ArticlePDF Available

Abstract and Figures

Study of the hydrothermal dynamics of Yellowstone Lake (Wyoming, USA) is important for identifying potential changes in sublacustrine hydrothermal systems in response to external perturbations from earthquakes, seiches, large waves, and seasonal effects. Remotely operated vehicle (ROV) submersible-based investigations of hydrothermal vents offshore from Stevenson Island reveal numerous non-constructional ~10-cm-diameter orifices with diffuse fluid flow at temperatures up to 174 °C. The vent field occurs in a large roughly conical depression on the lake floor at a water depth of ~120 m. The volatile-rich composition (CO2, H2S) of the vent fluids is preserved by using a novel isobaric sampling system that precludes degassing effects. In addition to high temperatures, the vent fluids have high CO2 and H2S, but low chloride (Cl) and major element concentrations largely indistinguishable from those in ambient lake water. These results are consistent with steam addition to the sublacustrine hydrothermal system. Kaolinite- and boehmite-rich alteration indicates acidic conditions and provides a low-permeability substrate that may contribute to the development of a steam-heated upflow zone. At the scale of individual vent areas (centimeters to meters), perturbations cause bursts of steam-rich fluids that locally expel and disperse sediment and contribute to the formation of vent orifices. Here we report on chemical and physical phenomena associated with the hottest and deepest sublacustrine hydrothermal vents in Yellowstone Lake. Results indicate that vapor-dominated sublacustrine systems are fundamentally different in hydrothermal alteration and hydrothermal dynamic characteristics than their liquid-dominated counterparts.
Content may be subject to copyright.
Geological Society of America
|
GEOLOGY
|
Volume 47
|
Number 3
|
www.gsapubs.org 1
Vapor-driven sublacustrine vents in Yellowstone Lake,
Wyoming, USA
Andrew P.G. Fowler1*, Chunyang Tan1, Christie Cino1, Peter Scheuermann1, Michael W.R. Volk1, W.C. Pat Shanks III2,
and William E. Seyfried, Jr.1
1Department of Earth Sciences, University of Minnesota, Minneapolis, Minnesota 55455, USA
2U.S. Geological Survey, Denver Federal Center, MS 973, Denver, Colorado 80225-0046, USA
ABSTRACT
Study of the hydrothermal dynamics of Yellowstone Lake (Wyoming, USA) is important for
identifying potential changes in sublacustrine hydrothermal systems in response to external
perturbations from earthquakes, seiches, large waves, and seasonal effects. Remotely operated
vehicle (ROV) submersible-based investigations of hydrothermal vents offshore from Stevenson
Island reveal numerous non-constructional ~10-cm-diameter orices with diffuse uid ow
at temperatures up to 174 °C. The vent eld occurs in a large roughly conical depression on
the lake oor at a water depth of ~120 m. The volatile-rich composition (CO2, H2S) of the vent
uids is preserved by using a novel isobaric sampling system that precludes degassing effects.
In addition to high temperatures, the vent uids have high CO
2
and H
2
S, but low chloride
(Cl) and major element concentrations largely indistinguishable from those in ambient lake
water. These results are consistent with steam addition to the sublacustrine hydrothermal
system. Kaolinite- and boehmite-rich alteration indicates acidic conditions and provides a
low-permeability substrate that may contribute to the development of a steam-heated upow
zone. At the scale of individual vent areas (centimeters to meters), perturbations cause bursts
of steam-rich uids that locally expel and disperse sediment and contribute to the formation
of vent orices. Here we report on chemical and physical phenomena associated with the hot-
test and deepest sublacustrine hydrothermal vents in Yellowstone Lake. Results indicate that
vapor-dominated sublacustrine systems are fundamentally different in hydrothermal altera-
tion and hydrothermal dynamic characteristics than their liquid-dominated counterparts.
INTRODUCTION
The Yellowstone volcanic-hydrothermal
system (Wyoming, USA) is the most recent
expression of a sequence of events that trace
back ~16 m.y. along the track of the Yellowstone
hotspot (Pierce and Morgan, 1992). Heat and
non-condensable gases derived from the volca-
nic system interact with deeply circulating mete-
oric water to produce geochemically diverse and
abundant hydrothermal activity (Hurwitz and
Lowenstern, 2014).
Sublacustrine hydrothermal activity in Yel-
lowstone Lake has been inferred for >100 yr
(Hayden, 1878), yet direct observation and sam-
pling of vents was rst achieved comparatively
recently in 1987 (Klump et al., 1988; Remsen et
al., 2002). Observations from lake-oor bathym-
etry surveys, breccia deposits in sediment cores,
and the rock record exposed around the lake
establish that sublacustrine hydrothermal activity
produced noteworthy hydrothermal explosions
in the past (Morgan et al., 2003, 2009; Wold et al.,
1977). Thus, understanding sublacustrine hydro-
thermal processes is important, considering that
more than four million people visit Yellowstone
annually (National Park Service, 2018). Hun-
dreds of active and inactive vents have now been
identied within the lake (Morgan et al., 2003),
and uids from many vents have been analyzed
to better understand sublacustrine hydrothermal
processes (Balistrieri et al., 2007; Gemery-Hill
et al., 2007; Klump et al., 1988).
The present study is a collaborative multi-
disciplinary effort (Hydrothermal dynamics of
Yellowstone Lake, HD-YLAKE) to understand
the response of the Yellowstone Lake hydrother-
mal system(s) to geological and environmental
perturbations (Sohn et al., 2017). Sublacustrine
hydrothermal vent uids are observed in the
“Deep Hole” area east of Stevenson Island, which
is the deepest (100–125 m) region of the lake.
The Deep Hole is located within a NW-trending
fracture zone 2 km east of Stevenson Island that
is dened by large conical depressions that form
linear arrays with en echelon offsets (Fig. 1).
*E-mail: afowler@umn.edu
CITATION: Fowler, A.P.G., et al., 2019, Vapor-driven sublacustrine vents in Yellowstone Lake, Wyoming, USA: Geology, v. 47, p. 1–4, https://doi.org /10.1130
/G45577.1
Manuscript received 11 September 2018
Revised manuscript received 28 December 2018
Manuscript accepted 3 January 2019
https://doi.org/10.1130/G45577.1
© 2019 Geological Society of America. For permission to copy, contact editing@geosociety.org. Published online XX Month 2019
Stevenson
Island
Deep Hole
1,220 Meters6100
0 250 Meters125
Vent YL17F01 and
co-located cores
YL17U03/YL17U04
Deep Hole vent field
Vent
YL16F12
Detail maps
Stevenson
Island
Yellowstone Lake
110.373025 W 110.363250 W 110.353475 W
44.498431 N44.506252 N44.514073 N
Figure 1. Location of Deep Hole sublacustrine
vents (east of Stevenson Island, Yellowstone
Lake, Wyoming, USA). Bathymetry and digi-
tal elevation data modified from Morgan et al.
