Bonneville Basin Shoreline

Quaternary Science Reviews 58 (2012) 1e6
Contents lists available at SciVerse ScienceDirect
Quaternary Science Reviews
journal homepage: www.elsevier.com/locate/quascirev
Bonneville basin shoreline records of a large lake during Marine Isotope Stage 16
Shizuo Nishizawa a, *, Donald R. Currey b,1, Andrea Brunelle c, Dorothy Sack d
a
Isogo, Yokohama 235-0023, Japan
Department of Geography, University of Utah, Salt Lake City, UT 84112, USA
c
Records of Environment and Disturbance Laboratory, Department of Geography, University of Utah, Salt Lake City, UT 84112, USA
d
Department of Geography, Ohio University, Athens, OH 45701, USA
b
a r t i c l e i n f o
a b s t r a c t
Article history:
Received 27 December 2011
Received in revised form
21 September 2012
Accepted 11 October 2012
Available online 7 November 2012
The occurrence of the 650,000-year-old Rye Patch Dam tephra within shoreline sedimentary sequences
suggests the presence of a large lake in the Bonneville basin in North America during Marine Isotope
Stage (MIS) 16. The observed shoreline sedimentology and stratigraphy indicate that the lake was
expanding when the air-fall ash landed onto the lake water. The minimum estimated surface area of the
650 ka lake almost equals that of Lake Bonneville during the Provo stage in late MIS 2. The magnitude of
the 650 ka large lake implies that the climatic and hydrologic conditions in the Bonneville basin during
early MIS 16 might have been comparable to those in late MIS 2.
Ó 2012 Elsevier Ltd. All rights reserved.
Keywords:
Bonneville
Shoreline
Marine Isotope Stage 16
Rye Patch Dam tephra
Great Basin
1. Introduction
The 135,000 km2 Bonneville drainage basin occupies the eastern
part of the North American Great Basin (Fig. 1). Being hydrologically
closed, lake size in the Bonneville basin is largely determined by
climate through the net balance between precipitation and evaporation. The magnitude of fluctuating paleolake sizes in the basin,
therefore, closely reflects changing effective moisture. Shifts in the
polar jet stream’s mean path, which was influenced by the
changing size of the Cordilleran and Laurentide ice sheets, have
been long considered as one of the primary causes of large amplitude Pleistocene lake size fluctuations in the Bonneville basin
(Antevs, 1948; Oviatt, 1997; Kaufman, 2003). Lake Bonneville, for
example, is estimated to have had a surface area of 50,500 km2 at its
maximum stage in the cool, wet, late MIS 2 climatic conditions
whereas Great Salt Lake in today’s semiarid environment has had
an average surface area of 4400 km2 over the last 150 years (Arnow
and Stephens, 1990; Currey, 1990). A close linkage between large
lake intervals in the Bonneville basin and periods of major glaciation is well illustrated in an investigation of the Burmester Core,
which indicates that a large lake occurred only during some of the
most extensive glaciations: MIS 16, 12, 6, and 2 (Oviatt et al., 1999).
* Corresponding author.
E-mail address: sn1734@geog.utah.edu (S. Nishizawa).
1
Deceased.
0277-3791/$ e see front matter Ó 2012 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.quascirev.2012.10.014
Few shoreline-based studies in the Bonneville basin provide
records of large lake intervals older than MIS 6 with reliable age
control. In contrast, a large number of studies have provided
shoreline records of many lakes in the western Great Basin from the
early to the late Pleistocene, by using tephrochronology (Morrison,
1991; Reheis, 1999; Reheis et al., 2002; Knott et al., 2008; Kurth
et al., 2011). The shoreline-based lacustrine records in the
western Great Basin played an important role in providing a basis
for the hypothesis that a long-term decline in lake size throughout
the Great Basin might have resulted from a regional drying and
warming trend since the middle Pleistocene (Reheis, 1999; Reheis
et al., 2002; Kurth et al., 2011).
This study presents shoreline evidence for a large MIS 16 Bonneville basin lake and estimates its size by analyzing nearshore
stratigraphy. Identifying the 650 ka Rye Patch Dam tephra in the
context of shoreline stratigraphy makes it possible to estimate
shoreline depositional environments and approximate surface area
of the early middle Pleistocene lake.
