(Geological Society of America Field Guides 1) David R. Lageson, Alan Lester, Bruce Trudgill - GSA Field Guide 1 - Colorado and Adjacent Areas - Geological Society of America (1999)
(Geological Society of America Field Guides 1) David R. Lageson, Alan Lester, Bruce Trudgill - GSA Field Guide 1 - Colorado and Adjacent Areas - Geological Society of America (1999)
(Geological Society of America Field Guides 1) David R. Lageson, Alan Lester, Bruce Trudgill - GSA Field Guide 1 - Colorado and Adjacent Areas - Geological Society of America (1999)
Edited by
David R. Lageson
GSA Field Guide Editor
Department of Earth Sciences
Montana State University
Bozeman, Montana 59717-0348
Alan P. Lester
Co-Editor
Department of Geological Sciences
University of Colorado
Boulder, Colorado 80309-0399
and
Bruce D. Trudgill
Co-Editor
Department of Geological Sciences
University of Colorado
Boulder, Colorado 80309-0399
Field Guide 1
1999
Copyright © 1999, The Geological Society of America, Inc. (GSA). All rights reserved. GSA grants per-
mission to individual scientists to make unlimited photocopies of one or more items from this volume
for noncommercial purposes advancing science or education, including classroom use. Permission is
granted to individuals to make photocopies of any item in this volume for other noncommercial, non-
profit purposes provided that the appropriate fee ($0.25 per page) is paid directly to the Copyright
Clearance Center, 222 Rosewood Drive, Danvers, MA 01923, USA, phone (978) 750-8400,
http://www.copyright.com (include title and ISBN when paying). Written permission is required from
GSA for all other forms of capture or reproduction of any item in the volume including, but not limited
to, all types of electronic or digital scanning or other digital or manual transformation of articles or any
portion thereof, such as abstracts, into computer-readable and/or transmittable form for personal or cor-
porate use, either noncommercial or commercial, for-profit or otherwise. Send permission requests to
GSA Copyrights.
Copyright is not claimed on any material prepared wholly by government employees within the scope of
their employment.
Printed in U.S.A.
Colorado and adjacent areas / edited by David R. Lageson, Alan Lester, and Bruce Trudgill.
p. cm. -- (Field guide ; 1)
Includes bibliographical references.
ISBN 0-8137-0001-9
1. Geology--Colorado--Guidebooks. 2. Geology--West (U.S.)--Guidebooks. 3.
Colorado--Guidebooks. 4. West (U.S.)--Guidebooks. I. Lageson, David R. II Lester,
Alan, 1960- III. Trudgill, Bruce, 1964- IV. Field guide (Geological Society of America) ;
1.
Cover: Andrew’s Peak in southwestern Rocky Mountain National Park. Photo by Diane C. Lorenz
10 9 8 7 6 5 4 3 2 1
ii
Contents
Preface. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . v
Note . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . vi
1. Bouncing boulders, rising rivers, and sneaky soils: A primer of geologic hazards and
engineering geology along Colorado’s Front Range . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
D. C. Noe, J. M. Soule, J. L. Hynes, and K. A. Berry
4. Hydrogeology and wetlands of the mountains and foothills near Denver, Colorado . . . . . . . . . . 51
K. E. Kolm and J. C. Emerick
5. Field trip to Manitou Springs, Colorado, with specific emphasis on the sediments of
Cave of the Winds and their relationship to nearby alluvial deposits and spring sediments . . . . 61
F. G. Luiszer
6. 200,000 years of climate change recorded in eolian sediments of the High Plains
of eastern Colorado and western Nebraska . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71
D. R. Muhs, J. B. Swinehart, D. B. Loope, J. N. Aleinikoff, and J. Been
8. Active evaporite tectonics and collapse in the Eagle River valley and the southwestern
flank of the White River uplift, Colorado . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 97
R. B. Scott, D. J. Lidke, M. R. Hudson, W. J. Perry, Jr., B. Bryant, M. J. Kunk, J. R. Budahn,
and F. M. Byers, Jr.
9. Coal mining in the 21st century: Yampa coal field, northwest Colorado . . . . . . . . . . . . . . . . . . 115
M. E. Brownfield, E. A. Johnson, R. H. Affolter, and C. E. Barker
10. Field guide to the continental Cretaceous-Tertiary boundary in the Raton basin,
Colorado and New Mexico . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 135
C. L. Pillmore, D. J. Nichols, and R. F. Fleming
12. Field guide for the Heart Mountain detachment and associated structures, northeast Absaroka
Range, Wyoming . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 177
D. H. Malone, T. A. Hauge, and E. C. Beutner
iii
Preface
This inaugural volume of the Geological Society of America (GSA) Field Guide series
represents an entirely new publication venue for the Society, one that emphasizes the “field
roots” of the earth sciences profession. The timing of this new series seems both appropri-
ate and ironic, considering that geology at the close of the twentieth century is immersed in
a level of technology that has drawn attention away from its field-oriented beginnings. In an
era when universities are altering their traditional field-based curricula in favor of more
marketable (computer-based) skills, it seems that going to the field and observing has
become a thing of the past. In contrast, we feel that a broad-based field perspective is cru-
cial to understanding most geological problems. Field observations provide for the germina-
tion of new ideas and are, ultimately, their testing ground.
The Reverend John Walker, who taught the first systematic course in geology at the
University of Edinburgh (1781–1803), had this to say to his students: “The way to knowl-
edge of natural history is to go to the fields, the mountains, the oceans, and observe, collect,
identify, experiment, and study.” Two hundred years later, there is still much to be said for
his approach. It is in this spirit that GSA has created this Field Guide series.
The new Field Guide series will initially showcase field excursions associated with
national meetings of the Society, maintaining the high standards of excellence for which
other GSA publications are known. The series may be expanded in the near future to show-
case selected field trips from sectional meetings as well as special field excursions.
Moreover, as professional earth scientists, we hold a responsibility to promote public out-
reach and education, to address those geological issues relevant to the needs of society, and
to disseminate information about the Earth in formats that do the greatest good for the
greatest number. Therefore, it is hoped that the Field Guide series will not only serve as a
valuable resource for our profession, but also serve as a venue for public education and out-
reach.
As we prepare to slide into the year 2000, we dedicate this first Field Guide volume to
those field-based earth scientists of the twentieth century who walked the outcrops and paid
their dues in the field.
DAVID R. LAGESON
GSA Field Guide Editor
Department of Earth Sciences
Montana State University
ALAN LESTER
Co-Editor
Department of Geological Sciences
University of Colorado
BRUCE TRUDGILL
Co-Editor
Department of Geological Sciences
University of Colorado
v
Note
This volume was compiled from field trips offered as part of the 1999
Geological Society of America Annual Meeting in Denver. The following field
trips could not be included in this volume because of unusually tight publication
deadlines.
South Park conjunctive use project: A combined look at geology and hydrology in
the South Park Basin, Colorado
M. F. McHugh, J. Jehn, and H. Eastman
Geology and paleontology of the gold belt back-country byway: Florissant Fossil
Beds and Garden Park fossil area
H. Meyer, T. W. Henry, D. Grenard, and E. Evanoff
vi
Geological Society of America
Field Guide 1
1999
C
ments with planning issues involving engineering geology. Since 70 17 Cr
Green i v er ee
k
the early 1970s, CGS geologists have reviewed plans for hun- Mountain
te R
25
dreds of subdivisions along the Colorado Piedmont. The authors Bear Cre 18
ek Bear Creek
la t
have provided emergency assistance and advice for incidents Lake Englewood
S o uth P
such as heaving bedrock (e.g., Noe and Dodson, 1997) for the
470 20 Littleton
nt
Chatfield Highlands
Lake Ranch
Figure 1. Index map showing the field-trip route in relation to cities and
Plu
other prominent natural and cultural features in the Denver Metropolitan 10 miles
m
Area. The locations of actual stops for the GSA Annual Meeting field trip
Cr
ee
are marked by bull’s-eye dots; dots with arrows denote “roll-by” locations. 10 km
k
Noe, D. C., Soule, J. M., Hynes, J. L., and Berry, K. A., 1999, Bouncing boulders, rising rivers, and sneaky soils: A primer of geologic hazards and engi-
neering geology along Colorado’s Front Range, in Lageson, D. R., Lester, A. P., and Trudgill, B. D., eds., Colorado and Adjacent Areas: Boulder, Colorado,
Geological Society of America Field Guide 1.
1
2 D. C. Noe et al.
This field trip guide may be used for self-guided parties. Boulder Baseline Reservoir
Lafayette
Although it contains general directions and locations of the stops,
Superior Louisville
we recommend bringing along a “close up” atlas and street guide .
Cr
of the Metropolitan Denver Area to aid in locating and driving l de r ee k
S o . B ou Cr
between the field trip stops. We have found these detailed street al
Co Broomfield
U D
maps to be a valuable tool for locating geologic-hazard areas in
Ku1 Rocky
rapidly developing, suburban areas. Flats
Northglenn
TKu3
Ku2 Thornton
PART 1. DENVER TO BOULDER PC
Golden Fault Westminster
Ralston
Reservoir
Stop 1 (Roll-by)—Geologic Setting Tv
(0.0-12.3 miles; 0.0-19.8 km) Arvada ek
TRJKl Ku1 r C re
North Table Cle a
Commerce
PZ Mountain
City
Directions: Leaving downtown Denver, drive north on Speer Tv
Boulevard. Exit right onto Interstate 25 (north), then exit onto
U.S. Highway 36 to Boulder (west). Golden
South Table
Mountain
A'
The foothills of the Colorado Piedmont are situated between Denver
the Denver Basin to the east and the Front Range uplift of the A Lakewood h erry
C
Cr
i v er
Rocky Mountains to the west. The Denver Basin contains nearly Green
ee
k
13,000 feet (4,000 m) of Pennsylvanian to Paleocene sedimen- Mountain
te R
Bear Cre
tary rocks (Fig. 2; Table 1). The basin is highly asymmetrical in ek Bear Creek
la t
U D Lake Englewood
its east-west cross-section (Fig. 3). Most of the Denver Metro-
S o uth P
PC Morrison
politan Area is underlain by the gently dipping, eastern limb. In
contrast, the foothills area is underlain by the steeply dipping, TKu3
Fr o
western limb. The Front Range uplift rises 1,500-2,500 feet Littleton
nt
above the Piedmont along the mountain front. The uplift contains Ku2
Ra
Plu
uplift, between Boulder and Colorado Springs. Of these, we will 10 miles
m
Cr
encounter the Golden Fault at Stop 15. The fault is considered to be
ee
10 km
k
inactive by most geologists, although some Quaternary movement
E X P L A N AT IO N
has been postulated for an associated fault graben near Golden (See Table 1 for For mation Abbreviations)
(Kirkham, 1977). We will encounter a second set of faults in the Tv Tertiary Volcanics (in TKd)
Boulder-Weld coal field area near Marshall, at Stop 5. These faults TKu3 Tertiary and Upper Cretaceous Continental (Tgm, TKd, Ka, Kl)
have a northeast-southwest trend, with normal displacements on Ku2 Upper Cretaceous Marginal Marine (Kfh)
Ku1 Upper Cretaceous Marine (Kp, Kn, Kcgg)
the order of a few feet to over 400 feet. They are interpreted to have TRJKl Triassic, Jurassic, and Lower Cretaceous (Kd, Jm, Jrc, TRPl)
formed contemporaneously with the deposition of coals during PZ Paleozoic (TRPl, Ply, IPf)
PC Precambrian
Cretaceous time, based on coal-thickness relationships (i.e., thicker
in grabens, thinner on horsts; see Weimer, 1977).
Unconsolidated, Quaternary deposits in the Denver area Figure 2. Index map showing the general bedrock geology of the Denver
include eolian (sand and loess), colluvial, and alluvial-terrace Metropolitan Area (modified from Trimble and Machette, 1979, and
deposits (Fig. 4). The alluvial terraces along the Front Range Pied- Costa and Bilodeau, 1982).
mont have been studied and subdivided by several authors, most
notably Hunt (1954) and Scott (1960). From oldest to youngest,
there are four well-defined, older alluvial-terrace deposits on rock- The general geology of the Colorado Piedmont has been
floored pediments: the Nussbaum, Rocky Flats, Verdos, and studied and mapped in detail. There have been several engineer-
Slocum Alluviums (Pleistocene). Additionally, there are five ing-geologic studies published for the area; however, few pub-
younger alluvial-fill terrace deposits: the Louviers and Broadway lished case-study investigations involving geologic hazards are
Alluviums (Pleistocene) and the Piney Creek Alluvium and post- available. The U.S. Geological Survey has published 1:24,000
Piney Creek alluvium (Holocene). These deposits are shown in a engineering geology maps for the entire field-trip area (Gardner,
schematic profile in Figure 5. A volcanic-ash bed in the Verdos 1969; Gardner and others, 1971; Miller and Bryant, 1976;
Alluvium has been dated at 0.6 m.y. (see Van Horn, 1976). McGregor and McDonough, 1980; Simpson and Hart, 1980), and
A WEST EAST A'
Front The Flatirons
Range and
Red Rocks Dakota
Hogback Table
Mountains Denver
Figure 3. Schematic cross-section through the
Tv Golden-Denver area showing the bedrock geology of
PC t TKu3 the Denver Basin and the Colorado Piedmont (modi-
ul
Fa Coal Beds fied from Wells, 1967; Van Horn, 1972; and Costa
n
l de (Laramie Fm.)
Ku2 and Bilodeau, 1982). See Figure 2 for cross-section
Go
location and explanation of geologic symbols.
Ku1
PC
PZ TRJKl
Not to Scale
4 D. C. Noe et al.
Green Cr
ee
Mountain k accompanied the rapid decline in silver prices around 1893.
t e Ri
Bear Cre Qe
ek Bear Creek Many mansions, commercial buildings, and churches were
Lake Englewood
la t
built during the late 1880s. The Colorado State Capitol was
th P
Morrison
Qal started in 1890. Many of these buildings are built from Col-
Qe
So u
Qe orado native stone and are the focus of an interesting tour for
Fr o
A century later, this region has seen rapid growth during the
N regional “energy-boom” years of the mid-1970s to early 1980s, a
nge
Qal Highlands
Chatfield Ranch “bust” from about 1984 to the early 1990s, and another boom
Lake accompanying national trends which continues to the present.
Qe
Plu
10 km
k
WEST EAST
FRONT
RANGE
Qrf Level of Major Stream
Pe di m en t Qv Lava Creek B Ash
Valleys
Pe dim en t (e.g., South Platte River)
Qs
Qlo
G ra d ie n t
o f Tr ib u ta
Pe dim en t Qb Qp Qpp Qb
r y S tr e a m
s
Not to Scale
Qrf Rocky Flats Alluvium (Aftonian/Nebraskan) Qlo Louviers Alluvium (Bull Lake)
Qv Verdos Alluvium (Yarmouth/Kansan) Qb Broadway Alluvium (Pinedale)
Qs Slocum Alluvium (Sangamon/Illinoian) Qp Piney Creek Alluvium (Holocene)
Qpp Post-Piney Creek alluvium (Holocene)
Figure 5. Schematic cross-section showing the Quaternary alluvial-terrace deposits in the Denver area (modified from Scott,
1963, and Madole, 1991).
Stop 3—Boulder Creek Valley Overlook (23.2 miles; 37.4 km) why a large, high-quality gravel resource was “unexpectedly”
discovered in the older, now-abandoned valley. We’ll see a
Directions: About a mile after passing the Louisville (McCaslin remarkably similar, pre-piracy situation later on the trip (at Ley-
Boulevard) exit on U.S. Highway 36, look for the “scenic over- den Gulch, Stop 11).
look” sign at the crest of the hill. Exit into the overlook area.
Our first stop is atop Davidson Mesa. It affords an excellent Stop 4 (Roll-by)—Table Mesa Area, Boulder
opportunity to overlook the valley of South Boulder Creek, the (27.1-29.2 miles; 43.6-47.0 km)
city of Boulder, the Flatirons (distinctive, reddish slabs of dipping
Fountain Formation conglomerate and sandstone), and the Front Directions: Continue along U.S. Highway 36 towards Boulder
Range beyond (Fig. 6). (northwest). Take the Table Mesa-Foothills Parkway exit, staying
The geologic display board at this site gives a good general to the right, and then turn left on Table Mesa Drive (west).
stratigraphic portrayal of the area. The mesa we’re standing on is For this roll-by, we will look at the engineering geology of
a classic example of topographic reversal. The Verdos-age allu- an urbanized, mountain-front community. The City of Boulder’s
vial-terrace gravel deposits that cap the mesa were deposited on Table Mesa area was subdivided and developed during the 1960s
this surface about 600 thousand years ago, when this area was a and 1970s. During development, many of the small, mountain-
valley. These gravel deposits now act to protect the soft sediments front stream drainages in the area were regraded, filled, or
below, while the flanking sediments have been eroded away to rerouted. Unfortunately, the resulting collection system was occa-
become sides of the mesa. sionally overwhelmed by runoff from summer thunderstorms.
Other topographic highs off to the north have been instru- Driving westward on Table Mesa Drive past Broadway, we will
mental in creating a stream piracy episode on South Boulder encounter Bear Creek, the major collecting stream for urban
Creek. The original channel of South Boulder Creek can be seen runoff from the area. The stream has undergone several episodes
trending off to the northeast, in a nearby valley just beyond Base- of periodic flash flooding because of the increased runoff due to
line Reservoir. Farther to the north stands Valmont Butte, an iso- urbanization. A number of flood-control improvements have
lated, basalt dike. At the western end of this butte, a been made along the drainage in recent years to reduce the effects
north-draining side stream had developed along a north-south, of aggradation and erosion. These include drop structures, which
strike valley in the soft shales of the Pierre Shale. This stream feature large, native boulders from the Fountain Formation, and
eroded headward (southward) along the edge of a more-resistant an oversized box culvert and bikeway beneath Broadway.
questa of Fox Hills Sandstone, and it had a significantly steeper Turning south on Lehigh Street, we will climb up a series of
gradient than that of South Boulder Creek. Its head eventually alluvial terraces of Quaternary age as we approach Shanahan Hill.
intercepted the original channel of South Boulder Creek and The present valley of Bear Creek contains the Piney Creek and
“captured” the larger stream, probably during a large flood event. post-Piney Creek alluviums. To the right (west), the scenic
In terms of land use, the piracy hypothesis explains why the cur- National Center for Atmospheric Research (designed by I.M. Pei)
rent valley of South Boulder Creek is a poor gravel resource, and is perched atop the area’s highest, oldest pediment terrace, under-
6 D. C. Noe et al.
lain by the Rocky Flats Alluvium. We begin a mile-long incline der and Longs Peak) to the stoplight. Turn right onto Broad-
underlain by the Slocum Alluvium. After passing a steeper slope way (south), which becomes State Highway 93. After a mile,
near the crest of the hill, underlain by the Pierre Shale, we finally turn left at the stoplight and drive eastward through the ham-
“top out” Shanahan Hill. There, the road meanders along a terrace let of Marshall to the corner of Marshall Road and Cherry-
underlain by the Verdos Alluvium. Circling to the east on Green- vale Road.
briar Boulevard, we descend the hill and reverse this sequence. Marshall is the discovery point for the Boulder-Weld
On Shanahan Hill, most of the Quaternary alluvial units are Coal Field, which stretches some thirty miles to the north-
alluvial fans that were deposited by debris flows and mud flows. east along the outcrop of the Upper Cretaceous Laramie
These deposits were flushed from the Flatirons above during local- Formation. The Laramie Formation is a sedimentary
ized, high-intensity downpour events. They typically consist of a sequence of interbedded shale, claystone and sandstone.
muddy sand matrix with interspersed boulders and cobbles. Note Coal seams are located throughout the formation. The large
the ubiquitous, large boulders that are now used for yard landscap- swales and ridges in the valley bottom to the south of Mar-
ing. These boulders give Boulder its name, and they present a for- shall Road are related to subsidence caused by both mining
midable challenge to the City’s home gardeners. The city has been and coal fires.
spared from catastrophic debris flows during historic times. The subsidence pattern in the open-space area to the
Throughout Colorado, however, there is a strong link between northeast of the road intersection reflects the room-and-pillar
wildfires and subsequent debris-flow activity (e.g., Coe, 1987; pattern of the workings directly below (Fig. 7). The mine
Soule, 1999; Kirkham et al., in preparation). Boulder’s debris-flow depth here is roughly 30-40 feet, and the extraction height was
safety may depend, to some degree, on careful fire suppression and 5-6 feet. The piston-like nature of these sinkholes is a func-
management in the city-owned forests above town. tion of the very brittle, low tensile-strength sandstone beds
The presence of cattails amongst the condominiums, on the that formed the roof of the mine. Walking across the area, it is
southern flank of Shanahan Hill, indicates that there is a shallow possible to detect steam and sulfur vapors and deposits. These
water table in the area. This is a result of groundwater “perching” are due to a smoldering coal fire under this site. The ditch on
within the Verdos Alluvium, on the top of the relatively imper- the hillside has undergone significant structural enhancement
meable Pierre Shale. Post-development lawn irrigation has added through this section to maintain its hydraulic integrity and
significantly to the overall, ambient level of perched groundwater proper grade.
in this area. Several of the condominiums now experience flood- The coal-mine fires and subsidence have imposed consid-
ing after multiple-day (monsoon) rainfalls, which raise the water erable land-use constraints on the Marshall area. We will see
table for days following the storms. examples of careful siting of residential buildings on a lot-by-
lot basis, based on carefully locating intact coal pillars and the
PART 2. BOULDER TO GOLDEN limits of mined areas. In Figure 7, note the location of the new
house (N) just outside of the subsidence area. The old, con-
Stop 5—Coal-Mine Fires and Subsidence, Marshall verted schoolhouse (S) has since been relocated to the west in
(32.7 miles; 52.6 km) order to increase its setback from the road. It, too, had to be
sited carefully to avoid workings of the Fox mine. Boulder
Directions: Follow Greenbriar Boulevard past the high County has attained much of the property in the area for use as
school (appropriately named Fairview, with its view of Boul- open space.
Bouncing boulders, rising rivers, and sneaky soils: Primer of geologic hazards and engineering geology 7
Stop 6 (Roll-by)—“Boulder’s Stonehenge,” a Water Tale of a grease fire in the kitchen and the unavailability of on-site
(34.5 miles; 55.5 km) water to fight the blaze. The position of the site at the edge of
the gravel cap had provided a stunning view; however, it had
Directions: Backtrack through Marshall, turn left onto Highway also doomed the project because no appreciable groundwater
93, and proceed toward Golden (south). could be located or recovered. The shales beneath the site are
Leaving Marshall, we will climb out of the valley of South too tight to yield water, and any perched water entering the
Boulder Creek. Notice the stone ruins at the crest of the hill, to granular alluvium from higher on the terrace drains away into
the right (west). These stone ruins are the remnants of a twice- the valley near the edge of the alluvial cap because of seepage
burned restaurant called the Matterhorn. The site is locally effects.
known as “Boulder’s Stonehenge”; Fig. 8). It is located atop a
mesa capped by Rocky Flats Alluvium, here underlain by a Stop 7 (Roll-by)—Mining Operations in Northern Jefferson
thick sequence of Pierre Shale. County (36.3-37.5 miles; 58.4-60.4 km)
Both times, the restaurant burned to the ground as the result
Directions: Continue along State Highway 93 toward Golden
(south).
As we cross the county line from Boulder County to Jefferson
County, you will notice an increase in mining and industrial activi-
ties (Fig. 9). On the right (west) side of State Highway 93 is a shale-
mining operation. The pit is several hundred feet deep and provides
R.
David Moffat began constructing a standard gauge line, the Den-
Fr Divide
ver, Northwestern & Pacific (later Denver & Salt Lake) Railway
Platte
.
Cr
as
der Company. The use of larger engines, rolling stock, and track
er
Boul
R
gauge (4-foot 8.5-inches wide) (about 1.43 m) required gentler
.
S. Boulder Cr.
WINTER
S.
Diversion Dam grades and larger-radius curves. Canyon routes, the traditional
PARK Gross choice of the narrow-gauge lines, were out of the question. This
DENVER
Reservoir Rocky
Moffat Flats presented a tantalizing problem for Moffat’s surveyors and engi-
Tunnel
Ralston
neers as they searched up and down the Front Range for favor-
Reservoir able routes and alignments.
Clear
C r. Moffat Moffat’s solution for surmounting the Rocky Mountain
N Treatment Front includes the Big Ten Curve (seen to your left, in Leyden
10 miles Plant
10 km Gulch) and the Tunnel District (seen ahead to your right). The
Big Ten is an innovative double-reverse curve that loops onto a
Figure 10. Schematic map of the features of the Denver water system in flat, Nussbaum-age (pre-Rocky Flats) alluvial terrace. Its name
the vicinity of Rocky Flats (Denver Water, 1994). The South Boulder
Diversion Conduit crosses Rocky Flats and links South Boulder Creek
arises from its tight, ten-degree radius of curvature. Note the sta-
and Ralston Reservoir. tionary hopper cars along the upper curve: they are filled with
ballast and serve as windbreaks, to keep trains from being blown
off the tracks during winter Chinook winds that sometimes
Environmental Technology Site in 1996. A contractor to the U.S. exceed 100 mph (161 kph). The Tunnel District consists of 27
Department of Energy (DOE) now manages this site. The site has tunnels in 13 miles, more than any other railway in the United
been the subject of numerous environmental assessments for seis- States over such a short distance. Eight of these tunnels penetrate
mic risk, site remediation, radioactivity hazards, and cleanup over “Flatirons” of steeply dipping Fountain Formation and Lyons
the years. The main facility, located to the left (east) of State High- Sandstone along the face of Eldorado Mountain (Fig. 11). Fur-
way 93, is readily visible along the field-trip route. Recently, ther to the west, the remaining tunnels penetrate Precambrian
DOE’s National Renewable Energy Laboratory has established Boulder Creek Granodiorite and quartz monzanite as the railroad
the National Wind Technology Center on Rocky Flats. It is located loops along the canyon walls high above South Boulder Creek.
about 1.5 miles (2.4 km) northwest of the Plant at the northwest The railroad then pierces the Continental Divide through the 6.2-
corner of the Rocky Flats DOE reservation. Sixteen experimental mile (10-km) long Moffat Tunnel.
wind turbines are currently in operation at this facility. Moffat’s transcontinental link was completed in 1934, several
A recent human activity at Rocky Flats has been capturing decades after his death. The D&SL railroad was absorbed by the
larger resistant stones from the Rocky Flats Alluvium for sale to Denver & Rio Grande Western Railroad, and later by the Southern
landscaping companies. Stockpiles of these usually can be seen Pacific and Union Pacific Railroads. For a closer look at this engi-
southwest of the intersection of State Highways 93 and 72. neering marvel, the reader is encouraged to ride the railroad’s
famous “Ski Train,” a one-day excursion that runs on certain sum-
Stop 9 (Roll-by)—The Moffat Road mer and winter weekends between Denver and Winter Park.
(39.7-42.2 miles; 63.9-67.9 km)
Stop 10—Water Flooding in the Mountain Canyons, Coal
Directions: Turn right onto State Highway 72, marked by a stop- Creek Canyon (42.6 miles; 68.6 km)
light, and head toward the mouth of Coal Creek Canyon (west).
Notice the railroad grade that climbs across the rugged face Directions: Continue driving westward on State Highway 72 and
of Eldorado Mountain, ahead to your right, as we turn westward enter Coal Creek Canyon. After passing underneath the railroad
onto State Highway 72 and continue across the Rocky Flats. This trestle, look for and enter the first large pullout on the left (south).
is “the Moffat Road,” one of Colorado’s most famous railroad [Note: the GSA bus will proceed further up the canyon to a safe
lines. We will have several opportunities along the trip route to turnaround, and then return to the pullout.]
look at the engineering geology of this railroad line in its diffi- All larger drainages along the Front Range head in the
cult transition between the Piedmont and the Rocky Mountains. mountains and fall rapidly from alpine elevations (above 9,500
First, a bit of history (from Bollinger and Bauer, 1962). Den- feet, or 2,900 m) through steep-gradient canyons. These debouch
ver, a mining town that aspired to be a great city, was in a difficult at the mountain front onto the piedmont, where stream gradients
spot at the turn of the century. It had been bypassed, both to the lessen substantially. There is typically 6,000 to 8,500 feet (1,830
north and south, by the transcontinental railroads. The numerous to 2,590 m) of relief from the highest mountain summits to the
railroad lines that radiated westward out of Denver were all nar- highest parts of the piedmont. The resulting orographic and
row gauge lines. These 3-foot-wide rail lines required small meterological effects on streamflow can be pronounced and
engines and small rolling stock to negotiate the narrow, sinuous extreme.
10 D. C. Noe et al.
At this locality near the mouth of Coal Creek Canyon we can sion on hill slopes, and deposition of prodigious amounts of sed-
see a likely paleosequence of flood deposits produced by events iment on gentler slopes above normally active stream channels
similar to the Big Thompson Flood. CGS (Soule and Costa, and on gentle to nearly level slopes near the river on the Pied-
unpublished data) obtained a C-14 date from this locality of 955 mont. An example of this compilation is shown in Figure 12.
+/- 80 ybp from what we interpreted to be the youngest of three Flood events of similar or possibly less magnitude have
flood deposits. Note the sizes of materials in the modern undoubtedly occurred in many Front Range drainages during the
streambed versus those in the bank cuts, and the accordance or Holocene. Soule, Costa, and Jarrett studied paleofloood deposits
discordance of surfaces immediately above the stream. CGS C- in fourteen drainages along the Front Range during the late sev-
14 dates from similar materials at other Front Range localities enties and early eighties. This was done to support a broad
range from about 10,000 ybp to 300 ybp. research proposal to model Front Range paleo- and modern
The greatest historical flood along the Colorado Front Range floods. The study was never completed because of fiscal “malnu-
occurred in the drainage of the Big Thompson River, Larimer trition.” Evidence for these occurrences include geomorphology
County, during the evening to nighttime of July 31-August 1, of paleoflood deposits, discharge estimates based on flows nec-
1976. This flood can be attributed to one thunderstorm event essary to move largest clasts entrained in paleoflood deposits,
when a large cell remained nearly stationary over the middle part depth of scour in streambeds, superposition of streamflows in
of the drainage basin for about 2.5 hours. This part of the basin channels, C-14 dating, study of demonstrable periglacial
extends from about 1 mile (1.6 km) east of Estes Park eastward to deposits, and meterological computations. Jarrett (1987) offered
the confluence of the Big Thompson River and North Fork Big field evidence for and proposed that the Big Thompson event has
Thompson River at Drake. Essentially all of the tributary streams a recurrence interval of 10,000 years, and that storms of the mag-
in this part of the basin were involved. Below Drake, relatively nitude to produce such an event must form below 8,500-foot alti-
little rain fell during the event. A peak discharge of the Big tude. Some debate about this continues, however.
Thompson River of about 39,000 cfs was computed at the moun-
tain front, about 4 miles west of Loveland. An estimated 139 per- Stop 11—Leyden Gulch Area (44.0-49.6 miles; 70.8-79.8 km)
sons were killed during this event and property loss of about $40
million resulted. Overbank flooding downstream from the moun- Directions: [Note: this section contains narratives for a roll-by
tain front occurred through the City of Loveland to the River’s followed by a stop. The roll-by portion begins after exiting Coal
confluence with the South Platte River. The Colorado Geological Creek Canyon.] Backtrack several miles along State Highway 72
Survey documented the geomorphic effects of this event (Soule, and turn right at the stoplight onto State Highway 93 (south).
et al., 1976; U.S.G.S., 1979). These included debris avalanches Proceed to the bottom of a long hill, along a distinctive, fin-like
and flows, deep scour of ephemeral streambeds, deep sheet ero- ridge of sandstone to your left. Turn left onto 82nd Avenue (east),
Bouncing boulders, rising rivers, and sneaky soils: Primer of geologic hazards and engineering geology 11
an unmarked gravel road that passes through the ridge’s water the mined-out stopes by looking northward along strike (Fig. 13).
gap. Proceed to the Public Service Company substation, which is In addition to the clay mines, several minor coal mines are located
enclosed by a chain-link fence. Park in front of the substation. in close proximity to these vertical beds (Fig. 14).
The valley below and to our right (south) as we backtrack onto Within 1/4 mile (0.4 km) to the east, these same beds go
the Rocky Flats surface is Leyden Gulch. It is a fairly prominent from vertical to horizontal at a depth of about 500-600 feet (152-
valley that parallels the pronounced break in slope at the base of 183 m). Below this portion of the valley and up the sides lie the
the foothills. The upper end of the gulch contains a steep channel Leyden Coal Mines (Fig. 14). These mines produced about 6 mil-
that is eroded to within 3/4 of a mile (1.2 km) of the main channel lion tons of coal in the first half of the 20th Century. Public Ser-
of Coal Creek. It has migrated between 200 and 250 feet (61 and vice Company of Colorado is currently exploiting the residual
76 m) northward in the last thirty years. Is this the next piracy void space as a natural gas storage facility. A system of charge
episode? What would happen if a flood like the ones just discussed wells and withdrawal wells access the mine. Several hundred feet
(in Stop 3 and Stop 11) were diverted into Leyden Gulch, through of saturated claystone and shale provide an incredibly effective
a densely populated area several miles east of here? seal to prevent leakage of the gas to the surface.
Descending into Leyden Gulch, we will travel along the base These tight claystone and shale beds also provide the seal
of a near-vertical exposure of lower Laramie Formation sandstone. that allows the landfill to the north to operate with no anticipated
Between these vertical sandstone ribs can be seen the collapsed groundwater impacts. Several older landfills have been operated
stopes of old, abandoned clay mines. After turning left (east) onto and reclaimed along the northern flanks of this valley, to the east
82nd Avenue, we will pass through the hogback in a water gap cut of our stop location.
by Leyden Creek. Some roof-crown arches can still be seen above All of these issues come into play as potentially harmful or
Cl
Stop 12 (Roll-by)—Tertiary Igneous Rocks near Golden Barb
ara
Le Cl
(52.1-54.3 miles; 83.9-87.4 km) y
d
en
Cl
Lf Leyden
Directions: Backtrack to State Highway 93 and turn left (south).
G
Cl
Mines
ul
ch
Resume driving toward Golden. Cl
Approaching the Golden area, we will pass several intrusive Co Cl
Cl
bodies and dikes that are located in the Pierre Shale and lava flows Gulch
Co Leyden
that are interbedded with the Tertiary part of the Denver Formation. Co
Cl
The rocks are mafic monzonite and its extrusive equivalent, alter- G G G
93
natively called mafic latite by Van Horn (1976) and shoshonite, a 1 mile
1 km
potassium-rich basalt, by Scott (1972) and Trimble and Machette G
(1979). A majority of the intrusives bodies in this area are west of,
Figure 14. Map showing the extent of coal (Co), clay (Cl), and gravel
or associated with, the Golden Fault (Van Horn, 1976). (G) mining and landfills (L) in Leyden Gulch (modified from Amuedo
The Ralston Dike is the largest of the intrusives and is and Ivey, 1978). Vertical clay-mine stopes in the area are up to 125 feet
located approximately 1/2 mile (0.8 km) west of State Highway (38 m) deep, and are still undergoing roof collapse. The extensive, deep
93, to your left. The dike is about 7,500 feet (2,290 m) long and Leyden Coal Mines are currently used for natural gas storage.
2,000 feet (610 m) wide. Its hollowed-out center is used as a
water-storage reservoir. The western side of the dike is about 400 240 feet (73 m) in total thickness. These lava-capped mesas are
feet (122 m) above the reservoir and the eastern side is only about another prime example of topographic reversal. The flows occu-
50 feet (15 m) above the reservoir. The Ralston Dike is being pied valleys when they were deposited, but they have resisted
mined for a high-quality aggregate which is used in the manufac- erosion and are now the highest ground.
ture of “Super Pave,” a durable pavement that is used in the con-
struction of airport runways and interstate highways. The Ralston PART 3. GOLDEN AREA
Quarry produces about 800,000 tons of aggregate per year.
South of the Ralston Dike and east of State Highway 93 are Stop 13—Lunch (57.9 miles; 92.2 km)
North and South Table Mountains, ahead and to the left (south-
east). Capping the Table Mountains, and interbedded with the Directions: After entering Golden on State Highway 93, turn left
Denver Formation, are lavas which flowed from vents, such as onto Washington Avenue and drive southeast. Parfet Park is
the Ralston Dike, about 63 to 64 million years ago. The shoshon- located on the left after about a mile (1.6 km), along Clear Creek.
ite flows can be separated into three separate flows that are about There are other parks nearby, along both banks of the creek.
Bouncing boulders, rising rivers, and sneaky soils: Primer of geologic hazards and engineering geology 13
We will stop in or near to the city of Golden for a box lunch. Stop 15—North Table Mountain Rockfall Area, Golden
Golden, Colorado’s territorial capitol from 1862-1867, is located (59.9 miles; 96.4 km)
along Clear Creek. North Table Mountain, South Table Mountain
(the profile of which is immortalized on labels of Coors beer, Directions: Drive north on State Highway 93 to the first stoplight
from the famous resident brewery), Lookout Mountain, and and turn right onto Iowa Street (northeast). Follow Iowa Street
Mount Zion (with the white “M” announcing Colorado School across Washington Avenue and turn right onto Ford Street (south-
of Mines) surround the city and provide a scenic backdrop. east). Pull into the church parking lot to the left, at the corner of
Ford Street and First Street.
Stop 14—Highway 6-58-93 Junction, Golden (59.0 miles; 95.0 km) We now re-cross the northern part of Golden to get a closer
view of the west-facing slope of North Table Mountain (Fig. 15).
Directions: Backtrack part way along Washington Avenue (north- The rockfall and landslide hazard area around North Table Moun-
west), turn left onto State Highway 58 (west), then right onto tain was mapped by the U.S. Geological Survey during the 1970s
State Highway 93 (north). Pull off where the shoulder is suffi- (Simpson, 1973a; 1973b). Jefferson County has adopted the
ciently wide. mapped hazard area as part of its geologic-hazards overlay, and
After circling through the northern part of Golden, we will has considered it to be a “no-build” area. This was challenged in
park and climb a small hillock near the mouth of Clear Creek, to the 1980s by a developer, who staged an actual rock-rolling
the northeast of the intersection of U.S. Highway 6 and State demonstration on the northwest flank of the mountain in order to
Highways 58 and 93 converge. Looking to the north and south, prove the fallacy of the outer (distal) hazard-area boundaries. The
we can see the distinctive hogback of the Dakota Sandstone at demonstration was curtailed (and the subdivision application was
both ends of the valley. But why is this hogback missing at the subsequently denied) after several boulders rolled beyond the
valley center? The answer is that the Golden Fault, a large, outer boundary, but not before one boulder had bounced at least
Laramide thrust fault, weaves along the valley just to the east of 10 feet (3 m) into the air and knocked a cross-bar off a high-ten-
the mountain front. Here, the fault has displaced nearly 8,000 feet sion power-line tower at the base of the mountain.
of the Paleozoic and Mesozoic, sedimentary section (Van Horn, In Figure 16, we see that the rockfall hazard area, as mapped
1976). The trace of the fault can be seen as it cuts up the side of by the USGS, exists in both unincorporated Jefferson County and
the first large alluvial terrace to the south of Clear Creek. There is the incorporated city of Golden. Unfortunately, protection of the
a distinct break in the vegetation between the Fountain Forma- public from the rockfall hazard does appear to stop at the juris-
tion to the west (covered with mountain mahogany shrubs) and dictional boundary in this case. The houses we see on the hillside
the Pierre Shale to the east (covered with grasses). above are located within the city of Golden, which has not
Looking directly west, we see a road cut for State Highway adopted Simpson’s map. This subdivision was approved and has
93 (Fig. 15). This portion of the highway was first constructed in expanded due the absence of any requirements for home-rule
1991. A small landslide formed in the west (opposite) face of the cities to follow the state-mandated, geologic-hazard review
cut shortly thereafter, and the resulting toe bulge closed the south- process for subdivisions.
bound lanes of the new highway. Within a year, the landslide had
developed a rim of head scarps with up to 12 feet of vertical slip- Stop 16—Clay Pits and Differential Subsidence, Golden
page, and had captured the surface flow of a small stream in (61.7 miles; 99.3 km)
Magpie Gulch. Early efforts to drain the landslide were unsuc-
cessful. A full geologic investigation subsequently revealed that Directions: Continue southeast on Ford Street, passing the Coors
the landslide sits directly atop the Golden Fault, with low-perme- Brewery on your left (note: mileage for side trip to Coors is not
ability Pierre Shale on the eastern side and a wedge of fractured, included). Shortly after the street splits into a one-way and
permeable Fountain Formation on the western side. In 1994, becomes Jackson Street, turn right onto 19th Street. After passing
three lines of rock anchors (about 40 anchors in total) were two stoplights, turn right onto the Colorado School of Mines
installed across the landslide. The Magpie Gulch stream was campus at Elm Street, then immediately to the left onto Campus
piped across the head scarps, longer horizontal drains were Road. Drive past the fraternity houses to the married-student
installed, and a remote data-logging unit was set up. To date, this housing area.
mitigative effort appears to have been successful. The combined At this stop, we will look at the challenges of multiple-
maintenance and mitigation operations for this incident are sequential land use as related to clay mining operations and
reported to have cost about 3 million dollars. reclamation along the western side of Golden. Here, an array of
The houses to the north of the landslide were built shortly open-stope clay pits along the east side of U.S. Highway 6 have
after the landslide was mitigated and all signs of its existence had been reclaimed for several uses with a variety of problems and
been “erased” from view. One can only imagine the concern of solutions, some more successful than others.
these residents had the landslide been active a few years later. Severe differential-settlement problems have been experi-
This is a good illustration of the often short-term public memory enced in the student housing at Colorado School of Mines, along
of geologic hazards. Campus Road. The fill used to reclaim this area has settled, per-
14 D. C. Noe et al.
RF
haps only a few percent, but the bounding sandstone ribs are sta-
High-Tension
ble. Major settlement and damage has occurred where structures, Power Line H
flatwork, and roadways straddled these highly variable units.