(2003) and Sohn et al. (2017). Coordinates are
relative to the World Geodetic System 1984
(WGS84).
Downloaded from https://pubs.geoscienceworld.org/gsa/geology/article-pdf/doi/10.1130/G45577.1/4617352/g45577.pdf
by University of Minnesota user
on 22 January 2019
2 www.gsapubs.org
|
Volume 47
|
Number 3
|
GEOLOGY
|
Geological Society of America
Hot liquid and gas bubbles exit through non-
constructional centimeter-scale orices in sedi-
ments that are sparsely populated by white bacte-
rial mats (Fig. DR1 in the GSA Data Repository
1
).
In situ geochemical sensor measurements show
that the hydrothermal uid is acidic, with an in
situ pH at the vent temperature, pH
(T)
, of 4.2–4.5,
and is notably reducing (−0.2 to −0.3 V) (Tan et
al., 2017). A maximum temperature of 174 °C
was measured with a titanium-sheathed tempera-
ture probe inserted ~10 cm into one vent (Tan et
al., 2017), making these the hottest sublacustrine
vent uids reported to date in Yellowstone Lake
or elsewhere.
In comparison, the highest measured vent-
uid temperature in New Zealand’s Lake Taupo
is 45 °C (de Ronde et al., 2002). While high-tem-
perature sublacustrine vent uids are also likely
present in New Zealand’s Lake Rotomahana
based on heat-ow measurements, water-column
temperature anomalies, shoreline hot-spring uid
geothermometry results, and vigorous shallow
vent activity (de Ronde et al., 2016; Stucker et
al., 2016; Tivey et al., 2016; Walker et al., 2016),
vents on the lake oor have not been directly mea-
sured or sampled.
In the present study, a newly designed gas-
tight (isobaric) sampling system, constructed
entirely of titanium and with real-time tempera-
ture monitoring (Wu et al., 2011), was used to
acquire hydrothermal-vent uid samples (See
File DR1 for detailed sampling and analytical
procedures). The isobaric capability means that
lake-bottom pressure is maintained up to the
point of subsampling at the surface, permitting
acquisition of unaltered gas-rich uid samples
from sublacustrine vents for the rst time.
RESULTS AND DISCUSSION
Absolute and relative concentrations of many
dissolved constituents in vent uids are remark-
ably similar to those of Yellowstone Lake water
(Fig. 2). However, elevated CO
2
and H
2
S con-
centrations in all samples and the noteworthy
correspondence between dissolved gases and
temperature (Table 1), combined with stable-
isotope systematics, provide clear evidence of
the dominating inuence of a hydrothermal
component. The high-temperature vent uids
have lower δD (Fig. 3) and Cl relative to lake
water (with as much as 17% less Cl in the high-
est-temperature sample), suggesting that steam
is being added to the sublacustrine system and
providing the heat to achieve high tempera-
tures while diluting conservative species, such
as Cl, in the entrained lake water. Steam-heated
CO2-rich waters are relatively common in sub-
aerial geothermal systems (Hedenquist, 1990),
1
GSA Data Repository item 2019081, File DR1, Fig-
ures DR1–DR6, and Video DR1 (brief footage of gas bub-
bles exiting a sublacustrine hydrothermal vent), is avail-
able online at http://www.geosociety.org /datarepository
/2019/, or on request from editing@geosociety.org.
but have not previously been thoroughly char-
acterized in a sublacustrine setting.
The proportion of steam required to heat
lake water to the sampling temperature of
150 °C for sample YL17F01 was estimated using
enthalpy values from steam tables (Wagner and
Kretzschmar, 2008), assuming binary conserva-
tive mixing, as follows:
=⋅ +− xxMixturecomponent A(1)component B
=⋅ +− xxMixturecomponent A(1)component B
. (1)
The enthalpy corresponding to a temperature
of 188 °C was used in the calculation, the tem-
perature of steam condensation for pure water
at the lake-oor pressure of 1.2 MPa (Wagner
and Kretzschmar, 2008). While the temperature
of the vapor source is unknown, using steam
enthalpy values at liquid-steam saturation cor-
responding to temperatures from 188 °C to
the critical point for pure water makes only a
minor (3%) difference in the calculated mix-
ing fraction. The result shows that condensation
of 19% steam into 4 °C lake water produces a
vent uid at a sampling temperature of 150 °C.
This mixing ratio is in close agreement with the
chloride dilution factor of 17% calculated for
sample YL17F01 and suggests minimal conduc-
tive cooling or mixing with lake water for this
sample following steam condensation.
Consistent with dissolved-species concentra-
tions, vent-uid oxygen isotope (δ18O) and hydro-
gen isotope (δD) values indicate limited reaction
between the lake-oor substrate and vent uids
or high water-to-rock ratios in an upow zone.
Should the sampled steam-heated lake water have
reacted extensively with lake sediments at low
water-to-rock ratios, vent uid δ
18
O would be
shifted to heavier values as is typical for thermal
waters relative to local meteoric water recharge
(Fig. 3). Instead, sublacustrine vent uids are
shifted to lower δ18O and δD values, suggestive
of mixing with isotopically light uid. The δ18O
and δD values of isotopically light sublacustrine
steam can be calculated by mass balance (Equa-
tion 1) using a 17% steam fraction and stable iso-
tope values listed in Table 1 for sample YL17F01.
The result (δ18O = −20.7‰ and δD = −148.1‰)
falls within the range of values of subaerial fuma-
roles in Yellowstone (Fig. 3).
The dominant gases in Yellowstone fuma-
roles after steam (H2O) are CO2 and H2S (Low-
enstern et al., 2015), and enrichments of these
gases in vent-uid samples relative to ambient
lake water (Fig. 2) further support our vapor-
source model. Ubiquitous gas bubbles observed
exiting vents (Video DR1) and in situ pH val-
ues are consistent with vent uids being CO
2
saturated (Tan et al., 2017). Using CO2 solubil-
ity relations in the CO2-H2O system (Duan and
Sun, 2003), CO2 in vent-uid sample YL17F01
is consistent with saturation at 180 °C and
1.2 MPa pressure. This temperature estimate lies
between the maximum measured vent tempera-
ture of 174 °C and the liquid-vapor transition
temperature of 188 °C for pure water at 1.2 MPa.
Yellowstone research drill hole Y-11 at the
subaerial Mud Volcano thermal area, completed
0.01
0.1
1
10
100
Cl SO4Ca KMgNaSiTCO2H2S
Concentration (mmol/kg)
YL16F12
YL17F01
BLW 1
Elevated in
vent fluid
Similar relative concentrations
in vent fluid and lake water
zero
Figure 2. Major and trace element concen-
trations (logarithmic scale) of two offshore
Stevenson Island (Yellowstone Lake, Wyo-
ming, USA) hydrothermal vent fluid samples
(YL16F12 and YL17F01) compared to typical
sample of bottom lake water (BLW 1). Refer to
Table 1 for values. TCO
2
—total carbon as CO
2
.