1.1. Previous studies of MIS 16 lake in the Bonneville basin
The possible existence of a deep lake in the Bonneville basin
during MIS 16 (680e620 ka) was previously suggested by Morrison
(1965), Eardley et al. (1973), Scott et al. (1982), McCoy (1987), Oviatt
et al. (1999), and Kaufman (2003). These studies inferred that the
regressive phase of the MIS 16 deep lake likely preceded the
2
S. Nishizawa et al. / Quaternary Science Reviews 58 (2012) 1e6
Fig. 1. Location of the Bonneville basin and lake surface area (after Nishizawa, 2010). The 50,500 km2 Lake Bonneville maximum extent (dark blue) (B5 shoreline of Burr and Currey,
1988, 1992), the 37,400 km2 650 ka lake (light blue), and the 4400 km2 Great Salt Lake (white). Localities discussed in the text: BCeBurmester sediment core; DCeDeep Creek dam;
LVeLittle Valley; AeLake Lahontan; BeLake Jonathan; CeLake Manly; DeWind River Basin; Eepre-Illinoian glaciation deposits; pre-Illinoian Laurentide ice sheet extent after Roy
et al. (2004). (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)
deposition of the Lava Creek B tephra 640 ka (Lanphere et al., 2004).
Morrison (1965), Scott et al. (1982), and McCoy (1987) located the
subaerially deposited Lava Creek B tephra overlying lacustrine
sediment in the Little Valley gravel pit near Promontory Point, Utah
(Fig. 1 LV). Oviatt et al. (1999) also identified Lava Creek B tephra
directly overlying a major lacustrine unit in the Burmester sediment core (Fig. 1 BC). In addition, the amino acid analyses on
mollusk and ostracode shells from the pre-Lava Creek B lacustrine
sediment broadly implied an MIS 16 age for a deep lake (McCoy,
1987; Oviatt et al., 1999; and; Kaufman, 2003). All these studies
indicate that a lake of unspecified magnitude (‘Lava Creek lake’ of
Oviatt et al., 1999) existed in the Bonneville basin prior to the
640 ka Lava Creek B tephra deposition.
2. Study site
The study site is located where a major paleostream (ancestor of
modern day Deep Creek) flowed into Lake Bonneville during the
last ice age (Fig. 1 DC). Around the Last Glacial Maximum, this
stream carried meltwater from the alpine glaciers in the Deep
Creek Range, forming numerous deltas as it flowed into Lake
Bonneville (Currey et al., 1983). The Deep Creek Range (Fig. 1) is
a high-angle rotational fault system with a 35 e44 tilt resulting in
a very steep western slope, and a less steep eastern slope on which
the study site is located (Miller et al., 1999). With its highest
elevation of 3684 m at Ibapah Peak, the Deep Creek Range is
underlain by an upper Precambrian and Palaeozoic shelf sequence,
intruded by granite rocks of the Late Cretaceous (75 Ma) and
EoceneeOligocene (39e35 Ma) ages (Miller et al., 1999; Osborn and
Bevis, 2001). The 1979 excavation and construction of the Deep
Creek Dam on Bureau of Land Management leased land
(4018.5280 N, 113 56.1720 W, 1420e1438 m a.s.l.) exposed stratigraphy that resulted from a series of lacustrine shoreline depositions
from the middle to late Pleistocene lake intervals. A gravel spit
complex on the dam’s left abutment is the most salient shoreline-
geomorphic feature at the dam site (Fig. 2), and was formed
during multiple large lake events that occurred from MIS 3 to MIS 2
(Nishizawa, 2010). On the right abutment, several exposures show
different lithofacies from those of the gravel spit complex on the
left abutment (Fig. 2). This study examines the stratigraphy and
sedimentology of the exposures on the right abutment.
3. Methods
Sedimentary sequences at the study site were measured and
described. Elevations at each measured stratigraphic section were
determined by a combination of 1:24,000 topographic maps,
a geodetic total station, benchmarks, and global positioning system.
All surveyed elevations are accurate within 1 m. In addition, all
observed lacustrine depositional features, except for marl facies
analyzed in this study, were interpreted to be in the shore-zone
from foreshore to backshore/lagoon (Currey and Sack, 2009).
Fig. 2. A 2004 photo of Deep Creek Dam/Reservoir (dry), looking north towards the
Bonneville Salt Flats, Silver Island Mountains, and Pilot Peak. Deep Creek flows from
the left to the right of the photo. The dark dot on the left abutment is a 1995 Jeep
Grand Cherokee for reference. An aspect of Fig. 3 is shown.
S. Nishizawa et al. / Quaternary Science Reviews 58 (2012) 1e6
Samples were collected for marl, mollusk shells, and tephra
layers. The marl and mollusk shell samples were radiocarbon dated
at Beta Analytic Inc. in Florida. Electron microprobe analyses of the
tephra samples were conducted by: F. Foit at the GeoAnalytical
Laboratory in the Department of Geology, Washington State
University, in 2002 (sample UT-02-ELPHK-02a, written personal
communication) (Table 1); and M. Perkins at the Department of
Geology and Geophysics, University of Utah, in 2002 (sample UT02-ELPHK-02b, written personal communication). Analytical
conditions and procedures at the GeoAnalytical Laboratory were
the same as those described in Hallett et al. (2001) and Martin et al.