South of the CSM campus, the stopes were used to effect N
T
Golden IN
(68.3 miles; 110.0 km) A
T
TH
N
M U
R
19th Street and turn right (southwest). At the stoplight, turn left
Co
onto 6th Avenue (U.S. Highway 6). Proceed about 4 miles (6.4
rp
km) and exit onto Indiana Street, turning right (south). Indiana
or
at
H
e
correlative with the nearby volcanic and intrusive rocks of the appear to be an earthflow or landslide deposit that originally
Table Mountains and the Ralston Dike discussed earlier on this formed as a bedrock failure.
trip. This conglomerate overlies the sandstones and shales of the When Jefferson County approved the subdivision, its
Denver Formation. Landslides of various ages (including active, approval was based on a geotechnical report submitted by the
modern ones) and colluvium derived from the underlying developer. This report indicated that the subdivision was accept-
bedrock or other surficial deposits underlie the side slopes of able from a geological and geotechnical standpoint. The geotech-
Green Mountain. nical engineer recommended several mitigation measures, such
Residential development on Green Mountain began in the as the installation of an underdrain below the sanitary sewer. The
1960s and continues to the present. Much of the remaining, unde- county imposed additional restriction on the development. For
veloped land is now in dedicated public open space. Residential example, the county required that site-specific slope stability
development here has been adversely affected by moderate-to- analyses be done for each proposed home site. Many of the rec-
severe drainage problems, erosion, natural-slope and fill instabil- ommendations made by the geotechnical engineer and required
ity, and soils-engineering problems. In the early years, ill-advised by the county were not followed as the area was developed. Later,
regrading and fill placements, mostly by one developer on the geotechnical studies performed subsequent to the renewed slope
east slopes of the Mountain, changed the area’s natural drainage movement found that the conditions modeled in the original
to the extent that serious local flooding and fill-failure problems report did not reflect the deep clay soils located in the ancient
occurred. Many of these conditions still exist. These were cause channel or failure surface.
for serious concerns by the City of Lakewood and Jefferson Three lines of earth anchors and drainage are currently being
County. constructed to stop or slow movement of the landslide. The
Residential development on Green Mountain continues anchors will be approximately 110 feet long (34 m) and extend
today with serious geology-related and geotechnical problems. In 40 feet (12 m) into the claystone bedrock. After installation of the
particular, younger landslides have formed in some of the older anchors, horizontal drains will be installed. The drains will be
landslides on Green Mountain, demonstrating that the older land- approximately 200 ft long (60 m) and will be placed on three lev-
slides are not necessarily stable. In 1990, Jefferson County els along with the anchors.
allowed development on an area of Green Mountain that had
been previously mapped as landslide (earthflow) deposits. In Stop 18 (Roll-by)—Major Flood-Control Structures in the
1998, renewed movement of an older landslide occurred. This Denver Metropolitan Area (77.3-78.3 miles; 124.4-126.1 km)
resulted in three homes being damaged beyond repair (Fig. 17)
and two other homes to be severely damaged. The landslide Directions: Backtrack to 6th Avenue (U.S. Highway 6) along
appears to fill a buried ancient channel or ancient landslide-fail- Archer Avenue, 1st Avenue, and Ellsworth Avenue. Turn left onto
ure surface that underlies West Bayaud and West Archer Avenues. the 6th Avenue freeway (west), then exit right onto Interstate 70
The soils that underlie the area of renewed slope movement con- (west) and State Highway C-470 (south). Bear Creek Reservoir
sist mainly of clay and sparse gravel. Because of the lack of lies to the east of C-470 near the Morrison exit. Continue south
gravel, the soils appear to be associated with displaced, greatly on C-470 and do not exit.
weathered claystone bedrock. The clay soils that are now moving The master stream in the Denver Metropolitan Area is the
16 D. C. Noe et al.
South Platte River. Its four major tributaries are Clear Creek, Street and turn right (south). Proceed for about 2 miles (3.2 km)
Bear Creek, Plum Creek, and Cherry Creek. The headwaters of and turn left onto Coal Mine Avenue (east). At the bottom of the
Clear Creek and Bear Creek are located in the mountains. The hill, turn left onto Owens Street (north). The second damage-
headwaters of Plum Creek and Cherry Creek are located in the area tour consists of taking a right onto Oak Court, left onto
Palmer Divide, south of Denver. All of these streams are subject Parfet Street, left onto Walker Drive, and right onto Owens
to the flooding scenarios discussed in a previous section (see Stop Street again (south).
10 narrative). Travelling southward on State Highway C-470, we This field trip mini-tour is designed to show the effects of
are afforded views of major flood-control structures that have differentially heaving bedrock on past, present and future devel-
been emplaced on these major streams. opment areas of the Upper Cretaceous Pierre Shale outcrop belt.
The South Platte River and Plum Creek (Chatfield Dam On Quincy Avenue, notice the long, parallel “speed bumps” that
and Lake), Cherry Creek (Cherry Creek Dam and Lake), and have severely deformed the pavement is several areas. After an
Bear Creek (Mount Carbon Dam and Bear Creek Lake) have introductory stop in the Harriman Park neighborhood, we will
flood-control works built on them as part of the U.S. Army take a driving tour to view examples of older (mid-1970s) and
Corps of Engineers’ Tri-Lakes Project. These impoundments newer (mid-1990s) damage. If the opportunity exists, we will
were completed in 1975, 1950, and 1982, respectively. Prior to look at a current development site where overexcavation and fill-
this, a small flood-control dam (Kenwood Dam) had been built replacement technologies are being used.
on Cherry Creek. It was subsequently determined to be grossly Differential ground heaving has adversely affected develop-
undersized and was removed. The present dams are all of the ment projects for over 25 years in southern Jefferson County, in
rolled-earth-fill type. Their reservoirs are used for recreation, an area that is underlain by the Pierre Shale and other Upper Cre-
and are surrounded by parklands. The Colorado Division of taceous formations. This geologic hazard is manifested by the
Parks and Outdoor Recreation manages Chatfield and Cherry progressive growth of long, somewhat parallel ridges separated
Creek State Parks, and the City of Lakewood manages Bear by relatively inert swales. Over 2 feet (0.6 m) of differential
Creek Lake Park. movement has occurred in some cases within a few years follow-
The earliest recorded major stream flooding in Denver ing development. Certain neighborhoods in the Pierre Shale out-
occurred on Cherry Creek on May 20, 1864, and the latest crop belt have performed well over the years, while others have
occurred in May 1973, on the South Platte River and other Pied- sustained tremendous amounts of damage to structures, roads and
mont streams. Flows of the South Platte of 55,000 cfs through utilities. One 110-home neighborhood has reportedly incurred
Denver were documented in June 1965, its greatest historic flood. more than 5 million dollars worth of damage and mitigation costs
The 1973 flood produced 18,000 cfs flows of the South Platte. after 15 years, and the ground is continuing to heave.
Floods have impacted numerous local communities over the Until about 1990, these problems were often attributed to the
years. Many of the more historic ones are documented in Fol- geological hazard of swelling soil. However, commonly used
lansbee and Sawyer (1948). One of the greatest floods occurred mitigative designs for swelling soil, such as drilled-pier founda-
in 1921 on the Arkansas River, which caused much damage in tions, floating-slab floors, and structural floors have been remark-
Pueblo and its immediate vicinity. An earlier large flood occurred ably unsuccessful to date. Subsequent research by the U.S.
in 1905 on Boulder Creek. Since these flood-control works were Geological Survey and the Colorado Geological Survey (Nichols,
installed, the Denver Metropolitan Area has mostly escaped seri- 1992; Noe and Dodson, 1997) indicates that the differential
ous damage caused by major flooding of their respective streams. movements occurs within near-vertical claystone bedrock
However, some streams such as Boulder Creek, Clear Creek, and beneath the ground, resulting from the combined effects of wet-
the upper reaches of Bear Creek, may be capable of producing ting and unloading surface (Fig. 18). The hazard has been called
flooding in populated areas. “heaving bedrock,” to alert engineers to a need to depart from
standard “swelling soil” considerations.
Stop 19—Heaving Bedrock in Southern Jefferson County As a result of CGS education efforts, Jefferson County
(80.9-86.1 miles; 130.2-138.6 km) convened a task force of nearly 75 stakeholders in 1994 to inno-
vate minimum standards for site exploration, evaluation, and
Directions: [Note: this section contains narratives for a com- design to mitigate heaving bedrock. New land-development
bined stop and roll-by tour. The GSA field trip will visit two regulations were enacted for the Dipping Bedrock Area in 1995
areas of historical damage; additional areas may be visited as (see Noe, 1997). Trenching has been implemented to evaluate
well.] Continue along State Highway C-470 past the U.S. High- the geometric complexities of the dipping bedrock. Overexca-
way 285 exit, then exit right (twice) onto Quincy Avenue (north, vation and fill replacement to at least 10 feet (3 m) beneath
then east). Descend a long hill and turn right onto Simms Street foundations are now specified as minimum site-construction
(south) at the stoplight, then right onto Marlowe Avenue. The standards if the geologic evaluation finds that the bedrock has a
first damage-area tour consists of taking a right onto Union potential for differential heave. The new Jefferson County regu-
Street, left onto Urban Way, and left onto Tanforan Avenue back lations have resulted in a significant reduction in damage to
to Marlowe Avenue (east). Take Marlowe Avenue back to Simms homes and infrastructure.
Bouncing boulders, rising rivers, and sneaky soils: Primer of geologic hazards and engineering geology 17
Noe, D.C., 1997, Heaving-bedrock hazards, mitigation, and land-use policy, Front Simpson, H.E., and Hart, S.S., 1980, Preliminary engineering geology map of the
Range Piedmont, Colorado: Environmental Geosciences, v. 4, no. 2, p. 48- Morrison quadrangle, Colorado: U.S. Geological Survey Open File Report
57 (reprinted as Colorado Geological Survey Special Publication 45, 1997). 80-654, 104 p.
Noe, D.C., and Dodson, M.D., 1995, The Dipping Bedrock Overlay District - an Soule, J.M., 1978, Geologic hazards study of Douglas County, Colorado: Col-
area of potential heaving bedrock hazards associated with expansive, orado Geological Survey Open-File Report 78-5, 16 plates.
steeply dipping bedrock in Douglas County, Colorado: Colorado Geological Soule, J.M., Rogers, W.P., and Shelton, D.C., 1976, Geologic hazards, geomor-
Survey Open-File Report 95-5, 32 p. phic features, and land-use implications in the area of the 1976 Big Thomp-
Noe, D.C., and Dodson, M.D., 1997, Heaving-bedrock hazards associated with son flood, Larimer County, Colorado: Colorado Geological Survey
expansive, steeply dipping bedrock in Douglas County, Colorado: Colorado Environmental Geology 10.
Geological Survey Special Publication 42, 80 p. Soule, J.M., 1999, Active surficial-geologic processes and related geologic haz-
NSA Engineering, Inc., 1997, Subsidence investigation, mitigation, and estimated ards in Georgetown, Clear Creek County, Colorado: Colorado Geological
cost/annual cost/risk evaluation for West Coal Mine Avenue extension: con- Survey Open-File Report 99-13, 6 p.
tract report for Jefferson County, 8 p. Trimble, D.E., and Machette, M.N., 1979, Geologic map of the greater Denver
Scott, G.R., 1960, Subdivision of the Quaternary alluvium east of the Front Range area, Front Range Urban Corridor, Colorado: U.S. Geological Survey Mis-
near Denver, Colorado: Geological Society of America Bulletin, Vol. 71, p. cellaneous Information Series, Map I-856-H.
1541-1543. U.S. Geological Survey, 1979, Storm and flood of July 31-August 1, 1976, in the
Scott, G.R., 1972, Geologic map of the Morrison Quadrangle, Jefferson County, Big Thompson River and Cache la Poudre River Basins, Larimer and Weld
Colorado: U.S. Geological Survey Map I-790-A. Counties, Colorado: U.S. Geological Survey Professional Paper 1115, 152 p.
Shroba, R.R. and Carrara, P.E., 1996, Surficial geologic map of the Rocky Flats Van Horn, R., 1972, Surficial and bedrock geology map of the Golden quadran-
Environmental Technology Site and vicinity, Jefferson and Boulder Counties, gle, Jefferson County, Colorado: U.S. Geological Survey Map I-761-A.
Colorado: U.S. Geological Survey Map I-2526. Van Horn, R., 1976, Geology of the Golden Quadrangle, Colorado: U.S. Geolog-
Simpson, H.E., 1973a, Map showing landslides in the Golden quadrangle, Jeffer- ical Survey Professional Paper 872, 116 p.
son County, Colorado: U.S. Geological Survey Miscellaneous Information Weimer, R.J., 1977, Stratigraphy and tectonics of western coals, in Murray, D.K.,
Series, Map I-761-B. ed., Proceedings of the symposium on the geology of Rocky Mountain
Simpson, H.E., 1973b, Map showing areas of potential rockfalls in the Golden Coal, 1976: Colorado Geological Survey Resource Series 1, p. 9-27.
quadrangle, Jefferson County, Colorado: U.S. Geological Survey Miscella- Wells, J.D., 1967, Geology of the Eldorado Springs quadrangle, Boulder and Jef-
neous Information Series, Map I-761-C. ferson Counties, Colorado: U.S. Geological Survey Bulletin 1221-D, 85 p.
Printed in U.S.A.
Geological Society of America
Field Guide 1
1999
Eric A. Erslev
Department of Earth Resources, Colorado State University, Fort Collins, Colorado 80523, United States
Karl S. Kellogg and Bruce Bryant
U.S. Geological Survey, MS 913, Denver Federal Center, Denver, Colorado 80225, United States
Timothy K. Ehrlich and Steven M. Holdaway*
Department of Earth Resources, Colorado State University, Fort Collins, Colorado 80523, United States
Charles W. Naeser
U.S. Geological Survey, MS 926A, Reston, Virginia 20192, United States
FIELD TRIP OVERVIEW Range of Colorado. The Front Range starts north of Canon City,
Colorado, and trends north-northwest to Golden, Colorado.
The field trip will traverse the highly asymmetrical Front North of Golden, the range takes a more northerly trend toward
Range to examine the west flank’s major Laramide thrusts, subse- the Wyoming border where it bifurcates into the north-trending
quent volcanic rocks and normal faults, as well as the east flank’s Laramie Range (Brewer and others, 1982) and the north-north-
higher-angle thrust and reverse faults and their associated fault- west-trending Medicine Bow Range.
propagation folds. Laramide to Holocene tectonics will be debated Specifically, the field trip will visit localities (Figs. 1, 2)
on the outcrop and during our evening soak at Hot Sulphur Springs. whose exposures address the following problems.
1) What were the styles of Laramide basement-involved
INTRODUCTION deformation?
2) Why are the Laramide structures of the eastern, northeast-
The Rocky Mountain province of the United States is a clas- ern, and western margins of the Front Range so dissimilar?
sic basement-involved foreland orogen. Deformation during the 3) What was the tectonic regime during mid-Tertiary igneous
Late Cretaceous to Eocene Laramide orogeny created an anasto- activity?
mosing system of basement-cored arches that bound the northern 4) What is the relationship between Laramide and Tertiary
and eastern margins of the Colorado Plateau and the elliptical deformation and the regional uplift of the southern Rockies and
sedimentary basins of the Rockies. The tectonic mechanism for adjoining High Plains?
Laramide deformation remains controversial, with proposed
mechanisms ranging from subcrustal shear during low-angle sub- REGIONAL SETTING
duction (Bird, 1988, 1998; Hamilton, 1988) to detachment of the
upper crust during plate collision to the west (Oldow and others, Laramide to Holocene deformation in the northern Front
1990; Erslev, 1993). The Rocky Mountains south of Wyoming Range was imposed on a region with a long and complex struc-
have the additional complication of a period of mid-Tertiary tural history (Figs. 2 and 3). The oldest rocks in north-central Col-
igneous activity and sedimentation that coincides with Neogene orado are supracrustal and plutonic rocks resulting from Early
extension along the Rio Grande rift. Proterozoic arc magmatism and related sedimentation. These
This field trip (Fig. 1) will explore the Laramide to Holocene rocks were accreted to the southern edge of the Archean Wyoming
structural development of the southern Rocky Mountains by craton over an interval of 130 m.y., beginning at about 1,790 Ma
examining the geologic record exposed in the northern Front (Reed and others, 1987). Basement rocks include complexly
folded and interleaved quartzo-feldspathic gneiss, amphibolite,
*Present address: Chevron, 1013 Cheyenne Drive, Evanston, Wyoming biotite schist, gneiss, and migmatite, commonly metamorphosed
82930, United States. to high T, low P upper-amphibolite facies assemblages. The
Erslev, E. A., Kellogg, K. S., Bryant, B., Ehrlich, T. K., Holdaway, S. M., and Naeser, C. W., 1999, Laramide to Holocene structural development of the
northern Colorado Front Range, in Lageson, D. R., Lester, A. P., and Trudgill, B. D., eds., Colorado and Adjacent Areas: Boulder, Colorado, Geological
Society of America Field Guide 1.
21
22 E. A. Erslev et al.
106 W WYOMING
COLORADO
F
287
R
PA
O
NO R T H
RK
Ca c he
N T
Wa l d e n
14 l a Po ud re
STOP
14 Ca ny o n B6
125 STOP B5
N A T ' L
Figure 1. Road map for field trip
NG
40 P A R K 25
showing major topographic features STOP
R A N G E
STOP B1 B2
and stop locations.
MIDDLE
E
Granby
Hot Sulfur
Kremmling
Springs Boulder
PA R K 40 N
STOP A5
9
Bl
G
u
40
RA
R
O
R U t e Pa s s
iv
N
G E
er
STOP A4
Golden
Va
STOP
E A2
70
DENVER
lle
70
y
Va i l Loveland STOP A1
STOP 6 Pass
A3 Keystone
Dillon
Fr i s c o
285
SAWATCH 25
biotite-rich rocks locally contain layers and lenses of marble, and the structural patterns around the major plutons led Reed and
quartzite, and conglomerate, indicating sedimentary protoliths. others (1987) to suggest that much of the deformation and meta-
Compositions and textures from less deformed and metamor- morphism was related to emplacement of the plutons rather than
phosed amphibolites and some quartzo-feldspathic gneisses in a subsequent tectonic and metamorphic events.
few areas suggest metavolcanic protoliths. In the Big Thompson Widespread mid-crustal reheating at 1.4 Ga (Shaw and oth-
and nearby Cache La Poudre River drainages on the east side of ers, 1999) was accompanied by re-equilibration at pressures sim-
the Front Range (Fig. 1), initial metamorphic pressures of 8-10 kb ilar to those of the main ~1.7 Ga mineral assemblage
were interpreted to be due to subduction-related metamorphism (Selverstone and others, 1997) as well as intrusions of dikes,
between 1758 Ma and 1726 Ma (Selverstone and others, 1997). stocks, and discordant plutons of weakly foliated to non-foliated
The supracrustal rocks are cut by calc-alkaline plutons, two-mica granite, hornblende dacite, granodiorite and gabbro.
chiefly granodiorite and monzogranite, but ranging from peri- This Berthoud Plutonic Suite (Tweto, 1987) is part of a continent-
dotite to granite. These plutons are part of the Routt Plutonic scale igneous event at about 1.4 Ga (Anderson, 1983). In the
Suite and range in age from 1,780 to 1,650 Ma (Tweto, 1987; southern Front Range, these rocks were cut by the 1,092-1,074
Reed and others, 1987; Reed and others, 1993). They are com- Ma (Unruh and others, 1995) Pikes Peak batholith which
monly foliated and concordant, suggesting syntectonic, synmeta- intruded at depths as shallow as 5 km (Barker and others, 1976).
morphic intrusion. Unfoliated, discordant plutons in some areas During the Proterozoic, the region was cut by an anastomos-
are older than foliated, concordant plutons in other areas, show- ing swarm of northeast-trending zones of recurrent ductile defor-
ing that deformation and metamorphism were not synchronous mation and cataclasis. These shears were zones of weakness that
throughout the region. The correspondence of intrusive and meta- locally controlled intrusive activity and mineralization during the
morphic ages, the high temperature-low pressure metamorphism, Late Cretaceous and Tertiary. For example, a swarm of shear
Laramide to Holocene structural development of the northern Colorado Front Range 23
zones through the central Front Range is parallel to the concen- sedimentary cover, but these strata are preserved in adjacent basins
tration of mostly Late Cretaceous to Eocene intrusions along the in which as much as 5 km of Pennsylvanian synorogenic strata
Colorado Mineral Belt (Tweto and Sims, 1963). Northwest- and accumulated (Fig. 3). The later Laramide arches partly coincide
north-northeast-trending fault zones of late Proterozoic age cut with the Paleozoic uplifts, but the Laramide arches have a more
the ductile shear zones and also were reactivated during Phanero- northerly trend. Some of the faults active in the late Paleozoic
zoic deformation (Tweto and Sims, 1963). were reactivated during Laramide and Neogene deformation.
During the early Paleozoic, thin platformal deposits of
quartz-rich sands and carbonate rocks covered the region. In the CENOZOIC TECTONIC PROBLEMS AND HYPOTHESES
late Paleozoic, northwest-trending mountain ranges and basins
formed during the basement-involved Ancestral Rocky Mountain 1) What were the styles of basement deformation during the
orogeny. In the area traversed by the field trip, erosion during Laramide?
Ancestral Rocky Mountain uplift removed the earlier Paleozoic The multitude of Laramide structural geometries has resulted
in a mirroring multitude of kinematic hypotheses. In the 1970s drilling at Laramide basin margins (Gries, 1983), and balancing
and 1980s, geoscientists were polarized into opposing horizontal- constraints (Stone, 1984; Erslev, 1986; Erslev and Rogers, 1993;
compression and vertical-tectonics schools. The vertical-tectonics Brown, 1988; Spang and others, 1985) have demonstrated that the
school was dominant in the 1970s, represented by upthrust Laramide was the result of lateral shortening due to horizontal
(Prucha and others, 1965) and block uplift models (Stearns, 1971; compression. High-angle, dip-slip faults do occur in numerous
1978; Matthews and Work, 1978). In the 1980s, incontrovertible locations in the Laramide foreland, however, indicating distinct
evidence for thrusting of Precambrian basement over Phanerozoic differences between the basement-involved Laramide and syn-
sediments (e.g., Smithson and others, 1979; Gries, 1983; Lowell, chronous thin-skinned Sevier (Schmidt and Perry, 1988) orogens
1983; Stone, 1985) swung opinion back towards models invoking of the Rocky Mountain foreland.
horizontal shortening and compression. Seismic profiles (Smith- The diversity of Laramide structural trends, with faults, folds,
son and others, 1979; Gries and Dyer, 1985), subthrust petroleum and arches trending in nearly every direction, has been attributed
Green Mountain
TERTIARY
200 m
Conglomerate
Fox Hills Sandstone 55 m Olive green to brown silty shale and sandstone; white and
CRETACEOUS
Carlisle Shale, Sandy limestone, gray silty sandstone, yellowish gray calcareous shale;
160 m gray limestone and calcareous shale; dark gray clayey shale; dark
Greenhorn L.,
Graneros Shale gray platy siltstone at base
L.
Yellowish gray sandstone, dark gray shale, yellowish brown
Cret Dakota Group 90 m
sandstone; conglomerate at base
90 m Red and green siltstone and claystone; some beds of
Morrison Fm. brown sandstone and gray limestone
Jurassic
Ralston Creek Fm. 27 m Purplish graysandstone and siltstone; yellowish gray silty sandstone
T -Perm Lykins Fm. 40 m Grayish red shale, sandy limestone, and grayish red and green siltsrone
Perm Lyons Sandstone 60 m Yellowish gray sandstone and conglomerate
Perm
and Fountain Fm.
500 m Grayish red calcareous sandstone and conglomerate
Penn
Figure 3 (this and opposite page). Stratigraphic columns for (a) northeastern flank of the Front Range at stop A1, after Scott (1972),
and (b) Middle Park and North Park (composite), after Izett (1968), Izett and Barclay (1973), and Tweto (1976).
Laramide to Holocene structural development of the northern Colorado Front Range 25
to multiple stages of differently oriented shortening and compres- These hypotheses may all be valid for individual areas within the
sion (Gries, 1983; Chapin and Cather, 1981; Bergh and Snoke, foreland but their regional significance is not clear.
1992), reactivation of pre-existing weaknesses in the basement
(Hansen, 1986; Blackstone, 1991; Chase and others, 1993; Stone, 2) Why are the Laramide structures of the eastern, northeast-
1986, 1995), transpressive motions (Wise, 1963; Sales, 1968), ern, and western margins of the Front Range so dissimilar?
rotation and indentation by the Colorado Plateau (Hamilton, Laramide structures of the Front Range basement arch are
1988), and detachment of the crust (Lowell, 1983; Brown, 1988; extremely diverse. This field trip begins and ends on the eastern
Kulik and Schmidt, 1988; Oldow and others, 1990; Erslev, 1993). flank of the range, where the ENE-directed Golden thrust system
(North Park)
Pa
Miocene
and
.
r th
Fm
Trouble- 450-
Troublesome Formation (Middle Park) - Siltstone, tuff,
No
some 500m
om
(Middle
o
Park)
Tr
Coalmont
Format
tion
ont
Forma
Upper Member
Coalm
2135m
Park
Formation
Middle Park
Windy Gap
Member
le
Volcanic
Midd
Pierre
Shale 1615m
Dark-gray marine shale and a few beds of fine-grained sandstone
west and southwest of Denver (Berg, 1962) changes northward data, 1998). Mid-Tertiary strike-slip deformation in the southern
into an array of WSW-directed, higher-angle reverse faults (Erslev, Rocky Mountains may have paralleled the current Rio Grande
1993). While earlier interpretations of vertical uplift and gravity rift, with transtensional tectonics forming mid-Tertiary pull-apart
sliding provided a temptingly integrated hypothesis (Fig. 4a; Boos basins and facilitating magmatic intrusion into the crust.
and Boos, 1957; Matthews and Work, 1978; Tweto, 1975,1979),
the faults are not vertical and horizontal shortening is clearly indi- 4) What is the relationship between Laramide and Tertiary
cated (Berg, 1962; Erslev, 1993; Erslev and Selvig, 1997; Weimer deformation and the regional uplift of the southern Rockies and
and Ray, 1997; Holdaway, 1998). Symmetric, concave-downward adjoining high plains?
upthrusts on the arch margins were proposed by Jacobs (1983; Fig. Central Colorado is part of a 1,000-kilometer-long topo-
4b), but major thrust overhangs have never been documented on graphic high named the Alvarado Ridge by Eaton (1986). It paral-
the northeastern side of the Front Range. Hypotheses of strike-slip lels the Rio Grande rift, which is closely related structurally and
deformation have also been suggested (Fig. 4b), but recent analyses temporally to regional extension in the Basin and Range Province
of minor faults generally indicate shortening perpendicular to the to the west. The high elevations can be attributed to numerous
strike of the major faults (Selvig, 1994; Holdaway, 1998). mechanisms, including thickening of the crust during Laramide
On the western side of the range, however, low-angle faults shortening and/or thinning of the mantle lithosphere during low-
displace Precambrian basement considerable distances over sed-
imentary strata. Large basement overhangs are documented in
several areas where post-Laramide stocks bow up hanging-wall
basement overhangs and expose underlying sedimentary strata.
The vertical tectonic school explained these fault windows as
landslides, which are common in areas where basement rocks
overlie Cretaceous shale. But the magnitude of the overhangs,
their consistency along strike, and associated shortening struc-
tures indicate that these faults are thrusts rooted in basement.
Asymmetrical “chip models” (Fig. 4c,d, Kluth and Nelson,
1988; Raynolds, 1997; Erslev, 1993) can explain the differences
between the major fault geometries on either side of the Front Range
arch. To date, however, the details of the structural transition between
east-directed thrusting on the southeastern margin of the arch and
west-directed thrusting on the northeastern margin of the arch have
not been fully addressed. In addition, the effects of changes in
Laramide shortening directions through time are not clear.
Figure 5. Seismically constrained (DOE, 1993) structure section through the eastern flank of the Front Range at the Rocky Flats Plant 20 km north of stop A1 (from Selvig, 1994).
Neogene Rio Grande extension. Eaton (1986) attributed uplift to
Neogene rifting coeval with extension and lithospheric thinning
along the crest of the ridge. The age of this topographic high has
been questioned (Gregory and Chase, 1992; Gregory, 1994) based
on the leaf morphology of late Eocene and early Oligocene floras
from Florissant in the southern Front Range. Gregory and Chase
interpret these flora to have lived at essentially the same altitude as
they are found today. Radical river incision and reorganization
throughout the Rocky Mountain region, however, strongly suggest
Neogene regional uplift (Steven and others, 1997), although these
features have also been attributed to climatic changes.
Driving instructions: Take I-70 west out of Denver to exit 259 for
Morrison. After exiting, turn south on Rte. 26 (under the interstate)
and take an immediate left into the parking lot. We will climb the
ridge east of the parking lot. After an overview at the top of the ridge,
participants can walk northeast down to the paved path which tra-
verses the Jurassic and Cretaceous units exposed in the ridge.
This locality exposes Jurassic Morrison Formation and Creta-
ceous Dakota Formation (Fig. 3a) folded above the Golden fault
(Fig. 2) near a critical transition in structural styles along the eastern
boundary of the Front Range. Permo-Pennsylvanian Fountain For-
mation and Precambrian rocks underlie the mountains to the west,
whereas younger Cretaceous and Tertiary rocks underlie the Denver
Basin to the east. Rapid exposure of the basement at the start of the
Laramide orogeny is indicated by the minimal time between the
youngest marine ammonite zone in the Pierre Shale (69 Ma) and
the basement clasts in the latest Cretaceous-Paleocene Arapahoe
Formation, which is overlain by the Denver Formation that includes
the 64 Ma basalt flows of Table Mountain. The widespread late
Eocene erosional surfaces and minimal elevation differences
between exposures of the Late Eocene Wall Mountain Tuff
(Leonard and Langford, 1994) allow for minimal post-Laramide
faulting along the southeastern margin of the Front Range.
Structural styles of Laramide faulting and folding change
radically along the flank of the range. Near the I-70 roadcut at
stop A1, an emergent splay of the Golden fault system lies
roughly at the base of the Dakota hogback. This thrust fault cuts
out much of the Benton Formation, all of the Niobrara Forma-
tion, and much of the Pierre Shale. The Golden fault system
probably connects with the Perry Park-Jarre Creek thrust to the
south, which puts Proterozoic basement against Paleocene rocks
and has a stratigraphic separation of over 2.8 km (Scott, 1963).
The Golden fault system has been interpreted as a series of paral-
lel, WSW-dipping reverse faults (Berg, 1962; Weimer and Ray,
1997) which thrust basement eastward over the western edge of
the southern Denver Basin. The highly asymmetrical shape of the
southern and central Denver Basin can be attributed to thrust
loading by this thrust system.
28 E. A. Erslev et al.
About 40 km north of stop A1, the eastern margin of the Front biotite-quartz-feldspar gneiss. Pervasive chlorite-epidote alter-
Range steps eastward, giving it a more northerly average orienta- ation is almost certainly related to contact metamorphism by the
tion. These steps are created by northwest-striking, northeast-dip- Montezuma stock. Numerous normal faults cut the Williams
ping reverse faults that expose basement rocks in the core of Range thrust zone and may be related to late-stage stock
fault-propagation folds (Erslev and Rogers, 1993). Relative to the emplacement or Neogene extension.
southern Denver Basin, the northern Denver Basin is both shallower
and more symmetric, with the basin axis farther from the range. STOP A3. Faulting in the Dakota Sandstone
STOP A2. Contact metamorphosed rocks in the Williams Driving instructions: Rejoin Rte. 6, drive 7.2 miles west to Dillon
Range thrust zone, which places Precambrian gneiss over Dam Road, just after the town of Dillon. Turn left, proceed 0.4 miles
Pierre shale (Fig. 6) and park in small turnout at west end of the first segment of the dam.
The Blue River valley follows part of the belt of sedimentary
Driving instructions: Take I-70 43 miles west to the Loveland rocks that extends along the west flank of the Front Range. To
Pass exit (Exit 216), continue 13.5 miles southwest on Rte. 6 over the northeast, on the west side of the Williams Fork Mountains,
Loveland Pass. Turn left on Gondola Road (Fig. 6), cross Mon- the Williams Range thrust defines the structural margin of the
tezuma Road, jogging slightly left onto North Fork Drive; park Front Range. The Williams Range thrust is buried beneath an
on east side of the circle at the end of drive. Note: this parking extensive landslide complex, except where the thrust is exposed
area is on private land and access permission must be granted along Interstate 70 about 2.5 km to the northeast of stop A3 (Kel-
by the owners. Follow a faint trail through the woods toward the logg, 1997a). There, the thrust dips about 35° east and contains a
cliffs on the ridge to the east. 5-meter-thick zone of brecciated Precambrian gneiss overlying
The Williams Range thrust forms the west-central structural Pierre Shale. The buried trace of the thrust climbs along the west
boundary of the Front Range (Fig. 2) and extends from Middle side of the range to the north and tops the range at Ute Pass,
Park just north of Kremmling to South Park, where it is probably which we will visit at stop A4.
continuous with the Elkhorn thrust to the south (Bryant and oth- The landslide deposits may be as thick as several hundred
ers, 1981). North of Middle Park, the low-angle, en echelon meters and contain blocks of Proterozoic rocks tens of meters
Never Summer thrust steps east from the Williams Range thrust long. They are deeply incised and no longer retain hummocky
and defines the east side of North Park. topography, suggesting a late Tertiary or early Pleistocene age
The age of thrusting on the west side of the Front Range is not (Kellogg, 1997a,b). Initial brecciation of these rocks may have
precisely known, although the onset of Laramide deformation in occurred during fault-bend folding as the Williams Range thrust
this area is generally regarded as the age of the 70 Ma Pando Por- flattened from relatively steep dips at depth, where Proterozoic
phyry near Minturn, about 30 km to the west (Tweto and Lovering, rocks occupy both the hanging wall and the footwall, to gentle
1977). Synorogenic, late Paleocene units in South Park are overrid- dips where the footwall is mostly Cretaceous shale (Kellogg,
den by the Elkhorn thrust, indicating that if the Elkhorn and 1997b). Evidence for pervasive fracturing of the basement hang-
Williams Range thrusts are synchronous, movement along the ing-wall rocks includes the smooth and rounded crest of the
Williams Range thrust probably continued into early Eocene. The Williams Fork Mountains to the northeast. This mountain crest is
Never Summer thrust cuts rocks of the Paleocene and Eocene Coal- in stark contrast to the rugged crest of the Gore Range, which is
mont Formation (Fig. 3b, O’Neill, 1981), but the age of the part of only slightly higher than the Williams Fork Mountains and is
the formation overridden by the fault is not precisely known. underlain by similar but less fractured Proterozoic rocks. Starting
Unlike the higher angle faults of the eastern margin of the in late Neogene time, incision of the Blue River undercut the shat-
Front Range, the Williams Range thrust is low angle to nearly tered Proterozoic rocks, which Kellogg (1997b) inferred caused
horizontal in most places. At this locality, the Montezuma stock, much of the west side of the Williams Fork Mountains to slide.
a monzogranite porphyry with an age of 38-39 Ma (Marvin and The Blue River valley is a 5-to-9-km-wide half graben,
others, 1989; McDowell, 1971) domed Williams Range thrust bounded on the west by the east-dipping Blue River normal fault
along normal faults, forming a thrust window (Fig. 6; Ulrich, of Neogene age. The fault defines the abrupt east margin of the
1963). A basement overhang of at least 9 kilometers is indicated Gore Range. A complex zone of faulted and fractured Protero-
by the distance from the thrust window to the frontal exposure of zoic rock 0.5-1.0 km wide lies in its footwall (Tweto and others,
the thrust to the west combined with the angle of the thrust plane 1970; West, 1978). The Blue River fault has a minimum dis-
and the thickness of the sedimentary section. Contact metamor- placement of 1.2 km based on the relief and the absence of
phism of both the cataclastic basement rocks in the hanging wall Phanerozoic sedimentary rocks in the Gore Range west of the
and the Pierre Shale in the footwall allows what may the best fault. The latest movement along the fault was probably no
exposure of a Laramide thrust in the entire Front Range! younger than Pliocene or early Pleistocene (West, 1978). An
The nearly horizontal Williams Range thrust is exposed at extensive apron of glacial deposits now covers most of the fault
the base of the cliff, where Pierre Shale hornfels is overlain by 2- trace and the valley floor west of the Blue River.
3 m of basement hornfels grading upward from cataclasite to The half graben is cut by numerous north-striking normal
Laramide to Holocene structural development of the northern Colorado Front Range 29
39 105 55'W
37' to Loveland
30"N Pass
Pspg
Pu
Qm
Pu
STOP Pu Pu
2A
Pu
Kp
Qm Pu
Kp
to Dillon 6 Qm
Gon d ola
Ro ad Qm Montezum
a
Qal Road
Mo
Pu Kp nte Figure 6. Geologic map
Qm zu of the Keystone area
Qm ma around stop A2, show-
Pu St ing a portion of the
oc Williams Range thrust
k window. Adapted from
Qm Ulrich (1963).
Kp
Tqm
Ql
Pu Tqm
Pu
0 2 KM
faults that are almost entirely east dipping, suggesting that the from several localities in the area indicate two distinct directions
Blue River graben is a west-tilted structure above a listric fault at (NNE and NE) of horizontal shortening in the Tertiary.
depth. Kellogg (1999) suggested that the west-directed Gore
fault, a reverse fault along the west side of the Gore Range with STOP A4. Williams Range thrust at Ute Pass (Fig. 8)
significant movement in both late Paleozoic and Laramide times,
is listric and provided the surface along which Neogene reactiva- Driving instructions: Return to Rte 6, which turns into Rte 9, and
tion occurred (Fig. 7). continue 19 miles north from I-70 underpass. Turn right on Ute
Faults in the outcrop of folded Dakota Formation include Pass Road and climb to the pass, stopping at the first large expo-
normal, strike-slip and thrust faults. Multiple slickenlines on the sure of fractured Precambrian rock beyond the pass.
thrust planes suggest multidirectional thrusting. Data collected This locality is just east of the Williams Range thrust (Fig. 8),
30 E. A. Erslev et al.
whose trace is covered in the gully just to the southwest. The low- time. In addition, young AFT dates adjacent to the graben gener-
angle attitude of the thrust was demonstrated by a small outcrop of ally become older away from the graben. The underlying
Pierre Shale in the valley to the north, which was exposed when assumptions and results are outlined in the following discussion.
the road over Ute Pass was improved to serve the mill for the Hen- At the beginning of Laramide deformation (about 70 Ma),
derson molybdenum mine (Tweto and Reed, 1973). As discussed the Proterozoic rocks of the Gore range were overlain by about
at the last stop, extensive fracturing of the Precambrian gneiss is 3.1 km of Phanerozoic sedimentary rocks under an epicontinental
possibly related to fault-bend folding of the hanging-wall rocks sea in which the Pierre Shale was being deposited. Assuming a
when they moved through a bend in the thrust (Kellogg, 1997b). thermal gradient of 25°/km and a sea-floor temperature of 4°C, a
A short distance to the west, a low-angle footwall imbricate of temperature of 110°C, the annealing temperature for fission
the Williams Range thrust places calcareous shale of the Niobrara tracks in apatite, would be attained at 4.2 km below sea level.
Formation above Pierre Shale. Farther west, a spectacular vista of Most of the apatite fission-tracks at depths greater than 4.2 km
the Gore Range and Blue River half graben is opposite a promi- below this sea floor would have been annealed in Late Creta-
nent roadcut in Pierre Shale containing numerous minor thrust ceous time, and the upper 1.1 km of basement rock would have
faults. Just above the road to the southeast, the Gunsight Pass been in the Laramide partial annealing zone and have had AFT
Member of the Pierre Shale (W.A. Cobban, personal commun., dates greater than 70 Ma.
1998) is prominently exposed in an overturned footwall syncline. The presence of Cretaceous sedimentary rocks on the low,
West of Ute Pass, Proterozoic granite and migmatite form northern part of the Gore Range demonstrates that less than 3.1
the high peaks of the Gore Range, which is bounded on the west km of rock has been stripped away from this area since the close
by the Gore fault (Fig. 7). This fault also marks the eastern mar- of the Laramide, yet AFT dates from basement rocks in this area
gin of the Central Colorado trough, which formed during the late are 20-30 Ma. AFT dates of apatite near Ute Pass are similar to
Paleozoic Ancestral Rocky Mountain orogeny. Post-Cretaceous those in the low part of the Gore Range, except for one older date
slip on the Gore fault is at least 300 m. The Blue River normal of 35 Ma from a down-faulted block forming the bottom of the
fault marks the east face of the Gore Range facing Ute Pass. The Williams Fork Valley (Tweto and Reed, 1973).
impression from the topography is that the Gore Range is a block AFT dates from a 1.4 km vertical interval in the high, central
that had a more or less flat surface prior to being uniformly ele- part of the Gore Range, where only Proterozoic basement rocks
vated and dissected. North of the high peaks, faults cut across the are exposed, are from 6 to 32 Ma (Fig. 9), with younger dates
range, which becomes is a smooth-topped ridge capped by about from the base of the Blue River fault scarp and older dates from
200 m of Dakota Sandstone and underlying strata (Fig. 9). Here, high altitudes on the west side of the range (Naeser and others,
total throw on the Blue River fault is only 100 to 200 m, a sub- 1999). Isothermochrons (surfaces of constant age) dip west,
stantial decrease from the minimum 1,200 m throw to the south. away from the east flank of the Gore Range, and are discordant to
the even-topped spurs that extend to the Blue River fault from the
Apatite fission-track geochronology crest of the range. If the ridge crests at the same altitude represent
the remains of an erosional surface, this surface must be younger
Apatite fission-track (AFT) dates from both sides of the Blue than 16 Ma (late Neogene), which is an AFT date from the high-
River graben (Fig. 9) show that uplift and cooling of the Protero- est sample along the east face of the range.
zoic basement rocks occurred in late Paleogene and Neogene Uplift and erosion alone cannot account for the young AFT
CRETACEOUS
ROCKS
JURASSIC-
0 PROTEROZOIC
PENNSYLVANIAN
ROCKS
MISSISSIPPIAN- BASEMENT
KM CENTRAL ROCKS
CAMBRIAN
COLORADO
-5 TROUGH
(PENNSYLVANIAN-PERMIAN)
-10
Figure 7. Schematic, east-west cross section across the Blue River graben showing the prevalent
east-facing normal faults and the suggestion that they sole into the Gore fault, a high-angle con-
tractional structure with significant displacement during the late Paleozoic. Adapted from Kel-
logg (1999).
