopic evidence for paleo-recharge conditions of groundwater. In: Paleoclimates
and Paleowaters: a Collection of Environmental Isotope Studies: Proceedings of an Advisory Group
Meeting, Vienna, 25-28 November 1980. International Atomic Energy Agency, Vienna, pp. 169-186.
Bromfield, C.S., Baker, A.A. and Crittenden, M.D., 1970. Geologic map of the Heber Quadrangle, Summit
and Wasatch Counties, Utah. US Geological Survey, Salt Lake City, UT, Map GQ-864.
Calkins, F.C. and Butler, B.S., 1943. Geology and ore deposits of the Cottonwood-American Fork area,
Utah. US Geol. Surv. Prof. Pap., 201,151 pp.
Cole, D.R., 1983. Chemical and isotopic investigation of warm springs associated with normal faults in
Utah. J. Volcanol. Geothermal Res., 16: 65-98.
Craig, H., 1961. Isotopic variations in meteoric water. Science, 133: 1702-1703.
Craig, H., 1963. The isotope geochemistry of water and carbon in geothermal areas. In: Nuclear Geology
on Geothermal Areas, 9-13 September 1963. Consiglio Nazionale delle Ricerehe, Laboratorio di
Geologia Nuclear, Pisa, 53 pp.
Dansgaard, W., 1964. Stable isotopes in precipitation. Tdlus, 16: 435-468.
Davis, F.D., 1983. Geologic map of the Central Wasatch Front. Utah Geological and Mineral Survey, Map
54, 2 sheets.
Deines, P. and Gold, D.P., 1973. The isotopic composition of carbonatite and kimberlite carbonates and
their bearing on the isotopic composition of deep-seated carbon. Geochim. Cosmochim. Acta, 37:
1709-1733.
Faure, G., 1986. Isotope Geology, 2nd edn. Wiley, New York, 589 pp.
58 A.L. Mayo, M.D. Loucks / Journal of Hydrology 172 (1995) 31-59
Feth, J. H. , Roberson, C.E. and Polzer, W.L., 1964. Sources of mineral constituents from granitic rocks,
Sierra Nevada, California and Nevada. US Geol. Surv. Water-Supply Pap., 1535-1, 70 pp.
Fontes, J. Ch. and Gamier, J.M., 1979. Determination of the initial 14C activity of the total dissolved
carbon: a review of the existing models and a new approach. Water Resour. Res., 15(2): 399-413.
Fournier, R.O., 1981. Application of water geochemistry to geothermal exploration and reservoir
engineering. In: L. Rybach and L.J.P. Muffler (Editors) Geothermal Systems: Principles and Case
Histories. Wiley, New York, pp. 109-143.
Fournier, R. O. and Potter, R.W., 1979. Magnesium correction to the Na - K- Ca chemical geothermometer.
Geochim. Cosmochim. Acta, 43: 1543-1550.
Fournier, R.O. and Rowe, J.J., 1966. Estimation of underground temperatures from silica content of water
from hot springs and wet steam wells. Am. J. Sci., 264: 685-697.
Fuhriman, D. K. and Merritt, L.B., 1976. Assessment of groundwater quality in the Mountaiuland Association
of Governments area. Mountainland Association of Governments, Salt Lake City, UT, 68 pp.
Fox, R.C., 1979. Midway hot springs gravity survey--preparat i on and interpretation of results. In: J. F.
Kohler (Editor), Geology, Characteristics and Resource Potential of the Low-temperature Geothermal
System near Midway, Wasatch County, Utah. Ut ah Geological and Mineral Survey, Salt Lake City,
UT, Report Investigation No. 142, Appendix.
Gat, J. R. , 1983. Precipitation, groundwater and surface waters, control of climatic parameters on their
isotopic compositions and their utilization as paleoclimatological tools. In: Paleoclimates and
Paleowaters: a Collection of Environmental Isotope Studies: Proceedings of an Advisory Group
Meeting, Vienna, 25-28 November 1980. International Atomic Energy Agency, Vienna, pp. 3-12.
Harmon, R.S., White, W.B., Drake, J.J. and Hess, J,W., 1975. Regional hydrogeochemistry of Nort h
American carbonate terrains. Water Resour. Res., 11(6): 963-967.
Holser, W. T. and Kaplan, I. R. , 1966. Isotope geochemistry of sedimentary sulfur. Chem. Geol., 1: 93-135.
Homes, W. F. , Thompson, K. R. and Enright, M., 1986. Water resources of the Park City area, Utah, with
an emphasis on ground water. Ut ah Department of Natural Resources, Technical Publication 85, 81 pp.
Huteheon, I. and Abererombie, H. , 1990. Carbon dioxide in elastic rocks and silicate hydrolysis. Geology,
18: 541-544.
Kaplan, I. R. and Rittenburg, S.C., 1964. Microbiological fraetionation of sulfur isotopes. J. Gen.
Microbiol., 34: 195-212.
Kohler, J. F. , 1979. Geology, characteristics and resource potential of the low-temperature geothermal
system near Midway, Wasatch County, Utah. Ut ah Geological and Mineral Survey, Salt Lake City,
UT, Report Investigation No. 142, 29 pp.
Lyon, G. L. and Hulston, J. R. , 1984. Carbon and hydrogen isotopic compositions from New Zealand
geothermal gases. Geochim. Cosmochim. Acta, 48:1161-1171.
Mayo, A. L. and Klauk, R. H. , 1991. Contributions to the solute and isotopic ground-water geochemistry,
Antelope Island, Great Salt Lake, Utah. J. Hydrol. , 127: 307-335.
Mayo, A. L. and Mitchell, J.C., 1988. Comparison of stable and unstable isotopic compositions from
thermal ground water in the Southern Idaho Batholith and adjacent Snake River Plain. Geol. Soc.