(2005). Table 1 shows the glass compositions of the tephra sample.
These were compared to the standard glasses in the Geoanalytical
Laboratory’s Pacific Northwest tephra database that contains
>1500 samples. The similarity coefficient of Borchardt et al. (1972)
was used as a discriminator.
4. Descriptions of stratigraphy
The base of the Deep Creek Dam right abutment exposure is
a >5-m-thick marl unit (unit 11, Fig. 3A) that contains moderately
abundant ostracode shells. This marl yielded a nonfinite radiocarbon age of >43,120 14C yr BP (Beta-169982). Unconformably
overlying this marl unit is a >3-m-thick greenish (in the web
version) tan, clayey, laminated mud with abundant large ostracode
shells (unit 10, Fig. 3A). On the wall facing southwest (Fig. 3B), this
greenish (in the web version) mud is overlain by a 35-cm-thick
layer of pebbles, which fine upwards, and range from angular at the
bottom to subrounded at the top of the unit (unit 9, Fig. 3B).
Overlying these pebbles is a 9-cm-thick thinly bedded to laminated
tephra, which is interbedded with five to seven fine sand layers
(unit 8, Fig. 3B) (1435 m a.s.l. at the measured section). The contacts
between units 8, 9, and 10 are sharp, but conformable.
Above unit 8, a reworked tephra gradually transitions upward
into 20e25-cm-thick, medium to thick-bedded ostracodal silty mud
(unit 7, Fig. 3B). A unit of 20-cm-thick oxidized subangular pebbles
(unit 3, Fig. 3B) unconformably overlies unit 7. Unit 3 separates the
3
underlying older units from an overlying 5-cm-thick, fine lacustrine
sand of MIS 2 age that contains moderately abundant Stagnicola
shells dated as 20,230 120 14C yr BP (Beta-189275) (unit 2, Fig. 3B).
Despite the substantial period of time separating unit 7 and MIS 2unit 2, no buried soils were observed. A thinly bedded 30-cm-thick
sandy marl with abundant ostracode shells unconformably overlies
the 20.2 14C ka sand (unit 2) and caps the exposure (unit 1, Fig. 3B).
A northwest facing exposure (Fig. 3C) at the same site lies across
a gully approximately 20 m to the southwest from the southwest
facing wall (see Fig. 3A). Unit 10 of greenish tan clayey mud, unit 8
of the laminated tephra, and unit 7 of ostracodal silty mud with
reworked tephra, are present at 1427e1431 m (unit 8 slopes up 2e
4 from southwest to northeast, see Fig. 3A). Here, however, the
layer of pebbles (unit 9, Fig. 3B) between the greenish (in the web
version) mud and tephra is absent (Fig. 3C). Instead, the tephra
directly overlies the greenish (in the web version) mud. This
contact between units 8 and 10 is sharp, but conformable. The
tephra unit here is also 9 cm thick and interbedded with fine sand
layers (unit 8, Fig. 3C). From unit 8 of the northwest facing exposure, samples were collected for tephra analyses (ELPHK-02a and
ELPHK-02b). Above unit 8, a reworked tephra gradually transitions
upward into 20e25-cm-thick ostracodal silty mud (unit 7, Fig. 3C).
Unconformably overlying these units is a 25-cm-thick unit of
gravelly mud (unit 6, Fig. 3C). This gravel consists of mostly angular
pebbles, which appear to have originated from local rhyolitic and
dacitic flows of unknown age. A 20-cm-thick mud unconformably
overlies unit 6 (unit 5, Fig. 3C). Above this mud is a 40-cm-thick unit
of angular pebbles in a muddy matrix (unit 4, Fig. 3A), which grades
upward into beach boulders of MIS 2 age that crop out at the
modern day surface. The sedimentology and stratigraphy of units 5
and 4 appear to indicate a regressive phase of an MIS 2 lake of an
unknown age.
5. Interpretation of stratigraphy
The water-laid tephra (unit 8) was identified as the Rye Patch
Dam (RPD) tephra by F. Foit (written personal communication,
Table 1
Electron microbeam analyses on Unit 8 tephra.