Laramide to Holocene structural development of the northern Colorado Front Range 31
106 07'30" W
EST O
F
R
Pu
C
R
A
N
lake
GE
39
Prairie Mtn. 50'
N
Pu
67 Kn
45
15 70 42
10
26
Kp Ute
Kph
45 Pass
36
(Stop A4)
Kd
10 Kph 30 (& pC)
Kpk
55
81
35
76 60
Kd Figure 8. Geologic map of the
Qal Ute Pass area around stop A4.
Kpg Adapted from Holt (1961).
Kn
15
25 Kn
Qtc
Kp
Kp 30
9 Tt
dates; the rocks along the axis of the Rio Grande rift must have Blue River region, therefore, cannot be obtained. The pattern of
been heated in late Tertiary time. Magma at high levels in the young AFT dates adjacent to the Blue River valley is similar to
crust probably caused the elevated heat flow, which wiped out the those from the Sawatch Range (Bryant and Naeser, 1980; Shan-
Laramide annealing or partial annealing zones. Quantitative non, 1987; Kelly and Chapin, 1995) and Sangre de Cristo Moun-
information on Laramide uplift from the fission-track data in the tains (Lindsay and others, 1986; Kelly and Chapin, 1995), which
32 E. A. Erslev et al.
also border the Rio Grande rift. The considerable Neogene heat- mid-Tertiary and Neogene age are faulted and either folded or
ing, uplift, and erosion along the northern Rio Grande rift also tilted during multiple deformation events spanning the Tertiary.
may be coeval with the Miocene tilting of the Front Range postu- The Mt. Bross reverse fault is exposed east of Mount Bross
lated by Steven and others (1997). and dips east, bringing Pierre Shale over Middle Park Formation
(Izett, 1968). Further east, the Pierre Shale is overlain by the Mid-
STOP A5. Green Mountain intrusive center in Pierre Shale dle Park Formation in a slight angular unconformity marked by
the truncation of a sandy unit within the Pierre Shale. The cliffy
Driving instructions: Return to Rte 9 and head north 12 miles to exposures above the unconformity expose the volcanoclastic
the northern end of Green Mountain Reservoir. Turn left on Windy Gap Volcanic Member of the Middle Park Formation and
Heney turnoff and drive to good exposures east of the dam. define the western limb of the Breccia Spoon syncline. This
A complex of laccoliths and sills surround a small stock on north-trending syncline appears to be folded by the WNW-trend-
the west side of Green Mountain. Dikes related to this center ing basement uplift west of Hot Sulfur Springs, suggesting
intrude the Pierre Shale 3 to 10 km east and north of Green Moun- changes in shortening directions through time. The Middle Park
tain. Some dikes in the northern part of the high Gore Range have Formation is overlain by the tuffaceous Rabbit Ears volcanic
been correlated with this center (Tweto and Lovering, 1977). rocks which are Oligocene to earliest Miocene in age (Izett,
Basaltic or andesitic lava flows to the east and south may have had 1968). These beds are overlain by the Miocene Troublesome For-
their source here at Green Mountain. The southernmost flows are mation, which were folded and subsequently overlain by tilted
in a tilted fault block that dips 25° south into a fault. Other flows Pliocene(?) basalt flows (Izett, 1968).
rest unconformably on gently east-dipping Cretaceous rocks. The aptly named Troublesome Formation is well dated by
This porphyritic latite or trachyte contains sanidine, ande- mammal fossils and volcanic ash with fission track ages between
sine, hornblende and augite phenocrysts. A fission-track analysis 20 and 13 Ma (Izett and Barclay, 1973). To the west near
on zircon gave an age of 29.9 +/- 2.4 Ma (Naeser and others, Kremmling, it blankets the Precambrian rocks in the hanging
1973). Faults of small displacement cut the intrusive rocks and wall of the Williams Range thrust. This suggests that the Trou-
adjacent rocks are altered to kaolinite. Pyrite and, rarely, spha- blesome Formation may have been deposited in localized basins
lerite and galena occur along the faults in the intrusive rock and caused by backsliding on this thrust fault. Fold axes in the Trou-
sedimentary country rock (Taggert, 1962). blesome Formation parallel the Williams Range thrust and are
defined by rotated strata with slight to locally steep (80°) dips
Driving instructions to evening accommodations at Hot Sulphur (Izett and Barclay, 1973). The origin of this folding is unclear.
Springs Resort: Return to Rte 9 and head north to Kremmling. Later tilting of Pliocene basalt flows is consistent with Neogene
Turn east on Rte. 40 at Kremmling, turning left on the first road in normal faulting seen to the south in the Blue River half graben.
Hot Sulphur Springs. The springs emanate from a fault separat- The cliffs across the river are made up of the Windy Gap Vol-
ing the Jurassic Morrison Formation from the Cretaceous-Pale- canic Member of the Middle Park Formation. The Middle Park
ocene Middle Park Formation. Enjoy! Formation and the correlative Coalmont Formation in North Park
are synorogenic with the Laramide orogeny. Up to 2 kilometers of
DAY 2 Middle Park Formation are exposed in this area. The lowest part of
the Middle Park Formation is the Cretaceous-Paleocene Windy
STOP B1. Overview of Middle Park, Mt. Bross fault, and the Gap Volcanic Member. The diversity of rock types within the
Breccia Spoon syncline Windy Gap Volcanic Member shows the complexity of early
Laramide events. It contains lenticular zones of fragmental
Driving instructions: Rejoin Rte 40 going east toward Granby. andesitic rocks that thicken to the south that were probably
Turn right on Grand County Road 55 at the east side of Hot Sul- extruded from volcanoes to the east and southeast along the Col-
phur Springs. Turn around at the cemetery and park at crest of orado Mineral Belt (Izett, 1968). Clasts of Precambrian basement
hill overlooking Hot Sulphur Springs. rocks in arkosic sedimentary rocks, both within and underneath the
This stop gives a panoramic view of structural relationships andesitic rocks, indicate unroofing of the Front Range in the early
in Middle Park. Middle Park is a section of the axial foreland Laramide. Arkosic sedimentary rocks become dominant at the top
basin that probably stretched along the entire western margin of of the Windy Gap Volcanic Member and then fine upward into silt-
the Front Range during the Laramide orogeny. To the south, syn- stones near the top of the Middle Park Formation (Izett, 1968).
orogenic strata are exposed in South Park but have been stripped
by erosion in the intervening Blue River half graben. Here and to STOP B2. Minor faulting in the Middle Park Formation
the north, the Middle Park basin is continuous with the North Park
basin, with the different names resulting from the topographic Driving instructions: Return to Rte 40 and travel east, turning
high separating the two topographic basins. This area shows the north on Rte 125, stopping at the large road cut at the bends in
complexity of the Tertiary structural evolution in the west flank of the road one kilometer north of its intersection with Grand
the Front Range. Igneous and volcanoclastic units of Laramide, County Road 408.
106 15' 106 00'
Xu
Tv
Ts
W
Ms Ti (30? Ma) Ms
il
Xu Gr
li
Xu STOP A5 X ee W Xu Xu Xu
nM Xu
a
tn illi
m
.R
es.
am Ts
s
Ms Tv Ms s
Ra
F
ng Prairie
A'
or
Qu Ms e Mountain
k
Ms Tv
X
Ms STOP A4
Mss Xu
Xu
thru
Ms
Xu Tvs
Bl
st fa
ue
Xu
Go
u
ive
Xu
lt
re
Ms
r
A
Qu 39 45'
Mt. Xu
PPs
PP
Ps I Powell
M
Xu
o
Qu
u
G Ms
n
or
ta
QuQu
QuQQu e
in
Qu
QuQu
uQu PIPs Xu
s
Xu Ms
Ms
PIP
PIPs
Ps Qu
no
rm
al
PIPS
Qu
Xu R
fau
an Xu
lt
Xu g e
Xu
STOP A3 X
Dillo
Sample localities and dates
n Re
fau
Ms
lt
s.
Vail Pass
30-40 Ma
20-30 Ma Ti
Xu
10-20 Ma Xu 44 Ma
<10 Ma Qu
Vail Pass
PIPs
Xu Ms
10 km Qu
Ti
Ti 39 30'
Xu Qu
Figure 9. Map showing locations of fission-track samples and ranges of data from the Gore Range and the Ute Pass
vicinity. Line of section shown on Fig. 10. Geology simplified from Tweto and others (1970).
34 E. A. Erslev et al.
Both the Windy Gap Volcanic Member, which forms the volcanic center may have caused localized thrust faulting and
cliffs one kilometer to the south, and the overlying arkosic sands folding, trapping the synclinal keel of the Miocene North Park
of the Middle Park Formation contain a multitude of slickensided Formation. Concentric thrust faults caused by gravity spreading
faults. Initial observations indicate multiple distinct populations of the Marysvale Volcanic Field in southern Utah (Merle and oth-
of thrust and strike-slip faults. At several localities in the Windy ers, 1993) could be analogous structures.
Gap Volcanic Member, thrust slickenlines trend from N-S to E-
W, with more east-west oriented slickenlines cut by more north- STOP B4. Laramide thrusting and mid-Tertiary igneous activ-
east-southwest slickenlines. At this locality, spectacular thrust ity at Cameron Pass
faults indicating E-W shortening appear to be cut by northeast-
and north-striking strike-slip faults. Driving instructions: Continue northeast on Jackson County Road
27 and turn right (east) at T intersection with Rte 14 toward Gould.
STOP B3. North Park syncline Stop at exposures of red sandstone 1 mile west of Cameron Pass.
The Cameron Pass area exposes the east-dipping Never
Driving instructions: Continue north on Rte 125 over Willow Summer thrust system of Laramide age overprinted by mid-Ter-
Creek Pass. At Rand, turn northeast on Jackson County Road 27. tiary intrusions of the Mt. Richthofen intrusive complex and later
Park on overlook (Owl Ridge) into North Park. high-angle faults. The roadcuts expose overturned Mesozoic
This locality is on the southwest margin of the North Park strata beneath Precambrian rocks brought up by a splay of the
syncline, a large, northwest-striking structure that is discordant to Never Summer thrust system. South of the road, the rugged
the major Laramide thrusts and folds on the eastern border of Nokhu Crags expose contact-metamorphosed Pierre Shale adja-
North Park. The North Park syncline parallels the Independence cent to the 29 Ma (K-Ar; Corbett, 1964) Mt. Richthofen batholith
Mountain thrust system that defines the north end of North Park, (O’Neill, 1981). Upward bowing by the batholith has formed a
where the thrust system truncates NNW-trending Laramide folds window in Never Summer thrust system in which Pierre shale is
in the Coalmont Formation (Blackstone, 1977). The core of the exposed 10 kilometers from the nearest exposure of the thrust
North Park syncline contains Miocene North Park Formation, front, where Precambrian rocks are thrust on top of the Coalmont
which is roughly equivalent to the Troublesome Formation to the Formation. The overhang may have been expanded by the intru-
south and Browns Park Formation to the west. sion of the Mt. Richthofen batholithic complex, but the Laramide
The origin of the North Park syncline is problematic. Early overhang still must have been quite substantial. The similarities
workers linked it to normal faulting on the northeast side of the between the Never Summer and the Williams Range (STOP A2)
structure. Later subsurface work indicates stratal shortening thrust windows are striking!
across the fault zone, suggesting a northeast-dipping thrust fault The Mt. Richthofen batholith is cut by high-angle, north-
or a transpressive structure (D. Stone, personal commun., 1999). striking faults with early strike-slip slickenlines overprinted by
The age of the fold is also problematic since it includes Miocene extensional dip-slip slickenlines. Chapin (1983) proposed that
strata deposited long after the end of Laramide shortening. several major north-striking fault zones along the western margin
Recent seismic profiles and well information indicate that of the Front Range formed during late Laramide right-lateral
the southwest side of the anticline is detached and suggest peri- strike-slip faulting. The fact that strike-slip slickenlines cut the 29
ods of both shortening and extension (Fig. 11. Preliminary stud- Ma Mt. Richthofen batholith suggests either unusually prolonged
ies of mid-Tertiary strike-slip faulting discussed earlier do Laramide shortening or a separate phase of post-Laramide strike-
suggest that north-south shortening and compression roughly slip faulting. The existence of similar strike-slip faults in southern
normal to the fold axis could have occurred during the mid-Ter- Colorado, where they cut the mid-Tertiary Cripple Creek intru-
tiary. Another possibility is that gravitational sliding away from a sions (T. Wawrzyniec, personal commun., 1996), and in northern
New Mexico, where they cut late Oligocene igneous rocks
(Erslev, unpub. data, 1998), suggests that strike-slip faulting may
A A' have been regionally important in mid-Tertiary time.
Strike-slip faulting on the western margin of the Front Range
Prairie Mountain
Mount Powell
Williams Fork River
27.4
f
24.7 thrust system that forms the northern boundary of North Park.
Blue River
22.4
Williams Range Right-lateral motion on the eastern margin of North Park could
Xu 21.4 thrust
Blue River
27.5
21.6 have moved the basin northward, providing the impetus for stuff-
14.4 Xu
12.8 ? Ms Tvs ing the northern end of the basin underneath the Independence
1 km Ms 35.9 Mountain thrust system. Accommodation of right-lateral shear
1 km
along thrusts would explain why evidence for right-lateral shear
has not been documented north of the Colorado-Wyoming border.
Figure 10. Cross section from the Williams Fork to Mt. Powell showing The causal relationships between mid-Tertiary sedimenta-
locations of apatite fission-track dates. 5X vertical exaggeration. tion, igneous activity and faulting are unclear. Laramide and/or
Figure 11. Uninterpreted (a) and interpreted (b) sketches of a seismic profile through North Park syncline near stop B3. Profile is oriented NE-SW from approximately 10 km SSE of Coalmont,
Colorado, to approximately 15 km ENE of Walden, Colorado. Well logs indicate that the Mesozoic section is thicker in the center than on the flanks, suggesting thrust faulting repeated and thick-
ened section. Later extension may have reactivated the low-angle thrusts, resulting in normal faulting in the hanging wall and tightening of the North Park syncline. Seismic data provided by Seis-
mic Exchange, Inc.; interpretation is that of the authors.
36 E. A. Erslev et al.
Figure 12. Simplified geologic map (after Braddock and others, 1988a, 1988b) of the eastern margin of
the Front Range northeast of Fort Collins, Colorado, with rose diagrams showing the trends of maxi-
mum compression directions indicated by minor faults (Holdaway, 1998)
earlier deformation may have produced a fundamental zone of Cordilleran orogeny to the southwest (Christiansen and Yeats,
crustal weakness along the current Rio Grande rift that was 1992). In either case, the resultant zone of structural and mag-
exploited by intrusion and strike-slip faulting in the mid-Ter- matic weakening may be responsible for subsequent Rio
tiary. The transtensional to transpressive environment indicated Grande rifting synchronous with Basin and Range extension.
by mid-Tertiary faulting may have also facilitated magmatic Thus, the coherence of the Colorado Plateau during Neogene
intrusion in the crust. Alternatively, thermal softening due to extension may not be due to the abnormal strength of its under-
mid-Tertiary igneous activity may have focused strike-slip lying lithosphere so much as the unusual weakness of the
faulting driven by the gravitational spreading of the extending lithosphere underlying the current Rio Grande rift.
Laramide to Holocene structural development of the northern Colorado Front Range 37
Figure 13. Cross section through the North Fork fault and fold (Holdaway, 1998). Line of section is shown in Fig. 12.
STOP B5. Precambrian mylonitized pegmatite in the Skin roclasts indicate high-angle reverse faulting at this outcrop, but
Gulch Shear Zone studies in progress at University of New Mexico suggest multiple
stages and ages of Precambrian ductile shearing on shear zones
Driving instructions: Continue east on Rte 14. 0.6 miles after the throughout the Front Range.
small tunnel through finer-grained, mylonitic basement rocks,
park on north side of road beyond the 35 MPH sign. STOP B6: North Fork Fault and associated fault-propagation fold
This exposure exhibits excellent shear textures that are read-
ily apparent in outcrop due to the coarse grain size of the Driving instructions: Continue east on Rte 14. At the T intersec-
mylonite’s pegmatite protolith. C-S surfaces and sigma porphy- tion with Rte 287, turn left (north). After 1.2 miles, take a sharp
38 E. A. Erslev et al.
Barker, Fred, Hedge, C.E., Millard, H.T., and O’Neil, J.R., 1976, Pikes Peak
left turn on to private property. Permission is needed for access, batholith: geochemistry of some minor elements and isotopes, and impli-
but the view from the nearby roadcut of Morrison Formation is cations for magma genesis, in Epis, R.S., and Weimer, R.J., eds., Studies
worthy of a stop. in Colorado field geology: Colorado School of Mines Professional Con-
The northwest-striking North Fork Fault causes a two kilo- tributions 8, p. 44-56.
Berg, R.R., 1962, Mountain flank thrusting in Rocky Mountain foreland,
meter right-lateral separation of the basement-Fountain Forma-
Wyoming and Colorado: American Association of Petroleum Geologists
tion unconformity (Fig. 12, Braddock and others, 1988a). This Bulletin, v. 46, p. 2019-2032.
structure is typical of the major, southeast-plunging fault-propa- Bergh, S., and Snoke, A., 1992, Polyphase Laramide deformation in the Shirley
gation folds in the northeastern Front Range (Fig. 13). Trenching Mountains, south central Wyoming foreland: Mountain Geologist, v. 29,
of this fault along the irrigation ditch shows that the fault dips p. 85-100.
Bird, P., 1988, Formation of the Rocky Mountains, a continuum computer model:
approximately 45° northeast, suggesting thrust/reverse motion.
Science, v. 239, p. 1501-1507.
Overlying strata are truncated by the fault at the level of the Per- Bird, P., 1998, Kinematic history of the Laramide orogeny in latitudes 35o-49o,
mian Ingleside Formation, whose bleaching adjacent to the fault western United States: Tectonics, v. 17, p. 780-801.
in the hanging-wall anticline suggests iron reduction during Blackstone, D.L., Jr, 1977, Independence Mountain thrust fault, North Park basin,
hydrocarbon migration through these rocks. At the level of the Colorado: Contributions to Geology, v. 16, p. 1-15.
Blackstone, D.L., 1991, Tectonic relationships of the southeastern Wind River
Cretaceous Dakota Group, the strata form a nearly continuous
Range, southwestern Sweetwater Uplift, and Rawlins Uplift, Wyoming:
fold, indicating fault-propagation folding where basement fault- Report of Investigation No. 47, Geological Survey of Wyoming, 24 p.
ing dies upward into cover folding. This fold is highly conical, Bolyard, D.W., 1997, Colorado Front Range Field Trip Road Logs, in Bolyard,
dying out to the southeast. D.W., and Sonnenberg, S.A., eds., Geologic history of the Colorado Front
Most minor faults in the northeastern margin of Front Range Range: Rocky Mountain Association of Geologists Field Trip Guidebook,
189 p.
indicate ENE shortening and compression, with almost no sug-
Boos, C.M., and Boos, M.F., 1957, Tectonics of eastern flank and foothills of
gestion of late extensional faulting common on the west side of Front Range: American Association of Petroleum Geologists Bulletin, v.
the Front range (Holdaway, 1998). At this locality, however, the 41, p. 2603-2676.
Ingleside Formation in the hanging wall anticline is elongated by Braddock, W.A., Connor, J.J., Swann, G.A., and Wolhford, D.D., 1988a, Geo-
normal faults. Bedding elongation during fault-propagation fold- logic map of the Laporte quadrangle, Larimer County, Colorado: U.S.
Geological Map GQ-1621.
ing probably generated these faults.
Braddock, W.A., Wohlford, D.D., and O’Connor, J.T., 1988b, Geologic map of
The thrust and reverse faults in the northeastern Front Range the Horsetooth reservoir quadrangle, Larimer County, Colorado: U.S.
are not directly responsible for the uplift of the range because Geological Survey Map GQ-1625.
their northeasterly dip brings the basin up relative to the range. Brewer, J.A., Allmendinger, R.W., Brown, L.D., Oliver, J.E., and Kaufman, S.,
Erslev and Selvig (1997) explained this geometry as the resulting 1982, COCORP profiling across the Rocky Mountain front in southern
Wyoming, Part 1, Laramide structure: Geological Society of America
from backthrusting off a blind, east-northeast-directed thrust in
Bulletin, v. 93, p. 1242-1252.
the basement. More recent work by Holdaway (1998) has shown Brister, B.S., and Gries, R.R., 1994, Tertiary stratigraphy and tectonic develop-
that an additional component of deep crustal wedging is neces- ment of the Alamosa basin (northern San Luis Basin), Rio Grande rift,
sary to uplift the range (Figs. 4d, 14). In this case, thrust on the south-central Colorado, in Keller, G.R., and Cather, S.M., eds., Basins of
east side of the Front Range may simply be surface tightening the Rio Grande rift—structure, stratigraphy, and tectonic setting: Boulder,
Colorado, Geological Society of America Special Paper 291, p. 39-58.
due to wedging at depth, a situation analogous to structures on
Brown, W.G., 1988, Deformation style of Laramide uplifts in the Wyoming fore-
the northeast side of the Wind River Mountains (Erslev, 1986). land, in Schmidt, C.J., and Perry, W.J., Jr., (eds.), Interaction of the Rocky
Mountain foreland and the Cordilleran thrust belt: Geological Society of
Driving instructions back to Denver: Follow Rte 14 south and America Memoir 171, p. 53-64.
east through Fort Collins to I-25 south. Bryant, B., and Naeser, C.W., 1980, The significance of fission-track ages of
apatite in relation to the tectonic history of the Front and Sawatch Ranges:
Geological Society of America Bulletin, v. 91, pt. 1, p. 156-164.
ACKNOWLEDGMENTS Bryant, B., Marvin, R.F, Naeser, C.W., and Mehnert, H.H., 1981, Ages of igneous
rocks in the South Park-Breckenridge region, Colorado, and their relation
The compilation of this field trip guide was a greatly assisted to the tectonic history of the Front Range uplift, in Shorter Contributions
by the meticulous road log of Dudley Bolyard (1997). Reviews to Isotope Research in the western United States, 1980: U.S. Geological
Survey Professional Paper 1199-C, p. 15-35.
by Dave Lageson, Jack Reed, Bob Scott, and Bruce Trudgill
Chapin, C.E., 1983, An overview of Laramide wrench faulting in the southern
improved this manuscript. Rocky Mountains with emphasis on petroleum exploration, in Lowell,
J.D., ed., Rocky Mountain foreland basins and uplifts: Denver, Rocky
REFERENCES CITED Mountain Association of Geologists, p. 169-179.
Chapin, C.E., and Cather, S.M., 1981, Eocene tectonism and sedimentation in the
Anderson, J.L., 1983, Proterozoic anorogenic plutonism in North America, in Colorado Plateau-Rocky Mountain area: Arizona Geological Digest, v.
Medaris, L.G., and others, eds., Proterozoic geology: Geological Society 14, p. 175-198.
of America Memoir 161, p. 133-154. Chapin and Cather, 1994, Tectonic setting of the axial basins of the northern and
Baldridge, W.S., Keller, G.R., Haak, V., Wendlandt, E., Jiracek, G.R., and Olsen, central Rio Grande rift, in Keller, G.R., and Cather, S.M. , eds., Basins of
K.H., 1995, The Rio Grande Rift, in Olsen, K.H., (ed.), Continental rifts: the Rio Grande rift—structure, stratigraphy, and tectonic setting: Boulder,
evolution, structure, tectonics: Elsevier, p. 233-275. Colorado, Geological Society of America Special Paper 291, p. 5-25.
Laramide to Holocene structural development of the northern Colorado Front Range 39
Chase, R., Genovese, P., and Schmidt, C., 1993, The influence of Precambrian Kelley, S.A., and Chapin, C.E., 1997, Internal structure of the southern Front
rock compositions and fabrics on the development of Rocky Mountain Range, Colorado, from an apatite fission-track thermochronology per-
Foreland folds, in Schmidt, C.J., Chase, R., and Erslev, E.A., (eds), Base- spective, in Bolyard, D.W., and Sonnenberg, eds., Geologic History of the
ment-cover kinematics of Laramide foreland uplifts: Geological Society Colorado Front Range, 1997 RMS-AAPG Field trip #7: Rocky Mountain
of America Special Paper 280, p. 45-72. Association of Geologists, Denver, p. 19-30.
Christiansen, R.L., and Yeats, R.S., 1992, Post-Laramide geology of the U.S. Kellogg, K.S., 1997a, Geologic map of the Dillon quadrangle, Summit and
Cordilleran region, in Burchfiel, B.C., Lipman, P.W., and Zoback, M.L., Grande Counties, Colorado U.S. Geological Survey Open-File Report 97-
eds., The Cordilleran Orogen: Conterminous U.S.: Boulder, Colorado, 738, scale 1:24,000.
Geological Society of America, The Geology of North America, v. G-3, p. Kellogg, K.S., 1997b, The Williams Range thrust near Dillon, Colorado - hanging
261-406. wall fracturing during Laramide thrusting preparing bedrock for major
Corbett, M.K., 1964, Tertiary igneous petrology of the Mt. Richthofen-Iron Neogene and Pleistocene landsliding: Geological Society of America
Mountain area, north-central Colorado: University of Colorado Ph.D. the- Abstracts with Programs, v. 29, no. 6, p. A163.
sis, 115 p. Kellogg, K.S., 1999, Neogene basins of the northern Rio Grande rift—partition-
DOE, 1993, Phase II geologic characterization data acquisition high resolution ing and asymmetry inherited from Laramide and older uplifts: Tectono-
deep seismic; revised final report: U.S. Department of Energy, Rocky physics, v. 305, p. 141-152.
Flats Plant, 155 p. Kluth, C.F., and Nelson, S.N., 1988, Age of the Dawson Arkose, southwestern
Eaton, G.P., 1986, A tectonic redefinition of the southern Rocky Mountains: Air Force Academy, Colorado, and implications for the uplift history of
Tectonophysics, v. 132, p. 163-193. the Front Range: The Mountain Geologist, v. 25, no. 1, p. 29-35.
Erslev, E. A., 1986, Basement balancing of Rocky Mountain foreland uplifts, Kulik, D.M., and Schmidt, C.J., 1988, Regions of overlap and styles of interaction
Geology, v. 14, p. 259-262. of Cordilleran thrust belt and Rocky Mountain foreland, in Schmidt, C.J.,
Erslev, E.A., 1993, Thrusts, back-thrusts, and detachment of Laramide foreland and Perry, W.J., Jr., eds., Interaction of the Rocky Mountain foreland and
arches, in Schmidt, C.J., Chase, R., and Erslev, E.A., (eds.), Laramide the Cordilleran thrust belt: Geological Society of America Memoir 171, p.
basement deformation in the Rocky Mountain foreland of the western 75-98.
United States: G.S.A. Special Paper 280, p. 339-358. Leonard, E.M., and Langford, R.P., 1994, Post-Laramide deformation along the
Erslev, E.A., and Rogers, J.L., 1993, Basement-cover kinematics of Laramide eastern margin of the Colorado Front Range - a case against significant
fault-propagation folds: in Schmidt, C.J., Chase, R., and Erslev, E.A., eds., deformation: The Mountain Geologist, v. 31, p. 45-52.
Basement-cover kinematics of Laramide foreland uplifts: Geological Lindsey, D.A., Andriessen, P.A.M., and Wardlaw, B.R., 1986, Heating, cooling,
Society of America Special Paper, p. 125-146. and uplift during Tertiary time, northern Sangre de Cristo Range, Col-
Erslev, E.A., and Selvig, B., 1997, Thrusts, backthrusts and triangle zones: orado: Geological Society of America Bulletin, v. 97, p. 1133-1143.
Laramide deformation in the northeastern margin of the Colorado Front Lowell, J.D., 1983, Foreland deformation, in Lowell, J.D., ed., Rocky Mountain
Range, in Bolyard, D.W., and Sonnenberg, S.A., Geologic history of the foreland basins and uplifts: Rocky Mountain Association of Geologists,
Colorado Front Range, Rocky Mountain Association of Geologists, Den- Denver, Colorado, 392 p.
ver, Colorado, p. 65-76. Marvin, R.F., Mehnert, H.H., Naeser, C.W., and Zartman, R.E., 1989, U.S. Geo-
Gregory, K.M., 1994, Paleoclimate and paleoelevation of the Florissant flora, in, logical survey radiometric ages-Compilation “C”, Part 5. Colorado, Mon-
Evanoff, E., ed., Late Paleocene geology and paleoenvironments of central Col- tana, Utah, and Wyoming: Isochron/West, no. 11, p. 1-42.
orado with emphasis on the geology and paleontology of Florissant Fossil Matthews, V., III, and Work, D.F., 1978, Laramide folding associated with base-
Beds National Monument: Guidebook for field trip Rocky Mountain Sec- ment block faulting along the northeastern flank of the Front Range, Col-
tion of the Geological Society of America at Durango Colorado, p. 57-66. orado, in Matthews, V., III (ed.), Laramide folding associated with
Gregory, K.M., and Chase, C.G., 1992, Tectonic significance of paleobotanically basement block faulting: Geological Society of America Memoir 151, p.
estimated climate of the late Eocene erosion surface, Colorado: Geology, 101-124.
v. 20, p. 581-585. McDowell, F.W., 1971, K-Ar ages of igneous rocks from the western United
Gries, R.R., 1983, North-south compression of the Rocky Mountain foreland States: Isochron/West no. 2, p. 1-16.
structures, in Lowell, J.D., and Gries, R.R., eds., Rocky Mountain fore- Merle, O.R., Davis, G.H., Nickelsen, R.P., and Gourlay, P.A., 1993, Relation of
land basins and uplifts: Rocky Mountain Association of Geologist, Den- thin-skinned thrusting of Colorado Plateau strata in southwestern Utah to
ver, Colorado, p. 9-32. Cenozoic magmatism: Geological Society of America Bulletin, v. 105, p.
Gries, R., and Dyer, R.C., 1985, Seismic exploration of the Rocky Mountain 387-398.
region: Denver, Colorado, Rocky Mountain Association of Geologists and Naeser, C.W., Izett, G.A., and White, W.H., 1973, Zircon fission-track ages from
Denver Geophysical Society, p. 1139-1142. some middle Tertiary igneous rocks in northwestern Colorado: Geological
Hamilton, W., 1988, Laramide crustal shortening, in Perry, W.J., and Schmidt, society of America Abstracts with Programs, v. 5, no. 6, p. 498.
C.J., eds., Interaction of the Rocky Mountain foreland and the Cordilleran Naeser, C.W., Bryant, Bruce, Kellogg, Karl, and Perry, W.J., Jr., 1999, Middle to
thrust belt: Geological Society of America Memoir 171, p. 27-39. late Tertiary cooling of the Gore and western Front Ranges, Central Col-
Hansen, W.R., 1986, History of faulting in the eastern Uinta Mountains, Colorado orado, from apatite fission-track data: Geological Society of America
and Wyoming, in Stone, D.S. (ed.), New interpretations of northwest Col- Abstracts with Programs, v. 31, National Meeting, Denver.
orado Geology: Rocky Mountain Association of Geologists, p. 19-36. Oldow, J.S., Bally, A.W., and Lallemant, H. G., 1990, Transpression, orogenic
Holdaway, S.M., 1998, Laramide deformation of the northeastern Front Range, float, and lithospheric balance: Geology, v. 18, p. 991-994.
Colorado: Evidence for deep crustal wedging during horizontal compres- O’Neill, J.M., 1981, Geologic map of the Mount Richthofen quadrangle and the
sion: M.S. thesis, Colorado State University, 146 p. western part of the Fall River Pass quadrangle, Grand and Jackson Coun-
Izett, G.A., 1968, Geology of the Hot Sulphur Springs quadrangle, Grand County, ties, Colorado: U.S. Geological Survey Miscellaneous Investigations
Colorado: U.S. Geological survey Professional Paper 586, 79 p. Series Map I-1291.
Izett, G.A., and Barclay, C.S.V., 1973, Geologic map of the Kremmling quadran- Prucha, J.J., Graham, J.A., and Nickelson, R.P., 1965, Basement-controlled defor-
gle, Grand County, Colorado: U.S. Geological Survey Geologic Quad- mation in Wyoming Province of Rocky Mountain foreland: American
rangle Map GQ 1115, scale 1:24,000. Association of Petroleum Geologist Bulletin, v. 49, p. 966-992.
Jacobs, A.F., 1983, Mountain front thrust, southeastern Front Range and northwest- Raynolds, R.G. , 1997, Synorogenic and post-orogenic strata in the central Front
ern Wet Mountains, in Lowell, J.D., ed., Rocky Mountain foreland basins Range, Colorado, in Bolyard, D.W., and Sonnenberg, S.A., eds, Geologic
and uplifts: Rocky Mountain Association of Geologists, p. 229-244. history of the Colorado Front Range, 1997 RMS-AAPG Filed Trip #7:
40 E. A. Erslev et al.
Denver, Colorado, Rocky Mountain Association of Geologists, p. 43-48. Association of Geologists, p. 19-36.
Reed, J.C., Jr., Bickford, M.E., Premo, W.R., Aleinikoff, J.N., and Pallister, J.S., Stone, D.S., 1995, Structure and kinematic genesis of the Calla wrench duplex:
1987, Evolution of the Early Proterozoic Colorado province: Geology, v. Transpressional reactivation of the Precambrian Cheyenne belt in the
15, no, 9, p. 861-865. Laramie basin, Wyoming: American Association of Petroleum Geologists
Reed, J.C., Jr., Bickford, M.E., and Tweto, O, 1993, Proterozoic accretionary ter- Bulletin, v. 79, p. 1349-1376.
ranes of Colorado and southern Wyoming, in Van Schmus, W.R. and Stearns, D.W., 1971, Mechanisms of drape folding in the Wyoming province:
Bickford, M.E., eds., Transcontinental Proterozoic provinces, in Reed, Wyoming Geological Association, 23rd Annual Field Conference,
J.C., Bickford, M.E., Houston, R.S., Link, P. K., Rankin, D.W., Sims, Wyoming Tectonics Symposium Guidebook, p. 82-106.
P.K., and Schmus, W.R., eds, Precambrian: Conterminous U.S.: Geologi- Stearns, D.W., 1978, Faulting and forced folding in the Rocky Mountain foreland:
cal Society of America Decade of North American Geology, v. C-2, p. in Matthews, V., III, ed., Laramide folding associated with basement block
211-228. faulting in the western United States: Geological Society of America
Sales, J.K., 1968, Cordilleran foreland deformation: American Association of Memoir 151, p. 1-37.
Petroleum Geologists Bulletin, v. 52, p. 2000-2015. Taggert, J.N., 1962, Geology of the Mount Powell quadrangle, Colorado: Har-
Schmidt, C.J., and Perry, W.J., Jr., 1988, Interaction of the Rocky Mountain fore- vard University unpublished Ph.D. thesis, 239 p.
land and the Cordilleran thrust belt: Geological Society of America Mem- Tweto, O., 1975, Laramide (Late Cretaceous-early Tertiary) orogeny in the South-
oir 171, 582 p. ern Rocky Mountains, in Curtis, B.F., ed., Cenozoic history of the south-
Scott, G.R., 1963, Bedrock geology of the Kassler quadrangle, Colorado: U. S. ern Rocky Mountains: Geologic Society of America Memoir 144, p. 1-44.
Geological Survey Professional Paper 421-B, p. 71-125, scale 1:24,000. Tweto, O., 1976, Geologic map of the Craig 1ox2o quadrangle, northwestern Col-
Scott, G.R., 1972, Geologic map of the Morrison quadrangle, Jefferson County, orado: U.S. Geological Survey Miscellaneous Investigations Series map I-
Colorado: U.S. Geological Survey Miscellaneous Geologic Investigations 972.
Map I-790-A, scale 1:24,000. Tweto, O., 1978, Northern rift zone guide 1, Denver-Alamosa, Colorado, in Haw-
Selvig, B.W., 1994, Kinematics and structural models of faulting adjacent to the ley, J.W., compiler, Guidebook to the Rio Grande rift in New Mexico and
Rocky Flats Plant, central Colorado: M.S. thesis, Colorado State Univer- Colorado: New Mexico Bureau of Mines and Mineral Resources Circular
sity, 133 p. 163, p. 13-32.
Selverstone, J., Hodgins, M., Shaw, C., Aleinikoff, J.N., and Fanning, C.M., 1997, Tweto, O., 1979, Geologic map of Colorado: U.S. Geological Survey, scale
Proterozoic tectonics of the northern Colorado Front Range, in Bolyard, 1:500,000.
D.W. and Sonnenberg, S.A., eds, Geologic history of the Colorado Front Tweto, O., 1987, Rock units in the Precambrian basement in Colorado: U.S. Geo-
Range, 1997 RMS-AAPG Field Trip #7: Rocky Mountain Association of logical Survey Professional Paper 1321, 54 p.
Geologists, Denver, p. 9-18. Tweto, O., Bryant, B., and Williams, F.E., 1970, Mineral resources of the Gore
Shaw, C.A., Snee, L.W., Selverstone, Jane, and Reed, J.C., Jr., 1999, 40Ar/39Ar Range-Eagles Nest Primitive Area and vicinity, Summit and Eagle Coun-
thermochronology of mesoproterozoic metmorphism in the Colorado ties, Colorado: U.S. Geological Survey Bulletin 1319-C, p. C1-C127.
Front Range: Journal of Geology, v. 107, p. 49-67. Tweto, O., and Lovering, T.S., 1977, Geology of the Minturn 15-minute quadran-
Smith, J.H., 1964, Geology of the sedimentary rocks of the Morrison quadrangle: gle, Eagle and Summit Counties, Colorado: U.S. Geological Survey Pro-
U.S. Geological Survey Miscellaneous Geologic Investigations Map I- fessional Paper 956, 96 p.
428, scale 1:24,000. Tweto, O., and Reed, J.C., Jr., 1973, Reconnaissance of the Ute Peak 15-minute
Smithson, S.B., Brewer, J.A., Kaufman, S., Oliver, J.E., and Hurich, C.A., 1979, quadrangle, Grand and Summit Counties, Colorado: U.S. Geological Sur-
Structure of the Laramide Wind River uplift, Wyoming, from COCORP vey Open-File Map, scale: 1:62,500.
deep reflection data and from gravity data: Journal of Geophysical Tweto, O., and Sims, P.K., 1963, Precambrian ancestry of the Colorado mineral
Research, v. 84, p. 5955-5972. belt: Geological Society of America Bulletin, v. 74, p. 991-1014.
Spang, J.H., Evans, J.P., and Berg, R.R., 1985, Balanced cross sections of small Unruh, D.M., Snee, L.W., Foord, E.E., and Simmons, W.B., 1995, Age and cooling
foldthrust structures: Mountain Geologist, v. 22, p. 3746. history of the Pikes Peak batholith and associated pegmatites [abs]: Geo-
Steven, T. A., 1975, Middle Tertiary volcanic field in the southern Rocky Moun- logical Society of America Abstracts with Programs, v. 27, no. 6, p. 468.
tains, in Curtis, B.F., ed., Cenozoic history of the Southern Rocky Moun- Ulrich, G.E., 1963, Petrology and structure of the Porcupine Mountain area, Sum-
tains: Geological Society of America Memoir 144, p. 75-94. mit County, Colorado: University of Colorado Ph.D. thesis, 205 p.
Steven, T.A., Evanoff, E., Yuhas, R.H., 1997, Middle and late Cenozoic tectonic Weimer, R.J., and Ray, R.R., 1997, Laramide mountain flank deformation and the
and geomorphic development of the Front Range of Colorado, in Bolyard, Golden fault zone, in Bolyard, D.W. and Sonnenberg, S.A., Geologic his-
D.W., and Sonnenberg, S.A., eds., Geologic history of the Colorado Front tory of the Colorado Front Range, 1997 RMS-AAPG Field Trip #7:
Range, 1997 RMS-AAPG Filed Trip #7: Rocky Mountain Association of Rocky Mountain Association of Geologists, Denver, p. 49-64.
Geologists, Denver, p. 115-124. West, M.W., 1978, Quaternary geology and reported surface faulting along the
Stone, D.S., 1984, The Rattlesnake Mountain, Wyoming, debate: a review and east flank of the Gore Range, Summit County, Colorado: Colorado
critique of models: The Mountain Geologist, v. 21, p. 37-46. School of Mines Quarterly, v. 73, no. 2, 66 p.
Stone, D.S., 1986, Seismic and borehole evidence for important pre-Laramide Wise, D.U., 1963, Keystone faulting and gravity sliding driven by basement uplift
faulting along the axial arch in northwest Colorado, in Stone, D.S. (ed.), of the Owl Creek Moutnains, Wyoming: Americal Association of Petro-
New interpretations of northwest Colorado Geology: Rocky Mountain leum Geologists, v. 80, p. 1397-1432.
Printed in U.S.A.
Geological Society of America
Field Guide 1
1999
Erslev, E. A., and Holdaway, S. M., 1999, Laramide faulting and tectonics of the northeastern Front Range of Colorado, in Lageson, D. R., Lester, A. P.,
and Trudgill, B. D., eds., Colorado and Adjacent Areas: Boulder, Colorado, Geological Society of America Field Guide 1.
41
42 E. A. Erslev and S. M. Holdaway
Figure 1. Simplified geologic map of the northern Front Range showing field trip stops and the loca-
tions of Figures 7 and 9.
This switch in fault dip directions along the eastern margin northeastern Front Range because these faults uplift the basin,
of the Front Range reveals a fundamental structural problem: not the range.
what is the relationship of the faults and folds in the foothills of Erslev and Selvig (1997) addressed the origin of this along-
the northeastern Front Range to the overall structure of the strike variability of fault dip direction by suggesting that the
Front Range? How did they contribute to the uplift of the Front west-dipping thrusts of the Golden fault system are the primary
Range basement-cored arch? These faults and folds exhibit the faults and that northeast-dipping faults are backthrusts above a
opposite shear sense of the regional structure, bringing the basin blind section of the Golden fault system that never emerges
side up and the mountain side down, in direct conflict with the from the basement (Fig. 3d). In this model, basin-ward tilt of
overall uplift of the range. Restoration of faulting exposed at the northeastern margin of the Front Range occurred by
the surface using simple vertical uplift (Fig. 3a), strike-slip (Fig. domino-style rotation of backthrust-bounded blocks in a wide
3b), or upthrust (Fig. 3c) models cannot explain the uplift of the zone of shear. Holdaway (1998) tested this model with regional
Figure 2. Evolution of structural interpretations of the
northeastern Front Range.