TABLE 1. AQUEOUS SAMPLE ANALYTICAL DATA, OFFSHORE STEVENSON ISLAND,
YELLOWSTONE LAKE, WYOMING, USA
SAMPLE Vent
YL16F12
Vent
YL17F01
Lake water
BLW 1
Latitude (WGS84) 44.510690 44.510725 44.511110
Longitude (WGS84) −110.356660 −110.356544 −110.356590
Date collected 17 August 2016 10 August 2017 19 August 2016
Depth (m) 114 11 5
Temperature (°C) 110 11 4 150 4
Cl (mmol/kg) 0.11 0.10 0.12
SO4 (mmol/kg) 0.10 0.11
Ca (mmol/kg) 0.11 0.12 0.13
K (mmol/kg) 0.039 0.039 0.042
Mg (mmol/kg) 0.09 0.09 0.10
Na (mmol/kg) 0.36 0.35 0.39
Si (mmol/kg) 0.45 0.42 0.20
TCO2 (mmol/kg) 7. 0 1 7. 8 0.63
H2S (mmol/kg) 2.1 1. 0
H2 (mmol/kg) < 0.031 <
δ18O ‰ (VSMOW) −15.53 −16.06 −15.12
δD ‰ (VSMOW) −124.7 −126.6 −122.0
Note: WGS84—World Geodetic System 1984; TCO2—total carbon expressed as CO2; VSMOW—
Vienna standard mean ocean water. Dash (–) represents not analyzed; < represents not detected.
Downloaded from https://pubs.geoscienceworld.org/gsa/geology/article-pdf/doi/10.1130/G45577.1/4617352/g45577.pdf
by University of Minnesota user
on 22 January 2019
Geological Society of America
|
GEOLOGY
|
Volume 47
|
Number 3
|
www.gsapubs.org 3
during 1967 and 1968 (White et al., 1975),
provides insight into rock alteration in a vapor-
dominated zone under similar pressure, tempera-
ture, and f
O2
conditions to those in the Deep Hole
off Stevenson Island. Hole Y-11 encountered a
165 °C subsurface steam zone where kaolinite
and pyrite are the most abundant alteration prod
-
ucts (White et al., 1971). Kaolinite in hole Y-11
rocks formed from acidic CO2-saturated steam
condensate that altered silicate minerals, while
H2S in the steam combined with rock-derived
Fe to form pyrite (White et al., 1971). Similar
to hole Y-11 and steam-heated areas elsewhere
(e.g., Hedenquist, 1990; Simmons and Browne,
2000), the vent substrate at the Deep Hole con-
tains ~90% kaolinite, with pyrite and pyrrhotite.
Additionally, boehmite, mixed-layer clay, dia-
toms, and quartz are present (Figs. DR2–DR5).
The occurrence of a dominantly kaolinite
substrate is critical for sustaining the steam zone.
In porous media, steam zones can develop when
a permeability contrast is imposed by the pres-
ence of a low-permeability cap (Ingebritsen and
Sorey, 1988; Schubert et al., 1980). Permeabil-
ity values for kaolinite (Al-Tabbaa and Wood,
1987; Olsen, 1966) are generally within the
range required for a steam cap (e.g., Raharjo et
al., 2016), particularly when contrasted to the
high permeability values expected for fracture-
controlled uid ow evidenced from the broader
structural setting of the Deep Hole vents. We
suggest that the occurrence of kaolinite and other
alteration minerals in surcial Deep Hole sedi-
ments may enhance steam zone development and
prevent gravitational ooding of the underlying
steam by overlying lake water.
Why do the offshore Stevenson Island vents
lack constructional amorphous silica features
found at some inactive sites in Yellowstone Lake
(Shanks et al., 2007)? The Si concentration in
the sampled vent uids is above ambient lake
water values (Table 1), but is undersaturated with
respect to amorphous silica at any temperature
or with respect to quartz above 74 °C. This is
consistent with steam-heated lake water enhanc-
ing silicate mineral dissolution rather than pre-
cipitation. Indeed, dissolution of diatomaceous
material that constitutes unaltered Yellowstone
Lake sediment may have contributed to the for-
mation of conical depressions at an earlier stage
in the formation of the Deep Hole vents (Shanks
et al., 2007).
Trace pyrrhotite is present in all samples (Fig.
DR5), however the high H
2
S
(aq)
/H
2(aq)
log ratio of
1.52 for uid sample YL17F01 is well within
the pyrite stability eld at any reasonable tem-
perature. Clearly, pyrrhotite formed under more
reducing conditions at some time in the past
based on its occurrence in vent muds (Fig. DR5),
and the current system has moved toward equi-
librium with pyrite based on the occurrence of
pyrite replacing pyrrhotite (Fig. DR2). The com-
mon association of pyrrhotite with steam zones
is noted in other areas and in epithermal min-
eral deposits where H2S(aq)/H2(aq) ratios are lower
(Browne and Ellis, 1970; Hedenquist, 1990; Sim-
mons and Browne, 2000), which suggests a simi
-
lar association in the Deep Hole area.
The ~16-cm-long YL17U03 push core, located
1 m from a vent orice (Fig. DR6), recovered a
clast supported semi-lithied mud breccia, likely
the buried equivalent of angular slabs (up to 10 cm
long) observed scattered around vent orices (e.g.,
Fig. DR1). The breccia is capped with a distinct
4–5-cm-thick layer of sediment with similar min-
eralogy to breccia clasts lower in the core, but is
coarser and contains large (up to 0.3 mm) hex-
agonal pyrrhotite crystals (Fig. DR2). The 4–5
cm sediment cap may be material exhaled from
deeper (centimeters to meters) within the sedi-
ments during a disturbance to the steam zone.
The occurrence of a sublacustrine steam zone
warrants investigation into potential hydrother-
mal explosion mechanisms. In a liquid-domi-
nated system at the boiling point, a pressure drop
triggers the downward propagation of a steam-
liquid interface that will manifest as a hydro-
thermal explosion if mechanical energy from
the volume change overcomes rock strength
(Browne and Lawless, 2001; Montanaro et al.,
2016; Morgan et al., 2009; Mufer et al., 1971;
Smith and McKibben, 2000). The hydrothermal
explosion potential is exacerbated by the addi-
tion of dissolved gases that lower the liquid sta-
bility limit (Hurwitz et al., 2016).