Sample
UT-02ELPHK-02a
SiO2
Analyses normalized
to 100 weight %
Standard deviations
of the analyses
TiO2 Al2O3 MgO CaO
MnO Fe2O3 Na2O K2O
Cl
Notes
71.66 0.45
14.47 0.39
1.53 0.03
3.32
4.53
3.49 0.13 Deep Creek Dam right abutment
0.38 0.05
0.13 0.05
0.14 N/A
0.12
0.16
0.13 0.02 Total # of shards analyzed ¼ 26
5 closest matches
Source
#1
70.90 0.43
14.80 0.35
1.42 0.08
3.36
5.20
3.50
#2
72.20 0.44
14.55 0.35
1.46 0.09
3.40
3.70
3.68 0.10
#3
71.72 0.33
14.46 0.38
1.34
3.44
4.95
3.30 0.10
#4
70.87 0.55
14.79 0.53
1.68
3.28
4.90
3.25 0.20
#5
72.70 0.44
14.40 0.39
1.49 0.04
2.22
4.60
3.40 0.10
0.93
0.95
0.88
0.91
0.97
0.99
0.98
0.97
0.99
0.67
0.87
0.82
0.92
0.92
0.99
0.99
0.95
0.95
0.93
0.97
Reference
RPD
Tule Lake Core at 65.38 m
(Rieck et al., 1992)
RPD
Bonneville basin S28 core
(Williams, 1994)
Willowdale, Oregon; Infotec
Research, Craig Skinner
Upper ash, Summer Lake-Ana River
Exposure locality; P. Wigand, Desert
Research Institute
Mt. Mazama
Southwest Caldera Wall, Crater
Pumice Castle Lake, Oregon (Davis, 1985)
Similarity coefficient for 5 closest matches
#1
#2
#3
#4
#5
Similarity coefficient
weighted average
0.96
0.95
0.94
0.94
0.94
0.99
0.99
0.99
0.99
0.99
0.96
0.98
0.73
0.82
0.98
0.98
0.99
0.99
0.98
0.99
0.90
0.90
0.98
0.74
1.00
*All analyses by F. Nick Foit, GeoAnalytical Lab, Washington State University.
RPD
RPD
Mt. Mazama
Pumice Castle
Same
Same
Same
Same
Same
as
as
as
as
as
#1
#2
#3
#4
#5
above
above
above
above
above
4
S. Nishizawa et al. / Quaternary Science Reviews 58 (2012) 1e6
Fig. 3. Deep Creek Dam right abutment, looking east, showing exposures and stratigraphic columns of southwest facing and northwest facing walls. All numbered stratigraphic
units are the same in panels A, B, and C (i.e., unit 10 in panel A is the same unit as unit 10 in B and C). Unit 8 is the RPD tephra.
2002) (Table 1) and M. Perkins (written personal communication,
2002). The RPD tephra is believed to have erupted during an event
that produced the Desert Spring ash-flow in the Cascade Range
(Hill, 1985; Rieck et al., 1992; Reheis, 1999; Reheis et al., 2002). The
RPD tephra was also identified in the Bonneville basin sediment
core S208 (0.5 cm thick at a depth of 154.1 m) by Williams (1993,
1994). The depositional environment of the S208 RPD tephra,
however, was not reported in either of Williams’ studies.
Martin et al. (2005) summarized estimated ages of the RPD
tephra that range from 630 ka to 690 ka. In the study of the Tule
Lake core, Rieck et al. (1992) estimated the age to be 630 ka. The
only available absolute date by laser fusion 40Ar/39Ar, 686 50 ka,
was reported by Fleck’s personal communication (2004) to Martin
et al. (2005). Following the improvement of precisions in the Lava
Creek B tephra age estimate (Lanphere et al., 2004), Sarna-Wojcicki
estimated the RPD tephra date to be 646 ka, based on its relative
position to the overlying 639 0.2 ka Lava Creek B tephra and the
underlying BrunheseMatuyama Chron boundary in the Tule Lake
core (in the 2004 personal communication to Martin et al., 2005).
Because of the large error associated with the 40Ar/39Ar RPD tephra
date, we followed Sarna-Wojcicki’s estimated age, 646 ka. A future,
more precise age estimate of the RPD tephra, will refine the timing
and duration of this large lake in the Bonneville basin: an older
date indicating a long lake interval starting from the MIS 16/17
boundary or early MIS 16; or a younger date indicating a short
mid-late MIS 16 lake interval that ended with Lava Creek B tephra
deposition.
The absence of observable unconformities separating units 7, 8,
9, and 10 suggests that a shallow to deep lacustrine environment at
the study site persisted during early MIS 16. Laminations of the
tephra unit and its subsequent reworking also indicate that the airfall tephra was likely deposited in shallow water above wave base.