Figure 3. Structural models and basement block restorations for the northeastern Front Range.
44 E. A. Erslev and S. M. Holdaway
balancing and found that whereas simple backthrusting can are shown on the geologic sketch map in Fig. 1. The first four
explain a portion of the uplift in this part of the range, it cannot stops are in the Horsetooth Reservoir Quadrangle (Braddock
by itself explain all of the uplift of the arch margin because the and others, 1988c), the middle stops are in the Masonville
amount of exposed faulting is not sufficient to account for all Quadrangle (Braddock and others, 1970) and the final two stops
of the observed uplift (Fig. 3d). He suggested that uplift of the are in the Carter Lake Quadrangle (Braddock and others,
northeastern margin of the Front Range arch was due to the 1988b).
insertion of a lower crustal wedge beneath the arch (Fig. 3e).
This hypothesis is consistent with the larger slip amounts on STOP 1. Northern dam embankment of Spring Creek dam
thrust faults (Elkhorn thrust, Williams Range thrust, Never on Horsetooth Reservoir
Summer thrust) bounding the western margin of the Front
Range arch. While the geometry of the deep crustal wedge is Driving instructions: Drive north on I-25 to the Harmony Road
unknown, this model can explain all of the uplift of the north- exit in south Fort Collins. Exit and head west on Harmony
eastern Front Range. Road, which will turn into County Road 38E, through Fort
Collins to the top of Spring Creek dam. Turn north on County
FIELD TRIP ROAD LOG AND STOP DESCRIPTIONS Road 23 and park on north side of dam. Cross the road and
walk west to the cliff overlooking Horsetooth Reservoir.
This field trip follows paved roads in Larimer County, Col- From the lip of Duncan’s ridge, a popular Fort Collins climbing
orado, from Horsetooth Reservoir west of Fort Collins to Carter cliff in the Lytle Sandstone west of the parking lot, the rugged and
Lake Reservoir southwest of Loveland. The location of stops irregular topography underlain by Proterozoic crystalline rocks lies
to the west and is overlain by a late Paleozoic to Mesozoic clastic
sequence (Fig. 4). This sequence contains three resistant sandstone
units, the Permian Ingleside Formation, Permian Lyons Formation
and Cretaceous Dakota Group, which form well-defined hogbacks
generally dipping 15° to 30° east toward the northern Denver Basin.
Strike-valleys are formed by the less resistant Fountain, Owl
Canyon and Triassic-Jurassic formations. The Dakota Group forms
a sandwich of units, with the Lytle and Plainview Sandstones below
the Skull Creek Shale, which is overlain by the Muddy Sandstone
which consists of the Fort Collins and Horsetooth Sandstones.
Figure 4. Stratigraphic column of Phanerozoic units in the Figure 5. Slickensided faults (arcs) and slickenlines (triangles) north-
northeastern Front Range (from Graham and Ethridge, 1995). east of Spring Dam indicating ENE-WSW shortening and compression.
Laramide faulting and tectonics of the northeastern Front Range of Colorado 45
Figure 6. Stereonets of minor fault poles and slickenlines for the northeastern Front Range with average eigenvector
orientations and eigenvalues. The 4 fault pole maxima correspond to conjugate thrust and strike-slip faults. The aver-
age slip direction (080-13) is given by the first eigenvector of the slickenline orientations.
To the south, the section of road built over the Skull Creek slip direction existed during Laramide deformation, they should
Shale collapsed in the Spring of 1995 when unusually wet con- be recorded by the fractures. Shear sense on the fractures is usu-
ditions saturated the road bed and accelerated slumping of the ally clearly indicated by Riedel fractures (Petit, 1987) which
road. Repair of the roadbed included the installation of a cul- intersect the major fault planes at a low angle, giving a rough-
vert to drain the Skull Creek strike valley, replacement of the ness to the plane in the direction of slip.
earth fill and regrading of the slope below the road. An intense Fault studies at CSU have focused on Laramide faults in
rainstorm in July 1997 did not damage the road, although it did strata deposited before the Laramide orogeny and after the
trigger numerous landslides and caused major flooding in Fort Ancestral Rocky Mountain orogeny. Once the faults are divided
Collins. into subsets, strain and stress orientations can be determined
After the overview at Duncan’s ridge, the trip will walk east using average slickenline orientations, conjugate relationships,
across the road to the Horsetooth sandstone which has excellent P-T dihedra (Allmendinger and others, 1989), and reduced
exposures of strike-slip fractures indicating near-horizontal stress tensor methods (Angelier, 1990). The possibility of mul-
shortening and compression trending N79E (Fig. 5). The silica- tiple compression directions is evaluated by using the Compton
cemented quartz arenites of the Ingleside, Lyons and Dakota for- (1966) method which calculates the ideal s1 orientation for
mations preserve slickensided fractures throughout the region. each fault based on the average angle between conjugate pairs.
Shear fractures in these units rarely follow bedding planes, sug- This method can discriminate different compression directions
gesting that they form on ideal, “Andersonian” planes symmetric when conventional conjugate, Angelier (1990) and P-T dihedral
to the stress axes. Fault surfaces show excellent slickenlines in methods just average the compression directions.
the fine-grained, low-porosity cataclasite which coat the sur- Holdaway (1998) compiled 2232 minor fault and slicken-
faces. The multitude of shear fractures suggest that these planes line orientations to determine the kinematic development of the
are strain hardened, annealing after slip to be stronger than the northeastern Front Range (Fig. 6). When the ideal s1 trends are
original sandstone. This means that early phases of slip can be plotted in rose diagrams for the area around Horsetooth Reser-
preserved along with later phases of slip. As a result, if multiple voir and Milner Mountain anticline (Fig. 7), they show a uni-
46 E. A. Erslev and S. M. Holdaway
Figure 7. Expanded geologic map of the area west and southwest of Fort Collins with rose
diagrams showing ideal s1 trends calculated by the Compton (1966) method assuming an
angle of 50° between conjugate minor faults.
form ENE compression direction. The general sparsity of multi- outcrop pattern defining a southeast-plunging fold axis. The
modal ideal s1 trends suggests one direction of regional com- Lyons Formation, which forms the ridge under the Perch, is
pression and shortening (Holdaway, 1998). both folded and faulted, with 5 fault splays defining tilt
domains. Farther west, the Precambrian basement-Fountain
STOP 2. The Perch Formation unconformity was broken by two major faults but
was not folded. This indicates that basement block movements
Driving instructions: Drive south back across the dam to at deeper structural levels transitions upward into arcuate fault-
County Road 38E and turn west. Park at the parking lot at the propagation folding.
Perch, which is on the first ridge crest after you drive around All workers in this area have proposed basement block
the southern arm of Horsetooth Reservoir. motion, with detailed analyses (LeMasurier, 1970; Erslev and
This stop shows the form of many of the basement-cored Rogers, 1993) showing only minor (20°) basement rotations in
fault-propagation folds in the northeastern Front Range. To the the hanging wall tip of the largest structure in the area, the Milner
east, the Dakota Group sandstones form a smoothly-arcuate Mountain anticline. This can be seen in the excellent geologic
Laramide faulting and tectonics of the northeastern Front Range of Colorado 47
maps of the region by Braddock and others (1970, 1988a, 1988b, Dakota Group sandstones show more rounded fold hinges,
1998c, 1989), which show planar basement contacts offset by probably due to distributed deformation in the shales and evap-
faults with minimal folding of the contact in the vicinity of the orites of the Lykins Formation. The Fletcher Hill fault is well-
faults. These faults progressively lose displacement upward in the exposed in a valley incised into the resistant schists to the
folded sedimentary strata, forming fault-propagation folds northeast. The Precambrian-Fountain gouge contact within the
(Erslev, 1991; Erslev and Rogers, 1993). Of the six major faults fault zone has a 60°NE dip, suggesting reverse faulting on the
that cut the Precambrian-Pennsylvania contact near Fort Collins, Fletcher Hill fault. At this locality, the lack of steeply-inclined
only the Milner Mountain fault cuts the Dakota Group. sedimentary beds at the fault suggests a localized sub-thrust
The strata form doubly-plunging anticlines, like Bellvue basement wedge, a style evident elsewhere in the Rocky Moun-
dome to the north, or, more commonly, southeasterly-plunging tain foreland. Because the formation of a wedge steepens the
anticlines cored by northwest-striking faults. Trenching of the fault plane, the 60°NE dip should be interpreted as a maximum
major faults near Fort Collins show that the faults dip 20° to 70° dip for the fault. Less than a kilometer to the north, the sedi-
northeast, consistent with the earliest interpretations of Ziegler mentary beds are overturned, indicating the lack of an underly-
(1917). Absolute slip directions, which have been interpreted as ing basement wedge.
everything from strike slip to dip slip, are hard to determine The possibility of basement folding was investigated at this
from the major faults because their strain-softened gouge zones locality and at Milner Mountain to the southeast (Erslev and
only preserve the last motion on the fault. In general, however, Rogers, 1993). In both places, schistose rocks strike parallel to
the largest faults strike perpendicular to compression directions the fault. This orientation should provide an optimal configura-
indicated by minor faults (Fig. 7), suggesting dip-slip on thrust tion for flexural slip on foliation planes in the basement. Dip
and reverse faults. measurements, measured from the west, were taken at regular
intervals from the fault. These orientations show a 50 meter
STOP 3. Basement hanging wall immediately adjacent to the
Fletcher Hill fault
Figure 9. Expanded geologic map of the area around Carter Lake Reservoir with rose diagrams showing ideal s1 trends calculated by the Comp-
ton (1966) method assuming an angle of 50° between conjugate minor faults.
zone of possible basement rotation associated with the fault stones form the ridge on which we are standing as well as the
zones (Fig. 8). Thus, extensive folding in these Laramide fault- ridge to the east across the fold axis. The Carter Lake anticline
propagation folds is largely restricted to the sedimentary strata. is slightly asymmetric and trends north along most of its length
until it abruptly becomes highly asymmetric and takes a north-
STOP 5. Overview of Carter Lake anticline westerly trend south of where we are standing.
Multiple fold trends are found throughout the northeastern
Driving instructions: Return south on County Road 27 and turn Front Range, but the Carter Lake Reservoir area best shows the
right at Masonville. Turn right on Rte 34 and then take the first interaction of distinct north- and northwest-trending structures.
left on County Road 29. Turn right on County Road 18 and then The variety of structural trends may be caused by multiple
left on County Road 31, which will take you to Carter Lake stages of differently oriented shortening and compression, an
Reservoir. Park at the northern abutment of the second dam and edge effect that refracted maximum shortening and compres-
climb to the ridge for an overview. Note: The ridge directly sion axes to become nearly perpendicular to the arch margin, or
above this spot is on public land, but the extension of the ridge reactivation of pre-existing weaknesses. Rose diagrams of ideal
to the south is on private land for which permission is required. s1 axis trends for individual localities in the Carter Lake Reser-
The area surrounding Carter Lake Reservoir exposes a voir area (Fig. 9) show dominantly uni-modal, east-west com-
zone of interaction between distinctly different structural pression perpendicular to north-trending structures.
trends. The top of this ridge is in the forelimb of the asymmet- Immediately north of the Carter Lake Reservoir area, Pre-
ric Carter Lake anticline from which we can view the intersec- cambrian dikes parallel several northwest-striking faults (Brad-
tion of north- and northwest-trending structures. The dock and others, 1970). This suggests that northwest-trending
Pennsylvanian Fountain Formation is exposed in the core of the structures in this area may be localized on pre-existing weak-
anticline to the east and the resistant Lyons and Ingleside Sand- nesses, which can explain their obliquity to east-west Laramide
Laramide faulting and tectonics of the northeastern Front Range of Colorado 49
shortening and compression. Several other northwest- and west- Wyoming, Part 1, Laramide structure: Geological Society of America
Bulletin, v. 93, p. 1242-1252.
northwest- striking faults are both oblique to Laramide contrac- Chapin, C.E., and Cather, S.M., 1983, Eocene tectonics and sedimentation in
tion directions and parallel to Precambrian foliations or the Colorado Plateau; Rocky Mountain area, in Lowell, J.D., ed., Rocky
mylonite zones in the northeastern Front Range (Braddock and Mountain foreland basins and uplifts: Rocky Mountain Association of
others 1988a, 1988b). This suggests that reactivation of pre- Geologists, p. 33-56.
existing basement weaknesses can explain at least some of the Compton, R.R., 1966, Analyses of Pliocene-Pleistocene deformation and
stresses in northern Santa Lucia Range, California: Geological Society
structural variability in this area. of America Bulletin, v. 77, p. 1361-1380.
Erslev, E.A., 1991, Trishear fault-propagation faulting: Geology, v. 19, p. 617-
STOP 6. South side of south dam, Carter Lake 620.
Erslev, E.A., 1993, Thrusts, backthrusts and detachment of Laramide foreland
Driving instructions: Continue south on County Road 31, park- arches, in Schmidt, C.J., Chase, R., and Erslev, E.A., eds., Laramide
basement deformation in the Rocky Mountain foreland of the western
ing at southeast corner of the third dam. United States: Geological Society of America Special Paper 280, p. 125-
This locality exposes two sets of thrusts with radically dif- 146.
ferent trends. This can be explained by localized effects, but the Erslev, E.A., in review, Multi-stage, multi-directional Tertiary shortening and
presence of multistage, multidirectional shortening elsewhere in compression in north-central New Mexico: submitted to GSA Bulletin.
the Laramide orogen (Gries, 1983; Chapin and Cather, 1983; Erslev, E.A., and Rogers, J.L., 1993, Basement-cover kinematics of Laramide
fault-propagation folds: in Schmidt, C.J., Chase, R., and Erslev, E.A.,
Bergh and Snoke, 1992; Erslev, in review) suggests that these eds., Laramide basement deformation in the Rocky Mountain foreland
faults may represent distinct periods of shortening. of the western United States: Geological Society of America Special
Paper 280, p. 339-358.
Driving instructions back to Denver: Continue south on County Erslev, E.A., and Selvig, B., 1997, Thrusts, backthrusts and triangle zones:
Road 31 and turn left on County Road 8E at the first intersection. Laramide deformation in the northeastern margin of the Colorado Front
Range, in Bolyard, D.W., and Sonnenberg, S.A., Geologic history of the
Go east until the road ends, turn right on County Road 23 and Colorado Front Range, Rocky Mountain Association of Geologists,
then left on Rte 56, which will take you through Berthoud to I-25. Denver, Colorado, p. 65-76.
Flores, R.M., Roberts, S.B., and Perry, W.J., 1995, Paleocene paleogeography
ACKNOWLEDGMENTS of the Wind River, Bighorn, and Powder River Basins, Wyoming,
Wyoming Geological Association Guidebook.
Graham, J., and Ethridge, F.G., 1995, Sequence stratigraphic implications of
This manuscript was improved by reviews by Dave Lage- gutter casts in the Skull Creek Shale, lower Cretaceous, northern Col-
son and Bruce Trudgill. orado: Mountain Geologist, v. 32, p. 81-94.
Gries, R.R., 1983, North-south compression of Rocky Mountain foreland struc-
REFERENCES CITED tures, in Lowell, J.D., ed., Rocky Mountain foreland basins and uplifts:
Rocky Mountain Association of Geologists, p. 9-32.
Allmendinger, R.W., Aydin, A., Engelder, T., and Pollard, D.D., 1989, Quanti- Holdaway, S.M., 1998, Laramide deformation of the northeastern Front Range,
tative interpretation of joints and faults: G.S.A. Short Course. Colorado: Evidence for deep crustal wedging during horizontal com-
Angelier, J., 1990, Inversion of field data in fault tectonics to obtain the regional pression: M.S. thesis, Colorado State University, 146 p.
stress - III. A new rapid direct inversion method by analytical means: Jacobs, A.F., 1983, Mountain front thrust, southeastern Front Range and north-
Geophysical Journal International, v. 103, p. 363-379. western Wet Mountains, in Lowell, J.D., ed., Rocky Mountain foreland
Berg, R.R., 1962, Mountain flank thrusting in the Rocky Mountain foreland, basins and uplifts: Rocky Mountain Association of Geologists, p. 229-
Wyoming and Colorado: AAPG Bull. v. 46, p. 2010-2032. 244.
Bergh, S., and Snoke, A., 1992, Polyphase Laramide deformation in the Shirley LeMasurier, W.E., 1970, Structural study of a Laramide fold involving shallow
Mountains, south central Wyoming foreland: Mountain Geologist, v. 29, seated basement rock, Front Range Colorado: Geological Society of
p. 85-100. America Bulletin, v. 81, p. 435-450.
Boos, C.M., and Boos, M.F., 1957, Tectonics of eastern flank and foothills of Matthews, V., III, and Sherman, G.D., Origin of monoclinal folding near Liver-
Front Range: A.A.P.G. Bull., v. 41, p. 2603-2676. more, Colorado: Mountain Geologist, v. 13, p. 61-66.
Braddock, W.A., Connor, J.J., Swann, G.A., and Wolhford, D.D., 1988a, Geo- Matthews, V., III, and Work, D.F., 1978, Laramide folding associated with base-
logic map of the Laporte quadrangle, Larimer County, Colorado: U.S. ment block faulting along the northeastern flank of the Front Range,
Geological Map GQ-1621. Colorado, in Matthews, V., III (ed.), Laramide folding associated with
Braddock, W.A., Calvert, R.H., Gawarecki, S.J., and Nutalaya, P., 1970. Geo- basement block faulting: Geological Society of America Memoir 151, p.
logic map of the Masonville quadrangle: U.S. Geol. Survey Geol. Quad. 101-124.
Map GQ-832. Petit, J.P., 1987, Criteria for the sense of movement on fault surfaces in brittle
Braddock, W.A., Calvert, R.H., O’Connor, J.T., and Swann, G.T., 1989, Geo- rocks: Journal of Structural Geology, v. 9, p. 597-608.
logic map of the Livermore quadrangle, Larimer County, Colorado: Prucha, J.J., Graham, J.A., and Nickelson, R.P., 1965, Basement-controlled
U.S. Geological Survey Map GQ-1618. deformation in Wyoming Province of Rocky Mountain foreland: Amer-
Braddock, W.A., Nutalaya, P., and Colton, R.B., 1988c, Geologic map of the ican Association of Petroleum Geologist Bulletin, v. 49, p. 966-992.
Carter Lake quadrangle: U.S. Geol. Survey Geol. Quad. Map GQ-1628. Weimer, R.J., and Ray, R.R., 1997, Laramide mountain flank deformation and
Braddock, W.A., Wohlford, D.D., and O’Connor, J.T., 1988b, Geologic map of the Golden fault zone, in Bolyard, D.W. and Sonnenberg, S.A., Geo-
the Horsetooth reservoir quadrangle, Larimer County, Colorado: U.S. logic history of the Colorado Front Range, 1997 RMS-AAPG Field Trip
Geological Survey Map GQ-1625. #7: Rocky Mountain Association of Geologists, Denver, p. 49-64.
Brewer, J.A., Allmendinger, R.W., Brown, L.D., Oliver, J.E., and Kaufman, S., Ziegler, V., 1917, Foothills structure in northern Colorado: Journal of Geology,
1982, COCORP profiling across the Rocky Mountain front in southern v. 25, p. 715-740.
Printed in U.S.A.
Geological Society of America
Field Guide 1
1999
ABSTRACT
This trip integrates the geomorphic, botanic, and hydrogeologic aspects of diverse
ground-water and surface-water systems of the Geneva Creek sub-basin near
Guanella Pass, Colorado. Fracture-flow crystalline bedrock systems, and alluvial, col-
luvial, and glacial systems are integrated for slope and riverine wetland structure and
function. Current research regarding the development of the Wetlands Integrated
Hydrologic Analysis method for delineating wetlands structure and function in the
southern Rocky Mountains is discussed with respect to various sites observed along
the field trip route.
INTRODUCTION Grant, turn right, following the all-weather road up the valley,
past the Geneva Basin Ski Area, to the top of Guanella Pass.
This field trip integrates the geomorphic, botanic, and Either route requires approximately 1.5 to 2 hours driving time.
hydrogeologic aspects of the diverse ground-water and surface-
water systems, and wetlands of the Geneva Creek subbasin, and GENERAL BACKGROUND
part of the North Fork of the South Platte River System from
Guanella Pass to Bailey, Colorado. Fractured-flow bedrock Wetlands are important features in the landscape, particu-
ground-water systems and flow through unconsolidated larly in mountain valleys where they occur in association with
deposits and soils occur in this landscape. The geomorphic sys- streams, rivers, and ground-water discharge at seeps and
tems and materials affecting the hydrogeology of the area in- springs. Wetland definitions are numerous and often compli-
clude weathering and pedogenic processes and deposits, cated, however, one proposed by Tarnocai (1979) works well
Holocene mass movements and deposits, and Pleistocene gla- for our purposes here: “Wetland is defined as land having the
cial and Holocene fluvial erosional and depositional features. water table at, near, or above the land surface or which is satu-
The route passes through three life zones: the alpine, subalpine, rated for a long enough period to promote wetland or aquatic
and montane. Along the way, many of the plant communities processes as indicated by hydric soils, hydrophytic vegetation,
typical of the southern Rocky Mountains are seen. The various and various kinds of biological activity which are adapted to the
plant communities mentioned in the following road log are dis- wet environment.” During the last decade or so, there has been
cussed in greater detail in Mutel and Emerick (1992). The wet- increased recognition of the ecological and hydrological impor-
lands observed along the field trip result from the combined tance of wetlands (see for example, the excellent text of Mitsch
effects of surface water and ground-water systems, as well as and Gosselink, 1993). For the past three years, we have been
local geology and soils. involved in studies to characterize the hydrologic functions of
The road log starts at Guanella Pass (see Figure 1), which wetlands, which has resulted in the Wetland Integrated Hydro-
can be reached from Georgetown, Colorado, located along Inter- logic Analysis (WIHA) method. This field trip provides an
state 70 west to Denver, Colorado. In Georgetown, follow the opportunity to demonstrate the use of the WIHA method.
roadsigns to the Guanella Pass road. Alternatively, Guanella pass Based on studies by Kolm (1996); Kolm, et al.(1996);
may be reached by driving west from Denver along US 285. At Kolm, et al. (1998); and Harper-Arabie and Kolm (1998), the
Kolm, K. E., and Emerick, J. C., 1999, Hydrogeology and wetlands of the mountains and foothills near Denver, Colorado, in Lageson, D. R., Lester,
A. P., and Trudgill, B. P., eds., Colorado and Adjacent Areas: Boulder, Colorado, Geological Society of America Field Guide 1.
51
52 K. E. Kolm and J. C. Emerick
Wetland Integrated Hydrologic Analysis (WIHA) method was tion, and to quantify and distribute wetland system parameters.
designed to identify the hydrologic functions that are present in Based upon the information that was collected and measured,
the wetland being evaluated. In addition to providing a standard the appropriate wetland structure, physical and chemical char-
methodology for wetland hydrologic analysis, the advantages acteristics, distribution, and continuity of stratigraphic and
of incorporating the conceptualization and characterization lithologic units (soil and rock) were determined for each field
approach of Kolm (1996) and Kolm, et al (1996) are that: 1) the site. The type, properties, and distribution of geomorphic mate-
method is non-invasive; 2) in using the approach, the subsur- rials, landforms, slope, and other geomorphic processes and
face framework and ground-water flow system can be estimated characteristics were interpreted with respect to the surface
with minimal time spent on-site; and 3) WIHA provides a clear water and ground-water.
basis for model design when used for mathematical modeling Each type of wetland system was then characterized and
of the wetland system. classified using surface and subsurface characterization, and
hydrogeologic and ground-water system characterization and
METHOD DEVELOPMENT quantification, as appropriate (Kolm et al, 1998). The impor-
tance of hydrogeology and ground-water systems was deter-
The WIHA method was developed in an iterative manner mined for each wetland class.
that began with a theoretical concept of the common wetland
systems in Colorado. This was followed by data collection, as RESULTS AND DISCUSSION
needed, which led to a better understanding of the wetland sys-
tems. This process resulted in the development of several mod- The basic steps involved in the Wetland Integrated Hydro-
els of wetland structure and hydrology that were tested with logic Analysis method include: 1) problem definition; 2) data
field studies (Kolm, et al., 1998). The field sites included five base development; 3) preliminary conceptualization; 4) surface
wetlands that were identified as being typical of wetland types and subsurface characterization; 5) ground-water system char-
in the southern Rocky Mountains. These included two that were acterization; and 6) wetland system characterization and classi-
initially classified according to Brinson (1993) as hydrogeo- fication based on function, including water supply (Figure 2).
morphic (HGM) slope wetlands and three that were initially Based upon the WIHA assessment, the hydrodynamics and pre-
classified as HGM riverine wetlands. dominant water source of a wetland can be characterized, and
Prior to the field studies, existing data, including both the the wetland hydrologic functions determined (Table 1). The
natural and anthropogenic features of the five wetland sites, WIHA method will be demonstrated during the field trip.
were collected and analyzed. Field reconnaissance was con- The three-dimensional hydrogeologic structure of the wet-
ducted, as necessary, to relate the preliminary analysis of the land system has a dominant control over hydrologic functions
information collected to field study site conditions. In areas (Kolm et al. (1998); Harper-Arabie and Kolm, 1998). Therefore,
where field data were sparse, basic photointerpretation and ter- the wetland classes in the southern Rocky Mountains should be
rain analysis techniques were applied to remote sensing data, based upon the wetland hydrogeologic structure, with the stan-
aerial photography, and topographic maps to acquire informa- dard HGM classes of Brinson (1993) regarded as subclasses:
Hydrogeology and wetlands of mountains and foothills near Denver, Colorado 53
ROAD LOG
Mileage
Hydrogeology: Groundwater recharge occurs zones, at the base of over-steepened valley walls
from precipitation distributed throughout these high (such as in the Mt. Spalding area to the east), and in
elevations. Recharge can be particularly high along stream channels where it is an important component
the ridgetops during the summer with the melting of of the surface water baseflow. It is hypothesized that
large snow accumulations; note the snow cornices on the directions of preferred ground-water flow are ori-
the east, southeast- and northeast-facing slopes. ented parallel with the 1) northwest-southeast and 2)
Ground-water flow paths include flow through frac- northeast-southwest alignment of observed stream
tured bedrock from ridge tops to valley bottoms or channels.
other discharge areas, as well as through soil or Wetlands: The wetlands observed at the crest of
weathering materials, usually as infiltration to the Guanella Pass are primarily of the 3-layer and 2-
water table. These shallow and fractured-flow sys- layer hydrogeologic structure class, and include frac-
tems are unconfined. Ground-water discharge ap- tured crystalline bedrock systems supporting the
pears as springs or seeps in gullies along fracture overlying saturated glacial and mass wasting materi-
Hydrogeology and wetlands of mountains and foothills near Denver, Colorado 55
2.0 The road cut exposes a cross section of glacial till; 4.0-4.2 Stop 4: Outwash plain and top of debris-flow fan
the materials are well graded and poorly sorted. This at an approximate elevation of 3139 m (10,300 ft.).
location is the first observance of bristlecone pine Geomorphology: This stop is on a glacial ter-
(Pinus aristida), indicating windy conditions and dry race/outwash plain; a younger debris avalanche/land-
and rocky soils. Ski runs are seen to the west. Note slide fan covers the top of the outwash plain.
the avalanche/debris avalanche chute to the north of Vegetation: Note the extensive willow communi-
the last ski run to the right. The willows indicate that ties indicating nearsurface ground water. Beavers
the soils in the chute are wet and the water table is prefer valley bottoms that are wider and relatively
near the surface. flat such as seen here. Aspen and willow are favorite
(0.3 mile) items for food and building materials. Beaver occu-
2.5 This is the entrance to Geneva Basin Ski Area; the Tar- pation in these sites is cyclic. When enough aspen
ryall Mountains are seen to the south (downvalley). and willow are present, beavers invade the area and
(0.2 mile) construct dams, raising the water table. This often
2.7 Note the willow-covered debris flow/avalanche kills conifers as a result of over-saturation of the root
materials in the valley that have flowed to the south. zone. Over time, the beavers remove all of the large
To the west, note rock fall (weathering) talus where aspen and willow, and eventually leave the area. The
the absence of soil prevents plant growth. abandoned dams disintegrate and the water table is
(0.5 mile) lowered, allowing reinvasion of conifers. Once the
3.2.1 Springs (discharge) are seen above the road from aspen and willow recover (perhaps in 10 to 20 years),
here to stop 3. the beaver return. There is evidence that beaver pop-
(0.4 mile) ulations have recently returned to this area.
3.6 Stop 3: Overlook of Geneva Creek, elevation ap- Hydrogeology: Groundwater tables are near the
proximately 3200 m (10,500 ft). surface in the outwash materials; the river channel
Geomorphology: Note the glacial moraine at tree- and stage probably control the groundwater level.
line up-valley toward stop 2. Below the moraine lie Groundwater discharge in the fan deposits is evident
debris avalanche/debris flow deposits derived from by the presence of willows.
the over-steepened valley walls. Below the overlook is Wetlands: The wetlands observed at this stop are
an outwash plain with a younger landslide/debris a mixture of 3-layer and 2-layer hydrogeologic sys-
avalanche fan stratigraphically and topographically on tems. Slope wetlands occur near the valley sides,
top. Directly beyond the road switchback is the sear grading into riverine wetlands adjacent to the stream.
(release area) of the debris avalanche/landslide. It (0.9-1.3 miles)
appears that the release area is a ground-water dis- 5.3 A lodgepole pine (Pinus contorta) forest appears on
charge zone, which, in turn, may have triggered the colluvium/till pockets; this indicates drier conditions
debris avalanche. Ground-water seeps through the till and past disturbance from a forest fire.
and colluvium located on the U-shaped, over-steep- (0.3 mile)
ened valley sides. The debris fan has pushed the river 5.6 Note the medial moraine on the other side of the
to the west side of the valley. Falling colluvium on the drainage; this deposit supports a sparse lodgepole
road indicates active creep. pine-Engelmann spruce forest. Outwash materials
Vegetation: The following plant community supporting a willow community indicate a shallow
relationships are observed here: a) spruce is located water table or a discharge zone.
on the sites where soils are relatively shallow and (0.4 mile)
more stable; b) aspen (Populus tremuloides) occurs 6.0 Note the burned area along the road that is covered
on unstable or disturbed sites, often near ground- with a young lodgepole forest. In the Rocky Moun-
water discharge areas; c) willows grow on ground- tains, lodgepole pine and aspen are socalled “succes-
water discharge areas and on wet disturbed sites. sional” tree species, invading forested areas that have
Willows are also growing on the outwash plain and been disturbed by fire, logging, or mass movements.
the debris flow fan (both deposits are wet areas). As time passes, and barring further disturbance, the
Hydrogeology: Groundwater is seeping from lodgepole and aspen are gradually replaced by
fractured bedrock and along the colluvium/bedrock Engelmann spruce, subalpine fir (Abies lasiocarpa),
contact. The roadcut has created a springline dis- and Douglasfir (Pseudotsuga menziesii), which form
charge area. Groundwater probably has triggered the climax forest stands.
debris avalanches; ground-water is currently draining (0.2 mile)
at these mass movement sites. 6.2 The main road intersects Bruno Gulch Road. Note
(0.4-0.6 mile) two levels of outwash plains. The willows (on wet
Hydrogeology and wetlands of mountains and foothills near Denver, Colorado 57
Figure 5. Schematic cross-section of the Geneva Valley showing the interaction of the surface
water, ground-water, and wetlands systems.
10.6 Whiteside Camp Ground is located here. Note the ern) with waste material from the tunnel excavation
ponderosa pine (Pinus ponderosa) and Douglas fir on the opposite bank.
on drier sites. The appearance of these conifer Vegetation: The wider valley of the North Fork
species and that of the blue spruce indicates that we of the South Platte River supports narrowleaved cot-
are leaving the subalpine life zone and entering the tonwood (Populus angustifolia). Intensive land use,
montane life zone. Phreatophytes include river birch including farming and grazing, has eliminated much
(Betula fontinalis), aspen, thin-leafed alder (Alnus of the natural riparian shrub communities that once
tenuifolia), and willow species. proliferated on both sides of the river channel.
(0.7 mile) Hydrogeology: Because of the water diversion,
11.3 Geneva Creek Picnic Ground. a more constant stream flow is maintained. The
(0.7 mile) South Platte River is still gaining by groundwater
12.0 Terrace and colluvial materials are seen along the baseflow and is the ultimate groundwater discharge
road. Soils are welldrained and thick, and support area from fractureflow in the surrounding crystalline
mountain mahogany (Cercocarpus montanus) shrub rocks. Near-surface water tables in the South Platte
communities. alluvium are indicated by cottonwoods. Turn around
(1.0 mile) and proceed down-valley (northeast) along US 285
13.0 Junction with U.S. 285 and the North Fork of the toward Denver, Colorado.
South Platte River at Grant, Colorado. Turn right (0.7 mile)
(toward the southwest). 14.5 Grant, Colorado (elevation 2642 m 8667 ft.). The cli-
(0.3-0.8 mile) mate at Grant is considerably different than that at
13.3-13.8 Excavated material from the Roberts Tunnel is on the Guanella Pass. Measurements from a weather station
south side of road. located here indicate an average annual temperature of
(0.0 miles) 3.7°C (38.6°F), and an average precipitation of 40 cm
13.8 Stop 7: Harold D. Roberts Tunnel at an approximate (15.6 inches) per year, 65 percent of which falls during
elevation of 2652 m (8700 ft). The Harold D. Roberts the months from June through November.
Tunnel was constructed in 1962 by the Army Corps of (0.6 mile)
Engineers as part of a trans-mountain water diversion 15.1 Terrace gravels can be observed on the south side of
project. The tunnel delivers water from Lake Dillon, the road.
located on the western slope of the Continental Divide, (0.5 mile)
to the South Platte watershed on the eastern slope. 15.6 U.S. 285 crosses a terrace.
Geomorphology: Note the river channel (mod- (1.1 miles)
Hydrogeology and wetlands of mountains and foothills near Denver, Colorado 59
Fred G. Luiszer
Department of Geological Sciences, University of Colorado, Boulder, Colorado 80309, United States
ABSTRACT
INTRODUCTION dence of vadose activity. Morgan (1950) suggested that the ori-
gin of Cave of the Winds was associated with circulation of
Manitou Springs, which is located 10 km west of Colorado CO2-rich groundwater through fractures and joints formed dur-
Springs (Fig. 1), is well known for the mineral springs that issue ing the Laramide Orogeny. Bianchi (1967) indicated that the
from several locations within the city (Fig. 2). Cave of the Cave of the Winds was exposed during an erosional event
Winds, which is 1.5 km north of Manitou Springs, was first dis- related to oscillating climatic conditions during the Pleistocene.
covered by Arthur B. Love in the early 1870’s. In 1880 and
1881 large portions of the cave became known from efforts of STOP 1: CAVE OF THE WINDS
George W. Snider, who dug open many of the sediment-filled
passages; soon thereafter the cave was commercialized. Strieby At Cave of the Winds there are several geological features
(1893) was the first to suggest that Cave of the Winds and the that will be looked at and discussed: the general geology of the
springs of Manitou Springs were genetically related. He corre- region, the Nussbaum Alluvium, and sediments in Cave of the
lated red clay found in limestone cavities near the springs to the Winds. For those using this guide as step-by-step guide should
red clay in Cave of the Winds. He also presented hypotheses note that the sequence of the stops is not important because all
regarding the origin of the CO2 and mineral water that were of the stops are within a few kilometers of each other. The only
remarkably sound considering the paucity of available data. In thing that should be remembered is that the process that is tak-
his famous paper, “Features of Limestone Caverns,” Bretz ing place under Manitou Springs is the same process that
(1942) states that Cave of the Winds is phreatic with no evi- formed Cave of the Winds millions of years ago.
Luiszer, F. G., 1999, Field trip to Manitou Springs, Colorado, with specific emphasis on the sediments of Cave of the Winds and their relationship to
nearby alluvial deposits and spring sediments, in Lageson, D. R., Lester, A. P., and Trudgill, B. D., eds., Colorado and Adjacent Areas: Boulder, Colorado,
Geological Society of America Field Guide 1.
61
62 F. G. Luiszer
The Cave of the Winds is developed for the most part in the
Ordovician Manitou Formation, composed primarily of
dolomitic marine limestone. In a few places in Cave of the
Colorado Winds and in other nearby caves, cave passages can be found
Denver that extend into the overlying Mississippian Williams Canyon
Formation and in even fewer places one to two meters into the
El Paso Mississippian Leadville Formation. Most of the cave passages
County are developed along joints and fractures that were produced
during the early Laramide Orogeny. Some of the cave passages
are developed along Pennsylvanian paleokarst features, which
25
are evident along the commercial trail. The cave passages tend
Cave of the Winds to have larger dimensions in upper and lower portions of the
Manitou Formation where the bedrock has a higher calcite con-
24 tent (Table 1).
Colorado Springs
Geology of Manitou Springs region
Manitou Springs
The springs of Manitou issue from, and Cave of the Winds
25 is hosted by, limestones of a block of Paleozoic sediments (Fig.
3). Historically, this block has been called the Manitou Embay-
Figure 1. Location of study area. ment. The Ute Pass Fault, a reverse fault that is the southwest
boundary of the embayment, juxtaposes Precambrian crys-
talline rocks and Paleozoic sediments. The northwest boundary
Downstream
anyon
Fountain Creek
Creek
k U. S. 24
ree
tonC
x
Ru Fountain Creek
Ruxton Avenue Twin Spring Manitou Avenue
Ruxton Creek
Stratton Spring
Shoshone Spring
Cheyenne Spring
Iron Geyser
Navajo Spring 0.0 0.5 1.0 km
Big Chief SpringOuray Spring
Figure 2. Location map of springs, wells, streams, and other sampling sites.
Field trip to Manitou Springs, with emphasis on Cave of the Winds sediments 63
between the Precambrian crystalline rocks and Paleozoic sedi- time the springs that are now located in Manitou would have
ments underlies an erosional surface. The eastern boundary is been located along the paleo-Fountain Creek above Cave of the
the Rampart Range Fault. The Paleozoic sediments in the block Winds. Subsequent discussion will show that there is a corro-
generally dip gently to the southeast. The 1.05 Ga Pikes Peak sive mixing zone below the springs of Manitou. Apparently, in
Granite and 1.7 Ga metamorphics crop out to the south, west, the past, below the paleo-Fountain Creek, there was a similar
and north of and underlie the embayment. The sediments in the mixing zone, which was dissolving Cave of the Winds. Aminos-
embayment consist of limestone, siltstone, sandstone, conglom- tratigraphy of snail shells present in the Nussbaum and other
erate, and shale (Fig. 3). Almost all of the subsurface water flow radiometrically dated alluvial terraces in the Manitou Springs
in the embayment occurs in the very permeable Manitou and area was used to date the Nussbaum Alluvium at ~1.9 Ma
Williams Canyon Formations and to a lesser extent the (Luiszer, 1997). The age of the Nussbaum Alluvium was then
Leadville Formation. used to calibrate the magnetostratigraphy of the sediments in
The origin of Cave of the Winds is very dependent upon the Cave of the Winds.
geometry of the Paleozoic sediments, the faults and Fountain
Creek. The Rampart Range Fault, which juxtaposes permeable Sediments at Cave of the Winds
limestone beds and impermeable shale, creates a very effective
barrier to the east. This barrier prevents any water that enters There are two categories of sediments at Cave of the
the limestone beds of the embayment from draining out to the Winds: chemical and detrital. The chemical sediments include
east. The northwestern exposed up-dip side of the embayment flowstone, stalactites, stalagmites, and helictites. While most of
acts as a collection area that channels meteoric water into the these speleothems consist of calcite, the helictites and frostwork
limestone beds. The limestone beds along the Ute Pass Fault act consist mostly of aragonite. Gypsum is also found in the cave in
as a collection gallery that channels ascending CO2-rich min- the form of selenite needles, flowers, crusts, and starbursts. The
eral water into the embayment. The water that enters the above speleothems are commonly found in caves and are not
embayment drains out at the lowest point possible, where Foun- indicative of early speleogenesis. The manganese- and iron-rich
tain Creek intersects the limestone. Thus, we find all of the sediments present throughout Cave of the Winds, which are evi-
modern springs located along Fountain Creek in Manitou dence of the mixing zone that existed at the cave, are much
Springs. more important for understanding the genesis of the cave. These
sediments are especially prominent in Thieves Canyon (Fig. 4).
Nussbaum Alluvium Though not as obvious, these same sediments can be seen along
the tourist trail in the Guides Rest area (Fig. 5) and outside of
The presence of the Nussbaum Alluvium, which caps the the cave on an outcrop in the northwest corner of the upper
ridge east of Cave of the Winds, indicates that in the past Foun- parking lot. Evidence of the modern counterpart can be seen in
tain Creek flowed to the east over Cave of the Winds. At that Manitou Springs, where Fe- and Mn-rich sediments are being
Figure 3. Geologic map of the Manitou Springs area with suggested extent of mixing zone.