Studies of hydrothermal explosion depos-
its at Waiotapu, New Zealand, suggest that the
existence of a steam zone perched beneath a
low-permeability cap provides a locus for the
accumulation of non-condensable and com-
pressible gases such as CO2 (Hedenquist and
Henley, 1985), which may enhance the vulner-
ability of such systems to hydrothermal explo-
sions. In addition to water-level changes modi-
fying the pressure regime, seismic events might
affect cap-rock permeability or provoke rapid
depressurization, releasing non-condensable
gases with sufcient force to cause rock frag-
mentation and debris ejection (Hedenquist and
Henley, 1985). Although differences exist in
hydrothermal explosion mechanisms for subla-
custrine hot-water and vapor zones, each is still
poorly understood in Yellowstone and similar
active hydrothermal terrains.
CONCLUSIONS
The identification of a vapor-rich zone
beneath the Deep Hole east of Stevenson Island
in Yellowstone National Park presents a previ-
ously unappreciated phenomenon associated
with sublacustrine hydrothermal activity. Vent-
uid chemistry, stable-isotope, and enthalpy
constraints are consistent with sublacustrine
steam mixing with lake water prior to vent-
ing. Vent sediments are dominated by kaolinite,
boehmite, and pyrite, and are compatible with
a CO2-saturated, H2S-bearing, and mildly acidic
steam condensate–lake water mixture. A steam
zone with accumulated gases may respond to
lake-level changes or seismic activity, perhaps
with localized hydrothermal explosions, or other
changes to hydrothermal dynamics of sublacus-
trine hydrothermal processes.
ACKNOWLEDGMENTS
This work was funded by U.S. National Science Foun-
dation (NSF) grants EAR-1515377 and OCE-1434798.
The authors thank Dave Lovalvo, the skilled engineers
onboard R/V Annie, and the Global Foundation for
Ocean Exploration for their efforts that contributed
to the successful outcome of the project. We also
thank Dr. Shijun Wu (Zhejiang University) for the
high-pressure gas-tight sampling system, and Dr. Rob
Sohn and members of the HD-YLAKE research team
for constructive advice. Part of this work was per-
formed at the Institute for Rock Magnetism (IRM)
at the University of Minnesota. The IRM is a U.S.
National Multi-user Facility supported through the
Instrumentation and Facilities program of the NSF,
Earth Sciences Division, and by funding from the
University of Minnesota. We thank Dr. Christopher
Mills, U.S. Geological Survey (Denver, Colorado,
USA), for analyzing water samples for hydrogen
and oxygen isotopes and optimizing precision and
accuracy of these results. We thank Stuart Simmons,
Cornel deRonde, and an anonymous reviewer for their
helpful comments.
fδD (‰ VSMOW)
Sublacustrine
steam source
(assuming YL17F01
= 17% steam)
-180
-170
-160
-150
-140
-130
-120
-22-20 -18-16 -14
δ
18
O (‰ VSMOW)
Y. Lake
YL16F12
YL17F01
Subaerial fumaroles
Subaerial hot springs
Local meteoric water
Rock
reaction
Figure 3. Oxygen and hydrogen isotope rela-
tionships for sublacustrine vent fluids from the
Deep Hole east of Stevenson Island, Yellow-
stone Lake, Wyoming, USA (yellow symbols)
and subaerial thermal waters and fumaroles
from throughout Yellowstone National Park
(black symbols). Yellowstone meteoric water
line is from Kharaka et al. (2002); subaerial
fumaroles and thermal waters are from Berg-
feld et al. (2014). The Y. Lake sample represents
a bottom-water sample collected near the Ste-
venson Island deep hydrothermal vents and
analyzed in this study (blue circle). The theo-
retical oxygen and hydrogen isotope values
for steam that contributes to sublacustrine
vents is calculated assuming sample YL17F01
is a mixture between bottom lake water and
17% steam (red symbol). VSMOW—Vienna
standard mean ocean water.
Downloaded from https://pubs.geoscienceworld.org/gsa/geology/article-pdf/doi/10.1130/G45577.1/4617352/g45577.pdf
by University of Minnesota user
on 22 January 2019
4 www.gsapubs.org
|
Volume 47
|
Number 3
|
GEOLOGY
|
Geological Society of America
REFERENCES CITED
Al-Tabbaa, A., and Wood, D.M., 1987, Some measure-
ments of the permeability of kaolin: Geotechnique,
v.37, p.499–514, https:// doi .org /10 .1680 /geot
.1987 .37 .4 .499.
Balistrieri, L.S., Shanks, W.C., III, Cuhel, R.L., Aguilar,
C., and Klump, J.V., 2007, The inuence of subla-
custrine hydrothermal vents on the geochemistry of
Yellowstone Lake, in Morgan, L.A., ed., Integrated
Geoscience Studies in the Greater Yellowstone
Area—Volcanic, Tectonic, and Hydrothermal Pro-
cesses in the Yellowstone Geoecosystem: U.S. Geo-
logical Survey Professional Paper 1717, p. 169–199.
Bergfeld, D., Lowenstern, J.B., Hunt, A.G., Shanks, W.P.,
III, and Evans, W.C., 2014, Gas and isotope chemis-
try of thermal features in Yellowstone National Park,
Wyoming (version 1.1): U.S. Geological Survey
Scientic Investigations Report 2011-5012, 28 p.,
https:// doi .org /10 .3133 /sir20115012.
Browne, P.R.L., and Ellis, A.J., 1970, The Ohaki-Broad-
lands hydrothermal area, New Zealand: Mineral-
ogy and related geochemistry: American Journal
of Science, v.269, p.97–131, https:// doi .org /10
.2475 /ajs .269 .2 .97.
Browne, P.R.L., and Lawless, J.V., 2001, Character-
istics of hydrothermal eruptions, with examples
from New Zealand and elsewhere: Earth-Science
Reviews, v.52, p.299–331, https:// doi .org /10 .1016
/S0012 -8252 (00)00030 -1.
de Ronde, C.E.J., et al., 2002, Discovery of active hy-
drothermal venting in Lake Taupo, New Zealand:
Journal of Volcanology and Geothermal Research,
v.115, p.257–275, https:// doi .org /10 .1016 /S0377
-0273 (01)00332 -8.
de Ronde, C.E.J., et al., 2016, Reconstruction of the
geology and structure of Lake Rotomahana and its
hydrothermal systems from high-resolution multi-
beam mapping and seismic surveys: Effects of the
1886 Tarawera Rift eruption: Journal of Volcanol-
ogy and Geothermal Research, v.314, p.57–83,
https:// doi .org /10 .1016 /j .jvolgeores .2016 .02 .002.
Duan, Z.H., and Sun, R., 2003, An improved model cal-
culating CO2 solubility in pure water and aqueous
NaCl solutions from 273 to 533 K and from 0 to
2000 bar: Chemical Geology, v.193, p.257–271,
https:// doi .org /10 .1016 /S0009 -2541 (02)00263 -2.