The observed sedimentology and stratigraphy suggest that units 7
and 8 were likely deposited in a foreshore zone. Unit 9 was interpreted as beach pebbles in a shore zone from foreshore to backshore/lagoon (see Currey and Sack, 2009). When the RPD tephra
(unit 8) was laid down in shallow water, it draped over these beach
pebbles (unit 9). Fining up pebbles of unit 9 also suggests that the
lake level was rising at the time of the RPD tephra deposition
650 ka. Because the Lava Creek B tephra appears to mark the end of
a large lake as reported in previous Bonneville basin studies (Oviatt
et al., 1999), it is likely that an MIS 16 lake of substantial size was
expanding w650 ka, and then shrinking w640 ka.
Estimating the exact size of the 650 ka lake through observed
shoreline elevations alone is difficult because details of basin
tectonics, and numerous hydroisostatic rebound events since MIS
S. Nishizawa et al. / Quaternary Science Reviews 58 (2012) 1e6
16, are not well understood. The Deep Creek Range, with the study
site located on its north end, comprises part of the 150-km-long
north-south-trending Snake RangeeDeep Creek Range fault system
(Miller et al., 1999). A high-angle rotational fault in the Deep Creek
Range is reported to have accumulated at least 12 km of rapid slip in
Oligocene time (37e34 Ma) and Miocene time (18e14 Ma) as
a result of the Tertiary extension of the Basin and Range province
(Miller et al., 1999). The Quaternary slip rate, especially from 650 ka
to today, is unknown. In addition, the surveyed elevations in this
study are reported with no adjustment for hydroisostatic rebound.
Current models for the Bonneville lake basin hydroisostatic
rebound are based on the MIS 2 Lake Bonneville chronology in
which the lake started expanding approximately 26,000 calendar
years ago (cal ka), and reached its maximum stage 18 cal ka (Oviatt,
1997; Bills et al., 2002).
Despite uncertainties involving basin tectonics and hydroisostatic rebound, approximating the magnitude of the 650 ka
lake from the observed shoreline elevations is useful for estimating MIS 16 hydrologic and climatic conditions in the Bonneville basin. In the Bonneville basin, once lake surface elevation is
determined, lake surface area can be estimated because of its
semi-linear relationship with lake surface elevation (Currey and
Oviatt, 1985; Nishizawa, 2010). This study follows Godsey et al.
(2005) by assuming a lake surface level to be 5 m at the time
of deposition of a surveyed stratigraphic unit in a foreshore zone.
Because unit 8, which contains thinly bedded to laminated tephra,
appears to have been deposited within the foreshore zone, we
interpreted that the observed maximum elevation of the 650 ka
shoreline sediment, 1435 m, would indicate a lake surface
elevation range of 1430e1440 m. It should also be noted that the
lake continued to expand after unit 8 was formed, as indicated in
the overlying unit 7. The actual lake size after the RPD tephra
deposition was, therefore, likely much larger than our estimate.
The lake surface elevation at 1440 m would have an area of
w37,400 km2, which is 74% of MIS 2 Lake Bonneville’s maximum
surface area, 50,500 km2 during the Bonneville stage (B5 shoreline of Burr and Currey, 1988, 1992) (Nishizawa, 2010) (Fig. 1). This
estimated 650 ka lake area almost equals that of Lake Bonneville
during the Provo stage in late MIS 2. The magnitude of the 650 ka
lake indicates that the hydrologic and climatic conditions in the
Bonneville basin during early MIS 16 were at least comparable to
those of late MIS 2.
6. MIS 16 large lake in the Bonneville basin and northern
hemisphere glaciation
On a regional scale, the 650 ka large lake in the Bonneville basin
is consistent with a number of MIS 16 large lakes reported in the
western Great Basin (Morrison, 1991; Reheis, 1999; Reheis et al.,
2002). Lake Lahontan (Fig. 1 A) appears to have left the highest
shoreline complex, which was correlated with the Rye Patch Dam
Alloformation (Reheis et al., 2002). This MIS 16 Lake Lahontan
would be the largest Pleistocene lake ever identified in the western
Great Basin. In Kobeh Valley, which lies between Lake Lahontan and
the Bonneville basin (Fig. 1 B), the youngest lake sediments and
highest beach gravels of Lake Jonathan are constrained by the RPD
tephra (Reheis et al., 2002). Similar to the MIS 16 Bonneville basin
shoreline facies, the RPD tephra marks the onset of the Lake
Jonathan-MIS 16-transgressive phase, which eventually resulted in
the lake overflowing and becoming part of the Lahontan drainage
system (Reheis et al., 2002). Moreover, the highest shorelines of
Lakes Columbus, Hubbs, Newark, and a lake in Pine Valley, suggest
that the largest Pleistocene lakes in these western Great Basin subbasins likely formed during MIS 16 (Reheis et al., 2002). In Death
Valley (Fig. 1 C), the Lava Creek B tephra was found interbedded
5
with lacustrine sediments, indicating the presence of a perennial
lake, Lake Manly, during a substantially wet phase in MIS 16 (Knott
et al., 2008). Because MIS 16 paleoshoreline data from Death Valley
are currently limited, an estimated dimension of MIS 16 Lake Many
varies from a large lake to several disconnected small lakes (Knott
et al., 2008). With this study’s finding from the eastern half of the
Great Basin added to the large lake reports from the western Great
Basin, it is highly likely that many major lake basins throughout the
Great Basin contained large to the largest Pleistocene lakes during
MIS 16.