Field trip to Manitou Springs, with emphasis on Cave of the Winds sediments 65
deposited as a result of the mixing of oxygen-rich near-surface upward sequence of silt, sand and gravel, which in parts of the
meteoric and metal-rich deep-seated waters. cave can be up to a few meters thick. Magnetostratigraphy of
Another chemical sediment that is apparently related to the the sediments in the Grand Concert Hall (Luiszer, 1997) indi-
mixing zone is calcite spar that is found with the Fe- and Mn- cates that major dissolution at Cave of the Winds started ~4.5
rich sediments (Fig. 4). The co-precipitation of these minerals Ma ago and that major detrital sedimentation ceased ~ 1.5 Ma
was probably a function of the depth of the mixing zone. Appar- ago (Fig. 7). In the last 1.5 Ma detrital sedimentation at the cave
ently the CO2-rich and the metal-rich waters mixed at depth has been limited to the introduction of sediments near the cave
where the hydrostatic pressure was high enough to keep the entrances from rainstorms and snowmelt. In this same interval
CO2 in solution. As the water ascended to the paleo-springs, the the chemical sedimentation has been dominated by the forma-
decreasing pressure allowed the CO2 to outgas. The resultant tion of the generic speleothems consisting of calcite, aragonite
pH increase caused the co-precipitation of the calcite and metal and gypsum.
oxides. Carbon dioxide bubble trails in Thieves Canyon and
near the Guides Rest are evidence of CO2 outgassing (Fig. 5). THE MANITOU SPRINGS STOPS
To the chagrin of cave management and explorers alike is
the detrital sediments that fill many of the passages to the ceil- The stops in Manitou Springs will allow participants to
ing throughout the cave. The oldest of these sediments is debris look at many of the springs with discussions that will empha-
from the solution of the limestone. This bed is normally only a size how the mixing zone located below Mantiou Springs con-
few centimeters thick, however, a ~10-cm thick bed of the trols the content of the different springs. The calcite, Mn-rich
debris does occur in Thieves Canyon (Fig. 4). Fe- and Mn-rich and Fe-rich sediments that precipitate from different springs
sediments, which represent an early stage of the mixing zone, and their relation to the sediments present at Cave of the Winds
overlie the solution debris. Overlying the Fe- and Mn-rich sed- will also be looked at and discussed.
iments is a thick sequence of finely bedded clay (Fig. 4). The
shear volume of the clay found in the cave along with the clay STOP 2: THE IRON SPRINGS
mineralogy indicates that the cave clay was derived from
nearby soils (Fig. 6), which were eroded and transported into Below Manitou Springs two major types of water are mix-
the cave by surface streams. Overlying the clay is a coarsening ing, one is a deep-seated, CO2-rich mineral water, the other a
Detrital clay
Interlayered solution
debris, detrital clay,
and manganese oxide.
Manganese oxide
(hollandite)
Solution debris
Snider Hall
Snider Pit
A
Hole 5
Hole 1
NT
Grand Concert Hall
Hole 3
Hole 4
Hole 6
A’
Heavenly
Hall
Silent Splendor
Manitou Grand
Cavern Entrance
(Sealed)
Ordovician
Avenue
Limestone Natural
sampling site Entrance
Meters
0 12.2 24.4
0 40 80
Feet Tunnel Entrance
Iron and manganese
oxide sampling site Thieves Canyon
Passage below
main cave (in gray.
Passage above
main cave (in gray).
near-surface meteoric water. Because of the nonlinear nature of major waters type can be further broken down into four distinct
the solubility of calcite in water with varying CO2 content, the sources: Williams Canyon Creek and subsurface waters that
mixing of waters that contain different amounts of and that are have chemistry similar to the Iron Geyser, the Cave of the
saturated with dissolved calcite results in a solution that is Winds Spring and the Seven Minute Spring (Luiszer, 1997).
undersaturated with calcite. The undersaturated water below There are several lines of evidence that indicate that the
Manitou Springs is dissolving ~70 tonnes of limestone every Iron Geyser represents water that has been modified solely by
year. Study of the springs at Manitou has shown that these two contact with the Pikes Peak Granite. The elements that are
Field trip to Manitou Springs, with emphasis on Cave of the Winds sediments 67
relatively abundant in the granite are sodium, potassium, geothermometer indicates that the water issuing from the Iron
lithium, iron, and silicon (Table 1). Accordingly, the water of Geyser obtained a temperature of 126 °C. Assuming an aver-
the Iron Geyser contains elevated amounts of all of these ele- age geothermal gradient of 30 °C/km, the calculated temper-
ments (Table 2). The high concentrations are probably related ature is equivalent to a depth of ~4 km. Radio carbon dating
to the elevated temperature at which the water-rock interac- indicates that some of the water issuing from the downtown
tion took place and the extended period of time that the water springs can be up to ~30 ka. The Iron Geyser may be much
was in contact with the rock. The quartz-with-no-steam-loss older, because the downtown springs are actually a mixture
68 F. G. Luiszer
Declination
Meters
Floor
270
180
-90
90
Angular limestone clasts
0
up to 50 cm contains silt,
Subchrons
clay, small (bat) and large
0
Time (Ma)
bones near base.
Polarity
Polarity
Chrons
Epoch
Appears to be artificial fill.
Flowstone
1.5
1 Brown, laminated,
micaceous silt contains
clay intraclasts.
Pleistocene
Brown, laminated,
Olduvai 2 micaceous silt interbedded
with mottled red clay.
Mostly silt near top and
clay near bottom. Beds
MATUYAMA
2.0 dip 20 degress east.
Reunion 3
2.5 4
Reddish brown clay
Figure 7. Paleomagnetic correlation contains red, brown,
purple, green and blue
and stratigraphy of Grand Concert Hall 5 mottles.
Hole 5. Paleomagnetic time scale
adapted from Harland and others
GAUSS
3.0
(1982).
Kaena 6
Mammoth
White clay
7
Pliocene
3.5
Cochiti 9
of near-surface modern water and water similar to the Iron and lead content of the iron is related to the ability of iron oxide
Geyser. to absorb these and other metals as it precipitates.
All of the iron springs are actively precipitating iron oxide
sediments. This is a result of the oxidation as the nearly oxy- STOP 3: THE WESTERN SPRINGS
gen-free spring water contacts the atmosphere and absorbs oxy-
gen. The iron-rich sediment contains a fairly high amount of The western springs include the Ute Chief Magnetic,
lead and arsenic, which is similar to the sediments at Cave of Gusher, Ute Chief, and the Creighton Springs. These springs
the Winds (Table 3). Apparently, water with a composition sim- have the lowest total dissolved solids (TDS) of any of the
ilar to the Iron Geyser is entering the mixing zone at depth springs in Manitou because they are composed of a large
below Manitou Springs. When this water mixes with the oxy- amount of near-surface waters. The high nitrate content of the
genated near-surface meteoric water iron oxide is precipitated western springs, especially the Gusher Spring (Table 2), indi-
in the dissolving cave system. This is why most of the other cates that Williams Canyon Creek is one of the near-surface
springs in Manitou have a low iron content. The high arsenic waters. Dye tracing has shown that it takes ~85 days for dye to
Field trip to Manitou Springs, with emphasis on Cave of the Winds sediments 69
600
7-minute East Spring
500
Eastern
Chloride (mg/L)
Springs
400
oxide layers intercalated with calcite that is precipitating THE MIXING ZONE
from the water that issues from the Shoshone Spring is more
evidence suggesting that manganese oxides are probably pre- All of the inputs and outputs of the Manitou Springs mix-
cipitating in the mixing zone. This correlates with the man- ing zone have either been measured directly or have been cal-
ganese sediments found in Cave of the Winds. The high culated (Luiszer, 1997). The mass balance of the system
manganese content of the Shoshone Springs has been attrib- indicates that calcium and magnesium are being added to the
uted to the precipitation and dissolution of manganese near water in the mixing zone from mixing corrosion. The calcula-
the boundary between the hypolimnion and epilimnion in the tions indicate that ~71 tonnes of limestone are being removed
mixing zone. every year from beneath the city of Manitou Springs. The mass
balance also indicates that iron and manganese are being lost in
STOP 5: THE EASTERN SPRINGS the mixing zone. The Cave of the Winds and the sediments
found within the cave are strong evidence that the processes
One of the four mixing end members appears to have a taking place today under Manitou Springs are the same
composition similar to the 7-minute East Spring. The elevated processes that took place at Cave of the Winds some 5 m.y. ago.
amounts of sulfate, chloride, and boron (Table 2), suggest that
the 7-minute East Spring has a different evolution than the other REFERENCES CITED
end members. Sulfate, chloride and boron are anions that are Bianchi, L., 1967, Geology of the Manitou-Cascade Area, El Paso County, Col-
commonly associated with marine sediments, suggesting that orado with a study of the permeability of Its crystalline rocks (M.S. The-
sis): Golden, Colorado School of Mines.
the Eastern Springs may have been modified by rock-water Bretz, J. H., 1942, Vadose and phreatic features of limestone Caverns: Journal
interaction with marine sediments. The low sulfur and chlorine of Geology, v. 50, no. 6, part 2, p. 675-811.
content of the marine limestone of the Manitou and Leadville Harland, W. B., and others, 1982, A geologic time scale: Cambridge, Great
Britain, Cambridge University Press, p. 66.
Formations removes these formations from consideration. Hawley, C. C., and Wobus, R. A., 1977, General geology and petrology of the
Another rock with which water could be interacting is the Precambrian crystalline rocks, Park and Jefferson Counties, Colorado:
Gleneyrie Member of the Fountain Formation, a marine shale Geological Survey Professional Paper 608-B, 77 p.
Luiszer, F. G., 1997, Genesis of Cave of the Winds, Manitou Springs, Colorado
that overlays the limestone beds. Another possibility is the (Ph.D. thesis): Boulder, Colorado, University of Colorado, 122 p.
Pierre Formation, a marine shale that abuts the Rampart Range Morgan, G. B., 1950, Geology of Williams Canyon area, north of Manitou
Fault east of Manitou Springs. More study would be needed to Springs, El Paso County, Colorado (Masters thesis): Golden, Colorado
School of Mines, 80 p.
enable the assignment of the marine influence to either of these Strieby, W., 1893, The origin and use of the natural gas at Manitou, Colorado:
marine shales or other unknown sources. Colorado College Studies, v. 4, p. 14-36.
Printed in U.S.A.
Geological Society of America
Field Guide 1
1999
Daniel R. Muhs
U.S. Geological Survey, MS 980, Box 25046, Federal Center, Denver, Colorado 80225, United States
James B. Swinehart
Conservation and Survey Division, Institute of Agriculture and Natural Resources, University of Nebraska, Lincoln, Nebraska
68588-0517, United States
David B. Loope
Department of Geosciences, University of Nebraska, Lincoln, Nebraska 68588-0304, United States
John N. Aleinikoff and Josh Been
U.S. Geological Survey, MS 980, Box 25046, Federal Center, Denver, Colorado 80225, United States
Muhs, D. R., Swinehart, J. B., Loope, D. B., Aleinikoff, J. N., and Been, J., 1999, 200,000 years of climate change recorded in eolian sediments of the
High Plains of eastern Colorado and western Nebraska, in Lageson, D. R., Lester, A. P., and Trudgill, B. D., eds., Colorado and Adjacent Areas: Boulder,
Colorado, Geological Society of America Field Guide 1.
71
72 D. R. Muhs et al.
EXPLANATION
Eolian
sand
NE
DU
HB, Hoover blowout (STOP 5); BH,
AN
Bignell Hill (STOP 6); NP, North Platte;
G
WDF
R
O
.M
O, Ogallala; BC, Eldred Camp and Blue HB
FT
FM W
Creek (STOPS 7 and 8). A
er
40° WRAY
Riv
DUNE BI KANSAS
DENVER FIELD
tte
Pla
uth
So
39°
Geochronological studies indicate that the uppermost loess Maat and Johnson (1996), the age of the Gilman Canyon For-
deposits on the Great Plains span the last interglacial-glacial mation is ~40,000 to ~22,000 14C yr BP. The Gilman Canyon
cycle (Fig. 2). Most recent age estimates of Great Plains loesses Formation, therefore, corresponds in time to the mid-Wisconsin
have been from localities in Nebraska. At Eustis, Nebraska, there interstadial period and has its equivalent in the deep-sea record
are numerous pre-Loveland loesses and intercalated paleosols, as oxygen isotope stage 3 (Fig. 2). Charcoal from spruce (Picea),
and a carbonate nodule from the Btk horizon of the youngest as well as bone, snails, and detrital organic matter found within
well-developed paleosol below the Loveland Loess gives a U- Peoria Loess give ages ranging from ~21,000 to ~10,000 14C yr
series age of 184,000 ± 5000 yr (analysis by B.J. Szabo, com- BP (Wells and Stewart, 1987; Martin, 1993; May and Holen,
municated to D.R. Muhs), which is a minimum-limiting age for 1993; Feng and others, 1994; Maat and Johnson, 1996), and TL
this buried soil. Thermoluminescence (TL) ages that average dating of Peoria Loess in Nebraska gives ages ranging from
about 163,100 yr have been made on the Loveland Loess itself, ~24,000 to ~12,000 cal yr BP (Pye and others, 1995; Maat and
also at Eustis (Maat and Johnson, 1996). The TL ages and the Johnson, 1996). Direct dating of probable Peoria Loess at a
underlying U-series age indicate that Loveland Loess could have locality in eastern Colorado using TL methods gives ages rang-
been deposited during the penultimate glacial period, equivalent ing from ~20,000 to ~15,000 cal yr BP (Forman and others,
to deep-sea oxygen isotope stage 6, in good agreement with TL 1995). All these ages indicate that Peoria Loess found in the
ages on Loveland Loess from the paratype locality in Iowa (For- Great Plains, as with Peoria Loess east of the Missouri River,
man and others, 1992b). The Sangamon Soil, therefore, could correlates to the late Wisconsin glacial period and deep-sea oxy-
have developed over a period from sometime after ~160,000 yr gen isotope stage 2 (Fig. 2). Maximum-limiting ages of Bignell
BP until deposition of the earliest Gilman Canyon Formation Loess are based on radiocarbon ages of organic matter from the
sediments, a timespan that may correspond to all of oxygen iso- Brady soil, and range from ~11,800 to ~8,000 14C yr BP (Mar-
tope stage 5 and perhaps part of stage 4 (Fig. 2). Based on radio- tin, 1993; Maat and Johnson, 1996; Muhs and others, 1999).
carbon ages of soil organic matter reported by Martin (1993), Direct dating of Bignell Loess using TL gives ages ranging from
May and Holen (1993), Maat and Johnson (1996) and Muhs and ~9,000 to ~3,000 cal yr BP (Pye and others, 1995; Maat and
others (1999), and TL analyses by Pye and others (1995) and Johnson, 1996).
200,000 years of climate change recorded in eolian sediments of Colorado and Nebraska 73
SPECMAP LOESS
δ18O (o/oo) STRATIGRAPHY GEOCHRONOLOGY
5900 ± 850
(TL on loess)
0 2
0 Modern soil 8800 ± 1050
1 (TL on loess)
Bignell Loess
10,490 ± 60
2 Brady Soil (12,400 cal yr BP)
25 (Radiocarbon on humic acids)
16,800 ± 2200
50 3 (TL on loess) Figure 2. Generalized loess stratigraphy
Peoria Loess
21,300 ± 2150
of the western Great Plains, with age
(TL on loess) estimates for deposits and soils and pos-
4 sible correlation to the deep-sea oxygen
75 30,770 ± 210
isotope record. Age data for western
Age (ka)
EOLIAN SAND STRATIGRAPHY IN THE CENTRAL time of eolian sand movement, but only one locality really has
GREAT PLAINS definitive evidence for this. In Texas, mid-Holocene eolian sand
is found in the stratigraphic record of dry valleys (Holliday,
The most extensive eolian sands in North America are 1989; 1995b). Elsewhere in the Great Plains, the record for
found in the central and southern Great Plains. Limited radio- mid-Holocene eolian sand deposition is scanty, and is based on
carbon ages and degree of soil development suggest that eolian indirect lines of evidence such as dunes with a degree of soil
sand sheet and dune deposition took place during the last glacial development that is greater than that on late Holocene eolian
period in Nebraska and Colorado. Late glacial eolian activity is sand, but less than that on late Pleistocene deposits (Madole,
supported by maximum-limiting ages of ~13,000 14C yr BP for 1995; Muhs and others, 1996). Although it seems likely, from
some of the largest barchanoid ridges in the Nebraska Sand independent evidence, that conditions were optimal for eolian
Hills (Swinehart and Diffendal, 1990) and evidence for eolian sand activity over the central Great Plains during the mid-
sand movement in the southwestern part of the dune field Holocene, extensive late Holocene eolian activity may have
(Loope and others, 1995), seen on this trip. In Colorado, radio- removed much of the geomorphic record.
carbon and soil evidence indicate that eolian sheet sands were Both radiocarbon and luminescence methods demonstrate
deposited over large areas during the last glacial period that eolian sands over much of the Great Plains have been active
(Madole, 1995; Muhs and others, 1996). in the past 3,000 yr (Ahlbrandt and others, 1983; Swinehart and
The mid-Holocene (~8000-5000 14C yr BP) has long been Diffendal, 1990; Madole, 1994, 1995; Holliday, 1995a, 1997a,
considered to be a dry period in central North America (Webb 1997b; Forman and others, 1992a, 1995; Loope and others,
and others, 1993), and would seem to be an optimum time for 1995; Muhs and Holliday, 1995; Arbogast, 1996; Muhs and
eolian sand movement. However, there are actually few records others, 1996, 1997a, 1997b; Wolfe and others, 1995; Stokes and
of paleoclimatic conditions from the Great Plains itself during Swinehart, 1997). In addition, most of these studies have strati-
this period, and there is not widespread evidence for mid- graphic data indicating multiple periods of eolian activity in the
Holocene eolian sand activity in the region. Radiocarbon ages late Holocene. The number of radiocarbon ages and their ana-
from a few localities in the Nebraska Sand Hills indicate some lytical uncertainties do not yet make it possible to test the
probable mid-Holocene eolian activity (Loope and others, hypothesis of regional synchroneity of activity. However, these
1995; Stokes and Swinehart, 1997). In Colorado, Forman and observations indicate that, contrary to earlier beliefs, eolian
co-workers (Forman and Maat, 1990; Forman and others, sands in this region can be active under an essentially modern
1992a, 1995) infer that the mid-Holocene was an important climatic regime.
74 D. R. Muhs et al.
103°45'
Kp Qes1 EXPLANATION
Ql
Qes1 Qal
Qal Floodplain alluvium (Holocene)
Qt2
Qes2 Eolian sand (late Holocene)
FOR LD
T FIE Dune crest
MORGAN DUNE
U U U UU
U
UUU
U
UU
U U
U
U U
U
U
U U
U
0 5
U U
U
U
U U U
U
UU U
U U U U UU U
U U UU UU UUU U U
Qes2 Qes2
U U
U
U
U U U U UU
U
U
U KILOMETERS
U
U U UU U U
UU
U
U
UU
U
U U U
UUU
U
U U U
U
UU U
UU
U
UU
U
U
UU
UU
Figure 3. Simplified geologic map of the Fort Morgan, Colorado area (STOP 1), based on unpublished aerial photo-
graph interpretation and field mapping by D.R. Muhs, and soil survey data of Spears and others (1968).
BETWEEN DENVER AND STOP 1 tains near Colorado Springs and joins the North Platte River
just east of the city of North Platte, Nebraska. Because the
Eolian deposits can be seen shortly after passing the town of river figures prominently in the origin of both eolian sand and
Hudson on I-76. East of Hudson, at the Kersey Road exit, there is loess in the region, our first stop will be a short landscape
a cut made for a railroad visible on the right. The sediments view of the South Platte River near Fort Morgan. Fort Morgan
exposed in this cut have been studied by Forman and others itself and much of I-76 in this area are built on what has been
(1992a, 1995) and Madole (1995). Between Hudson and Keenes- called the "Broadway" (Scott, 1978), "Kersey" (Holliday,
burg, the landscape you see is covered with 2-3 m of late Wiscon- 1987) and "Qt1" (Muhs and others, 1996) terrace, of probable
sin (Peoria) loess. A few miles east of Keenesburg, as you climb late-glacial age. It is as much as 8 km wide in places (Fig. 3).
the hill, the Fort Morgan dune field will be visible, on both sides of To the south, compound parabolic dunes of the Fort Morgan
I-76. The landforms here are low-relief parabolic dunes and eolian dune field, with their minimally developed soils (A/AC/C pro-
sand sheets, and can be easily distinguished by the abundance of files) are visible and overlie this terrace. To the northwest and
Artemisia (sage) cover and frequent blowouts. The soils here northeast, more subdued dunes and sand sheets of the Sterling
belong mostly to the Valent series, and have simple A/AC/C pro- dune field, with their relatively well developed soils
files. Between mileposts 45 and 46, just before entering Roggen (A/Bt/Bk/C profiles), can be seen. Immediately north of Fort
(exit 48), partially active dunes can be seen on both sides of the Morgan, the highest part of the landscape is an alluvium-man-
highway, but particularly on the north side. These dunes were tled upland overlain by loess.
active during the 1930s drought; just after passing through Roggen, Isotopic analyses (discussed at STOP 4) indicate that
there is a good view to the north of other dunes that were also South Platte River sediments were a contributing, but not sole
active in the 1930s and are barely stable now. After passing by the source of loess in eastern Colorado (Aleinikoff and others,
town of Wiggins, I-76 traverses part of the Broadway-Kersey-Qt1 1999). Geochemical and isotopic analyses demonstrate that
terrace (see discussion below); just outside of Wiggins, the terrace the most likely source of sediment for both late Pleistocene
can be viewed to the north, eolian sand to the south. and Holocene eolian sand in the Fort Morgan and Wray dune
fields (Fig. 1) was the South Platte River (Muhs and others,
STOP 1: SOUTH PLATTE RIVER NEAR 1996).
FORT MORGAN Stratigraphic and radiocarbon studies by Madole (1994)
show that the most recent episodes of eolian sand movement
The major drainage for northeastern Colorado is the in the Fort Morgan dune field occurred in the past ~1500 yr
South Platte River, which heads in the Colorado Rocky Moun- (Fig. 4), and helped build the compound parabolic dunes
200,000 years of climate change recorded in eolian sediments of Colorado and Nebraska 75
FRIEHAUFS HILL
MILLIRON
DRAW
BIJOU
1m OUTLET
HILLROSE
1370 ± 80
Eolian
Paleosol
sand
shown in Fig. 3. Stratigraphic studies through augering show STOP 2: GEOMORPHOLOGY OF THE WRAY DUNE
that the late Holocene parabolic dunes are underlain by eolian FIELD
sheet sands that also occur at the surface in interdune areas
and have well-developed soils (Muhs and others, 1996). These This stop is an overview of landforms in the Wray dune
older sheet sands have maximum-limiting radiocarbon ages of field, the largest eolian sand body in Colorado and southwestern
~27,000 yr BP, suggesting they were probably deposited dur- Nebraska. Sediments in this dune field range from medium to
ing the last glacial period. fine sand, and show a general northwest-to-southeast fining,
consistent with dune orientations that indicate northwesterly
BETWEEN STOPS 1 AND 2 winds at the time of formation (Fig. 5). At this stop, we are near
the northern edge of the Wray dune field, but the landscape here
From the city of Brush (immediately east of Fort Mor- has some of the finest examples of parabolic dune forms in
gan) to ~25 km southwest of Julesburg, on the eastern Colorado. At the intersection of the county line and
Colorado/Nebraska state line, I-76 is built on late Holocene highway 385, we are standing on the left arm (as you look
sands of the Fort Morgan dune field. This dune field parallels downwind) of a small parabolic dune that is part of a much
the South Platte River (which provided the source sediments) larger compound parabolic dune (Fig. 6). Other small dunes
for more than 100 km northeast of Brush (Fig. 1). Near Jules- that are part of this megadune are visible to the southwest and
burg, we turn south on U.S. 385. Note the exposure of car- U.S. Highway 385 itself cuts through two arms of such a dune.
bonate-cemented sand and gravel in Miocene Ogallala Group Simple parabolic dunes are also found in the area, and are of
rocks behind the gas station on the right immediately after approximately the same dimensions as those that make up the
turning onto highway 385. Lugn (1968) and Hunt (1986) con- compound dunes (Fig. 6). Soils on both the simple and com-
sidered sediments derived from Ogallala Group rocks to be pound parabolic dunes belong to the Valent series, an Ustic Tor-
the most important source of eolian sands on the Great ripsamment with a simple A/AC/C profile, indicating relatively
Plains, a concept no longer supported, at least for dunes in young deposits, similar to those in the Fort Morgan area.
northeastern Colorado (Muhs and others, 1996). About 10 km Interdune areas in this region are occupied by eolian sand
south of the turnoff onto 385, we ascend the High Plains sur- sheets, such as the one with the center-pivot irrigation system visi-
face on Ogallala Group rocks. Scott (1978) mapped loess on ble to the southwest. These eolian sands have much better devel-
much of this surface. New field work by Muhs and others oped soils with A/Bt/Btk/C profiles, indicating a considerably
(1999) confirms the presence of this loess, but it is patchy and greater age than late Holocene. Although stratigraphic studies have
frequently only about a meter thick. About 9 km south of yet to confirm it, it is likely that, as with the Fort Morgan dune
Holyoke, we enter the Wray dune field. field, these older eolian sheet sands underlie the parabolic dunes.
76 D. R. Muhs et al.
103° 102°
Imperial, Nebraska
COLORADO drift direction
1.6
Parabolic
40°30' dune overlook 1.8
5
2.
Figure 5. Map of the Wray dune field,
0
Colorado and Nebraska, locations of 2
2.
2.
STOPS 2, 3, and 5, and mean particle
sizes for late Holocene eolian sands.
Hoover
Particle size data are plotted geograph- Blowout
ically for the first time here, but are
from Muhs and others (1996).
40°00' Wray
1.6 Quarry
1.8
KANSAS
EXPLANATION
Eolian sand
Sample locality
Paleowind
from dunes
2.0 2 0 30
Mean particle
2. 2.5
2.0
KILOMETERS size (φ)
N 102°18' Compound
parabolic
PHILLIPS COUNTY
YUMA COUNTY
Sand dune
Figure 6. Aerial photograph of a portion
of the northern part of the Wray dune sheet
field (STOP 2), showing simple and
40°26'
compound parabolic dunes and eolian
sheet sands. Note erosion of sheet sands
since last cultivation of area in circular
field.
Simple
parabolic dunes
0 1
KILOMETERS
200,000 years of climate change recorded in eolian sediments of Colorado and Nebraska 77
STOP 3: STRATIGRAPHY AND SEDIMENTOLOGY OF extremely low and not much different from the underlying sand
A PARABOLIC DUNE NEAR WRAY, COLORADO (Fig. 7), but it is easily identifiable by its darker colors (10YR
5/2, dry), the presence of roots, and abundant krotovina (infilled
Having seen the geomorphology of the Wray dune field at rodent-sized burrows). Its combination of minimal horizon
its northern end, at the next stop we will examine the stratigra- development (i.e., lack of B horizon formation) but relatively
phy and sedimentary structures in the left arm (looking down- great thickness suggests that the final stages of deposition of
wind) of a parabolic dune in a quarry exposure immediately eolian sand were slow, such that pedogenesis kept pace with
north of the North Fork of the Republican River, near the town sedimentation. The contact between the buried soil and the
of Wray, Colorado (Figs. 7, 8). The lowermost unit is a younger eolian unit above it is unusually sharp, suggesting that
medium-to-fine sand that is thinly bedded and has apparent the soil profile, thick as it is, has been truncated by later defla-
gentle (5-7°) dips to the east. About 3 m of this eolian sand is tion.
exposed and it is characterized by prominent clay lamellae (also The middle eolian unit is light brown or light yellowish
called "clay bands," cf., Gile, 1985 and "dissipation structures," brown (10YR 6/3, 6/4, dry) sand about 5 m thick and also has
cf. Ahlbrandt and Fryberger, 1980) in the lower part of the beds that dip gently (7-10°) to the east (Fig. 7). Particle size is
exposure. At least nine bands, 0.5-1.5 cm thick, can be found extremely variable and ranges from coarse to very fine sand.
over an ~80-cm depth zone. The lowermost eolian unit has a "Pin stripe" or "wood grain" ripple strata 1-3 mm thick are com-
minimally developed A/AC/C profile, but one which is more mon, particularly in the upper part of the unit. Secondary struc-
than a meter thick. Organic matter content in this soil is tures are also common. Root casts are visible in the upper beds
Cross-bedded 300
eolian sand
Depth (cm)
400
500
Figure 7. Upper: stratigraphy and structures of the left arm of a parabolic dune exposed in a quarry at
Wray, Colorado (STOP 3), and location of trench shown in detail. Lower: detail of trench stratigraphy
shown above, and organic matter content (done by the Walkley-Black method) of eolian sand and
soils shown as a function of depth. Previously unpublished data of the authors.
78 D. R. Muhs et al.
102°14' 102°13'
Qes
Qes Quarry
Qal
40°05'
er
iv
R
Republican
Fork
Nor th To
Wray
To Ql
0 1
KILOMETERS
EXPLANATION
40°03'
U.S. Qal: Alluvium
385 Qes: Eolian sand
Ql: Loess
Ql To: Ogallala
Group
Figure 8. Aerial photograph of the Wray, Colorado area with surficial geology super-
imposed and location of quarry (STOP 3). Previously unpublished geologic map-
ping by D.R. Muhs.
and suggest that some vegetation was present at the time of sed- dune suggests that all three eolian units were probably
imentation. In addition, possible bison hoof prints (cf. Loope, deposited in the late Holocene. The lowermost paleosol closely
1986) can be seen about 0.5 m above the lower paleosol and resembles those in a similar stratigraphic position in the Fort
possibly in the some of the finer-grained strata upsection. The Morgan dune field that have radiocarbon ages of ~800 to 1400
middle eolian unit is capped by a thin (10-15 cm), dark (10YR yr BP (Madole, 1994, 1995). The thick, lower paleosol, and the
5/2, dry) paleosol with an A/AC/C profile that has higher presence of root casts in primary structures above it, suggest
organic matter content than the underlying sand (Fig. 7). A that vegetation colonization with soil development and sedi-
meter or less of very young eolian sand occurs above the thin mentation were competing processes. Despite this, the volume
paleosol and could be historic, because only the simplest of A of sand that was transported in late Holocene time is consider-
horizons has developed in the upper 5 cm or so. able, given the thickness seen in section here and the size of the
The stratigraphy and degree of soil development in this dune itself.
200,000 years of climate change recorded in eolian sediments of Colorado and Nebraska 79
South of Wray, the North Fork of the Republican River has Mean particle sizes of loess in northeastern Colorado vary
cut magnificent cliffs into Ogallala Group rocks, underlain by from fine silt in the western part of the area to coarse silt in the
White River Group rocks (Fig. 8). These cliffs, which are visi- eastern part of the area (Fig. 9). This eastward coarsening is in
ble to the south of STOP 3, are mantled with loess but little or part a function of decreasing clay content to the east, which is
no eolian sand. Dune sand of the Wray dune field apparently as high as 30-40% in the western part of the region and less
did not cross the North Fork of the Republican River. However, than 10% near Julesburg (Muhs and others, 1999). The latter
loess thickness here is considerable, and is the subject of the workers suggested that much of the clay in eastern Colorado
next stop. loess may have been eroded from high-clay bedrock units such
as the Pierre Shale, which are exposed more widely in the west
STOP 4: LOESS STRATIGRAPHY AT BEECHER than they are in the east, and where deflation hollows have been
ISLAND, COLORADO mapped (Colton, 1978). This clay may have been transported as
silt-sized clay aggregates. Silt-sized grains in eastern Colorado
Loess in northeastern Colorado is the westernmost part of loess require other sources, which we discuss below.
an almost continuous loess blanket in the North American mid- A roadcut near Beecher Island, Colorado (Fig. 9), has one
continent. Loess is distributed widely but discontinuously to the of the thickest exposures of loess yet reported for eastern Col-
southeast of the South Platte River, with only isolated occur- orado. Buried soils can be found in this section and are identifi-
rences to the north of this major drainage (Figs. 1 and 9). The able on the basis of morphology, organic matter maxima,
thickest loess we have observed in eastern Colorado is about 12 CaCO3 minima, and clay maxima (Fig. 10). At this locality,
m. Thicknesses of 2-5 m are more typical, and in the northeast- grey-green calcareous clays of unknown origin (not visible in
ernmost part of the state, thin loess occurs in a patchy distribu- the roadcut, but accessible by augering) are overlain by eolian
tion on the surface of Ogallala Group bedrock. (?) sands and silts in which a strongly expressed buried soil
er
Riv
6.0φ
GE
te
Plat
RAN
φ
5.0
ide
h
ut
So
Div
φ
6.0
6.5
7.0
Fort Morgan
φ
al
inent
40°
Cont
Beecher
Island
-
Denver
T
ON
FR
+
7.0
φ
6.5
φ
39°
EXPLANATION
φ 0 50
7.0
Late Pleistocene White River Mean particle Sample
Loess glacial extent localities
Group size (φ) KILOMETERS
Figure 9. Distribution and mean particle size data for Peoria loess of northeastern Colorado, distribution of rocks of the
White River Group, and extent of last-glacial (Pinedale) glaciers on the east side of the Continental Divide. Particle
size data are from Muhs and others (1999), but are plotted geographically for the first time here; plus/minus symbols
indicate anomalous samples. Distribution of White River Group rocks from Scott (1978); extent of Pinedale glaciers
from Madole and others (1998).
80 D. R. Muhs et al.
Stratigraphy Organic matter (%) CaCO3 equivalent (%) Clay content (%)
1,480 ± 60 0 0.5 1 1.5 0 2 4 6 8 10 0 5 10 15 20 25 30
6,680 ± 60 Modern soil 0 0 0
8,160 ± 60
"Beecher Loess"
9,250 ± 60
200 200 200
Buried soil
11,090 ± 60 (Brady soil?)
11,810 ± 50 400 400 400
Depth (cm)
Peoria
Loess
600 600 600
Figure 10. Stratigraphy, AMS radiocarbon ages, organic matter content, CaCO3-equivalent content,
and clay content as a function of depth in the loess section at Beecher Island, Colorado (STOP 4).
From Muhs and others (1999).
developed. This buried soil has a subangular-to-angular blocky (~13,000 and 10,000 cal yr BP). We had originally assumed that
structure with well-expressed clay films, and up to 22% clay in this younger loess was the Bignell Loess of Nebraska and
the Bt horizon. Approximately 10 m of what is interpreted to be Kansas (seen at the type locality, STOP 6), but the radiocarbon
Peoria Loess overlies this buried soil, although a possible thin age of the deepest part of the modern soil suggests either that it
buried soil, with strong, coarse prismatic structure and 12-14% is slightly older or that deposition of Bignell Loess was time-
clay, is found at a depth of ~8 m. Between a depth of ~1.5 m transgressive.
and 3.5 m, there is a thick, buried soil with an The complex origin of silts in loess of eastern Colorado,
A1/A2/AB/Bw1/Bw2/C profile (Fig. 10) that may be equivalent alluded to earlier, can be demonstrated by examination of Pb-
to the Brady soil of Nebraska and Kansas. This buried soil is in isotopic data from Peoria Loess at Beecher Island. One likely
turn overlain by stratified eolian silt and sand with a modern source of silt-sized particles during the last glacial period in
soil, characterized by an A/Bw1/Bw2/C profile, in its upper Colorado is glaciogenic silt derived from alpine glaciers of
part. Land snails (Succinea grosvenori Lea) can be found in Pinedale age in the Front Range. Rock flour from Front Range
both the upper part of the Peoria Loess and in the younger loess glaciers likely would have been carried by the South Platte
above the uppermost buried soil. River and its tributaries to the Great Plains and deposited in
AMS radiocarbon dating of carefully extracted humic acids what are now sediments of the Qt1-Broadway-Kersey terrace.
from paleosols, following the methods given in Abbott and Such an origin was proposed for western Great Plains loess by
Stafford (1996) and reported by Muhs and others (1999), pro- Bryan (1945), Frye and Leonard (1951), Swineford and Frye
vides a chronology of loess deposition at Beecher Island (Fig. (1951), and Pye and others (1995). A less obvious source of
10). Humic acids from the upper part of the lowermost buried loess is the White River Group of Eocene-Oligocene age (Fig.
soil give a radiocarbon age of 20,520 ± 90 14C yr BP, and those 9), which is rich in silt-sized particles. These two sources can be
from the A1 and A2 horizons of the buried soil (between 1.5 distinguished from one another because the Pb-isotopic compo-
and 3.5 m depth) give ages of 11,090 ± 60 and 11,810 ± 50 14C sitions of K-feldspars are distinctly different (Aleinikoff and
yr BP (~13,000 and ~13,700 cal yr BP), respectively. Collec- others, 1999). K-feldspars in modern South Platte River silts,
tively, the radiocarbon ages indicate that Peoria loess deposition derived from Precambrian crystalline rocks of the Front Range
occurred between about 20,000 and 12,000 14C yr BP. AMS (ages of 1.0, 1.4, and 1.7 Ga), have 206Pb/204Pb values of 17.0 to
radiocarbon age determinations were also made on humic acids 17.7, whereas silt-sized K-feldspars of the White River Group
from the A, Bw1, Bw2, and upper C horizons of the modern (age of volcanism is ~34 Ma) have 206Pb/204Pb values that are
soil, in order to provide a minimum age for the youngest, sandy much more radiogenic, and range from 18.1 to 19.6. In addi-
loess at this locality. These ages are consistently younger up tion, U-Pb ages of silt-sized zircons from these two sediment
through the modern soil profile, and suggest that the youngest groups are distinctly different (Proterozoic vs. Tertiary).
loess was deposited between about 11,000 and 9,000 14C yr BP At Beecher Island, K-feldspars from Peoria Loess have Pb-
200,000 years of climate change recorded in eolian sediments of Colorado and Nebraska 81
Beecher Island
stratigraphy 206Pb/204Pb Delta 13C (o/oo)
1,480 ± 60 1,480 ± 60
6,680 ± 60 16 17 18 19 20 6,680 ± 60 -30 -25 -20 -15 -10
8,160 ± 60 0 8,160 ± 60 0
Modern soil Modern soil
9,250 ± 60 9,250 ± 60
400 400
C4 plants
Peoria Peoria
Loess Loess
600 600
800 800
Peoria Peoria
Loess 1000 Loess 1000
1200 1200 C3
20,520 ± 90 20,520 ± 90 plants
Buried soil Buried soil
1400 1400
Figure 11. Pb isotopic composition of K-feldspars, ranges of these values in modern South Platte
River sediments and sediments of the White River Group, and carbon isotopic composition of soil
organic matter and detrital organic matter in loess at Beecher Island. Pb isotope data from Aleinikoff
and others (1999); carbon isotope data from Muhs and others (1999).
isotopic compositions that span the entire range of ratios mea- erating greater amounts of outwash. The Pb-isotopic data from
sured in both possible sources (Fig. 11). This indicates that Beecher Island suggest the occurrence of an earlier cycle of
loess was derived from both silt from the South Platte River warming and cooling between peak late Pinedale glaciation and
(and therefore could be glaciogenic) and silt from the White final deglaciation (Fig. 11).
River Group. The isotopic ratios vary systematically within the Carbon isotopic composition of organic matter in buried
loess section at Beecher Island and within another section found soils and loess show changes in dominant vegetation types over
farther west, near Last Chance, Colorado (Aleinikoff and oth- the last glacial-interglacial cycle (Fig. 11). The lowermost
ers, 1999). In both sections, loess just above the ~20,000 yr old buried soil, dated at ~20,000 14C yr BP, has d13C values of -
paleosol has Pb isotopic compositions within the range of val- 19‰ to -23‰, indicating a probable mix of C3 and C4 vegeta-
ues measured in K-feldspars from the South Platte River. tion. In the overlying Peoria Loess, detrital organic matter has
Upsection, the ratios increase (to values corresponding to those d13C values ranging from about -23‰ to -25‰, indicating a
found in the White River Group) and decrease twice. Aleinikoff dominance of C3 vegetation at the time of loess fall. However,
and others (1999) suggested that under relatively cold condi- the A horizon of the ~11,000 14C yr BP buried soil and the loess
tions within the last (Pinedale) glacial period, valley glaciers of and modern soil above it have d13C values of -16‰ to -17‰,
the Front Range advanced and glaciogenic silt derived from indicating a dominance of C4 vegetation. Overall, the carbon
Proterozoic crystalline rocks was entrained within the ice, with isotopic compositions indicate a dominance of C3 vegetation
relatively little silt released to streams during short summer from about 20,000-12,000 14C yr BP, and a dominance of C4
ablation periods. Concomitantly, vegetation, which even now is vegetation after ~12,000 14C yr BP. Two vertebrate faunal local-
fairly sparse on rocks of the White River Group, decreased, ities near Beecher Island studied by Graham (1981) provide
thereby destabilizing the volcaniclastic sediments and making additional details of the environment at the time of loess depo-
them more susceptible to erosion. Although there may have sition. Fossil ungulates and rodents recovered from Peoria
been some eolian erosion directly from sediments of the White Loess at these localities led Graham (1981) to conclude that
River Group, it is more likely that reduced vegetation cover grassland was the predominant vegetation in eastern Colorado
would allow greater fluvial erosion and delivery to tributaries during the time of Peoria Loess deposition. The combined fau-
of the South Platte River. As conditions became warmer during nal and carbon isotope data indicate that, during the last glacial
the Pinedale glacial period, vegetation cover increased in east- period, eastern Colorado supported a cool grassland, perhaps
ern Colorado, including that on sediments of the White River similar to that found today in southern Canada, Montana, or the
Group, and Pinedale glaciers of the Front Range receded, gen- Dakotas.
82 D. R. Muhs et al.
METERS
STOP 5: HOOVER BLOWOUT, WRAY DUNE FIELD, 0 EXPLANATION
NEBRASKA Eolian sand
(Holocene)
Paludal or lacustrine
In the eastern part of the Wray dune field a deep deflation 5 290 ± 60 yr sediments (Pleistocene)
360 ± 60 yr
hollow called the Hoover blowout (Madole, 1995; Muhs and A/AC/C soil profile
7870 ±
others, 1997a) exposes ~8 m of eolian sand within the nose area Bioturbated 240 yr
soil profile
of a northwest-trending compound parabolic dune (Fig. 12).
The eolian sands contain two buried soils with A/AC/C profiles
13,130 ±
and are underlain by pond or lacustrine sands, silts, and clays. 295 yr
Fossil mollusk shells (Stagnicola palustris) from the paludal
sediments gives a radiocarbon age of 13,130 ± 295 14C yr B.P.
(DIC-2198) and humus from the lowermost buried soil gives a Figure 12. Stratigraphy and radiocarbon ages in eolian sands at the
Hoover blowout (STOP 5). Data from Madole (1995) and Muhs and
radiocarbon age of 7870 ± 240 14C yr B.P. (DIC-2270)
others (1997b).