Gemery-Hill, P.A., Shanks, W.C., III, Balistrieri, L.S., and
Lee, G.K., 2007, Geochemical data for selected riv-
ers, lake waters, hydrothermal vents, and subaerial
geysers in Yellowstone National Park, Wyoming and
vicinity, 1996–2004, in Morgan, L.A., ed., Integrated
Geoscience Studies in the Greater Yellowstone
Area—Volcanic, Tectonic, and Hydrothermal Pro-
cesses in the Yellowstone Geoecosystem: U.S. Geo-
logical Survey Professional Paper 1717, p.365–426.
Hayden, F.V., 1878, Survey of the territories—A report of
progress of the exploration in Wyoming and Idaho
for the year 1878, Part II: Yellowstone National Park:
Geology, thermal springs, topography: U.S. Geo-
logical and Geographical Survey Annual Report 12,
503 p., https:// doi .org /10 .3133 /70038941.
Hedenquist, J.W., 1990, The thermal and geochemical
structure of the Broadlands-Ohaaki geothermal sys-
tem, New Zealand: Geothermics, v.19, p.151–185,
https:// doi .org /10 .1016 /0375 -6505 (90)90014 -3.
Hedenquist, J.W., and Henley, R.W., 1985, Hydrother-
mal eruptions in the Waiotapu geothermal system,
New Zealand: Their origin, associated breccias,
and relation to precious metal mineralization: Eco-
nomic Geology and the Bulletin of the Society of
Economic Geologists, v.80, p.1640–1668, https://
doi .org /10 .2113 /gsecongeo .80 .6 .1640.
Hurwitz, S., and Lowenstern, J.B., 2014, Dynamics of
the Yellowstone hydrothermal system: Reviews of
Geophysics, v.52, p.375–411, https:// doi .org /10
.1002 /2014RG000452.
Hurwitz, S., Clor, L.E., McCleskey, R.B., Nordstrom, D.K.,
Hunt, A.G., and Evans, W.C., 2016, Dissolved gases
in hydrothermal (phreatic) and geyser eruptions at
Yellowstone National Park, USA: Geology, v.44,
p.235–238, https:// doi .org /10 .1130 /G37478 .1.
Ingebritsen, S.E., and Sorey, M.L., 1988, Vapor-dom-
inated zones within hydrothermal systems: Evo-
lution and natural state: Journal of Geophysical
Research, v.93, p.13,635–13,655, https:// doi .org
/10 .1029 /JB093iB11p13635.
Kharaka, Y.K., Thordarson, J.J., and White, L.D., 2002,
Isotope and chemical compositions of meteoric
and thermal waters and snow from the greater Yel-
lowstone National Park region: U.S. Geological
Survey Open-File Report 02-194, 75 p., https://
doi .org /10 .3133 /ofr02194.
Klump, J.V., Remsen, C.C., and Kaster, J.L., 1988, The
presence and potential impact of geothermal activ-
ity on the chemistry and biology of Yellowstone
Lake, Wyoming, in DeLuca, M.P., and Babb, I., eds.,
Global Venting, Midwater, and Benthic Ecological
Processes: National Oceanic and Atmospheric Ad-
ministration National Undersea Research Program
Research Report 88-4, p. 81–98.
Lowenstern, J.B., Bergfeld, D., Evans, W.C., and Hunt,
A.G., 2015, Origins of geothermal gases at Yellow-
stone: Journal of Volcanology and Geothermal Re-
search, v.302, p.87–101, https:// doi .org /10 .1016
/j .jvolgeores .2015 .06 .010.
Montanaro, C., Scheu, B., Gudmundsson, M.T., Vog-
fjörd, K., Reynolds, H.I., Dürig, T., Strehlow, K.,
Rott, S., Reuschlé, T., and Dingwell, D.B., 2016,
Multidisciplinary constraints of hydrothermal ex-
plosions based on the 2013 Gengissig lake events,
Kverkfjöll volcano, Iceland: Earth and Planetary
Science Letters, v.434, p.308–319, https:// doi .org
/10 .1016 /j .epsl .2015 .11 .043.
Morgan, L.A., et al., 2003, Exploration and discovery
in Yellowstone Lake: Results from high-resolution
sonar imaging, seismic reection proling, and
submersible studies: Journal of Volcanology and
Geothermal Research, v.122, p.221–242, https://
doi .org /10 .1016 /S0377 -0273 (02)00503 -6.
Morgan, L.A., Shanks, W.C.P., III, and Pierce, K.L.,
2009, Hydrothermal Processes above the Yellow-
stone Magma Chamber: Large Hydrothermal Sys-
tems and Large Hydrothermal Explosions: Geo-
logical Society of America Special Paper 459, 95 p.,
https:// doi .org /10 .1130 /2009 .2459 (01).
Mufer, L.J.P., White, D.E., and Truesdell, A.H., 1971,
Hydrothermal explosion craters in Yellowstone Na-
tional Park: Geological Society of America Bulle-
tin, v.82, p.723–740, https:// doi .org /10 .1130 /0016
-7606 (1971)82 [723: HECIYN]2 .0 .CO;2.
National Park Service, 2018, National Park Service Visi-
tor Use Statistics, volume 2018.
Olsen, H.W., 1966, Darcy’s law in saturated kaoli-
nite: Water Resources Research, v.2, p.287–295,
https://irma.nps.gov/Stats/SSRSReports /Park % 20
Specic % 20Reports /Annual %20Park %20Recre-
ation%20Visitation%20(1904% 20 -% 20 Last %20
Calendar %20Year.
Pierce, K.L., and Morgan, L.A., 1992, The track of the
Yellowstone hot spot—Volcanism, faulting, and up-
lift, in Link, P.K., et al., eds., Regional Geology of
Eastern Idaho and Western Wyoming: Geological
Society of America Memoir 179, p. 1–54, https://
doi .org /10 .1130 /MEM179 -p1.
Raharjo, I.B., Allis, R.G., and Chapman, D.S., 2016, Vol-
cano-hosted vapor-dominated geothermal systems
in permeability space: Geothermics, v.62, p.22–32,
https:// doi .org /10 .1016 /j .geothermics .2016 .02 .005.
Remsen, C.C., et al., 2002, Sublacrustrine geothermal
activity in Yellowstone Lake: Studies past and
present, in Anderson, R.J., and Harmon, D., eds,
Yellowstone Lake: Hotbed of Chaos or Reservoir
of Resilience? Proceedings of the 6th Biennial
Scientic Conference on the Greater Yellowstone
Ecosystem, 8–10 October 2001, Yellowstone Na-
tional Park, p. 192–212.
Schubert, G., Straus, J.M., and Grant, M.A., 1980, A
problem posed by vapour-dominated geothermal
systems: Nature, v.287, p.423–425, https:// doi
.org /10 .1038 /287423a0.