The discovery of MIS 16, and other early Pleistocene lake
shoreline features that lie higher than those of the MIS 2 highstand
in the western Great Basin, have led Reheis (1999), Reheis et al.
(2002), and Kurth et al. (2011) to suggest a long-term decrease in
effective moisture available to the region from the middle to the
late Pleistocene. Reheis et al. (2002) attributed this drying trend to
interactions among such factors as climate, tectonics, drainage
changes, and changes in moisture sources. The exact mechanism or
processes of how these factors interacted with each other over
time, however, remain to be explored. Meanwhile, currently
available lake shoreline data in the Bonneville basin from the early
to the late Pleistocene, except for MIS 2, is too insufficient to
examine this long term drying trend.
Occurrences of exceptionally large MIS 16 lakes in the Great
Basin are consistent with records of North American continental ice
sheets. In North America, the MIS 16 Cordilleran and Laurentide ice
sheets expanded more than they did during MIS 2, indicating
a greater ice volume during MIS 16. Rocky Mountain glacial
deposits, for example, indicate that one of the greatest glacial
advances occurred during MIS 16 (Chadwick et al., 1997; Pierce,
2004). Glacial outwash deposits containing the Lava Creek B
tephra typically correlate well with the most extensive moraines in
the Rocky Mountains, e.g. the Sacagawea Ridge glaciation in the
Wind River Basin, Wyoming (Chadwick et al., 1997; Hall and
Jaworowski, 1999) (Fig. 1 D). By using paleomagnetic measurements, till compositional data, and the Lava Creek B tephra beds
from glacial sedimentary sequences in Iowa, Nebraska, Kansas, and
Missouri (Fig. 1 E), Roy et al. (2004) suggest that the MIS 16 Laurentide ice sheet formed some of the seven pre-Illinoian (before
MIS 10) advances that extended further south than those of the Last
Glacial Maximum (Fig. 1). This apparent correlation between the
large continental ice sheets and large lakes in the Great Basin seems
to support Antevs’ model (1948) that pluvial lake level changes
during ice ages in the Great Basin, including the Bonneville basin,
were mainly caused by shifts in the polar jet stream’s mean path,
which was influenced by the changing size of the Cordilleran and
Laurentide ice sheets.
On the hemispheric scale, MIS 16 left a pronounced d18O profile
in the Quaternary deep ocean sediment records (Raymo, 1994).
Lang and Wolff (2011) characterized MIS 16 as a particularly strong
glacial in both planktonic and benthic d18O, yet rather weak in sea
surface temperature records. They suggested that the isotopic
strength of MIS 16 could indicate more of an ice volume record than
a climate signal. MIS 16 also marks the beginning of glaciations,
driven by the 100,000 year orbital cycle, with longer duration and
more ice volume than those driven by the preceding 41,000 year
orbital cycle (Ruddiman, 2001).
Here, we have presented evidence of the first measurable
Quaternary large lake in the Bonneville basin, which appears to
have coincided with MIS 16dthe first long and extensive
Quaternary glaciation. This finding is supported by the proposed
linkages between fluctuating lakes in the Great Basin, migrating
mean position of the polar jet stream, expanding/contracting ice
sheets, and orbital cycles (Antevs, 1948; Oviatt, 1997; Oviatt et al.,
1999).
6
S. Nishizawa et al. / Quaternary Science Reviews 58 (2012) 1e6
Acknowledgements
Authors thank: Nick Foit and Mike Perkins for tephra analyses;
Nancy Carruthers, Molly Hanson, and Takashi Nishizawa for field
assistance; and Shawn Blissett for improving the manuscript. Don
Currey was partially supported by National Science Foundation
grants SBR-9817777 and EAR-980924. Two anonymous reviewers’
comments significantly improved the depth and context of the
manuscript.
References
Antevs, E., 1948. Climatic changes and pre-white man. In: The Great Basin, with
Emphasis on Glacial and Postglacial Times, vol. 38. Bull. of the Univ. of Utah,
pp. 167e191.
Arnow, T., Stephens, D., 1990. Hydrologic Characteristics of the Great Salt Lake,
Utah: 1847e1986. U.S. Geological Survey Water-Supply Paper 2332, 32 pp.