(Madole, 1995). Impressions that we interpret as bison hoof
marks (cf. Loope, 1986) can be found in the middle eolian unit
and transect primary bedding structures. At a depth of ~4.5 m, Gilman Canyon Formation contains abundant evidence of bur-
within the middle eolian unit, we recovered a long bone frag- rowing in the form of krotovina. Overlying the Gilman Canyon
ment attributable to Bison. The bone gives a carboxyl age of Formation is one of the thickest (~48 m) exposures of Peoria
290 ± 60 14C yr B.P. (490-0 cal yr B.P.) and a collagen age of Loess yet described from North America. This loess has distinct
360 ± 60 14C yr B.P. (515-290 cal yr B.P.) (Muhs and others, laminae, particularly in the upper 20 m of the unit, which Feng
1997a). From all of these radiocarbon ages, we infer that eolian and others (1994) suggest may represent annual layers. At the
sedimentation began sometime after ~13,000 14C yr B.P. and top of the Peoria Loess, there is a distinctive paleosol (10YR
was episodic, based on the presence of the two buried soils. The 3/2 dry colors in the A horizon) with an A/Bw1/Bw2/BC/C pro-
two most recent episodes of eolian sedimentation began some- file called the Brady soil, first reported by Schultz and Stout
time after ~7870 14C yr B.P., and one episode apparently (1945). The Brady soil is overlain by about 2 m of Bignell
occurred within the past ~500 cal yr. Thus, the data indicate that Loess (2.5Y 6/3, dry) and a modern soil.
the eastern Wray dune field, like the Fort Morgan dune field and There have been many geochronological studies of the
the Nebraska Sand Hills, has been active in the past 1000 cal yr, loesses at Bignell Hill (Fig. 13). Maat and Johnson (1996)
and could even have been active in historic time. reported an age of 30,970 ± 780 14C yr BP on organic matter
from the oldest of the two paleosols from the Gilman Canyon
STOP 6: BIGNELL HILL, NEBRASKA Formation. Muhs and others (1999) used the humic acid extrac-
tion method of Abbott and Stafford (1996) in an attempt to min-
Bignell Hill, Nebraska, is one of the most famous loess imize contamination from both older (reworked) and younger
sections in the midcontinent of North America. The locality has carbon, and obtained an age of 40,600 ± 1100 14C yr BP for the
been studied by Quaternary geologists for more than 50 years, same paleosol. Feng and others (1994), Maat and Johnson
and results of this work have been reported by Schultz and (1996) and Muhs and others (1999) also dated the uppermost
Stout (1945), Frye and Leonard (1951), Dreeszen (1970), John- Gilman Canyon paleosol. The earlier studies reported an age of
son (1993), Feng and others (1994), Maat and Johnson (1996) 28,130 ± 610 14C yr BP; Muhs and others reported an age of
and Muhs and others (1999). It is the type locality for the Brady 30,770 ± 210 14C yr BP. Maat and Johnson (1996) reported con-
soil, developed in the uppermost Peoria Loess, and the Bignell cordant total-bleach and partial-bleach TL ages of 28,300 ±
Loess, which overlies the Brady soil (Schultz and Stout, 1945). 5100 cal yr BP and 28,200 ± 8400 cal yr BP, respectively, for
Bignell Hill contains what may be the thickest (>50 m) late the Gilman Canyon Formation. All age estimates for the upper-
Quaternary loess section in North America. At the northern end most paleosol are in good agreement with one another within
of the roadcut, a reddish-brown paleosol can sometimes be seen analytical uncertainties and possible radiocarbon-to-calendar
cropping out on the road surface. Above this, and visible in the year calibration uncertainties (which are unknown for this time
roadcut itself (to about 4.5 m above road level), is a well devel- period). However, there is a significant difference (almost
oped brown (10YR 5/4, dry) paleosol with a Btk/Bt/C profile 10,000 yr) in the ages reported for the lowermost Gilman
almost a meter thick that has developed in eolian (?) silty sands Canyon paleosol. Part of the difference may be due to the por-
(Fig. 13). Both paleosols are undated, but they may correlate to tion of the paleosol that was sampled in the two studies as well
some part of the Sangamon interglacial period. Overlying these as the extraction methods themselves.
two paleosols is the Gilman Canyon Formation, consisting of There are fewer ages of the Peoria Loess at Bignell Hill.
two loesses that both contain minimally developed but organic- Feng and others (1994) found a charcoalized twig of Picea
rich soil A horizons that are darker (10YR 5/2 and 4/3, dry) than glauca from the upper part of the Peoria Loess, ~3.5 m below
the underlying and overlying loess (10YR 6/3, 6/4, dry). The the base of the Brady soil, and reported an age of 11,880 ± 90
200,000 years of climate change recorded in eolian sediments of Colorado and Nebraska 83
Bignell Loess
Modern
soil
5900 ± 850 (TL)
METERS (covered above) 8800 ± 1050 (TL)
0
21,300 ± 2150 (TL) 9110 ± 110 2Ab
Brady Soil
10,580 ± 130 2Bw1b
(beta 14C) 2Bw2b
10,070 ± 60 2BCb
Peoria Loess
10,490 ± 60
Peoria Loess
1 (AMS 14C)
~60 METERS
2
16,800 ± 2200 (TL)
28,130 ± 610
(beta 14C)
30,770 ± 210 (covered below)
Btk1b
30,970 ± 780 SOIL A HORIZON
Btk2b (beta 14C)
Eolian silt and sand
(AMS 14C)
Btb LOESS OR
? EOLIAN SAND
Bone
ROAD LEVEL (covered below)
AT BASE OF
HILL 163,000 ± 17,000 (TL)
[by correlation to Eustis, NE]
(covered below)
Figure 13. Stratigraphy, soils and age estimates of deposits at Bignell Hill, Nebraska (STOP 6).
Stratigraphy from Maat and Johnson (1996), soils data from the present authors, and age data from
Maat and Johnson (1996) and Muhs and others (1999).
14C yr BP. This radiocarbon age translates to a calendar-year Loess and 6100 ± 900 (total bleach) and 4500 ± 2200 cal yr BP
age of ~13,600-14,200 yr BP. Maat and Johnson (1996) (partial bleach) for the upper Bignell Loess. The TL ages
reported basal Peoria Loess TL ages of 21,700 ± 3200 (total obtained by both methods are concordant and stratigraphically
bleach) and 20,900 ± 3000 cal yr BP (partial bleach). These consistent with the radiocarbon ages of the underlying Brady
same workers also reported TL ages for the uppermost Peoria soil. Collectively, all the age estimates indicate that the Bignell
Loess (~4 m above Feng and others’ sample depth) of 17,900 ± Loess was deposited during the Holocene, in agreement with
2500 (total bleach) and 12,400 ± 5000 (partial bleach) cal yr TL ages of this unit reported from elsewhere in Nebraska (Pye
BP. and others, 1995).
The Brady soil and overlying Bignell Loess have received The radiocarbon and TL results from Bignell Hill indicate
the most attention for geochronological studies at Bignell Hill, that while there is general agreement in the timing of last-
starting with Dreeszen (1970). Johnson (1993) and Maat and glacial loess deposition in various parts of the midcontinent,
Johnson (1996) redated the upper and lower parts of the Brady there are differences in detail. Loess deposition in Iowa, Illinois,
soil and obtained ages of 9110 ± 110 14C yr BP and 10,580 ± and areas eastward was, to a great extent, a function of source
130 14C yr BP, respectively. Muhs and others (1999) also sediment availability from the Laurentide ice sheet via the Mis-
redated the upper and lower Brady soil and obtained ages of sissippi and Missouri Rivers, and the timing of loess deposition
10,070 ± 60 and 10,490 ± 60 14C yr BP (11,007-12,054 and closely followed the history of movement of the ice sheet (see
12,176-12,590 cal yr BP), respectively. Maat and Johnson Grimley and others, 1998). Based on the ages from Bignell Hill,
(1996) reported TL ages of 8300 ± 1200 (total bleach) and Peoria Loess deposition in western Nebraska could have begun
10,400 ± 2300 (partial bleach) cal yr BP for the lower Bignell earlier (sometime just after ~30,000 14C yr BP) and continued
84 D. R. Muhs et al.
until significantly later (~10,500 14C yr BP) than in areas to the km long and are easily visible on Landsat imagery. Barchans,
east (Ruhe, 1983; Curry and Follmer, 1992; Grimley and others, linear dunes, parabolic dunes, dome-like dunes, and sand sheets
1998). Furthermore, eolian silt of Holocene age, such as the all are found in the Nebraska Sand Hills. Measurement of slip-
Bignell Loess, has not been reported from areas east of the Mis- faces of stabilized dunes and high-angle foreset bed dip
souri River. These observations suggest that sources of loess in azimuths indicate that paleowinds originated from the north-
the central Great Plains are unrelated to the specific dynamics west and north, similar to modern wind regimes (Warren, 1976;
of the Laurentide ice sheet, a conclusion supported by recent Ahlbrandt and Fryberger, 1980). In the next two stops we will
isotopic data found in Aleinikoff and others (1998, 1999). examine the field evidence for at least two episodes of eolian
sand movement that were dramatic enough to dam drainages
STOP 7: ELDRED CAMP, NEBRASKA SAND HILLS within the western Nebraska Sand Hills (Figs. 1 and 14).
At Eldred Camp (STOP 7), the steep groundwater gradient
The 50,000 km2 Nebraska Sand Hills area (Fig. 1) is the between Crescent Lake and the springs at the head of Blue
largest dune field (active or stabilized) in North America. Creek (1:115) contrasts with the 1:450 gradient of Blue Creek
Diverse stabilized eolian landforms are found in the Nebraska and the 1:1100 slope of the groundwater table north of the lake
Sand Hills (Smith, 1965; Ahlbrandt and Fryberger, 1980; (Fig. 14). Crescent Lake, the southernmost of hundreds of lakes
Swinehart, 1990). Barchanoid-ridge dunes are as much as 50 in the western Sand Hills, lies 3 km north of and 25 m higher
Figure 14. Western Nebraska Sand Hills showing lakes and present Snake Creek and Blue Creek
drainages. Section A-A' was drawn down the primary Snake Creek and Blue Creek valleys and across the
Sand Hills (dashed line) along a probable trace of the dune dammed valley. The position of the buried
valley is well known only in the area of Crescent and Swan lakes (Fig. 15).
200,000 years of climate change recorded in eolian sediments of Colorado and Nebraska 85
Figure 15. At the southwest margin of the Sand Hills, springs at the head of Blue Creek (lower right) emerge from a
sand dam, the southernmost of numerous sand bodies that block this major paleovalley. Configuration of the buried
valley system is based on vibracores, rotary drilled test holes, outcrops, long axes of lakes, and the dimensions of
Blue Creek valley. The high water table behind the dam creates lakes in interdune positions. None of the lakes
presently has natural surface water inlets or outlets. Note position of the intermittent stream course just east of the ter-
minal dune dam that acted as a spillway when the level of Crescent Lake was about 2 m higher.
than the springs at the head of Blue Creek (Figs. 14 and 15). Re-establishment of Blue Creek as a through-going stream
This portion of the valley of Blue Creek was not formed by would require that many dune dams be removed. If the south-
headward sapping; dune sand clearly dammed a through-going ernmost dune dam were overtopped, then the southernmost lake
extension of Snake Creek that occupied this valley (Loope and or cluster of a lakes would drain, and the head of Blue Creek
others, 1995; Mason and others, 1997). would migrate upstream several kilometers to the base of the
Although dune sand is presently mounded across the posi- next dam. A counterintuitive conclusion of our work is that a
tion of the paleovalley, an abandoned spill point for the system positive change in the water budget could cause a drastic drop
lies on Ogallala Group bedrock about 1 km east of the sand- in the water table throughout the catchment area due to over-
filled paleovalley, at an elevation 2 m higher than the adjacent topping and removal of the dune dams.
lake surface (Fig. 15). A sinuous channel below this spill point
that is cut into Ogallala Group rocks testifies to overflow during STOP 8: BLUE CREEK DUNE DAM, NEBRASKA SAND
a former lake high stand. Catastrophic drainage of the lake HILLS
immediately behind the dam did not take place because the
spillover is floored by partially lithified material, not dune sand. Across broad areas of the Sand Hills, interdune surfaces
Seepage through the dune dam must have taken place at a suffi- intersect the groundwater table, forming extensive wetlands.
cient rate to prevent massive overflow and deep entrenchment. Paradoxically, the part of the sand sea with the least precipita-
86 D. R. Muhs et al.
tion—the western Sand Hills—contains the greatest number of Based on rotary drilling, vibracores, piston cores, outcrops,
lakes. Estimates of the total number of interdune lakes in the the orientation of the long axes of lakes, and the dimensions of
region vary from 1500 to 2500. Blue Creek’s valley, we project a 1200 to 2200-m-wide, 30-m-
Eastward-flowing water courses on the dune- and lake-free deep buried fluvial paleovalley between Swan Lake and the
tableland west of the Sand Hills disappear when they reach the head of Blue Creek, 8 km southeastward (Fig. 15). Sixteen
western margin of the sand sea (Fig. 14). At the southern edge radiocarbon ages on various materials in cores from three dif-
of the Sand Hills, Blue Creek, a perennial, spring-fed stream ferent lake basins (Figs. 16, 17, 18) establish a chronological
that occupies a valley cut into Miocene bedrock, emerges from framework for correlation and interpretation (Loope and others,
dune sand (Figs. 14 and 15). This “terminal dune dam” is the 1995; Mason and others, 1997).
southernmost of scores of valley-blocking sand bodies; modern At least two distinct episodes of blockage are required to
lakes and Holocene lacustrine and wetland sediments occupy explain the history of sedimentation in Swan, Blue and Cres-
the parts of the paleovalleys not filled by dune sand. cent Lakes. Recently, we obtained five radiocarbon ages from
200,000 years of climate change recorded in eolian sediments of Colorado and Nebraska 87
basal lacustrine/wetland sediments in cores taken near Krause Basin and Colorado Plateau summarized by Thompson and oth-
ranch (Fig. 14), about 60 km north-northwest of Crescent Lake ers (1993). Lake and pollen data from the north-central U.S.
(Sweeney and others, 1998). These sites were part of the ances- also support a middle Holocene period of minimum effective
tral Blue Creek drainage basin prior to the latest Wisconsin moisture (Webb and others, 1993). Holliday (1989) gave evi-
blockage. The radiocarbon ages range from 12,160 to 12,360 dence for a prolonged drought and widespread eolian activity
14C yr BP and suggest dune blockage of the drainage just prior on the Southern High Plains between 6500 and 4500 14C yr BP,
to 12,000 radiocarbon years ago. It appears that a significant while Stokes and Swinehart (1997) presented direct evidence of
arid interval well after the last glacial maximum led to a major middle Holocene eolian activity in the northern Sand Hills
episode of dune blockage in the Blue Creek drainage. We pos- based on an optically stimulated luminescence age of ~5700 cal
tulate that the blockage at Swan Lake formed prior to 10,600 yr BP.
14C yr BP and possibly as early as 12,000 14C yr BP since the We postulate that as the Sand Hills area became increas-
10,600 14C age comes from the upper part of a 2-m-thick peat. ingly arid and vegetation became sparse, dune sand would
The dune dams that created Blue and Crescent basins were have covered a larger and larger proportion of the interfluves.
emplaced during remobilization of this part of the dune field Infiltration of precipitation into the permeable dune sand
during the mid-Holocene (about 6000 14C yr BP). would have reduced the magnitude of runoff events, and as
The second blockage event most likely reflects the middle rainfall diminished further, surface flow would have ceased
Holocene period of minimum effective moisture in the Great when subsurface flow through the unconsolidated, highly per-
88 D. R. Muhs et al.
geography, Palaeoclimatology, Palaeoecology, v. 110, p. 345-358. accumulation of lake and marsh sediments: Geological Society of
Forman, S. L., and P. Maat, 1990, Stratigraphic evidence for late Quaternary America Bulletin, v. 107, p. 396-406.
dune activity near Hudson on the piedmont of northern Colorado: Geol- Lugn, A.L., 1968, The origin of loesses and their relation to the Great Plains in
ogy v. 18, p. 745-748. North America, in C.B. Schultz and J.C. Frye, eds., Loess and related
Forman, S.L., Goetz, A.F.H., and Yuhas, R.H., 1992a, Large-scale stabilized eolian deposits of the world: Lincoln, NE, University of Nebraska Press,
dunes on the High Plains of Colorado: Understanding the landscape p. 139-182.
response to Holocene climates with the aid of images from space: Geol- Maat, P.B., and Johnson, W.C., 1996, Thermoluminescence and new 14C age
ogy, v. 20, p. 145-148. estimates for late Quaternary loesses in southwestern Nebraska: Geo-
Forman, S.L., Bettis, E.A., III, Kemmis, T.J., and Miller, B.B., 1992b, Chrono- morphology, v. 17, p. 115-128.
logic evidence for multiple periods of loess deposition during the late Madole, R.F., 1994, Stratigraphic evidence of desertification in the west-cen-
Pleistocene in the Missouri and Mississippi River valley, United States: tral Great Plains within the past 1000 yr: Geology, v. 22, p. 483-486.
Implications for the activity of the Laurentide Ice Sheet: Palaeogeogra- Madole, R.F., 1995, Spatial and temporal patterns of late Quaternary eolian
phy, Palaeoclimatology, Palaeoecology, v. 93, p. 71-83. deposition, eastern Colorado, U.S.A.: Quaternary Science Reviews, v.
Forman, S.L., Oglesby, R., Markgraf, V., and Stafford, T., 1995, Paleoclimatic 14, p. 155-177.
significance of Late Quaternary eolian deposition on the Piedmont and Madole, R.F., VanSistine, D., and Michael, J.A., 1998, Glaciation in the upper
High Plains, Central United States: Global and Planetary Change, v. 11, Platte River drainage basin, Colorado: U.S. Geological Survey Geologic
p. 35-55. Investigations Series I-2644, scale 1:300,000.
Frye, J.C., and Leonard, A.B., 1951, Stratigraphy of the late Pleistocene loesses Martin, C.W., 1993, Radiocarbon ages on late Pleistocene loess stratigraphy of
of Kansas: Journal of Geology v. 59, p. 287-305. Nebraska and Kansas, central Great Plains, U.S.A.: Quaternary Science
Gile, L.H., 1985. The Sandhills project soil monograph: Las Cruces, Rio Reviews. v. 12, p. 179-188.
Grande Historical Collections, New Mexico State University. Martinson, D.G., Pisias, N.G., Hays, J.D., Imbrie, J., Moore, T.C., Jr., and
Gosselin, D.C., Sibray, S., and Ayers, J., 1994, Geochemistry of K-rich alkaline Shackleton, N.J., 1987, Age dating and the orbital theory of the ice ages:
lakes, western Sand Hills, Nebraska, USA: Geochimica et Cosmochim- Development of a high-resolution 0 to 300,000-year chronostratigraphy:
ica Acta, v. 58, p. 1403-1418. Quaternary Research, v. 27, p. 1-29.
Graham, R.W., 1981, Preliminary report on late Pleistocene vertebrates from Mason, J.P., Swinehart, J.B., and Loope, D.B., 1997, Holocene history of lacus-
the Selby and Dutton archeological/paleontological sites, Yuma County, trine and marsh sediments in a dune-blocked drainage, southwestern
Colorado: University of Wyoming Contributions to Geology, v. 20, p. Nebraska Sand Hills, U.S.A.: Journal of Paleolimnology, v. 17, p. 67-83.
33-56. May, D.W., and Holen, S.R., 1993, Radiocarbon ages of soils and charcoal in
Greeley, R., and Iversen, J.D., 1985, Wind as a geological process on Earth, late Wisconsinan loess, south-central Nebraska: Quaternary Research,
Mars, Venus, and Titan: Cambridge University Press, 333 p. v. 39, p.55-58.
Grimley, D.A., Follmer, L.R., and McKay, E.D., 1998, Magnetic susceptibility Muhs, D.R., and Holliday, V.T., 1995, Evidence of active dune sand on the
and mineral zonations controlled by provenance in loess along the Illi- Great Plains in the 19th century from accounts of early explorers: Qua-
nois and central Mississippi River valleys, Quaternary Research, v. 49, ternary Research v. 43, p. 198-208.
p. 24-36. Muhs, D.R., Stafford, T.W., Jr., Cowherd, S.D., Mahan, S.A., Kihl, R., Maat,
Holliday, V.T., 1987, Geoarchaeology and late Quaternary geomorphology of P.B., Bush, C.A., and Nehring J., 1996, Origin of the late Quaternary
the middle South Platte River, northeastern Colorado: Geoarchaeology, dune fields of northeastern Colorado: Geomorphology v. 17, p. 129-149.
v. 2, p. 317-329. Muhs, D. R., Stafford, T. W., Jr., Swinehart, J.B., Cowherd, S.D., Mahan, S.A.,
Holliday, V.T., 1989, Middle Holocene drought on the southern High Plains: Bush, C.A., Madole, R.F., Maat, P.B., 1997a, Late Holocene eolian
Quaternary Research, v. 31, p. 74-82 activity in the mineralogically mature Nebraska Sand Hills: Quaternary
Holliday, V.T., 1995a, Late Quaternary stratigraphy of the Southern High Research, v. 48, p. 162-176.
Plains, in E. Johnson, ed., Ancient peoples and landscapes: Lubbock, Muhs, D. R., Stafford, T. W., Jr., Been, J., Mahan, S.A., Burdett, J., Skipp, G.,
TX, Museum of Texas Tech University, p. 289-313. and Rowland, Z.M., 1997b, Holocene eolian activity in the Minot dune
Holliday, V.T., 1995b, Stratigraphy and paleoenvironments of late Quaternary field, North Dakota: Canadian Journal of Earth Sciences, v. 34, p. 1442-
valley fills on the Southern High Plains: Geological Society of America 1459.
Memoir 186, 136 p. Muhs, D.R., Aleinikoff, J.N., Stafford, T.W., Jr., Kihl, R., Been, J., Mahan, S.A.,
Holliday, V.T. 1997a, Origin and evolution of lunettes on the High Plains of and Cowherd, S.D., 1999, Late Quaternary loess in northeastern Col-
Texas and New Mexico: Quaternary Research, v. 47, p. 54-69. orado, I: Age and paleoclimatic significance: Geological Society of
Holliday, V. T., 1997b, Paleoindian geoarchaeology of the Southern High America Bulletin, in press.
Plains: Austin, University of Texas Press, Austin. Pye, K., Winspear, N.R., and Zhou, L.P., 1995, Thermoluminescence ages of
Hunt, C.B., 1986, Surficial deposits of the United States. New York, Van Nos- loess and associated sediments in central Nebraska, USA.: Palaeogeog-
trand Reinhold Company, 189 p. raphy, Palaeoclimatology, and Palaeoecology, v. 118, p. 73-87.
Ingram, H.A.P., 1982, Size and shape in raised mire ecosystems: a geophysical Ross, J.A., 1991, Geologic map of Kansas: Kansas Geological Survey, Univer-
model: Nature, v. 297, p. 300-303. sity of Kansas Map M-23, scale 1:500,000.
Johnson, W.C., 1993, Surficial geology and stratigraphy of Phillips County, Ruhe, R.V., 1983, Depositional environment of late Wisconsin loess in the mid-
Kansas, with emphasis on the Quaternary Period: Kansas Geological continental United States, in Wright, H.E., Jr., and Porter, S.C., eds.,
Survey Technical Series 1, 66 p. Late-Quaternary environments of the United States Volume 1, The late
Kuzila, M.S., Mack, A.M., Culver, J.R., and Schaefer, S.J., 1990, General soil Pleistocene: Minneapolis, University of Minnesota Press, p. 130-137.
map of Nebraska: Conservation and Survey Division, Institute of Agri- Schultz, C.B., and Stout, T.M., 1945, Pleistocene loess deposits of Nebraska:
culture and Natural Resources, University of Nebraska, Lincoln and American Journal of Science, v. 234, p. 231-244.
U.S. Department of Agriculture Soil Conservation Service, scale Scott, G.R., 1978, Map showing geology, structure and oil and gas fields in the
1:1,000,000. Sterling 1° x 2° quadrangle, Colorado, Nebraska and Kansas: U.S. Geo-
Loope, D.B., 1986, Recognizing and utilizing vertebrate tracks in cross-section: logical Survey Miscellaneous Investigations Series Map I-1092, scale,
Cenozoic hoofprints from Nebraska: Palaios, v. 1, p. 141-151. 1:250,000.
Loope, D.B., Swinehart, J.B., and Mason, J.P., 1995, Dune-dammed paleoval- Smith, H.T.U., 1965, Dune morphology and chronology in central and western
leys of the Nebraska Sand Hills: Intrinsic versus climatic controls on the Nebraska: Journal of Geology, v . 73, p. 557-578.
200,000 years of climate change recorded in eolian sediments of Colorado and Nebraska 91
Spears, C.F., Amen, A.E., Fletcher, L.A., Healey, L.R., 1968, Soil survey of the last glacial maximum: Minneapolis, University of Minnesota Press,
Morgan County, Colorado: Washington, D.C., U.S. Government Print- p. 468-513.
ing Office, 102 p. Toth, J., 1962, A theoretical analysis of ground-water flow in small drainage
Stokes, S., and J.B. Swinehart, 1997. Middle- and late-Holocene dune reactiva- basins: Proceedings Hydrology Symposium 3, Groundwater, Queens
tion in the Nebraska Sand Hills, USA: The Holocene, v. 7, p. 263-272. Printers, Ottawa, p. 75-96.
Sweeney, M.R., Swinehart, J.B., and Loope, D.B., 1998, Testing the hypothesis Warren, A., 1976, Morphology and sediments of the Nebraska Sand Hills in
for latest Wisconsin blockage of streams at the west margin of the relation to Pleistocene winds and the development of eolian bedforms:
Nebraska Sand Hills [abs.]: Proceedings of the Nebraska Academy of Journal of Geology, v. 84, p. 685-700.
Sciences, 118th Annual Meeting, Lincoln, NE, p. 50. Webb, T., III, Bartlein, P.J., Harrison, S.P., and Anderson, K.H., 1993, Vegeta-
Swineford, A., and Frye, J.C., 1951, Petrography of the Peoria Loess in Kansas: tion, lake levels, and climate in eastern North America for the past
Journal of Geology, v. 59, p. 306-322. 18,000 years, in Wright, H.E., Jr., Kutzbach, J.E., Webb, T., III, Ruddi-
Swinehart, J.B., 1990, Wind-blown Deposits, in A. Bleed and C. Flowerday, man, W.F., Street-Perrott, F.A., and Bartlein, P.J., eds., Global climates
eds., An atlas of the Sand Hills: Resource Atlas No. 5a, University of since the last glacial maximum: Minneapolis, University of Minnesota
Nebraska-Lincoln, p. 43-56. Press, p. 415-467.
Swinehart, J.B., and Diffendal, R.F., Jr., 1990, Geology of the pre-dune strata, Wells, P.V., and Stewart, J.D., 1987, Spruce charcoal, conifer macrofossils, and
in A. Bleed and C. Flowerday, eds., An atlas of the Sand Hills: Resource landsnail and small-vertebrate faunas in Wisconsinan sediments on the
Atlas No. 5a, University of Nebraska-Lincoln, p. 29-42. High Plains of Kansas, in Johnson, W.C., ed., Quaternary environments
Swinehart, J.B., Dreeszen, V.H., Richmond, G.M., Tipton, M.J., Bretz, R., of Kansas, Kansas Geological Survey Guidebook Series 5, p. 129-140
Steece, F.V., Hallberg, G.R., and Goebel, J.E., 1994, Quaternary geo- Winter, T.C., 1976, Numerical simulation analysis of the interaction of lakes and
logic map of the Platte River 4° x 6° quadrangle, United States: U.S. groundwater: U.S. Geological Survey Professional Paper 1001, 45 p.
Geological Survey Miscellaneous Investigations Series Map I-1420 Wolfe, S.A., Huntley, D.J., and Ollerhead, J., 1995, Recent and late Holocene
(NK-14), scale 1: 1,000,000. sand dune activity in southwestern Saskatchewan: Current Research
Thompson, R.S., Whitlock, C., Bartlein, P.J., Harrison, S.P. and Spaulding, 1995B; Geological Survey of Canada, p. 131-140.
W.G., 1993, Climatic changes in the western United States since 18,000 Wright, H.E., Jr., Almendigner, J.C., and Gruger, J., 1985, Pollen diagram from
yr B.P., in Wright, H.E., Jr., Kutzbach, J.E., Webb, T. III, Ruddiman, the Nebraska Sandhills and the age of the dunes: Quaternary Research,
W.F., Street-Perrott, F.A., and Bartlein, P.J., eds., Global climates since v. 24, p. 115-120.
Printed in U.S.A.
Geological Society of America
Field Guide 1
1999
Léo F. Laporte
Professor Emeritus, Earth Sciences Department, University of California, Santa Cruz, California 95064
ABSTRACT
BIOGRAPHICAL BACKGROUND dom facts that he learned, including such dubiously useful infor-
mation as the densities of various materials.
Simpson was born in Chicago on 16 June 1902. He was the Simpson had just a few close friends in childhood, chiefly a
third and last child of Helen J. (Kinney) and Joseph A. Simpson, neighborhood chum, Bob Roe (Fig. 1) and his sister, Anne,
having been preceded in the world by his sisters, Margaret (1895- whom he would marry years later (Fig. 2). In old age, Simpson
1991) and Martha (1898-1984). His father was an attorney who reminisced about his childhood and noted that being more intelli-
handled railroad claims, then became involved in land specula- gent, shorter, and redheaded had guaranteed antagonism from his
tion and mining in the West, which resulted in the family’s reset- peers. He was also afflicted with an eye condition that made it
tlement in Denver while Simpson was still an infant. His mother difficult to follow the flight of a ball—a serious handicap for vir-
was born in Iowa and, owing to the premature death of her tually all sports. His father, sister Martha, and Bob Roe all
mother, had been brought up in Hawaii by her grandparents who enjoyed the outdoors, so Simpson preferred to spend much of his
were lay missionaries. Simpson’s Scots ancestry and missionary recreational time exploring the Rocky Mountain landscape,
background led to a strict fundamental Presbyterian upbringing, which undoubtedly fed his interest in natural history.
which he turned his back on by his early teens. As a boy, Simpson Simpson attended Denver elementary and high schools.
was curious about everything. He talked his parents into subsi- Despite losing a year or so, because of eye ailments and appen-
dizing his purchase of the now classic 11th edition of the Ency- dicitis, he managed to skip grades and graduated from East Den-
clopaedia Britannica, which he then read straight through. It ver high school, close to his 16th birthday. In the fall of 1918 he
became the foundation of what was to become a huge personal entered the University of Colorado at Boulder. He called Denver
research library, and he was still using it at the end of his long his home from 1903 until 1923 when he married and began his
life. As a boy he also kept a notebook in which he recorded ran- graduate studies at Yale University.
Laporte, L. F., 1999, Walking tour of George G. Simpson’s boyhood neighborhood, in Lageson, D. R., Lester, A. P., and Trudgill, B. D., eds., Colorado
and Adjacent Areas: Boulder, Colorado, Geological Society of America Field Guide 1.
93
94 L. F. Laporte
Figure 2. Simpson and his wife Anne in their New York City apartment
Figure 1. George G. Simpson (right) around age 10, just about when he in the late 1940s. They both lived on Milwaukee Street as children. It
decided he “didn’t want to give up being naughty.” He is with his boy- was only years later, after both had divorced their first spouses, that they
hood companion Bob Roe on their homemade raft. It was Bob Roe who married and remained so for more than 46 years, until Simpson died in
brought Simpson home to meet his sister Anne, when both were still 1984.
very young.
WALKING TOUR accumulated enough money through a series of odd jobs in the
neighborhood—mowing lawns, selling pop, and putting coal into
The Capitol Heights neighborhood is two miles east of the cellars. He then proceeded to read all 28 volumes straight
downtown Civic Center. A bus from downtown on the East Col- through! “I think it gave me my first conception of the world of
fax Ave. line stops at Milwaukee St. Walk four and one-half learning as a whole, my first definite feeling for organized facts,
blocks south to 1048 Milwaukee St. and my first inkling of how to go systematically about finding out
STOP 1—1048 Milwaukee St. The 1908 telephone book such facts.”1
indicates that the Simpson family was living here then, but the According to his oldest sister Margaret, when the streets
first home they had in the neighborhood was on Vine St. (nine were torn up to put in electric lighting her brother picked up all
blocks to the west), later torn down for the Botanical Park. Simp- sorts of rocks and put them in a make-shift museum in the attic.
son lived several years here while in the lower grades of elemen- He charged a penny admission, and when he later had fossils he
tary school. A favorite past-time of his was to balance along the charged two-cents.2 He also had a small menagerie that included
top of the fence that ran the length of the property on the north lizards, horned toads, turtles, and ants—the latter raised in Mason
side. Two of his household chores were to shovel coal into the jars and occasionally let loose to observe their behavior when lib-
basement and empty the furnace ashes into the backyard ash pit. erated.
The family later moved into a larger house diagonally across the STOP 3—Northwest corner of Milwaukee & 11th St. Site
street at 1069 Milwaukee St. (Continue one-half block to the of the home-made stand where Simpson and Bob Roe sold cold
north, almost to the intersection with 11th St.) lemonade, pop, and buttermilk. Proceeds went to support Simp-
STOP 2—1069 Milwaukee St. Here Simpson was allowed son’s half-share of the purchase of the Encyclopaedia Britannica.
to claim the whole attic for himself where he built a model of (Continue north a short distance—less than one-half block— on
Machu Picchu, discovered in 191l, that he copied from pictures in Milwaukee St.)
the National Geographic. From the lenses of a discarded stere- STOP 4—1110 Milwaukee St. Home of his playmate Bob
opticon he also constructed a make-shift telescope that he put on Roe and the latter’s sister Anne (1904-1991), whom Simpson
the peak of the roof to look at the stars and moon. Here, too, he would marry in 1938 in New York City. Anne has said that she
taught himself the international (“Morse”) code and would sig- cannot remember a time when she didn’t know Simpson, because
nal with a few other amateurs in the neighborhood. Planning to Bob Roe brought him home when she was about four years old
go to sea, he earned a 2nd class radio operator’s license as a teen and sick in bed, and they both read to her. They were all “great
and soon had an offer to ship out to South America but declined, chums.” They had a group called “The Eight,” chiefly organized
deciding he would be better off finishing school first. by Anne that met regularly in high school and early college years,
When Simpson was about 9 years old, he cajoled his parents usually Sunday afternoons at the Roes or Simpsons, and they
into guaranteeing half the cost of the new 11th edition of the would sing with Anne at the piano. At one point Simpson asked
Encyclopaedia Britannica (1910-11) that he saw advertised in the Anne to marry him (he was 18, she 16). After first saying yes,
local newspaper, if he in turn could come up with his half share. she called the next day and said no. Later, they corresponded on
Before long, to his parent’s surprise and financial dismay, he had and off when Simpson was at the University of Colorado. Corre-
Walking tour of paleontologist George G. Simpson’s boyhood neighborhood 95
spondence dwindled when he was at Yale and she at Denver Uni- STOP 7—Stevens Elementary School. The school is
versity, at least in part because Anne had a brief teen-age flirta- named after Edward Stevens who was principal during the years
tion with religious fundamentalism of which Simpson strongly when Simpson attended. Simpson skipped several grades and
disapproved. Anne says her intellectual interest in those days was after seventh grade, at age 11, he went directly to the East Den-
stimulated by Simpson’s conversation about such ideas as the ver Latin High School (long since torn down). One day Simp-
fourth dimension and non-Euclidean geometry. Simpson saw son was called in by Principal Stevens and asked what he did
Anne whenever he came back to Denver, and he and her brother during recess. Simpson replied that it was marbles just then, but
Bob always remained good friends.3 other times it might be trading cards or flying kites. Later,
(Return to 11th St. and go one block west to the intersection Simpson was approached by three brothers, two of whom held
with Fillmore St.) him while the third and youngest brother punched him. “I asked
STOP 5—Capitol Heights Presbyterian church. Simp- what the matter was, and they said ‘the principal called them in
son’s mother was raised by her grandparents who were lay for misbehaving and said why can’t you be regular boys like
missionaries in Hawaii, after her own mother died when she Simpson, why can’t you be more like him.’”7
was quite young. She was brought up in a strongly religious Simpson claimed that “as large as school loomed in my
setting that carried over into her adult life, for Simpson life, my real mental growth was entirely outside it and I still
claimed that as a child he “commonly attended three services feel that I never learned anything before college that I would
on Sunday at the Capitol Heights Presbyterian Church as well not have learned just as quickly and well without ever entering
as the midweek prayer meeting, and in addition had family a school….As a child I was inclined to be solitary, or to prefer
prayers and psalm recitations”4 Simpson’s father had a Pres- the constant companionship of one or two close friends to a
byterian background as well, for his own father was a Welsh larger circle of acquaintances, and in fact this is still true of
Presbyterian minister. Simpson was made a formal member me. I never cared for the orthodox sports…partly due to a
of the church at the age of nine, but soon after de-converted sense of physical inferiority, for even when as strong as my
when he decided in a fit of childish peevishness that he “did playmates I was not as skillful at games & was always a little
not want to forsake forever being naughty.” One outcome of small for my age. But it was due much more to my charac-
this decision was Simpson’s being put with half-a-dozen other ter…My favorite outdoor occupations were such things as
boys in a special Sunday school class that met in the church roller skating, bicycling, or hiking. Because of my dislike for
tower, out of sight and hearing of the more devout. The local gangs and team play and my interest in books and handiwork,
physician was put in charge of this unruly group and rather I suffered a great deal from a sense of being different from
than follow the prescribed biblical lessons told of his adven- other boys, and therefore inferior to them.…but real as this
tures in South America. Simpson also pumped the mechanism unhappiness was, it was only a small part of my life, even
on the church organ to maintain the air pressure when it was then, and my childhood was very happy on the whole. My
being played, while his father played the flute and one of his abiding interests were always books and the childish equiva-
sisters sang in the choir.5 lent of science, & from these my chief happiness & most
(Continue two blocks west on 11th St. to the intersection important development came.”8
with Clayton St.)
STOP 6—Fire House. Simpson’s route to and from school
passed the fire house and he often stopped by to talk to the fire- NOTES
men, all of whose names he knew as well as those of the horses
that pulled the fire wagons. “One of the horses was killed in the 1. Autobiographical Notes, Simpson Papers, American Philosophi-
street one day, and Stevens [the school principal] tried to keep cal Society.
us in school until the body was removed so we could not see 2. Interview with Margaret Peck Simpson, 6 June 1987, Glendale,
our friend lying dead, but I sneaked out. The wife of one of the Calif.
3. Interview with Anne Roe Simpson, 17-18 December 1985, Tuc-
firemen was found dead one day in her room, and although I of son, Ariz.
course knew that everyone died when they were old, I somehow 4. Simpson’s autobiography, Concession to the Improbable, 1978,
had not really felt that someone I knew and was no older than Yale Univ. Press, p. 20-21, 25-26.
my very active parents could also die. I had nightmares for a 5. Ibid.
while over that.”6 6. Letter from G.G. Simpson to LFL, 16 July 1980.
7. Interview with G.G. Simpson, 2 February 1979, Tucson,
(Continue two blocks west on 11th St. to the intersection Ariz.
with Columbine St. Turn right—north—to Stevens Elementary 8. Autobiographical Notes, Simpson Papers, American Philosophi-
School.) cal Society.
Printed in U.S.A.
Geological Society of America
Field Guide 1
1999
Active evaporite tectonics and collapse in the Eagle River valley and
the southwestern flank of the White River uplift, Colorado
R. B. Scott, D. J. Lidke, M. R. Hudson, W. J. Perry, Jr., Bruce Bryant, M. J. Kunk, J. R. Budahn, and F. M. Byers, Jr.
U.S. Geological Survey, MS 913, Denver Federal Center, Denver, Colorado 80225, United States
ABSTRACT
This field trip presents field evidence for Neogene evaporite tectonism, dissolution
of evaporites, and related collapse in Eagle River valley and along the southwestern
flank of the White River uplift. In the Eagle collapse center, Pennsylvanian evaporite
flowed to form anticlinal diapirs, dissolved, and disrupted a lower Miocene basaltic
plateau originally at elevations as high as 3.35 km by tilting, faulting, and sagging to
elevations as low as about 2.1 km. Also in the Eagle collapse center, the 30 x 10-km,
homoclinal Hardscrabble Mountain sank into evaporite during Triassic and Permian
collapse followed by Neogene(?) tilting and collapse, based on seismic reflection data.
Along the southwestern flank of the White River uplift in the northwestern part of the
Carbondale collapse center, parts of the Grand Hogback monocline have collapsed
northeastward toward a series of strike-elongate extrusive diapirs. The volume of evap-
orite removed from the Eagle and Carbondale collapse centers during the Neogene
(about 2,250 km3 from an area of roughly 4,500 km2) was calculated by measuring the
departure of collapsed basalts from an assumed original basalt plateau. Regional Neo-
gene uplift and incision of the Rocky Mountains, which locally began about 8-10 Ma,
probably triggered dissolution and collapse. Presently the Colorado River removes a
dissolved-solids load of about 1.4 x 109 kg per year from the two collapse centers.
Scott R. B., Lidke D. J., Hudson M. R., Perry, W. J., Jr., Bryant B., Kunk M. J., Budahn J. R., and Byers, F. M., Jr., 1999, Active evaporite tectonics and
collapse in the Eagle River valley and the southwestern flank of the White River uplift, Colorado, in Lageson, D. R., Lester, A. P., and Trudgill, B. D., eds.,
Colorado and Adjacent Areas: Boulder, Colorado, Geological Society of America Field Guide 1.
97
98 R. B. Scott et al.
Denver
40˚30'
106˚
FR
Map Area
ON
T
BL
RA
UE
COLORADO
NG
RI
VE
E
R
VA
LL
EY
WHITE RIVER
GO
DO R
RA
UPLIFT LO
RE
CO
Figure 1. Map showing geologic and State Bridge
geographic features. Field trip stops are Grand Hogback
3
RA
5 1
NG
numbered. Modified from Scott and oth- 4 E RI
VE
EA G L
E
ers (1998). Glenwood Canyon R Vail
PICEANCE BASIN Eagle
2 I-70
I-70 Glenwood HARDSCRABBLE Minturn
RIVER
MT.