Shanks, W.C., III, Alt, J.C., and Morgan, L.A., 2007,
Geochemistry of sublacustrine hydrothermal depos-
its in Yellowstone Lake—Hydrothermal reactions,
stable-isotope systematics, sinter deposition, and
spire formation, in Morgan, L.A., ed., Integrated
Geoscience Studies in the Greater Yellowstone
Area—Volcanic, Tectonic, and Hydrothermal Pro-
cesses in the Yellowstone Geoecosystem: U.S. Geo-
logical Survey Professional Paper 1717, p.201–234.
Simmons, S.F., and Browne, P.R.L., 2000, Hydrothermal
minerals and precious metals in the Broadlands-
Ohaaki geothermal system: Implications for under-
standing low-suldation epithermal environments:
Economic Geology and the Bulletin of the Society
of Economic Geologists, v.95, p.971–999, https://
doi .org /10 .2113 /gsecongeo .95 .5 .971.
Smith, T., and McKibben, R., 2000, An investigation of
boiling processes in hydrothermal eruptions, in Pro-
ceedings of the World Geothermal Congress, Kyushu,
Japan, 28 May–10 June: Auckland, New Zealand, In-
ternational Geothermal Association, v. 1, p. 699–704.
Sohn, R., Harris, R., Linder, C., Luttrell, K., Lovalvo,
D., Morgan, L., Seyfried, W.E., Jr., and Shanks, P.,
2017, Exploring the restless oor of Yellowstone
Lake: Eos (Washington, D.C.), v.98, https:// doi .org
/10 .1029 /2017EO087035.
Stucker, V.K., de Ronde, C.E.J., Scott, B.J., Wilson, N.J.,
Walker, S.L., and Lupton, J.E., 2016, Subaerial and
sublacustrine hydrothermal activity at Lake Ro-
tomahana: Journal of Volcanology and Geother-
mal Research, v.314, p.156–168, https:// doi .org
/10 .1016 /j .jvolgeores .2015 .06 .017.
Tan, C., Cino, C.D., Ding, K., and Seyfried, W.E., Jr., 2017,
High temperature hydrothermal vent uids in Yel-
lowstone Lake: Observations and insights from in-
situ pH and redox measurements: Journal of Volca-
nology and Geothermal Research, v.343, p.263–270,
https:// doi .org /10 .1016 /j .jvolgeores .2017 .07 .017.
Tivey, M.A., de Ronde, C.E.J., Caratori Tontini, F.,
Walker, S.L., and Fornari, D.J., 2016, A novel heat
ux study of a geothermally active lake—Lake
Rotomahana, New Zealand: Journal of Volcanol-
ogy and Geothermal Research, v.314, p.95–109,
https:// doi .org /10 .1016 /j .jvolgeores .2015 .06 .006.
Wagner, W., and Kretzschmar, H.-J., 2008, International
Steam Tables: Properties of Water and Steam Based
on the Industrial Formulation IAPWS-IF97 (second
edition): Berlin, Heidelberg, Springer-Verlag, 392 p.
Walker, S.L., de Ronde, C.E.J., Fornari, D.J., Tivey, M.A.,
and Stucker, V.K., 2016, High-resolution water col-
umn survey to identify active sublacustrine hydro-
thermal discharge zones within Lake Rotomahana,
North Island, New Zealand: Journal of Volcanol-
ogy and Geothermal Research, v.314, p.142–155,
https:// doi .org /10 .1016 /j .jvolgeores .2015 .07 .037.
White, D.E., Mufer, L.J.P., and Truesdell, A.H., 1971,
Vapor-dominated hydrothermal systems compared
with hot-water systems: Economic Geology and the
Bulletin of the Society of Economic Geologists, v.66,
p.75–97, https:// doi .org /10 .2113 /gsecongeo .66 .1 .75.
White, D.E., Fournier, R.O., Mufer, L.J.P., and Trues-
dell, A.H., 1975, Physical results of research drill-
ing in thermal areas of Yellowstone National Park,
Wyoming: U.S. Geological Survey Professional
Paper 892, 70 p., https:// doi .org /10 .3133 /pp892.
Wold, R.J., Mayhew, M.A., and Smith, R.B., 1977, Bathy-
metric and geophysical evidence for a hydrothermal
explosion crater in Mary Bay, Yellowstone Lake, Wyo -
ming: Journal of Geophysical Research, v.82, p.3733–
3738, https:// doi .org /10 .1029 /JB082i026p03733.
Wu, S.-J., Yang, C.-J., Pester, N.J., and Chen, Y., 2011, A
new hydraulically actuated titanium sampling valve
for deep-sea hydrothermal uid samplers: IEEE
Journal of Oceanic Engineering, v.36, p.462–469,
https:// doi .org /10 .1109 /JOE .2011 .2140410.
Printed in USA
Downloaded from https://pubs.geoscienceworld.org/gsa/geology/article-pdf/doi/10.1130/G45577.1/4617352/g45577.pdf
by University of Minnesota user
on 22 January 2019
Article
Hydrothermal explosions are significant potential hazards in Yellowstone National Park, Wyoming, USA. The northern Yellowstone Lake area hosts the three largest hydrothermal explosion craters known on Earth empowered by the highest heat flow values in Yellowstone and active seismicity and deformation. Geological and geochemical studies of eighteen sublacustrine cores provide the first detailed synthesis of the age, sedimentary facies, and origin of multiple hydrothermal explosion deposits. New tephrochronology and radiocarbon results provide a four-dimensional view of recent geologic activity since recession at ca. 15–14.5 ka of the >1-km-thick Pinedale ice sheet. The sedimentary record in Yellowstone Lake contains multiple hydrothermal explosion deposits ranging in age from ca. 13 ka to ~1860 CE. Hydrothermal explosions require a sudden drop in pressure resulting in rapid expansion of high-temperature fluids causing fragmentation, ejection, and crater formation; explosions may be initiated by seismicity, faulting, deformation, or rapid lake-level changes. Fallout and transport of ejecta produces distinct facies of subaqueous hydrothermal explosion deposits. Yellowstone hydrothermal systems are characterized by alkaline-Cl and/or vapor-dominated fluids that, respectively, produce alteration dominated by silica-smectite-chlorite or by kaolinite. Alkaline-Cl liquids flash to steam during hydrothermal explosions, producing much more energetic events than simple vapor expansion in vapor-dominated systems. Two enormous explosion events in Yellowstone Lake were triggered quite differently: Elliott’s Crater explosion resulted from a major seismic event (8 ka) that ruptured an impervious hydrothermal dome, whereas the Mary Bay explosion (13 ka) was triggered by a sudden drop in lake level stimulated by a seismic event, tsunami, and outlet channel erosion.
Article
Full-text available
Yellowstone Lake, far from any ocean, hosts underwater hot springs similar to those on mid-ocean ridges. A research team is investigating the processes that drive the lake's hydrothermal systems.