Bills, B.G., Wambeam, T.J., Currey, D.R., 2002. Geodynamics of lake Bonneville. In:
Gwynn, J.W. (Ed.), Great Salt Lake: an Overview of Change. Special Publication
of the Utah Department of Natural Resources, pp. 7e32.
Borchardt, G., Aruscavage, P., Millard, H., 1972. Correlation of the Bishop Ash,
a Pleistocene marker bed, using instrumental neutron activation analysis.
J. Sediment. Petrol. 42, 301e306.
Burr, T., Currey, D., 1988. The Stockton bar. In: Machette, M. (Ed.), In the Footsteps of G.K.
Gilbertdlake Bonneville and Neotectonics of the Eastern Basin and Range Province.
Utah Geological and Mineral Survey Miscellaneous Publication 88e1, pp. 67e73.
Burr, T., Currey, D., 1992. Hydrographic Modeling at the Stockton Bar. Utah
Geological Survey Miscellaneous Publication 92e3. 207e219.
Chadwick, O., Hall, R., Phillips, F., 1997. Chronology of Pleistocene glacial advances in
the central Rocky Mountains. Geol. Soc. Am. Bull. 109, 1443e1452.
Currey, D., 1990. Quaternary paleolakes in the evolution of semidesert basins, with
special emphasis on Lake Bonneville and the Great Basin, U.S.A. Paleogeogr.
Paleoclim. Paleoecol. 76, 189e214.
Currey, D., Oviatt, C., Plyler, G., 1983. Lake Bonneville Stratigraphy, Geomorphology,
and Isostatic Deformation in West-central Utah, vol. 62. Utah Geological and
Mineral Survey Special Study, pp. 63e82.
Currey, D., Oviatt, C., 1985. Durations, average rates, and probable causes of Lake
Bonnevile expansions, stillstands, and contractions during the last deep-lake
cycle: 32,000 to 10,000 years ago. Geographical J. Korea 10, 1085e1099.
Currey, D., Sack, D., 2009. Hemiarid lake basin: geomorphic patterns. In: Parsons, A.J.,
Abrahams, A.D. (Eds.), Geomorphology of Desert Environments, pp. 489e514.
Davis, J., 1985. Correlation of late Quaternary tephra layers in a long pluvial
sequence near summer lake. Oregon. Quat. Res. 23, 38e53.
Eardley, A., Shuey, R., Gvosdetky, V., Nash, W., Picard, M., Grey, D., Kukla, G., 1973.
Lake cycles in the Bonneville basin. Utah Geol. Soc. Am. Bull. 84, 211e216.
Godsey, H., Currey, D., Chan, M., 2005. New evidence for an extended occupation of
the Provo shoreline and implications for regional climate change, Pleistocene
Lake Bonneville, Utah, USA. Quat. Res. 63, 212e223.
Hall, R., Jaworowski, C., 1999. Reinterpretation of the Cedar Ridge section, Wind
river range, Wyoming: implications for the glacial chronology of the Rocky
Mountains. Geol. Soc. Am. Bull. 111, 1233e1249.
Hallett, D., Mathewes, R., Foit, F., 2001. Mid-Holocene Glacier Peak and Mount St.
Helens We tephra layers detected lake sediments from southern British
Columbia using high resolution techniques. Quat. Res. 55, 284e292.
Hill, B., 1985. Petrology of the Bend Pumice and Tumalo Tuff, a Pleistocene Cascade
Eruption Involving Magma Mixing. M.S. thesis, Oregon State Univ.
Kaufman, D., 2003. Amino acid paleothermometry of Quaternary ostracodes from
the Bonneville Basin. Utah Quat. Sci. Rev. 22, 899e914.
Knott, J., Machette, M., Klinger, R., Sarna-Wojcicki, A., Liddicoat, J., Tinsley III, J.,
David, B., Ebbs, V., 2008. Reconstructing late Pliocene to middle Pleistocene
Death Valley lakes and river systems as a test of pupfish (Cyprinodontidae)
dispersal hypotheses. In: Reheis, M., Hershler, R., Miller, D. (Eds.), Late Cenozoic
Drainage History of the Southwestern Great Basin and Lower Colorado River
Region: Geologic and Biotic Perspectives. Geol. Soc. of Am. Spec. Pap., vol. 439,
pp. 1e26.
Kurth, G., Phillips, F., Reheis, M., Redwine, J., Paces, J., 2011. Cosmogenic nuclide and
uranium-series dating of old, high shorelines in the western Great Basin, USA.
Geol. Soc. Am. Bull. 123, 744e768.