Rifle
RO
Springs
EASTE AD O
LOR
E
AR
RN
CO Carbondale
NG
NG
RA
MAR FO
GIN
108˚30'
UNC OF RK
H
OMP THE
TC
AHG AN R
RE CE
WA
HIG S
HL
SA
TR
AN
Cameo D Aspen
AL
0 50
GRAND MESA
39˚ KILOMETERS
• hanging-wall strata of a large, south-dipping normal fault where dissolution and flow of evaporite control tectonism.
that dip steeply southward, whereas footwall strata that dip Not only can evaporite dissolution control the morphology of
steeply north, forming a puzzling tepee-shaped structure the land surface, style of structures, rate of tectonism, and
across the fault; nature of geologic hazards, but it also strongly affects the
• a tight syncline in a young conglomerate in contact with chemistry of ground and surface waters indicating the rate of
evaporite; modern dissolution (Scott and others, 1998).
• local anticline-syncline pairs that deform the otherwise
consistently steep-dipping Grand Hogback monocline; GEOLOGIC SETTING
• isolated low-angle normal faults along the southwest flank
of the White River uplift. Western Colorado geology provides evidence for a com-
plex history that includes Early and Middle Proterozoic tectonic
Here are some of the factors and concepts we used to solve and magmatic events, Ancestral Rocky Mountain crustal short-
these seeming enigmas: ening and uplift, Laramide crustal shortening, uplift, and mag-
matism, middle Tertiary extension and magmatism, and
• Halite and gypsum have unique physical properties: they Neogene uplift, extension, and evaporite tectonism. The conti-
readily flow under gravitational forces, they are about 10% nental crust in Colorado was produced by the accretion of mag-
less dense than most clastic sedimentary rocks, and they matic arcs and inter-arc basins 1.8-1.655 Ga, followed by
are readily soluble. These properties allow a great degree 1.45-1.00-Ga Middle Proterozoic magmatism that is interpreted
of freedom for structural interpretation compared to inter- to have been produced mostly by anatectic melting of preexist-
pretation of conventional sedimentary rocks. ing continental crust (Reed and others, 1993) (Figs. 2A and
• One of the most valuable tools to use in study of evaporite 2B). A thin Paleozoic shelf sequence covered the region during
tectonism is that of a datum plane, from which the amount Cambrian through Early Mississippian time. In Colorado and
of deformation can be estimated. Fortunately, we have such northern New Mexico, these strata and underlying Proterozoic
a datum, a 10-25-Ma basaltic plateau emplaced before the rocks were eroded from uplifts formed during shortening begin-
latest phase of evaporite deformation (Scott and others, ning in Early or Middle Pennsylvanian time. This deformation
1998; Kirkham and others, 1997b). that formed the Ancestral Rocky Mountains and basins proba-
• Finally, nowhere does rock/water interaction play a more pro- bly resulted from subduction along the southwest margin of the
found tectonic and hydrologic role than in evaporite tectonics North American plate (Ye and others, 1996).
Active evaporite tectonics and collapse in Eagle River valley and White River uplift, Colorado 99
During the Pennsylvanian and Permian, clastic sediments 900 m of synclinal subsidence of Tertiary basaltic flows at State
shed from the Ancestral Rockies accumulated in flanking basins Bridge (Fig. 1). Benson and Bass (1955) reported a young
to thicknesses of more than 5 km locally (Fig. 2). Coarse diapiric salt anticline along the Eagle River that tilted Tertiary
arkosic sediments ringed the margins of fine-grained basin basaltic flows, and Bass and Northrop (1963) attributed a tight
deposits. Locally thick sections of Middle Pennsylvanian evap- syncline in conglomerate in a canyon northwest of Glenwood
orites were deposited in sub-basins, which were at least in part Springs to “subsidence into caverns in the underlying thick gyp-
controlled by intrabasin deformation (De Voto and others, sum beds.” Mallory (1966, 1971), Freeman (1971, 1972), and
1986), probably related to continued Ancestral Rocky Mountain Piety (1981) recognized that evaporite tectonism persisted from
shortening. The original stratigraphic distribution and thickness Late Pennsylvanian to the Quaternary from investigations of
of evaporites is uncertain because flow of evaporites began diapiric evaporite anticlines, sink holes, back-tilted Pleistocene
shortly after deposition; at present active diapirs contain the terraces, and changes in thicknesses of post-evaporite units in
greatest thicknesses of evaporites. the Carbondale area. Ogden Tweto (1977) called for volumi-
Although hundreds of meters of halite have been encoun- nous flow of evaporite that began soon after denser Permian-
tered in deep wells, only gypsum, anhydrite, and gypsiferous Pennsylvanian arkosic sediments covered the evaporite. He also
siltstone crop out at the surface. The principal late Paleozoic attributed young geomorphic features near the Eagle River to
uplifts in central Colorado were the Front Range Highland, evaporite tectonics. Post-10 Ma collapse by evaporite dissolu-
which occupied much of the area of the present Front Range tion and flow modified the Laramide Grand Hogback (Murray
and Gore Range, and the Uncompahgre Highland that lay 1966, 1969; Soule and Stover, 1985; Stover, 1986; and Unruh
southwest and west of the present Sawatch Range (Fig. 1). The and others, 1993).
Central Colorado trough lies between those two uplifts and
includes two deep sub-basins, the Eagle basin to the northeast CURRENT RESEARCH
and a basin centered near Carbondale to the southwest.
Permian, Triassic, and Jurassic fluvial, eolian, and shore- Since 1993, CGS geologists have mapped a series of 7.5'
line deposits covered the area as the Ancestral Rockies gradu- quadrangles near and southeast of the junction of the Roaring
ally eroded. Intertonguing marine and continental rocks of Fork and Colorado Rivers. Clear evidence of salt dissolution,
Cretaceous age, locally nearly 3 km thick, blanketed the South- flow, and resulting collapse has been confirmed by detailed
ern Rocky Mountain region. The onset of Laramide crustal mapping in the Glenwood Springs-Carbondale area (Kirkham
shortening in the Late Cretaceous was accompanied by calc- and others, 1995, 1997a; Kirkham and others, 1996a, 1996b;
alkaline intrusions and volcanism (Tweto, 1975). In the Carroll and others, 1996; Streufert and others, 1997a, 1997b;
Piceance basin southwest of the White River uplift, the late and Kirkham and Widmann, 1997). In addition to many previ-
Paleocene and Eocene fluvial Wasatch Formation accumulated ously recognized evaporite-tectonic features, these more recent
as erosion of Laramide highlands continued (Johnson and May, studies documented abundant evidence of more subtle Neogene
1980) (Fig. 1). By middle Eocene the lacustrine Green River deformation including: complex patterns of faults ranging from
Formation accumulated in the Unita and Piceance basins. Dur- orthogonal to parallel sets that cut Tertiary basaltic rocks, syn-
ing the late Eocene and Oligocene, widespread calc-alkaline clinal sags that range from linear to arcuate shapes, structural
volcanic and magmatic activity spread across western Col- depressions that are circular, elliptical, rectangular, trough-like,
orado, concentrated in a belt from the San Juan Mountains in and half-graben in shape, local elongate closed basins that con-
the southwest to the Front Range in the northeast. Lower to tained lakes and lacustrine sediments, chaotically brecciated
Upper Miocene basalts flowed out onto an erosion surface that strata above areas of suspected collapse, and Tertiary basaltic
extended from Grand Mesa to as far east as the Gore Range rocks folded to form monoclines and broad warps.
(Larson and others, 1975). During this period, extensional In late 1995, USGS geologists began our mapping project
deformation displaced the old surfaces and produced new along the Eagle and Colorado Rivers near Interstate 70 (Fig. 1),
basins (Tweto, 1975, 1979), most notably along the trend of the and cooperative research began on topical issues recognized by
Rio Grande rift. Much of the present topographic relief in Col- both USGS and CGS mapping. Because USGS geologists also
orado resulted from differential regional uplift and dissection found areas of suspected collapse, we placed an emphasis upon
during the Neogene (Steven and others, 1997; Steven and oth- supporting research into the timing and chemistry of basalts and
ers, 1995; T.A. Steven, USGS emeritus, oral commun., 1997). study of seismic reflection lines in the vicinity of suspected col-
lapse.
PREVIOUS RESEARCH
CONCEPTUAL MODEL FOR NEOGENE EVAPORITE
For nearly a half century, evidence of salt tectonism has TECTONISM IN WESTERN COLORADO
been studied in western Colorado. Hubert (1954) assigned some
300 m of relief on Tertiary basaltic flows near Eagle to “adjust- We have created a conceptual model of evaporite tectonism
ment of the underlying gypsum,” and Wanek (1953) recognized that involves the recognition of a datum to quantify Neogene
100 R. B. Scott et al.
Meters
L. Cretaceous Dakota Sandstone 49 Medium-bedded to massive, gray ss.
Gray ss.; green, gray, and purple clyst.; gray ls.
U. Jurassic Morrison Formation 76
M. Jurassic Entrada Sandstone 18 Massive, cross-bedded, yellowish gray ss.
U. Triassic Chinle Formation 21 Reddish-brown, reddish-orange, and purplish-gray slts.,
mdst., and ss.; locally cgl. near base
L. Permian,
U. and M.
Maroon Formation 1280 Grayish-red and reddish-brown ss.slts. and cgl.
Pennsylvanian
Middle
Pennsylvanian Minturn Fm. Grit, cgl., ss., sh., and some ls. and dol.;
1900
chiefly gray, reddish-gray in upper part and near base
Middle
Belden Fm. Gray to black sh., ls., and minor ss.
Pennsylvanian 60
L. Miss. Leadville Limestone 45 Gray ls. and dol.
Upper Gilman Ss., Dyer Dol., and Parting Fm.
Devonian Chaffee Group 60
M. Ordovician Harding Sandstone 25 White, gray, and green ss. and gray sh.
Brown, red, green, and greenish-gray sandy dol. and
Upper Peerless Formation 20 dolomitic shale
Cambrian Sawatch Quartzite 67
Medium-grained white quartzite
Proterozoic rocks Migmatite and gneissic granitic rocks
Figure 2 (this and following page). Stratigraphic relations in field trip area. A. Simplified stratigraphic
column from the northeast flank of Eagle basin, modified after Tweto and Lovering (1977). B. Sim-
plified stratigraphic column from White River uplift and southeastern Piceance basin, modified after
Johnson and others (1990), Kirkham and others (1995, 1997), Bryant and others (1998), Scott and
Shroba (1997), Scott and others (1999), and Shroba and Scott (1997, 1999). Thicknesses are in meters.
tectonism, the timing of uplift, and the unloading of evaporite probably remnants of a basaltic plateau. If so, then removal of
by river incision. The resulting processes of diapirism, flow, and hundreds of meters of evaporite by dissolution could have
dissolution of evaporite cause large-scale geologic collapse allowed older basalt to collapse to near modern river levels.
(Scott and others, 1998; Kirkham and others, 1997). Both sur- Larson and others (1975) have established that much of the
face and ground waters play important roles in this tectonism. region between Grand Mesa on the southwest, the western part
of the Gore Range north of Vail on the east, the White River
Basaltic plateau datum uplift on the north, and areas southeast of Carbondale along the
Roaring Fork River was covered by basaltic flows that erupted
During the early phase of research, CGS and USGS geolo- between about 25 and 10 Ma (Fig. 1). Although relatively small
gists were trying to explain the presence of an isolated remnant volumes of basaltic volcanism continued to 4 ka (Giegengack,
of a 24-Ma basalt near the level of the Colorado River. Geo- 1962), volcanism younger than about 8 Ma appears to postdate
morphic concepts dictate that older basalts, such as this 24-Ma significant stream incision of the plateau.
basalt, should be at higher levels and only younger basalts We made the assumption that the plateau provides a datum
should be able to flow to lower elevations along young river from which to measure the amount of basalt collapse during
valleys. In an unrelated conversation, Tom Steven (USGS emer- Neogene evaporite tectonism (Scott and others, 1998). Areas
itus, oral commun., 1996) pointed out that many of the 10-25 between these plateau remnants contain 25-10-Ma flows that
Ma basalts in the region stand at remarkably constant elevations are commonly at elevations significantly lower than 3.0 km, and
between 3.0-3.4 km, well over a kilometer above the Colorado a growing body of independent detailed structural evidence sug-
River. USGS geologists realized that these elevated basalts were gested that these areas had undergone collapse. For example, 9
Active evaporite tectonics and collapse in Eagle River valley and White River uplift, Colorado 101
Meters
Pliocene or Conglomerate of 250
Bouldery cobble and pebble cgl.
Miocene Canyon Creek
Olivine basalt and trachybasalt
Tertiary
Mississippian Leadville Ls. 60 Bluish gray, coarse- to fine-grained ls. and dol.
L. Ordovician Manitou Fm. 40 Dol., ls. cgl., calcareous sh.,ss. and ls.
Thinly bedded dol., dolomitic ss., dolomitic sh.,
Upper Dotsero Fm. 30 dolomitic cgl., and algal ls.
Cambrian Sawatch White to yellowish-gray, qzt.; beds of brown sandy dol. in upper
Quartzite 165
part; local arkosic quartz pebble cgl. near base
Gneissic granitic rocks, sillimanite mica gneiss, amphibolite,
Proterozoic Rocks and felsic gneiss
km southeast of Eagle, a 22.3 ± 0.1-Ma flow has been traced underlain by evaporite, particularly where structures are similar
from an elevation of about 2.7 km to 2.3 km (Kunk and Snee, to those in areas of known collapse. A conservative estimate of
1998; Lidke, 1998). the present datum elevation of 3.0 km was used, although parts
We assumed that areas containing remnants of basalts of the uncollapsed plateau reach 3.4 km. Using these criteria,
below the datum have undergone collapse. Where evidence of a an estimated area affected by collapse was calculated to be
basalt datum is absent, evaporite dissolution, flow, and collapse about 4,500 km2 (Fig. 2). Where sub-datum basalts are reason-
probably occurred in areas that either exposure evaporite or are ably abundant, as in the Carbondale collapse center and in the
102 R. B. Scott et al.
COLORADO
Bl
u e
R
ive
r
Study Area
Vail
Glenwood
Eagle
Ea
Springs
gl
e
Ri
ve
Carbondale Eagle
r
ive
R Carbondale Collapse
Collapse
o
ad
Center Center
or
ol
R
C
oa
r in
g
Fo
Aspen
kr
Grand
R
ive
Junction
r
r
ive
R
Gunnison
N
0 50
Kilometers
Figure 3. Map showing areas affected by evaporite flow, dissolution, and col-
lapse. The two major collapse centers near Carbondale and Eagle coincide
roughly with the two areas underlain by the greatest thickness of evaporite
(De Voto and others, 1986).
central and northern part of the Eagle collapse center (Fig. 3), 165 basalt samples from the area have been analyzed for over
the basalts could be contoured. In areas of suspected collapse 40 major, minor, and trace elements by x-ray diffraction and
in the absence of basalts, contours were extrapolated following instrumental neutron activation analysis, and over 90 40Ar/39Ar
patterns of collapse. About 2,250 km3 of collapse was calcu- basalt samples have been dated. About 8 separate major groups
lated from these contours. of basalts have been recognized from data using chondrite-nor-
Where basalt flows drape over significant topographic malized La/Yb and Hf/Ta ratios. Basalts >20 Ma have La/Yb
relief, a question needs to be answered: Were basalts erupted on ratios of 16-20, basalts 10-11 Ma have La/Yb ratios of 5-8,
a nearly horizontal surface, or were they erupted on an irregular basalts 8-9 Ma have La/Yb ratios of 8-10, and basalts 3-4 Ma
topography and simply flowed downhill? K.A. Hon (USGS, have La/Yb ratios of 10-13. Two or three subgroups of each
oral commun., 1996) confirmed that the pahoehoe-type flow major group have been identified by differences in Ba, Sr, Rb,
structures in pre-10-Ma basaltic flows are indicative of flow on and Th. The differences in basalt compositions are assumed to
relatively horizontal surfaces, not on significant slopes. Also, be the results of differences in partial melting, crystal fraction-
initial paleomagnetic study of the basalts indicates that the ation and, in particular, crustal assimilation. Few flows can be
basalts were deposited on a nearly horizontal surface that was correlated farther than tens of kilometers. Although remnants
subsequently tilted about a horizontal axis. of basalt flows of statistically identical ages are found within
How far did the basaltic lavas flow? Are there only a few 10-20 km of one another, trace-element data do not support cor-
major vents or large numbers of eruptive vents on the plateau? relation. This requires that fairly closely spaced volcanic cen-
Accurate correlation of basalts, using high-field-strength trace- ters erupted magmas with different degrees of partial melting,
elemental signatures, 40Ar/39Ar geochronology, petrography, types of fractionation, or levels of crustal contamination. Thus,
and paleomagnetic data, has been used to establish both the the plateau probably was partly covered with discontinuous
stratigraphy and the continuity of basalts on the plateau. About flows from scattered vents.
Active evaporite tectonics and collapse in Eagle River valley and White River uplift, Colorado 103
Initiation of Neogene evaporite tectonism the annual load, including discharge from saline hot springs
near the town of Glenwood Springs (Cappa and Hemborg,
Both USGS and CGS geologists consider that regional 1995; Barrett and Pearl, 1978), is nearly 1.2x109 kg (Bauch and
Neogene uplift triggered more rapid downcutting in the Rocky Spahr, 1998; Butler, 1996; Liebermann and others, 1989; U.S.
Mountains that, in turn, prompted renewed evaporite tectonism Department of the Interior, 1997). These dissolved-solids loads
in the Eagle and Carbondale areas. Basaltic flows older than consist largely of Na+, Cl-, Ca+2, and SO4-2 ions. These rates do
about 10 Ma are stacked upon one another in normal strati- not include unmeasured saline ground water discharge into the
graphic order, suggesting that the plateau basalts erupted on Colorado River between Glenwood Springs and Cameo and
low-relief topography (Larson and others, 1975) and predate unmeasured loads from several streams, some of which drain
significant renewed uplift in the White River uplift area. Timing the relatively saline Green River Formation.
of uplift in the Eagle and White River uplift area probably is These data clearly indicate that evaporite dissolution is
slightly older than the 7.8-Ma age of a local basaltic flow col- active, and therefore, evaporite-induced collapse must also be
lected 7.5 km south-southwest of the junction of the Colorado active. If the rate of dissolution has been constant, if the annual
and Eagle Rivers (Kunk and others, 1997; Kunk and Snee, load at Cameo represents the total load added exclusively from
1998) because the flow rests on river gravels 200 m below older the collapse centers, and if the volume of evaporite removed is
basalts (Streufert and others, 1997b; Kirkham and others, 2,250 km3, then the length of time during which the process has
1997). Also, young (~5-Ma) apatite fission track ages have operated would be about 3.7 Ma. However, Neogene downcut-
recently been determined in the Gore Range (C.W. Naeser, ting probably began about 8-10 Ma, so presumably the rate of
USGS, written commun., 1997), suggesting late Miocene uplift dissolution has increased during this period rather than remain
in the area. Finally, Steven and others (1997) proposed late constant.
Miocene uplift in the Front Range of the Rocky Mountains.
Although local variations in the timing and degree of uplift GEOLOGIC COLLAPSE IN THE FIELD TRIP AREA
probably exist, they are all part of a regional Neogene uplift
event in the southern Rocky Mountains. This discussion will address broader observations of col-
Increased stream gradients would have increased the rate lapse in the general areas to be visited on this trip. In the fol-
of downcutting as uplift proceeded; this, in turn, removed suffi- lowing Abbreviated Road Log, more site-specific information
cient overburden to triggered local diapiric upwelling of evap- is provided.
orite in deeper river valleys. Once sufficient overburden had
been removed by stream erosion, either pre-existing diapirs or Eagle collapse center
diapirs initiated by that unloading would be activated. When the
evaporite encountered shallow ground water or surface water, Throughout much of the Eagle collapse center, the Eagle
dissolution removed material from the upper part of the diapir Valley Evaporite is widely exposed or present at relatively shal-
and the rate of removal of material increased significantly. low depths (Fig. 4). In the central part of the collapse center
Therefore, the rate of new exposure of evaporite along the crest along I-70, between the towns of Wolcott and Gypsum, evi-
of anticlinal diapirs probably rapidly increased, “unzipping” dence for Neogene salt tectonics includes large amplitude,
their tops. In both the Eagle and the Carbondale collapse cen- trough-like collapse features and linear to arcuate, discontinu-
ters, the major gypsum-cored diapiric anticlines coincide with ous, high-angle faults that locally form horsts and grabens, both
stream valleys. Replacement of dissolved evaporitic material of which deform early Miocene basaltic flows. Broad, east- to
required progressive flow of ductile evaporite from adjacent northeast-trending anticlines that follow the Eagle River proba-
highlands toward river valley diapirs, causing progressive col- bly reflect Neogene upwelling of the evaporite at depth, related
lapse of these highlands as material was removed. Irregular and to unloading from downcutting of the Eagle River (Lidke, 1998,
relatively rapid dissolution of exposed or shallowly buried evap- 1999; unpublished mapping, Mark Hudson and Richard Moore,
orite formed local sinkholes. Gypsum 7.5' quadrangle, 1998). The westernmost anticline is
clearly younger than the deformed 22.3-Ma basalts (Tweto,
Geologic significance of dissolved-solids load of streams 1977), and it is probably younger than about 8 Ma based on
timing of initial downcutting of the Colorado River.
Study of the dissolved-solids load of the upper Colorado The timing of initial uplift of the basalt plateau in the Eagle
River drainage provides a critical component of the evaporite Valley area can be estimated by dividing the elevation of
collapse story, allowing us to estimate the length of time the assumed uncollapsed basalt above the river level by the rate
process has been active. About 90 km downstream from the col- average of incision of the rivers. The elevations of basalts at
lapse centers (Fig. 1), the mean annual dissolved-solids load at Castle Peak are about 1.44 and 1.49 km above the Eagle and
Cameo is 1.3x109 kg (Fig. 1). The mean annual dissolved-solids Colorado Rivers, respectively. The rate of incision based on the
loads from individual streams upstream from the collapse cen- Lava Creek B volcanic ash just below the junction of the Eagle
ters is less than 5x107 kg. However, within the collapse centers, and Colorado Rivers is 0.14 m x 10-3 yr. (Izett and Wilcox,
104 R. B. Scott et al.
P O O
107 106 30'
P M
White River
uplift P M
Tb Tb
ER
M IV
Tb R Tb State Bridge M
P M Tb M
Tb
T
DO
T
RA
O
Castle Peak Tb
LO
CO
Figure 4. Map of the area of Stops 1, 2, O Tb
and 3 showing the generalized bedrock M
Tb
geology of the Eagle Valley area modi- O M 39 45'
E P
HORN RANCH
fied from Tweto and others (1978). T = E
ANTICLINE Wollcott
Tertiary lacustrine strata; Tb = Tertiary Tb
Stop 3 C' M P
basaltic flows; M = Mesozoic strata; P = O A'
E Tb P
Permian and Pennsylvanian strata; E = EAGLE RIVER STOP1
Eagle Valley Formation and Pennsylvan- E C Vail
Eagle E
ian Eagle Valley Evaporite; O = Paleo- Gypsum Tb A E
Tb
zoic and Precambrian rocks older than P
Eagle Valley Evaporite. Two large gyp-
B'
sum-cored anticlines are shown along the E
M P
Eagle River and a trough-like feature is P HARDS STOP 2
CRA
shown east of Eagle by a dashed line. BB
LE M
Modified from Scott and others (1998). Tb MT
39 30'
E
E
M
B
P 0 20
Tb
Tb KILOMETERS
E
M
E
M
1982). The rate of tectonic uplift of about 0.18 m x 10-3 yr. was ally within the evaporite. The isolated northeast-trending Horn
based on the Derby Peak fauna in the Flat Tops area of the Ranch anticline is interpreted to be a Neogene structure related
White River uplift (Colman, 1985). The timing of uplift based to diapiric upwelling of the evaporite in response to unloading
on the incision rate is 8.0-8.3 Ma and the timing based on uplift during Neogene downcutting by the Eagle River.
rate is 10.3-10.6 Ma, both of which are close to the 7.8 Ma age Southwest of Eagle, Tweto and others (1978) interpreted
of nearby basalts deposited on incised river gravels 7.5 km Hardscrabble Mountain to be a 30 x 10-km block of gently north-
southwest of the junction of the rivers. dipping Triassic to Cretaceous strata that has been downdropped
A northwest-trending sag, or trough-like Neogene collapse into Eagle Valley Evaporite (Fig. 4). At Stop 2 we observe the
feature (Lidke, 1998), expressed in 22.3-Ma basaltic flows 5 km northern fault contact of Hardscrabble Mountain with evaporite.
southwest of Wolcott (Fig. 5) will be visited at Stop 1. Evidence Tweto (1977) proposed that the evaporite flowed diapirically
for pre-Neogene deformation of the pre-basalt strata is seen in around the mountain. Seismic reflection data (Fig. 6) confirm his
south-dipping flows that overlie northeast-dipping Pennsylvan- conclusion and add significantly to it. Faults that drop the mountain
ian to Jurassic strata. Also, the evaporite has been intensely into the Pennsylvanian evaporite do not offset the older Paleozoic
deformed beneath the less deformed younger Paleozoic and strata. Only a small, unrelated Laramide or Permian-Pennsylvanian
Mesozoic strata. Clearly the evaporite decoupled structurally thrust fault offsets the older strata below the center of the moun-
from the overlying strata, but the genesis and timing of this tain. Evaporite deformation probably occurred at least twice, once
deformation is problematic. during deposition of the unusually thick sequence of Triassic-Per-
A few kilometers east of Eagle, linear to arcuate, north- mian State Bridge Formation and again sometime afterward, pos-
west- to north-trending faults displace basalt flows and clastic sibly during the Neogene. The unusual thickness of the State
basin facies rocks, and younger strata have either fault or intru- Bridge reflects the likelihood that the block was subsiding within
sive contacts with the Pennsylvanian evaporite (Lidke, 1999) the evaporite as State Bridge deposition occurred. The tilt of the
(Fig. 4). Many of these Neogene faults are attributed to evapor- Mesozoic homoclinal block to the north requires post-State Bridge
ite dissolution and collapse because they cannot be traced later- deformation, probably Neogene.
A A'
3 Bellyache
Ridge
Tb
K
J
TR
Kilometers
IP ee
PIPm
2
IP ee
IP e
Thickness of IP ee unknown
IP ee
0 2
KILOMETERS
North Limb
C'
SW
NE
Stop 3
2.5
Tb
Kilometers
Tb
Tb Tb
Ts
PIPm Ts Tb Q
IP e IP e
IP e
IP ee
1.5
IP ee
IP ee
0
1 Km
Figure 5. Cross sections shown in the Eagle collapse center. A. Section A-A' at Stop 1 trends northeast across a trough-like sag
developed in basaltic flows and underlying strata, which reflects removal of evaporite. Tb = Tertiary basalt flows; K = Creta-
ceous Dakota Sandstone; J = Jurrasic Morrison Formation and Entrada Sandstone; TRP Triassic and Permian Chinle and State
Bridge Formations; PIPm = Pennsylvanian and Permian Maroon Formation; IPe = Eagle Valley Formation; IPee = Eagle Val-
ley Evaporite. B. Section C-C' at Stop 3 trends northeast across the collapsed north limb of the Eagle River anticline. Symbols
are the same as for A except Q = Quaternary deposit, Ts = Tertiary sediment.
Figure 6. Seismic depth section through Hardscrabble Mountain along section B-B' visited at Stop 2.
Strata exposed along section in the structural block are the Triassic-Permian State Bridge Formation
and Triassic Chinle Formation. Note that the nominal vertical scale is in feet.
Active evaporite tectonics and collapse in Eagle River valley and White River uplift, Colorado 107
The evaporite-cored Eagle Valley anticline follows the River, about 26 km east of its junction with the Colorado River,
Eagle River where it bends to a westward course between Eagle (Fig. 4) has been identified as Lava Creek B by correlation
and Gypsum. The north limb of that diapiric anticline displays a based on its signature of high field-strength trace elements. We
10° northward dip on 22.7-Ma basalts (Larson and others, propose that 25-30 m of the underlying Eagle Valley Evaporite
1975), clearly recording Neogene evaporite tectonism. In the dissolved since 0.66 Ma collapsing Lava Creek B ash that
vicinity of Gypsum, flow remnants are folded into a series of amount.
large synclines and smaller anticlines and are cut by minor At State Bridge in the northern part of Figure 4, a large
southwest-dipping normal faults (unpublished mapping, Mark synclinal sag of basalt between Castle Peak and the mountain
Hudson and Richard Moore, Gypsum 7.5' quadrangle, 1998). north of State Bridge poses a structural puzzle. Unlike other
We will visit this deformation at Stop 3. structures in the collapsed areas, much of this structure does not
About 17 km northeast of Gypsum, Castle Peak is capped overlie known evaporite deposits. Tweto and others (1978)
by basalt at an elevation of 3,437 m and near Gypsum the basalt placed the boundary between the basin evaporitic facies and
is at an elevation of 2,135 m, recording an apparent collapse of more proximal coarser clastic facies largely south of this struc-
about 1,300 m. Several faults in this area appear to terminate ture. However, the presence of a saline warm spring in the area
laterally in Eagle Valley Evaporite. A channel filled with Neo- suggests the presence of evaporite at depth. Also the absence of
gene(?) river gravels dips gently southwest on a limb of the contractional structures in the basalts leaves the probability of
highest syncline, in the same direction as the basaltic rocks that withdrawal of evaporite by flow and dissolution as the most
appear to underlie the gravels. In this area, underlying Meso- likely explanation for the sag.
zoic and Paleozoic strata dip consistently to the northeast, sug-
gesting that deformation of the basalt also unfolded the Collapse structures along the southwestern flank of the
preexisting dip of these older rocks. These relationships suggest White River uplift
that diapirism may have already been operating to some degree
before the basalts were emplaced. Northwest of Glenwood Springs, a string of elliptical evap-
Volcanic ash identified by Izett and Wilcox (1982) as the orite diapirs intrudes Permian to Devonian rocks along the
0.66-Ma Lava Creek B (redated by Izett and others, 1992), the southwestern flank of the White River uplift (Fig. 7). South of
youngest of three (0.66, 1.27, and 2.02 Ma) far-traveled ashes the diapirs, the adjacent Maroon and Eagle Valley Formations
from the Yellowstone caldera complex (Izett and Wilcox, 1982), form bench-like sags that depart significantly from the rela-
overlies terrace gravel about 85 m above the Colorado River tively steep southwest dips of the Grand Hogback monocline.
about 4.5 km southwest of its junction with the Eagle River North of the diapirs, older Paleozoic strata and even Proterozoic
(Fig. 4) (Streufert and others, 1997a). The ash elevation is con- rocks are also deformed by evaporite tectonism.
sistent with that of other reported occurrences of the Lava North of the Colorado River at Stop 4, Canyon Creek cuts
Creek B at 85-90 m above stream level in northwestern Col- through a tight syncline in a Tertiary conglomerate (Bryant and
orado (Izett and Wilcox, 1982; Scott and Shroba, 1997). An ash others, 1998) (Fig. 8). The upright southern limb of the syncline
on or within main-stream terrace deposits along the Eagle is in contact with a narrow zone of Pennsylvanian evaporite
108˚
Tb
M O
P E WHITE RIVER UPLIFT
O
O Tb Figure 7. Map showing generalized
GR A
E P
K
ND
HO E
Creek; Tb = Tertiary basalt flows; T =
ON
GB RIFLE O
CANY
A'
39 37'30"
Pev 55
40 60 MC
Pe Pe
10 Pb Pev
50 Q MC
PPm
65 Pev 55
20 75 35
Pev 40
PPm
25 45 Pe
Q 45 Q
45
30 20 PPm
40
Tcc
80 45
30 Tcc
35
80
Tcc
80 35 STOP 4
Pe Tcc
PPm 60 60
35
85
Pe
55
40
65
55 Pev
Creek
70 80
KPQ 60 60
45
Canyon
85 80
25 15 55
40 70 70 PPm
10
20
15
25 35
Q
Q I-70
40
PPm
35
A
A'
Tcc Pev
A Pev
PPM
Pev
PPm
Pe
0 1 2 km
that, in turn, is in contact with overturned Pennsylvanian clastic area around the Eagle collapse center is estimated to have
basin facies rocks that dip steeply northward. The clast compo- begun between 7.8 and 10.6 Ma.
sition of the conglomerate suggests that the deposit is locally On the southwest flank of the White River uplift, the mon-
derived and young, probably Pliocene. Deposition of the Ter- oclinal limb of the hogback collapsed adjacent to diapiric intru-
tiary conglomerate on Pennsylvanian strata presently beneath sions of evaporite, diapiric intrusion is associated with normal
the conglomerate would require the highly unlikely geometry faults, and a low-angle normal fault in lower Paleozoic and Pro-
of nearly completely overturned Pennsylvanian rocks. Based on terozoic rocks was caused by diapiric intrusion.
locally nonconformable contacts and pod-shaped bodies of In the Eagle collapse center, 22-23-Ma basalt flows are
evaporite in the area, we conclude that the bodies of evaporite draped and faulted over an elevation change of at least 790 m,
have intrusive, not depositional, boundaries and that both flow- form graben-like depressions, and are tilted away from anticli-
and dissolution-induced collapse occurred (Bryant and others, nal crests along the Eagle River. The Lava Creek B volcanic ash
1998). The bench-like anticlinal and synclinal structures in the from Yellowstone has apparently collapsed about 25 m in the
Grand Hogback monocline south of the conglomerate near the last 0.66 m.y. Hardscrabble Mountain sank along an elliptical
mouth of Canyon Creek resulted from collapse as flow and dis- fault during Triassic-Permian deposition and probably still
solution of material undermined the monocline and allowed sinks into the Pennsylvanian Eagle Valley Evaporite.
tight synclinal sagging of the conglomerate as it collapsed into
dissolving evaporite. This explanation is similar to that pro- ABBREVIATED ROAD LOG
posed by Bass and Northrop (1963). A 10° low-angle normal
fault that places Mississippian Leadville Limestone above On our drive to the Eagle Basin, we cross the Front Range,
Cambrian Sawatch Quartzite with 250 m of dip separation is a major Laramide uplift modified by younger events, then the
probably related to the diapir complex exposed only 300 m to Blue River valley, an en echelon fault-valley extension of the
the south. Rio Grande rift, and finally the Gore and Tenmile Ranges that
About 24 km northwest of the Canyon Creek stop, another show fission-track and geomorphologic evidence of renewed
example of collapse along the Hogback monocline occurs at uplifted during Neogene time. Numerous road logs are avail-
Stop 5 near Rifle Falls (Scott and others, 1999). A bench-like able for the route along I-70 to the Eagle Basin (such as
structure similar to that at Canyon Creek is well exposed south Kirkham and others, 1996a; Reed and others, 1988). Our highly
of a pod-shaped diapir (Fig. 9). In this case, the diapir is 0.5 km abbreviated road log starts at Vail Pass; a detailed log will be
south of a major south-dipping normal fault displaying several available for attendees and for those who request it from the
hundreds of meters of dip separation. Not only does the foot- first author.
wall of the fault dip steeply southward, but also the footwall
strata locally dip steeply northward, forming a perplexing Cumulative miles
tepee-like structure. Because the north-dipping segment of the
footwall is locally restricted to areas adjacent to the diapir, we 0.0 Mileage begins on I-70 at exit 190 (Vail Pass rest
conclude that the origin of the “tepee” is related to intrusion of area). Continue westward.
evaporite across the fault into lower strata in the footwall. 20.2 Here we see an abrupt facies change from massive,
In the adjacent Horse Mountain quadrangle to the west of resistant, coarse-grained arkosic rocks of the proximal
Rifle Falls, similar bench-like structures are found (unpublished facies of the Pennsylvanian Minturn Formation into
mapping, W.J. Perry, 1998). However about 20 km to the south- siltstone of the Eagle Valley Formation that contains
east of Rifle Falls, in the northeastern part of the New Castle gypsiferous evaporite of the distal facies of the Eagle
quadrangle (Scott and Shroba, 1997), a different type of struc- Valley Evaporite.
ture seems to be associated with evaporite diaprism. In this 32.9 On the left is the landslide complex of Bellyache
case, a 2.5-km-wide overturned, northeast-dipping sequence of Ridge, expressed by hummocky topography.
the Maroon and the Eagle Valley Formations occurs south of a 33.4 At exit 157 to Wolcott, turn off I-70 and turn left under
large evaporite diapir that is exposed about 1 km north of the I-70. Proceed up Bellyache Ridge Road to enter the
New Castle quadrangle. landslide complex, which formed in the dip slope of
northerly-dipping Cretaceous Benton Shale and the
CONCLUSIONS underlying Cretaceous Dakota Sandstone and Jurassic
Morrison Formation.
We assume that 25-10-Ma plateau basalts formed a datum 37.6 At the intersection, stay on Bellyache Ridge Road.
from which the area and volume of collapse can be calculated. 40.0 At the electronic gate, enter a gated community.
The area affected by collapse in the Carbondale and Eagle col- 40.7 Stop 1. South-dipping basalt caps northeast-dipping
lapse centers is estimated to be 4,500 km2 and the volume of Mesozoic strata. The northwest view in foreground
evaporite removed is estimated to be 2,250 km3. shows dark-colored, south-dipping lower Miocene
Uplift of the basalt plateau and downcutting of rivers in the basaltic flows that overlie reddish-colored, northeast-
36
Jm dP IPe
Qs
Je IPe
PIPm
IPb
Qa
Ml
17
75 IPb
Km
PIPm
Jm
IPe
Kd
IPe
Je dP 27
N
39 40'
2Km
A
South A'
North
Qs
IPb Ml
Jm Je dP
2 IPe Dc
O-C
Kilometers
PC
IPee
PIPm
IPe
1
IPe
Figure 9. Map showing detailed geology in the Rifle Falls area and section B-B' at Stop 5. Qs = Landslide deposit,
Km = Mancos Shale, Jm = Morrison Formation, Je = Entrada Sandstone, TR P = Chinle Formation and State Bridge Forma-
tion, PIPm = Maroon Formation, IPe = Eagle Valley Formation, IPee = Eagle Valley Evaporite, IPb = Belden Formation,
Ml = Leadville Limestone, Dc = Chaffee Group, OC = Ordovician and Cambrian strata,PC = Proterozoic rocks.
Active evaporite tectonics and collapse in Eagle River valley and White River uplift, Colorado 111
dipping Mesozoic sedimentary rocks along the crest of Hardscrabble Mountain mapped by Tweto and others
Bellyache Ridge. The south-dipping flows define the (1978) suggests that initial collapse and subsidence of
northern limb of a west-northwest-trending, trough- this structure may have occurred during the Permian.
like sag defined by the basaltic flows (Figs. 4 and 5). Return to I-70.
Below, and to the south-southwest, basaltic flows in 75.8 Turn west on I-70.
the central part of the sag can be seen at lower eleva- 80.3 On the right, folds are well displayed within Eagle Val-
tions. Basalts in the central part of the sag overlie the ley Evaporite. Disharmonic folding is common within
lower part of the Eagle Valley Formation and the Eagle evaporites of the Eagle basin and reflects plastic flow
Valley Evaporite. These basaltic flows yielded within the thick deposits. Widmann (1997) demon-
40Ar/39Ar isotopic ages of about 22-23 Ma (Kunk and strated that fold axes are parallel to the long axes of
others, 1997; Kunk and Snee, 1998; Lidke, 1998). The evaporite upwellings and to coincident major
south-dipping flows on Bellyache Ridge are at an ele- drainages within the western Eagle Basin.
vation of about 2,750 m, whereas continuations of 83.0 At Gypsum exit 140, turn right on good dirt road.
these flows in the trough of the sag are at about 2,320 84.3 At the intersection, go left.
m, providing evidence for at least 430 m of collapse of 86.9 At road intersection, go left.
the basalt and underlying strata. If the original eleva- 87.1 At road intersection, turn left on BLM road 8460.
tion of the basalt was that of Castle Peak and the White 87.2 Note the overview on the right of a 1 x 0.5 km closed
River Uplift to the northwest at about 3,350 m, then basin that is interpreted to reflect near-surface dissolu-
over 1 km of collapse occurred here. tion of the underlying evaporite.