Article
Full-text available
Hydrothermal explosions frequently occur in geothermal areas showing various mechanisms and energies of explosivity. Their deposits, though generally hardly recognised or badly preserved, provide important insights to quantify the dynamics and energy of these poorly understood explosive events. Furthermore the host rock lithology of the geothermal system adds a control on the efficiency in the energy release during an explosion. We present results from a detailed study of recent hydrothermal explosion deposits within an active geothermal area at Kverkfjöll, a central volcano at the northern edge of Vatnajökull. On August 15th 2013, a small jökulhlaup occurred when the Gengissig ice-dammed lake drained at Kverkfjöll. The lake level dropped by approximately 30 m, decreasing pressure on the lake bed and triggering several hydrothermal explosions on the 16th. Here, a multidisciplinary approach combining detailed field work, laboratory studies, and models of the energetics of explosions with information on duration and amplitudes of seismic signals, has been used to analyse the mechanisms and characteristics of these hydrothermal explosions. Field and laboratory studies were also carried out to help constrain the sedimentary sequence involved in the event. The explosions lasted for 40–50 s and involved the surficial part of an unconsolidated and hydrothermally altered glacio-lacustrine deposit composed of pyroclasts, lavas, scoriaceous fragments, and fine-grained welded or loosely consolidated aggregates, interbedded with clay-rich levels. Several small fans of ejecta were formed, reaching a distance of 1 km north of the lake and covering an area of approximately 0.3 km2, with a maximum thickness of 40 cm at the crater walls. The material (volume of approximately 104 m3) has been ejected by the expanding boiling fluid, generated by a pressure failure affecting the surficial geothermal reservoir. The maximum thermal, craterisation and ejection energies, calculated for the explosion areas, are on the order of 1011, 1010 and 109 J, respectively. Comparison of these with those estimated by the volume of the ejecta and the crater sizes, yields good agreement. We estimate that approximately 30% of the available thermal energy was converted into mechanical energy during this event. The residual energy was largely dissipated as heat, while only a small portion was converted into seismic energy. Estimation of the amount of freshly-fragmented clasts in the ejected material obtained from SEM morphological analyses, reveals that a low but significant energy consumption by fragmentation occurred. Decompression experiments were performed in the laboratory mimicking the conditions due to the drainage of the lake. Experimental results confirm that only a minor amount of energy is consumed by the creation of new surfaces in fragmentation, whereas most of the fresh fragments derive from the disaggregation of aggregates. Furthermore, ejection velocities of the particles (40–50 m/s), measured via high-speed videos, are consistent with those estimated from the field. The multidisciplinary approach used here to investigate hydrothermal explosions has proven to be a valuable tool which can provide robust constraints on energy release and partitioning for such small-size yet hazardous, steam-explosion events.
Article
ROV investigation of hydrothermal fluids issuing from vents on the floor of Yellowstone lake revealed temperatures in excess of 170 °C — the highest temperature yet reported for vent fluids within Yellowstone National Park (YNP). The study site is east of Stevenson Island at depth of approximately 100–125 m. In-situ pH and redox measurements of vent fluids were made using solid state sensors designed to sustain the elevated temperatures and pressures. YSZ membrane electrode with Ag/Ag2O internal element and internal pressure balanced Ag/AgCl reference electrode were used to measure pH, while a platinum electrode provided redox constraints. Lab verification of the pH sensor confirmed excellent agreement with Nernst law predictions, especially at temperatures in excess of 120 °C. In-situ pH values of between 4.2 and 4.5 were measured for the vent fluids at temperatures of 120 to 150 °C. The slightly acidic vent fluids are likely caused by CO2 enrichment in association with magmatic degassing effects that occur throughout YNP. This is consistent with results of simple model calculations and direct observation of CO2 bubbles in the immediate vicinity of the lake floor vents. Simultaneous redox measurements indicated moderate to highly reducing conditions (− 0.2 to − 0.3 V). As typical of measurements of this kind, internal and external redox disequilibria likely preclude unambiguous determination of redox controlling reactions. Redox disequilibria, however, can be expected to drive microbial metabolism and diversity in the near vent environment. Thus, the combination of in-situ pH and redox sensor deployments may ultimately provide the requisite framework to better understand the microbiology of the newly discovered hot vents on Yellowstone lake floor.
Article
Although volcano-hosted vapor-dominated geothermal systems are rare, West Java, Indonesia, has five such systems: Kamojang, Darajat, Patuha, Telagabodas, and Wayang Windu. This paper seeks an explanation for such vapor-dominated geothermal systems through heat and mass transfer numerical modeling based on the characteristics of these five systems. Specifically, the 3D model TOUGH2 is used to explore heat input and permeability conditions that lead to vapor-dominated systems. Our generic model consists of a reservoir, a host rock, and a caprock. The size of the reservoir is 16 km2 and 2.8 km thick, with 10% porosity and a permeability of 10−13 m2. The initial condition of the model is fully liquid saturated at a normal geotherm with hydrostatic pressure. An intrusive heatpulse of 8 MW/km2 for 9 kyr is applied to the base to the reservoir and the resulting geothermal system characteristics are mapped out in permeability space. A vapor-dominated system is produced when the hostrock permeability is 5 × 10−17 m2 or less accompanied by the caprock permeability varying from 10−17 to 4 × 10−16 m2. The caprock permeability assures that the escaping steam exceeds the incoming liquids; the hostrock permeability prevents flooding until the heat source is turned off. Increasing the basal heat input to 12 MW/km2 shifts the maximum host-rock permeability for vapor systems to 10−16 m2, but there is minimal change in the maximum caprock permeability. In all models the maximum duration of the vapor reservoirs is 1.5 kyr, at which point the reservoir either dries out or floods. The timescale for vapor reservoir stability is an order of magnitude less than that for the formation or cooling of liquid-dominated geothermal reservoirs. The factors contributing to the occurrence of the five vapor-dominated reservoirs in West Java are intense heating due to prolonged active volcanism, an absence of shear faulting, and the restrictive range of low permeability in the host and caprocks surrounding a relatively permeable reservoir.
Article
Multiphase and multicomponent fluid flow in the shallow continental crust plays a significant role in a variety of processes over a broad range of temperatures and pressures. The presence of dissolved gases in aqueous fluids reduces the liquid stability field toward lower temperatures and enhances the explosivity potential with respect to pure water. Therefore, in areas where magma is actively degassing into a hydrothermal system, gas-rich aqueous fluids can exert a major control on geothermal energy production, can be propellants in hazardous hydrothermal (phreatic) eruptions, and can modulate the dynamics of geyser eruptions. We collected pressurized samples of thermal water that preserved dissolved gases in conjunction with precise temperature measurements with depth in research well Y-7 (maximum depth of 70.1 m; casing to 31 m) and five thermal pools (maximum depth of 11.3 m) in the Upper Geyser Basin of Yellowstone National Park, USA. Based on the dissolved gas concentrations, we demonstrate that CO2 mainly derived from magma and N2 from air-saturated meteoric water reduce the nearsurface saturation temperature, consistent with some previous observations in geyser conduits. Thermodynamic calculations suggest that the dissolved CO2 and N2 modulate the dynamics of geyser eruptions and are likely triggers of hydrothermal eruptions when recharged into shallow reservoirs at high concentrations. Therefore, monitoring changes in gas emission rate and composition in areas with neutral and alkaline chlorine thermal features could provide important information on the natural resources (geysers) and hazards (eruptions) in these areas.