Lang, N., Wolff, E., 2011. Interglacial and glacial variability from the last 800 ka in
marine, ice and terrestrial archives. Clim. Past 7, 361e380.
Lanphere, M., Champion, D., Christiansen, R., Izett, G., Obradovick, J., 2004. Revised
ages for tuffs of the Yellowstone Plateau volcanic field: assignment of the
Huckleberry Ridge Tuff to a new geomagnetic polarity event. Geol. Soc. Am.
Bull. 114, 559e568.
Martin, J., Patrick, D., Kihm, A., Foit, F., Grandstaff, D., 2005. Lithostratigraphy,
tephrochronology, and rare earth element geochemistry fossils at the Classical
Pleistocene Fossil Lake Area, south central Oregon. J. Geol. 113, 139e155.
McCoy, W., 1987. Quaternary aminostratigraphy of the Bonneville basin, western
United States. Geol. Soc. Am. Bull. 98, 99e112.
Miller, E., Dumitri, T., Brown, R., Gans, P., 1999. Rapid Miocene slip on the Snake
RangeeDeep Creek range fault system, east-central Nevada. Geol. Soc. Am. Bull.
111, 886e905.
Morrison, R., 1965. New Evidence on Lake Bonneville Stratigraphy and History from
Southern Promontory Point, Utah. U.S. Geol. Surv. Prof. Pap. 525-C, C110eC119.
Morrison, R., 1991. Quaternary stratigraphic, hydrologic, and climatic history of the
Great Basin, with emphasis on Lakes Lahontan, Bonneville, and Tecopa. In:
Morrison, R. (Ed.), Quaternary Nonglacial Geology: Conterminous U.S. The
Geology of North America K-2. Geological Society of America, Boulder, Colorado, pp. 283e320.
Nishizawa, S., 2010. The Bonneville Lake Basin Shoreline Records of Large Lake and
Abrupt Climate Change Events. Ph.D. dissertation, Univ. of Utah.
Osborn, G., Bevis, K., 2001. Glaciation in the Great Basin of the United States. Quat.
Sci. Rev. 20, 1377e1410.
Oviatt, C., 1997. Lake Bonneville fluctuations and global climate change. Geology 25,
155e158.
Oviatt, C., Thompson, R., Kaufman, D., Bright, J., Forester, R., 1999. Reinterpretation
of the Burmester Core, Bonneville basin, Utah. Quat. Res. 52, 80e184.
Pierce, K., 2004. Pleistocene glaciations of the Rocky Mountains. In: Gillespie, A.,
Porter, S., Atwater, B. (Eds.), The Quaternary Period in the United States.
Developments in Quaternary Science, vol. 1. Elsevier, pp. 63e76.
Raymo, M., 1994. The initiation of northern hemisphere glaciation. Annu. Rev. Earth
Planet. Sci. 22, 353e383.
Reheis, M., 1999. Highest pluvial-lake shorelines and Pleistocene climate of the
western Great Basin. Quat. Res. 52, 196e205.
Reheis, M., Sarna-Wojcicki, A., Reynolds, R., Repenning, C., Mifflin, M., 2002. Pleistocene to middle Pleistocene lakes in the western Great Basin: ages and
connections. In: Hershler, R., Madsen, D., Currey, D. (Eds.), Great Basin Aquatic
History. Smithsonian Contribution to the Earth Sciences, vol. 33. Smithsonian
Institution Press, Washington, D.C., pp. 53e108.
Rieck, H., Sarna-Wojcicki, A., Meyer, C., Adam, D., 1992. Magnetostratigraphy and
tephrochronology of an upper Pliocene to Holocene record in lake sediments at
Tule Lake, northern California. Geol. Soc. Am. Bull. 104, 409e428.
Roy, M., Clark, P., Barendregt, R., Glassmann, J., Enkin, R., 2004. Glacial stratigraphy
and paleomagnetism of late Cenozoic deposits of the north-central United
States. Geol. Soc. Am. Bull. 116, 30e41.
Ruddiman, W., 2001. Earth’s Climate Past and Future. W.H. Freeman and Company,
New York.
Scott, W., Shroba, R., McCoy, W., 1982. Guidebook for the 1982 Friends of the
Pleistocene, Rocky Mountain Cell, Field Trip to Little Valley and Jordan Valley,
Utah. U.S. Geol. Surv.. Open-File Rep. 82e845.
Williams, S., 1993. Tephrochronology and Basinal Correlation of Ash Deposits in the
Bonneville Basin, Northwest Utah. M.S. thesis, Univ. of Utah.
Williams, S., 1994. Late Cenozoic tephrostratigraphy of deep sediment cores from
the Bonneville Basin, Northwest Utah. Geol. Soc. Am. Bull. 105, 1517e1530.