Am. Abstr. Prog., 20(6): 430.
Mayo, A.L., Muller, A.B. and Ralston, D. R. , 1985. Chemical and isotopic composition of ground water in
the Meade thrust allochthon, southeastern Idaho, and its relevance to the stratigraphie and structural
control of ground-water flow. J. Hydrol. , 76(1-2): 21-65.
Mayo, A. L. , Nielsen, P.J., Loucks, M. and Brimhall, W. H. , 1992. The use of solute and isotopic chemistry
to identify flow patterns and factors which limit acid mine drainage in the Wasatch Range, Utah.
Groundwater, 30(2): 243-249.
18
Merlivant, L. and Jouzel, J., 1983. Deuterium and O in preci pi t at i on--a global model from oceans to ice
caps. In: Paleoclimates and Paleowaters: a Collection of Environmental Isotope Studies: Proceedings of
an Advisory Group Meeting, Vienna, 25-28 November 1980. International Atomic Energy Agency,
Vienna, pp. 65-66.
Milligan, J. H. , MarseU, R. E. and Bagley, J. M. , 1966. Mineralized springs in Ut ah and their effect on
manageable water supplies. Ut ah State University, Ut ah Water Resources Laboratory, Rep. W623-6,
50 pp.
A.L. Mayo, M.D. Loucks / Journal of Hydrology 172 (1995) 31-59 59
Moser, H. , Stichler, W. and Trimborn, P., 1983. Stable isotope studies on paleowaters. In: Paleoclimates
and Paleowaters: a Collection of Environmental Isotope Studies: Proceedings of an Advisory Group
Meeting, Vienna, 25-28 November 1980. International Atomic Energy Agency, Vienna, pp. 201-204.
Mundorff, J.C., 1971. Nonthermai springs of Utah. Ut ah Geological and Mineral Survey, Salt Lake City,
UT, Water-Resources Bull. 16, 70 pp.
Newton, R. C. and Sharp, W.E., 1975. Stability of forsterite + CO2 and its bearing on the role of CO2 in the
mantle. Earth Planet. Sci. Lett., 26: 239-244.
Parkhurst, D. L. , Thorstenson, D. C. and Plummer, L. N. , 1980. PHREEQE- - a computer program for
calculating mass transfer for geochemical reactions in ground water. US Geol. Surv. Water Resour.
Invest., 80-96, 193 pp.
Plummer, L. N. , Jones, B. F. and Truesdell, A. H. , 1976. WATEQF- - a FORTRAN IV version of WA-
TEQF, a computer program for calculating chemical equilibrium of natural waters. US Geol. Surv.
Water Resour. Invest., 76-13, 61 pp.
Rightmire, C. and Hanshaw, B., 1973. Relationship between the carbon isotopic composition of soil CO2
and dissolved carbonate species in ground water. Water Resour. Res., 9(4): 958-967.
Roark, D. M. , Holmes, W. F. and Shlosar, H. K. , 1991. Hydrology of Heber and Round Valleys, Wasatch
County, Utah, with emphasis on simulation of ground-water flow in Heber Valley. Ut ah Department of
Natural Resources, Salt Lake City, UT, Technical Publication 101, 93 pp.
Shieh, Y. N. and Taylor, H.P., 1969. Oxygen and carbon isotope studies of contact metamorphism of
carbonate rocks. J. Petrol., 10: 307-331.
Taylor, H.P., 1977. Water/rock interactions and the origin of H2 0 in granitic batholiths. J. Geol. Soc.
London, 133: 509-558.
Thode, H. G. , Monster, J. and Dunford, H.B., 1961. Sulphur isotope geochemistry. Geochim. Cosmochim.
Acta, 25: 150-174.
Toth, D. K. , 1963. A theoretical analysis of groundwater flow in small drainage basins. J. Geophys. Res., 68:
4795-4812.
Truesdell, A. H. and Hulston, J. R. , 1980. Isotopic evidence on environments of geothermal systems. In: P.
Fritz and J. Ch. Fontes (Editors), Handbook of Environmental Isotope Geochemistry, Vol. 1, The
Terrestrial Environment, A. Elsevier, Amsterdam, pp. 179-226.
Turner, F. J. , 1968. Metamorphic Petrology. McGraw-Hi l l , New York, 403 pp.
US Environmental Protection Agency (]SPA), 1981. Methods for chemical analysis of water samples. EPA
6000-4-81-020, US Environmental Protection Agency, Cincinnati, OH.
Ut ah Department of Natural Resources, 1987. Water quality in the Cottonwood Canyon' s watershed:
impacts from abandoned metal mines, a preliminary assessment. Ut ah Department of Natural Re-
sources, Salt Lake City, UT, 66 pp.
Wexsteen, P., Jaffe, F. C. and Mazor, E., 1988. Geochemistry of cold CO2-rich springs of the Scuol-Tarsp
region, lower Engadine, Swiss Alps. J. Hydrol. , 69: 954-973.
White, D. E. , 1957. Thermal waters of volcanic origin. Geol. Soc. Am. Bull., 68: 1637-1658.
White, D. E. and Chuma, N. J. , 1987. Carbon and isotopic mass balance models of Oasis Val l ey-Fort y Nine
Mile Canyon ground water basin, southern Nevada. Water Resour. Res., 23(4): 571-582.
White, D. E. , Hem, J. D. and Waring, G. A. , 1963. Dat a of geochemistry, Chapter F. chemical composition
of subsurface waters. US Geol. Surv. Prof. Pap., 440F, 67 pp.
Winkler, H. G. F. , 1976. Petrogenesis of Metamorphic Rocks, 4th edn. Springer, Berlin, 334 pp.

Vous aimerez peut-être aussi