Article
Lake Rotomahana is a crater lake in the Okataina Volcanic Centre (New Zealand) that was significantly modified by the 1886 Tarawera Rift eruption. The lake is host to numerous sublacustrine hydrothermal vents. Water column studies were conducted in 2011 and 2014 along with sampling of lake shore hot springs and crater lakes in Waimangu Valley to complement magnetic, seismic, bathymetric and heat flux surveys. Helium concentrations below 50 m depth are higher in 2014 compared to 2011 and represent some of the highest concentrations measured, at 6 × 10− 7 ccSTP/g, with an end-member 3He/4He value of 7.1 RA. The high concentrations of helium, when coupled with pH anomalies due to high dissolved CO2 content, suggest the dominant chemical input to the lake is derived from magmatic degassing of an underlying magma. The lake shore hot spring waters show differences in source temperatures using a Na–K geothermometer, with inferred reservoir temperatures ranging between 197 and 232 °C. Water δ18O and δD values show isotopic enrichment due to evaporation of a steam heated pool with samples from nearby Waimangu Valley having the greatest enrichment. Results from this study confirm both pre-1886 eruption hydrothermal sites and newly created post-eruption sites are both still active.
Article
The Broadlands-Ohaaki geothermal system is a boiling hydrothermal system hosted by a sequence of Quaternary felsic volcanic rocks and Mesozoic metasediments. More than 50 wells have been drilled (400 to >2,600 m deep) to assess the geothermal potential for the production of electricity. Fluids and precipitates sampled from wells, along with descriptions of the alteration minerals in more than 500 drill cores, provide a three-dimensional picture of the distribution of fluid types and secondary minerals. Interpretation of these features and the distribution of gold and silver highlight the relationship between alteration and mineralization in an active, low-sulfidation epithermal environment. Quartz, illite, K feldspar (adularia), albite, chlorite, calcite, and pyrite are the main hydrothermal minerals that occur in the deep central upflow zone at ≥250°C and >600 m depth. These minerals form through recrystallization of the volcanic host rocks and incorporation of H 2O, CO 2, and H 2S in the presence of a deeply derived chloride water containing ~1,000 mg/kg Cl and ~26,400 mg/kg CO 2. At the same time, and on the periphery of the upflow zone, illite, smectite, calcite, and siderite form through hydrolitic alteration in the presence of CO 2-rich steam-heated waters that contain <30 mg/kg Cl and ~13,000 mg/kg CO 2. Upward and outward from the deep central upflow zone, mineral patterns reflect the shift from rock-dominated to fluid-dominated alteration and the prevailing influence of boiling, mixing, and cooling on fluid-mineral equilibria. Accordingly, the abundance of quartz and K feldspar increase toward the upflow zone, whereas clay abundance increases toward the margin of the upflow zone (with smectite dominating at <150°C and illite dominating at >200°C); the abundance of chlorite, pyrite, and calcite varies here, but albite is absent. Geothermal production wells with high fluid fluxes are the main sites of precious-metal mineralization. The deep chloride water (with or without minor amounts of vapor) enters the well at depths >500 m and undergoes a pressure drop that causes boiling. As a result, precious metals precipitate and accumulate as scales on back-pressure plates or as detritus in surface weir boxes; these deposits contain <10 to >1,000 mg/kg Au, <100 to >10,000 mg/kg Ag and ~10 to ~1,000 mg/kg As and Sb, each. Within production wells, platy calcite deposits as a scale at the site of first boiling near the fluid feed point, while crustiform-colloform-banded amorphous silica deposits in surface pipe work. By contrast, the hydrothermally altered host rocks contain low concentrations of gold, ranging from <0.01 to 1.0 mg/kg Au (68 analyses), and these correlate positively with arsenic (<100 to ~5,000 mg/kg) and antimony (<10 to ~500 mg/kg). Reaction path modeling using SOLVEQ and CHILLER shows that calcite, K feldspar, gold, and amorphous silica deposit in sequence from a chloride water that cools along an adiabatic boiling path (300°to 100°C), analogous to fluid flow in a production well. By contrast, calcite, quartz, K mica, and pyrite deposit from a chloride water that cools due to mixing with CO 2-rich steam-heated waters; dilution prevents precipitation of precious metals. Thus field observations and reaction path modeling demonstrate that boiling is the main process influencing the deposition of precious metals. The results of this study show how peripheral hydrolytic alteration by CO 2-rich steam-heated waters relate to propylitic and potassic alteration by chloride waters in the epithermal environment of a hydrothermal system. Both the distribution of alteration mineral assemblages associated with the different water types and the broad-scale distribution of temperature-sensitive smectite and illite reflect the location of the upflow zone. On a local scale, the occurrence of platy calcite, crustiform-colloform silica, and K feldspar in veins indicates the existence of boiling conditions conducive to precious-metal deposition.
Article
A new technique for measuring conductive heat flux in a lake was adapted from the marine environment to allow for multiple measurements to be made in areas where bottom sediment cover is sparse, or even absent. This thermal blanket technique, pioneered in the deep ocean for use in volcanic mid-ocean rift environments, was recently used in the geothermally active Lake Rotomahana, New Zealand. Heat flow from the lake floor propagates into the 0.5 m diameter blanket and establishes a thermal gradient across the known blanket thickness and thereby provides an estimate of the conductive heat flux of the underlying terrain. This approach allows conductive heat flux to be measured over a spatially dense set of stations in a relatively short period of time. We used 10 blankets and deployed them for 1 day each to complete 110 stations over an 11-day program in the 6 x 3 km lake. Results show that Lake Rotomahana has a total conductive heat flux of about 47 MW averaging 6 W/m2 over the geothermally active lake. The western half of the lake has two main areas of high heat flux; 1) a high heat flux area averaging 21.3 W/m2 along the western shoreline, which is likely the location of the pre-existing geothermal system that fed the famous Pink Terraces, mostly destroyed during the 1886 eruption 2) a region southwest of Patiti Island with a heat flux averaging 13.1 W/m2 that appears to be related to the explosive rift that formed the lake in the 1886 Tarawera eruption. A small rise in bottom water temperature over the survey period of 0.01 °C/day suggests the total thermal output of the lake is ~ 112-132 MW and when compared to the conductive heat output suggests that 18-42% of the total thermal energy is by conductive heat transfer.