40.9 Turn around at the 6903 sign and return to I-70. 87.3 Turn around at jeep road to left.
48.3 Turn west on I-70 87.6 Turn right on jeep road just past fence.
51.0 Enter Red Canyon and begin driving through red beds 88.1 Turn right through fence
of the Triassic-Permian State Bridge and underlying 88.3 Stop 3. We will take a short walk (about 500 m) to
Permian-Pennsylvanian Maroon Formations. The axis examine basalt and underlying basal Tertiary sedi-
of the broad, northeast-plunging, Horn Ranch anticline ments in area of maximum collapse (Figs. 4 and 5).
coincides with the canyon here. This anticline is prob- Park and walk right from road down to wash to north.
ably related to diapiric rise of the underlying Eagle At wash turn left and walk southwest downstream. The
Valley Evaporite in response to rapid Neogene down- top basalt flow is well exposed at sharp left turn in
cutting by the Eagle River. wash. This locality illustrates the faulted and tilted
52.5 The valley opens ahead and is underlain by the Eagle character of basalt flows at their lowest elevation of
Valley Evaporite. 2070 m within the Gypsum 7.5' quadrangle. To the
58.2 Take exit 147 to Eagle, turn left at stop sign, proceed northeast the basalts rise through a series of faults and
south across I-70 and the Eagle River to Eby Creek folds to elevations as high as 2865 m. To the southwest
Road (U.S. 6) just beyond bridge, and turn right. the basalts climb the north limb of the Eagle River
58.5 Turn left on Broadway and proceed through downtown anticline to a maximum elevation of 2380 m. This
Eagle. wash exposes at least two basalt flows repeated by a
58.8 Turn left at 5th Street, go one block, and turn right on fault. The basalt in the top of the highest flow is rela-
Capitol Street which becomes Brush Creek Road tively fresh and contains cooling joints and elongate
about 1 mile ahead. vesicles that indicate a dip of about 27° NE. Walking
67.0 Stop 2. Hillside to left exposes spires of intensely frac- farther southwest down the wash, note a lower altered
tured Jurassic Entrada Sandstone just south of major flow with an oxidized base that overlies northeast-dip-
south-dipping normal fault revealed by reflection seis- ping siltstone. Continuing down the wash, the basalt
mic line (Figs. 4 and 6). This line reveals the Hard- sequence is repeated across a poorly exposed north-
scrabble Mountain homocline to be a collapse feature northwest striking fault. The local correlation of the
bounded on south and north sides by normal faults basalt flows across the fault is supported by paleomag-
dipping toward and beneath the homocline and vanish- netic data, which reveal that the top flow in each fault
ing in the Eagle Valley evaporite. The thickness of the block, as well as on the north limb of the Eagle River
Eagle Valley Evaporite interval beneath the homocline, anticline 2.5 km to the southwest, carries reversed
revealed by the seismic data, is relatively thin. The polarity remanent magnetization. All lower flows at
extraordinary thickness (~1.5 km) of Eagle Valley For- these localities carry normal polarity remanent magne-
mation and Evaporite to the north and south, indicates tization. A 22-Ma K-Ar date reported by Larson and
major flow of evaporite out from beneath the homo- others (1975) for the basalt on the west edge of the
cline, flow which probably continues. The extraordi- Gypsum quadrangle also carries reversed polarity
nary thickness of State Bridge Formation on magnetization and thus is probably correlative with the
112 R. B. Scott et al.
top basalt flow at this stop. Northeast-dipping Tertiary unusual structures. 1) The most obvious is a structural
sedimentary rocks are present below the basal basalt terrace on the south flank of Storm King Mountain.
in the second fault block, and they are relatively well The length of this terrace is similar to the known
exposed along the wash and at its confluence with a extent of the conglomerate of Canyon Creek. 2) In the
larger southeast-trending wash. This sequence is about vicinity of the stop, a local basin formed in the Penn-
40 m thick and consists of arkosic pebbly sandstone, sylvanian rocks, in which the conglomerate of Canyon
an interval of well-rounded river cobbles, and basal Creek was deposited. Continued collapse associated
sandstone and interbedded claystone that overlie Eagle with continuing dissolution and diapirism deformed
Valley Formation. Southwest of the larger wash, basalt the conglomerate into a syncline. This interpretation is
is downdropped across another northwest striking practically the same as that given by Bass and
fault. Retrace path back to vehicles and return to I-70. Northrup (1963), who first recognized and mapped the
92.8 Turn west on I-70 conglomerate of Canyon Creek. 3) To the north, not
101.9 Entrance to Glenwood Canyon. visible from here, a low-angle normal fault places
111.3 Take exit 121 to Grizzly Creek rest stop and have Leadville Limestone over Sawatch Quartzite. Either
lunch. upwelling or removal of material related to the evapor-
112.1 Return to I-70 and head west. ite diapir probably caused this rare occurrence of low-
116.4 At Glenwood Springs, note saline hot springs that flow angle normal faulting. Return to I-70.
from the Leadville limestone on the right. 127.7 Turn west on I-70. Half a mile west at 2-3 o’clock is a
119.7 Enter a narrow canyon where Maroon Formation dips hill with horizontal beds of light-gray sandstone of the
40-60° SW. Schoolhouse Member of the Maroon Formation near
123.9 At exit 109, turn right onto frontage road. the west end of the structural terrace we crossed at
124.2 Turn left up Canyon Creek Road. The locally unex- Stop 4, but did not see.
posed Maroon Formation is folded into a structural 145.9 At Exit 90, turn right on State Highway 13 and drive
terrace with irregular and gentle dips in marked con- north through Rifle.
trast to 40-60° SW dips in the canyon of the Colorado 149.9 Turn right on State Highway 325 toward Rifle Gap.
River (Figs. 7 and 8). At the first exposures we see on 154.0 Pass through Rifle Gap and turn right at Rifle Gap
the right, the beds are vertical to overturned, and we Reservoir dam on route 325.
see a transition between the Maroon and Eagle Valley 159.5 Rifle Falls campground is on the right; dips of the
Formations. Farther north we cross a poorly exposed Maroon Formation approach horizontal.
contact with a probable evaporite diapir. 159.8 The Maroon has a slight northward dip. As we
125.9 Stop 4. The tight syncline in the well-indurated con- approach exposures of the gypsum, notice that we
glomerate of Canyon Creek, exposed in the valley of drive through a shallow syncline and the Maroon dips
Canyon Creek, creates a baffling structure (Fig. 8). steeply to the south close to the contact with the evap-
The steeply dipping beds of the conglomerate are orite. Because the Eagle Valley Formation, which
right-side-up, yet they overlie steeply dipping over- overlies the evaporite at most localities, is absent here,
turned beds of Eagle Valley Evaporite. To the west, the pod-shaped gypsum body is assumed to be a
angular discord between the attitudes of Pennsylvan- diapiric intrusion. We consider the structural terrace
ian rocks and trends of faults and the attitudes of the we just drove through to the south to be analogous to
overlying conglomerate is marked. The evaporite the structure described at Canyon Creek (Fig. 9).
probably forms an intrusive rather than depositional 161.1 Stop 5. At the entrance to the narrow canyon cut by
contact with the base of the conglomerate. The con- East Rifle Creek, the Leadville Limestone forms the
glomerate was probably deposited close to its sources light bluish-gray strata at the top of the cliffs. The
by high-gradient streams because here this unit con- darker exposures below consist of the Upper Devonian
tains subangular to round clasts of sandstone and lime- Chaffee Group that dips 45 degrees northward. Map-
stone from the Maroon and Eagle Valley Formations, ping in the quadrangle (Scott and others, 1999) indi-
and 2-3 km west of here, many of the clasts are lime- cates that a south-dipping normal fault with an offset
stone, dolomite, sandstone, and quartzite from Cam- of 400-650 m occurs at the base of the cliffs (Fig 9). If
brian through Mississippian units, which form the so, then why are middle Paleozoic strata dipping
surface of the southern part of the White River plateau. northward? Note that upstream, the attitude of middle
The syncline becomes shallower 5.5 km west of here. Paleozoic strata shallows abruptly to dip 2° southward.
The east end of the syncline is on the slope across the Note also that the north-dipping strata on the upthrown
valley to the east. block occur only adjacent to the diapir of evaporite.
We conclude that after 10 Ma, collapse and We conclude that at depth some of the evaporite
diapiric movement of evaporite resulted in three injected across the normal fault into the upthrown
Active evaporite tectonics and collapse in Eagle River valley and White River uplift, Colorado 113
block, making this local tepee-like structural feature. Paper 750-D, p. D80-D83.
We also conclude that the local structural terrace, Freeman, V.L., 1972, Geologic map of the Ruedi quadrangle, Pitkin and Eagle
Counties, Colorado: U.S. Geological Survey Quadrangle Map GQ-
which departs from the Grand Hogback geometry, is 1004, scale 1:24,000.
related to evaporite flow, dissolution, and resulting Giegengack, R.F., Jr., 1962, Recent volcanism near Dotsero, Colorado: Boulder,
collapse. Colorado, University of Colorado, unpublished M.S. thesis, 69 p.
Hubert, J.F., 1954, Structure and stratigraphy of an area east of Brush Creek,
ACKNOWLEDGMENTS Eagle County, Colorado: Cambridge, Mass., Harvard University, unpub-
lished M. S. thesis, 104 p.
Izett, G.A., and Wilcox, R.E., 1982, Map showing localities and inferred distri-
When we began our work, Colorado Geological Survey butions of the Huckleberry Ridge, Mesa Falls, and Lava Creek ash beds
colleagues Bob Kirkham, Randy Streufert, and Chris Carroll (Pearlette family ash beds) of Pliocene and Pleistocene age in the west-
led field trips to familiarize us with local stratigraphic details ern United States and southern Canada: U.S. Geological Survey Miscel-
and evaporite-related collapse structures. A series of informal laneous Investigations Series Map I-1325, scale 1:4,000,000.
Izett, G.A., Pierce, K.L., Naeser, N.D., and Jaworowski, C., 1992, Isotopic dat-
seminars with CGS geologists encouraged a valuable exchange ing of Lava Creek B tephra in terrace deposits along the Wind River,
of ideas and data. Nancy Driver and Nancy Bauch, USGS Water Wyoming: Implications for post 0.6 Ma uplift of the Yellowstone
Resources Division hydrologists with the Upper Colorado River hotspot: Geological Society of America Abstracts with Programs, v. 24,
Basin project of the National Water-Quality Assessment Pro- no. 7, p. 102.
gram, provided us with critical dissolved-solids load data that Johnson, R.C., and May, Fred, 1980, A study of the Cretaceous-Tertiary uncon-
formity in the Piceance Basin, Colorado: The underlying Ohio Creek
provided us with the modern rate of dissolution. USGS emeri- Formation (Upper Cretaceous) redefined as a member of the Hunter
tus geologist Tom Steven inspired us with his fresh ideas and Canyon or Mesaverde Formation: U.S. Geological Survey Bulletin
depth of knowledge of Colorado geology. Anne Harding pro- 1482-B, 27 p.
vided valuable editorial assistance, and Tom Cooper provided Johnson, S.Y., Schenk C.J., and Karachewski, J.A., 1988, Pennsylvanian and
cartographic help. Ren Thompson, Dave Lageson, and Bruce Permian depositional systems in the Eagle Basin, northwest Colorado:
in Holden, G.S., ed., Geological Society of America 1888-1988 Centen-
Trudgill provided constructive reviews. nial Meeting Field Trip Guidebook: Colorado School of Mines Profes-
sional Contributions 12, p. 156-175.
REFERENCES CITED Johnson, S.Y., Ander, D.L., and Tuttle, M.L., 1990, Sedimentology and petro-
leum occurrence, Schoolhouse Member, Maroon Formation (Lower
Barrett, J.K., and Pearl, R.H., 1978, An appraisal of Colorado’s geothermal Permian), northwestern Colorado: American Association of Petroleum
resources: Colorado Geological Survey Bulletin 39, 224 p. Geologists Bulletin, v. 72, no. 2, p. 135-150.
Bass, N.W., and Northrop, S.A., 1963, Geology of the Glenwood Springs quad- Kirkham, R. M., and Widmann, B.L., 1997, Geologic map of the Carbondale
rangle and vicinity, northwestern Colorado: U.S. Geological Survey quadrangle, Garfield County, Colorado: Colorado Geological Survey
Bulletin, 1142-J, 74 p. Open-File Report 97-3, scale 1:24,000.
Bauch, N.J., and Spahr, N.E., 1998, Salinity trends in surface waters of the Kirkham, R. M., Streufert, R.K., and Cappa, J.A., 1995, Geologic map of the
Upper Colorado River Basin, Colorado: Journal of Environmental Qual- Glenwood Springs quadrangle, Garfield County, Colorado: Colorado
ity, v. 27, p. 640-655. Geological Survey Open-File Report 95-3, scale 1:24,000.
Benson, J.C., and Bass, N.W., 1955, Eagle River anticline, Eagle County, Col- Kirkham, R.M., Bryant, Bruce, Streufert, R.K., and Shroba, R.R., 1996a, Field-
orado: American Association of Petroleum Geologists Bulletin, v. 39, trip Guidebook on the geology and geologic hazards of the Glenwood
no. 1, p. 103-106. Springs area, Colorado, in Thompson, R.A., Hudson, M.R., and Pill-
Bryant, Bruce, Shroba, R.R., and Harding, A.E., 1998, Revised preliminary more, C.L., eds., Geologic excursions to the Rocky Mountains and
geologic map of the Storm King Mountain quadrangle: U.S. Geological beyond: Colorado Geological Survey Special Publication 44, CD-ROM,
Survey Open-File Report 98-472, scale 1:24,000. 38 p.
Butler, D.L., 1996, Trend analysis of selected water-quality data associated with Kirkham, R. M., Streufert, R.K., Hemborg, T.H., and Stelling, P.L., 1996b, Geo-
salinity-control projects in the Grand Valley, in the lower Gunnison logic map of the Cattle Creek quadrangle, Garfield County, Colorado:
River basin, and at Meeker Dome, western Colorado: U.S. Geological Colorado Geological Survey Open-File Report 96-1, scale 1:24,000.
Survey Water-Resources Investigations Report 95-4274, 38 p. Kirkham, R.M., Streufert, R.K., and Cappa, J.A., 1997a, Geologic map of the
Cappa, J.A., and Hemborg, H.T., 1995, 1992-1993 low-temperature geothermal Glenwood Springs quadrangle, Garfield County, Colorado: Colorado
assessment program, Colorado: Colorado Geological Survey Open-File Geological Survey Map Series 31, 22 p., scale 1:24,000.
Report 95-1, 19 p. Kirkham, R.M., Streufert, R.K., Scott, R.B., Lidke, D.J., Bryant, Bruce, Perry,
Carroll, C.J., Kirkham, R.M., and Stelling, P. W., 1996, Geologic map of the W.J., Jr., Kunk, M.J., Driver, N.E., and Bauch, N.J., 1997b, Active salt
Center Mountain quadrangle, Garfield County, Colorado: Colorado dissolution and resulting geologic collapse in the Glenwood Springs
Geological Survey Open-File Report 96-2, scale 1:24,000. region of west-central Colorado: Geological Society of America
Colman, S.M., 1985, Map showing tectonic features of late Cenozoic origin in Abstracts with Programs, v. 29, no. 6, p. A-416.
Colorado: U.S. Geological Survey Miscellaneous Geologic Investiga- Kunk, M.J., and Snee, L.W., 1998, 40Ar/39Ar age-spectrum data of Neogene and
tions Series Map I-1556, scale 1:1,000,000. younger basalts in west-central Colorado: U.S. Geological Survey
De Voto, R.H., Bartleson, B.L., Schenk, C.J., and Waechter, N.B., 1986, Late Open-File Report 98-243, p. 112.
Paleozoic stratigraphy and syndepositional tectonism, northwestern Kunk, M.J., Kirkham, Robert, Streufert, R.K., Scott, R.B., Lidke, D.J., Bryant,
Colorado, in Stone, D.S., ed., New interpretations of northwest Col- Bruce, and Perry, W.J., 1997, Preliminary constraints on the timing of
orado geology: Rocky Mountain Association of Geologists-1986 Sym- salt tectonism and geologic collapse in the Carbondale and Eagle col-
posium, p. 37-49. lapse centers, west-central Colorado: Geological Society of America
Freeman, V.L., 1971, Permian deformation in the Eagle basin, Colorado, in Abstracts with Programs, v. 29, no. 6, p. A-416.
Geological Survey Research 1971: U.S. Geological Survey Professional Larson, E.E., Ozima, Minoru, and Bradley, W.C., 1975, Late Cenozoic basic
114 R. B. Scott et al.
volcanism in northwestern Colorado and its implications concerning County, Colorado: U.S. Geological Survey Miscellaneous Investigations
tectonism and the origin of the Colorado River system, in Curtis, B.F., Series, scale 1:24,000, in press.
ed., Cenozoic history of the southern Rocky Mountains: Geological Soule, J.M., and Stover, B.K., 1985, Surficial geology, geomorphology, and
Society of America Memoir 155, p. 155-187. general engineering geology of parts of the Colorado River Valley,
Lidke, D.J., 1998, Geologic map of the Wolcott quadrangle, Eagle County, Col- Roaring Fork River Valley, and adjacent areas, Garfield County, Col-
orado: U.S. Geological Survey Miscellaneous Investigations Series Map orado: Colorado Geological Survey Open-File Report 85-1.
I-2656, scale 1:24,000. Steven, T.A., Hon, K., and Lanphere, M.A., 1995, Neogene geomorphic evolu-
Lidke, D.J., 1999, Geologic map of the Eagle quadrangle, Eagle County, Col- tion of the central San Juan Mountains near Creede, Colorado: U.S.
orado: U.S. Geological Survey Miscellaneous Field Studies Map MF-in Geological Survey Miscellaneous Investigations Series Map I-2504.
press, scale 1:24,000. Steven, T.A., Evanoff, Emmett, and Yuhas, R.H., 1997, Middle and Late Ceno-
Liebermann, T.D., Mueller, D.K., Kircher, J.E., and Choquette, A.F., 1989, zoic tectonic and geomorphic development of the Front Range of Col-
Characteristics and trends of streamflow and dissolved solids in the orado, in Bolyard, D.W., and Sonnenberg, S.A., eds., Geologic history
Upper Colorado River Basin, Arizona, Colorado, New Mexico, Utah, of the Colorado Front Range: Rocky Mountain Association of Geolo-
and Wyoming: U.S. Geological Survey Water Supply Paper 2358, 64 p. gists 1997 RMS-AAPG Field Trip #7, p. 115-124.
Mallory, W.W., 1966, Cattle Creek anticline, a salt diapir near Glenwood Stover, B.K., 1986, Geologic evidence of Quaternary faulting near Carbondale,
Springs, Colorado, in Geological Survey Research, 1966: U.S. Geolog- Colorado, with possible associations to the 1984 Carbondale earthquake
ical Survey Professional Paper 550-B, p. B12-B15. swarm, in Rogers, W.P., and Kirkham, R.M., eds., Contributions to Col-
Mallory, W.W., 1971, The Eagle Valley Evaporite, northwest Colorado-A orado seismicity and tectonics-A, 1986 update: Colorado Geological
regional synthesis: U. S. Geological Survey Bulletin 1311-E, p. E1-E37. Survey Special Publication 28, p. 295-301.
Murray, F.N., 1966, Stratigraphy and structural geology of the Grand Hogback Streufert, R.K., Kirkham, R.M., Shroeder, T.S., and Widmann, B.L., 1997a,
monocline, Colorado: Boulder, Colo., University of Colorado, unpub- Geologic map of the Dotsero quadrangle, Eagle and Garfield Counties,
lished Ph.D. dissertation, 219 p. Colorado: Colorado Geological Survey Open-File Report 97-2, scale
Murray, F.N., 1969, Flexural slip as indicated by faulted lava flows along the 1:24,000.
Grand Hogback Monocline: Journal of Geology, v. 77, p. 333-339. Streufert, R.K., Kirkham, R. M., Widmann, B.L., and Shroeder, T.S., 1997b,
Piety, L.A., 1981, Relative dating of terrace deposits and tills in the Roaring Geologic map of the Cottonwood Pass quadrangle, Eagle and Garfield
Fork valley, Colorado: Boulder, Colo., University of Colorado, unpub- Counties, Colorado: Colorado Geological Survey Open-File Report 97-
lished M.S. thesis, 209 p. 4, scale 1:24,000.
Reed, J.C., Jr., Bryant, Bruce, Sims, P.K., Beaty, D.W., Bookstrom, A.A., Tweto, Ogden, 1975, Laramide (Late Cretaceous-early Tertiary) orogeny in the
Grose, T.L., Mallory, W.W., Wallace, S.R., and Thompson, T.B. 1988, southern Rocky Mountains, in Curtis, B.F., ed., Cenozoic history of the
Geology and mineral resources of central Colorado, in Holden, G.S., Southern Rocky Mountains: Geological Society of America Memoir
ed., Geological Society of America 1888-1988 Centennial Meeting, 144, p. 1-44.
Denver, Colorado: Colorado School of Mines Professional Contribu- Tweto, Ogden, 1977, Tectonic history of west-central Colorado, in Veal, H.K.,
tions 12, p. 68-121. ed., Exploration frontiers of the central and southern Rockies: Rocky
Reed, J.C., Jr., Bickford, M.E., and Tweto, Ogden, 1993, Proterozoic accre- Mountain Association of Geologists-1977 Symposium, p. 11-22.
tionary terranes of Colorado and southern Wyoming, in Van Schmus, Tweto, Ogden, 1979, Geologic Map of Colorado: U.S. Geological Survey,
R.R., and Bickford, M.E., eds., Transcontinental provinces, in Reed, 1:500,000 scale.
J.C., Jr., Bickford, M.E., Houston, R.S., Link, P.K., Rankin, D.W., Sims, Tweto, Ogden, and Lovering, T.S., 1977, Geology of the Minturn 15-minute
P.K., and Van Schmus, W.R., eds., Precambrian of the conterminous quadrangle, Eagle and Summit Counties, Colorado: U.S. Geological
United States, v. C-2 of Geology of North America: Geological Society Survey Professional Paper 956, 96 p.
of America Decade of North American Geology, p. 211-228. Tweto, Ogden, Moench, R. H., and Reed, J.C., Jr., 1978, Geologic map of the
Scott, R.B., and Shroba, R.R., 1997, Revised preliminary geologic map of the Leadville 1o x 2o quadrangle, northwestern Colorado: U.S. Geological
New Castle quadrangle, Garfield County, Colorado: U.S. Geological Survey Miscellaneous Investigations Series Map I-999, 1:250,000 scale.
Survey Open-File Report 97-737, scale 1:24,000. Unruh, J.R., Wong, I.G., Bott, J.D., Silva, W.J., and Lettis, W.R., 1993, Seismo-
Scott, R.B., Lidke, D.J., Shroba, R.R., Hudson, M.R., Kunk, M.J., Perry, W.J., tectonic evaluation, Rifle Gap Dam, Silt Project, Ruedi Dam, Frying-
Jr., and Bryant, Bruce, 1998, Large-scale active collapse in western Col- pan-Arkansas Project, northwestern Colorado: unpublished report
orado: Interaction of salt tectonism and dissolution, in Brahana, J.V., prepared by William R. Lettis and Associates and Woodward-Clyde
Eckstein, Yoram, Ongley, L.K., Schneider, Robert, and Moore, J.E., Consultants for U.S. Bureau of Reclamation, 154 p.
eds., Gambling with Groundwater—Physical, chemical, and biological U.S. Department of the Interior, 1997, Quality of water, Colorado River
aspects of aquifer-stream relations: Proceedings of the joint meeting of Basin, Progress Report no. 18, Bureau of Reclamation, Salt Lake City,
the XXVIII Congress of the International Association of Hydrologists Utah.
and the annual meeting of the American Institute of Hydrologists, Las Wanek, L.J., 1953, Geology of an area east of Wolcott, Eagle County, Colorado:
Vegas, Nevada, USA, 28 September-2 October, 1998, p. 195-204. Boulder, Colo., University of Colorado, unpublished M.S. thesis, 62 p.
Scott, R.B., Shroba, R.R., and Egger, A.E. 1999, Geologic map of the Rifle Widmann, B.L., 1997, Evaporite deformation in the Dotsero, Gypsum, and Cot-
Falls quadrangle, Garfield County, Colorado: U.S. Geological Survey tonwood Pass 7.5' quadrangles, Eagle County, Colorado: Golden, Colo.,
Miscellaneous Investigations Series, scale 1:24,000, in press. Colorado School of Mines, unpublished M.S. thesis, 93 p.
Shroba, R.R., and Scott, R.B., 1997, Geology of the Rifle quadrangle, Garfield Ye, Hongzhuan, Royden, Leigh, Burchfiel, Clark, and Schuepbach, Martin,
County, Colorado: U.S. Geological Survey Open File Report 97-852, 1996, Late Paleozoic deformation of interior North America: The
scale 1:24,000. greater ancestral Rocky Mountains: American Association of Petroleum
Shroba, R.R., and Scott, R.B., 1999, Geology of the Silt quadrangle, Garfield Geologists Bulletin, v. 80, p. 1397-1432.
Printed in U.S.A.
Geological Society of America
Field Guide 1
1999
Brownfield, M. E., Johnson, E. A., Affolter, R. H., and Barker, C. E., 1999, Coal mining in the 21st century: Yampa coal field, northwest Colorado, in
Lageson, D. R., Lester, A. P., and Trudgill, B. D., eds., Colorado and Adjacent Areas: Boulder, Colorado, Geological Society of America Field Guide 1.
115
116 M. E. Brownfield et al.
R. 94 W. R. 93 W. R. 92 W. R. 91 W. R. 90 W. R. 89 W. R. 88 W. R. 87 W. R. 86 W. R. 85 W.
T. 9 N.
ROUTT CO.
MOFFAT CO.
T. 8 N.
107 30' 107
40 37' 30''
Colorado
13
T. 7 N.
Lay
40 Craig Bear
Hayden River
Milner 40
5
X
Mount T. 6 N.
McGregor
Harris
5
2 X6
1
Trapper Mine
3 4
Mesaverde Group Foidel Cr.
Wi T. 5 N.
llia Canyon
ms
Fo X
Iles Mtn. rk 6 X 3Eckman
Mo
un
tai Park
ns
4
X 2
13 Fish Creek X
Canyon Edna T. 4 N.
Mine 1
X Oak Creek
Explanation 40 15'
T. 3 N.
Active coal mine X Field trip stops
RIO BLANCO CO.
1. Eagle mine 1. Oak Creek 131
2. Trapper mine 2. Edna Mine
3. Seneca II-W mine 3. Eckman Park
4. Yoast mine 4. Fish Cr. Canyon 0 5 mi
5. Seneca II mine 5. Mt. Harris
6. Foidel Creek mine 6. Trapper Mine
Figure 1. Location of the Yampa coal field, Moffat, Rio Blaco, and Routt Counties, Colorado. The
outcrop of the Mesaverde Group (shaded) delineates the boundaries of the coal field. Approximate
locations of existing mines and trip stops are shown.
bed about 600 feet below the top of the Trout Creek Sandstone the Eagle No 5 Mine, and the Eagle No 9 Mine, both located
was extensively mined. Two rail lines were constructed south of about 12 miles southwest of Craig. The Eagle mines are tem-
McGregor where several mines were developed in the Wadge porarily closed.
and Wolf Creek coal beds. Later a strip mine was developed just Two mine-mouth power plants produce electricity in the
south of McGregor that supplied coal for one of the early coal- coal field, the Craig plant, located about four miles south-
fired power plants in northwest Colorado. southwest of Craig, and the Hayden plant, located about five
Large strip mines were developed during the 1950’s in the miles east of Hayden. Tri State Generation and Transmission
eastern part of the coal field, notably the Edna Mine located Association, Inc. own the Craig plant. Most coal burned at the
about three miles north of Oak Creek, and continued into the plant comes from the Trapper Mine located about three miles to
1970’s with the Energy Fuels mines in the southern part of the south (Fig. 1). Public Service of Colorado, PaciCorp, and
Twentymile Park. Both of these mines have ceased operations Salt River Project jointly own the Hayden plant. All of the coal
and are now in the reclamation stage. burned at this plant comes from the Seneca II-W and Yoast
Currently, only a few large mines operate in the Yampa coal mines located two miles to the south.
field (Fig. 1). In the eastern part of the coal field, three major During 1997, the Yampa coal field accounted for about 40
mines are the Foidel Creek Mine, located in the southern part of percent of Colorado’s total coal production (Resource Data
Twentymile Park, and the Seneca II-W, and Yoast Mines, International, 1998). The cumulative coal production for the
located several miles south and southwest of Mount Harris. In Yampa coal field, from 1864 through 1997, is 266.19 million
the western part of the coal field, three notable mines are the tons (Tremain and others, 1996; and, Resource Data Interna-
Trapper Mine (surface), located about five miles south of Craig, tional, 1998).
Coal mining in the 21st century: Yampa coal field, northwest Colorado 117
GEOLOGICAL SETTING is now central North America (Fig. 2). The seaway stretched
from Mexico to Alaska in what is now the central part of the
Cretaceous Paleogeography North America, the width of the seaway extended from central
Utah to eastern Nebraska. A stable cratonic platform bordered
During the Cretaceous Period, a large, north-trending epi- the seaway on the east, and the tectonically active Sevier oro-
continental seaway, the Western Interior Seaway, occupied what genic belt bordered the seaway on the west. Sediments moving
Sea
o
45
r
Interio
ds
o
45
lan
igh
Yampa
rH
coal field
vie
Se
rn te
Wes
o
30
o
30
o
30 Paleolatitude
Alluvial plain Coastal Peat Western Interior
and highlands plain swamps Seaway
Thrust fault
Figure 2. Paleogeography map of the central part of North America during the late Campanian Stage (72-79) of the
Late Cretaceous Period. The Yampa coal field is shown in relation to the western shoreline, coastal plain, and peat
swamps associated with the Western Interior Seaway. Modified from Roberts and Kirschbaum (1995).
118 M. E. Brownfield et al.
eastward from the highlands were deposited along the fluctuat- overlying Paleocene Fort Union Formation and the Fort Union
ing shoreline resulting in a complex package of Cretaceous sed- from the overlying Wasatch Formation. The oldest Cretaceous
imentary formations. unit exposed in the vicinity of the Yampa coal field is the Man-
cos Shale, which forms areas of low relief just south of the coal
Stratigraphy of Upper Cretaceous Sedimentary Rocks in field. The youngest Cretaceous unit exposed in the vicinity of
Northwestern Colorado the coal field is the Lance Formation, which forms low hills just
north of the coal field.
In ascending order, the Upper Cretaceous lithostratigraphic
units in northwestern Colorado (Fig. 3) are Mancos Shale, Mesaverde Group in the Yampa coal field
Mesaverde Group (Iles Formation and Williams Fork Forma-
tion), Lewis Shale, Fox Hills Sandstone, and Lance Formation. Holmes (1877) first applied the named Mesaverde Group
Major unconformities separate the Lance Formation from the to a sequence of sandstone, mudrock (siltstone, mudstone,
shale, and claystone), and coal in southwestern Colorado. Fen-
neman and Gale (1906b), noting a similar sequence of rocks in
the Yampa coal field, extended the name Mesaverde into north-
western Colorado. Fenneman and Gale did not subdivide the
Western Yampa coal field Mesaverde into formations, but they did describe two regional
sandstones, which they named the Trout Creek and the Twen-
Wasatch Fm.
Paleocene
between the Trout Creek and the Twentymile, and the upper
group contains all of the coal above the Twentymile.
Hancock (1925) later subdivided the Mesaverde Group into
the Iles Formation and overlying Williams Fork Formation, and
Lance Fm.
the lower part of those rocks between the Trout Creek and the
Twentymile, and reintroduced an earlier name, Holderness
member, for those rocks in the Williams Fork above the Twen-
Upper Cretaceous
lower coal group the Tertiary. Toward the west, the Mesaverde becomes increas-
ingly more fluvial to the point that in central Utah equivalent
strata are composed almost exclusively of conglomerate.
1000 Ft
DETAILED STRATIGRAPHY
OF THE UPPER CRETACEOUS
Iles Formation
Figure 3. Stratigraphic column of upper Cretaceous and Tertiary rocks
in the Yampa coal field showing depositional environments and coal The Iles Formation (Fig. 3) of the Mesaverde Group was
groups. Modified after Bass and others (1955). named by Hancock (1925) for coal-bearing rock exposed at Iles
Coal mining in the 21st century: Yampa coal field, northwest Colorado 119
NW SE
LEWIS SHALE
NE
TS
HO
RE
LIN
Three White sandstones ET
RA
NS
GR
ES
SIO
N
Big White sandstone
Twentymile Sandstone
Sub-Twentymile sandstone
P
OU
GR
ION
ES
E SS
GR
M
E
ER
ELIN
H OR
TS
NE
Tow Creek Sandstone Depositional environments
Alluvial, coastal plain,
and mire
MANCOS SHALE Upper shoreface
and estuarine
Lower shoreface and
offshore marine
Figure 4. Generalized northwest-southeast cross section of the Mesaverde Group in the Yampa coal
field showing the stratigraphic positions of major sandstone units and related major regressions and
transgressions. Modified after Seipman (1985).
Mountain (Fig. 1) in the western part of the Yampa coal field. tongue (Fig. 5) that ranges in thickness from about 400 feet just
The Iles Formation is conformable with the Mancos Shale west of Oak Creek to about 100 feet in the western part of the
below and the Williams Fork Formation above. The Iles con- coal field capped by the Trout Creek Sandstone.
sists of sandstone interbedded with mudrock (siltstone, mud- Three persistent sandstone units in the Iles Formation
stone, shale, and claystone), carbonaceous shale, and coal and deserve special mention as guides for correlation within the
ranges in thickness from about 1,300 feet in the west (Hancock, coal field. They are the Tow Creek Sandstone Member at the
1925) to 1,500 in the eastern and central (Bass and others, base, the double ledge sandstone unit about 400 feet above the
1955) parts of the coal field. base, and the Trout Creek Sandstone Member at the top of the
The lower two-thirds of the Iles Formation consists of mas- formation.
sive ledge-forming beds of sandstone interbedded with Tow Creek Sandstone Member. The Tow Creek Sandstone
mudrock, carbonaceous shale, and coal. This sequence of rocks Member of the Iles Formation (Figs. 3, 4, and 5) is the basal
form steep cliffs that rise above the broad lowland formed on unit in the Iles in the eastern and central part of the Yampa coal
the Mancos Shale along the southern and western boundaries of field. Crawford and others (1920) named the Tow Creek for
the Yampa coal field. Most of the coal beds within the lower exposures between Milner and Bear River (Fig. 1). The Tow
coal group are in the upper part of this sequence and occur Creek consists of light-gray to white sandstone ranging in thick-
about 400 feet above the base of the Iles Formation (Bass and ness from 35 to 125 feet (Bass and others, 1955). This unit is
others, 1955). Three principal coal beds or zones were recog- prominent west of Fish Creek canyon (Fig. 1) in the south cen-
nized by Fenneman and Gale (1906) in the lower coal group at tral part of the coal field.
Oak Creek and are referred to as No. 1, No. 2, and No. 3 coal Double Ledge Sandstone. Bass and others (1955) infor-
zones (Bass and others, 1955). The upper part of the Iles For- mally named the double ledge sandstone. The double ledge
mation consists of a mudrock sequence capped by cliff forming sandstone unit consists of 1 to 3 beds of light-gray and white
sandstone. The mudrock sequence is a transgressive marine cliff-forming sandstone and interbedded siltstone unit about
West East
Lower White River coal field Danforth Hills coal field Western Yampa coal field
Eocene
Depositional environments
Fm.
Alluvial, coastal plain,
Wasatch
Paleocene
and mire
Upper shoreface
Fm.
and estuarine
Fort Union
Lance Fm.
Lewis Shale
?
Lewis Shale
?
Wasatch Fm.
upper coal group
(part)
1000 Ft
Barren interval Twentymile Ss. Mbr.
Fort Union Fm.
Ohio Creek Fm. ?
Paleocene-Eocene
marine shale
Upper Cretaceous
?
Williams Fork Fm.
Black Diamond
Yampa ? coal group
Iles Fm.
Mesaverde Group
River
Iles Fm.
Lower
Upper Cretaceous
lower unit coal group
Line of Tow Creek Ss. Mbr.
Danforth Hills
section Sego Ss. Mbr. rimrock ss
(part)
(part)
Mancos Shale (part)
Mancos Shale
Figure 5. Generalized cross section showing part of the Upper Cretaceous and Tertiary rocks in the Lower White
River, Danforth Hills, and Yampa coal fields with interpreted depositional environments.
Coal mining in the 21st century: Yampa coal field, northwest Colorado 121
400 to 460 feet above the base of the Iles Formation. The dou- middle coal group (Fig 3). Throughout the eastern part of the
ble ledge ranges in thickness from 130 to 250 feet. The unit coal field the principal coal beds are, in ascending order, the
forms prominent ledges visible along the steep south facing Wolf Creek, Wadge, and Lennox (Bass and others, 1955). These
cliffs Williams Fork Mountains in the south central part of the beds occur in the lower 400 feet of the Williams Fork, immedi-
Yampa coal field. North of Milner, the unit forms a single white ately above the Trout Creek Sandstone. Data compiled by the
ledge where it was useful in determining displacement on a authors suggest that the number of coal beds in the middle coal
fault. group increases westward in the Yampa coal field (Johnson and
Trout Creek Sandstone. Fenneman and Gale (1906) others, in press). The upper part of the middle coal group is
named the Trout Creek Sandstone Member of the Iles Forma- characterized by mudrock capped by cliff-forming sandstone.
tion (Figs. 3) for exposures along Trout Creek in the eastern The upper third of the Williams Fork contains the upper coal
part of the coal field. The Trout Creek is a cliff-forming unit of group and consists of sandstone, mudrock, carbonaceous shale,
regional extent that transitionally overlies a sequence of marine and coal.
shale (Figs. 4 and 5) containing Exiteloceras jenneyi (Izett and A regional cross section by Roehler (1987) that extends
others, 1971) and is overlain by coal-bearing rocks. Fenneman from the vicinity of Mount Harris, Colorado northwest to Rock
and Gale (1906) and Bass and others (1955) reported the Trout Springs, Wyoming shows that the Williams Fork is roughly
Creek to be about 100 ft thick in the eastern part of the coal equivalent to the upper part of the Ericson Sandstone and the
field and Johnson (1987) reported the unit as ranging from 67 to Almond Formation. Southwest of the coal field, the Williams
79 ft thick in the western part of the coal field. Siepman (1985), Fork is also used in the Danforth Hills (Fig. 5) coal field (Han-
in his regional study of the unit, determined that it ranges from cock, 1925, Hancock and Eby, 1930)
140 ft thick near Mount Harris to 31 ft thick just west of Oak Five persistent stratigraphic units in the Williams Fork For-
Creek. The unit has an average thickness of about 67 ft. Toward mation have proven very useful as guides for correlation within
the northwest, the Trout Creek pinches out in the subsurface of the coal field. They are: 1) the Yampa bed, 2) sub-Twentymile
the Sand Wash Basin (Siepman, 1985; Roehler, 1987). Region- sandstone in the middle coal group, 3) the Twentymile Sand-
ally, the Trout Creek (Fig. 5) is well exposed in the Danforth stone Member, 4) the Big White sandstone in the upper coal
Hills (Hancock and Eby, 1930) and pinches out just east of the group, and 4) the Three White sandstones at the top of the for-
Lower White River coal field (Hail, 1978, Barnum and Gar- mation.
rigues, 1980). Toward the south, the Trout Creek is equivalent Yampa Bed. The term Yampa bed, introduced by Brown-
to the Rollins Sandstone in the southern Piceance Basin of field and Johnson (1986), is a regionally extensive tonstein
west-central Colorado (Collins, 1976). (altered ash-fall tuff) in the lower part of the Williams Fork For-
mation. When exposed on the surface or observed in drill core,
Williams Fork Formation the unit is a white to grayish white, massisve, claystone. In the
subsurface, the unit becomes an important regional marker bed
Hancock (1925) named the Williams Fork Formation for that is easily identified on drill hole geophysical logs. The
coal-bearing strata in the upper part of the Mesaverde Group. Yampa bed ranges in thickness from less than one foot to six ft.
The Williams Fork rests conformably on the top of the Trout In the central and western parts of the coal field, the unit lies
Creek Sandstone Member of the Iles Formation. The upper con- between 100 and 260 ft above the top of the Trout Creek Sand-
tact of the Williams Fork with the overlying Lewis Shale is also stone respectively. However, in the eastern part of the coal field,
conformable. The Williams Fork Formation consists of the stratigraphic separation is less than 20 ft, and on some geo-
mudrock interbedded with sandstone and lessor amounts of car- physical logs, the Yampa bed appears to rest directly on the
bonaceous shale and coal. These rocks are well exposed in the Trout Creek. The age of the Yampa bed, 72.5 mybp ± 5.1 my,
Williams Fork Mountains (Fig. 1). was determined using K-Ar methods on andesine.
In the western part of the coal field, Hancock (1925) The Yampa bed has been identified in the Danforth Hills
reported that the Williams Fork is approximately 1,600 ft thick; (Brownfield and others, in press) and in the subsurface of the
Johnson (1987) reported the formation to be about 1,880 ft Piceance and Sand Wash Basin (Brownfield and Johnson,
thick in the Round Bottom quadrangle southwest of Craig. In 1986).
the eastern part of the coal field, Bass and others (1955) Sub-Twentymile Sandstone. The term sub-Twentymile
reported the formation to range from 1,600 ft thick near Mount sandstone was introduced by Kitely (1983) for a sandstone unit
Harris to nearly 2,000 ft thick at the western margin of their that lies about 150 ft below the base of the Twentymile Sand-
study area. Four USGS coal exploration drill holes located in stone in the eastern-central part of the coal field (Fig. 4). Most
the western part of the coal field show an average thickness of of what is known about this unit comes from the subsurface, but
1,915 ft for the Williams Fork. it probably has many physical characteristics in common with
The lower two-thirds of the Williams Fork Formation con- the Trout Creek and Twentymile Sandstones. Based on infor-
sists of massive ledge-forming light gray and white sandstone mation compiled by the authors, the unit averages about 30 ft
interbedded with mudrock, carbonaceous shale, and coal of the thick and contains one to three sandstone bodies. Masters
122 M. E. Brownfield et al.
(1966) defined a sandstone unit, the Hayden Gulch sandstone, Lewis Shale is about 435 ft. Toward the west, the lower and
at this same stratigraphic level, and this sandstone is undoubt- middle sandstones pinch out, but the upper sandstone continues
edly equivalent to the sub-Twentymile. Siepman (1983) recog- to the edge of the coal field. Toward the east, the three sand-
nized the sub-Twentymile, and included it at the top of his stones continue for several miles east of the Yampa River before
sub-Twentymile unit. The sub-Twentymile pinches out toward pinching out.
the east, and toward the west, it looses its distinctiveness by
splitting into a number of smaller sandstone bodies. DEPOSITIONAL SETTING OF THE MESAVERDE
Twentymile Sandstone. The Twentymile Sandstone Mem- GROUP
ber of the Williams Fork Formation was named by Fenneman
and Gale (1906) for exposures in Twentymile Park (Fig. 1) in During the Late Cretaceous, the western edge of the West-
the eastern part of the coal field. The stratigraphic distance ern Interior Seaway was continually being modified by sedi-
between the top of the Trout Creek Sandstone and the base of ment influx from tectonically active areas to the west.
the Twentymile ranges from about 900 to 1,100 ft (Bass and According to Haun and Weimer (1960), as much as 11,000 ver-
others, 1955). The Twentymile typically forms distinctive tical feet of sediment were deposited in the seaway during this
cliffs of yellowish gray to white sandstone. Unlike the Trout time. Along the western margin of the Cretaceous Western Inte-
Creek, the Twentymile often comprises two to three sandstone rior Seaway, the Mesaverde deposition was characterized by a
units separated by finer grained material. Bass and others series of westward transgressions and eastward regressions of
(1955) reported that the Twentymile is about 100 to 200 ft the epicontinental seaway that resulted in the cyclic deposition
thick in the eastern part of the coal field. Siepman (1985) of marine and non-marine lithofacies (Weimer, 1960; Zapp and
reported that the unit ranges from 184 ft thick in Fish Creek Cobban, 1960). The sediment source area was located to the
Canyon in the eastern part of the coal field to 28 ft thick at west in the Sevier orogenic belt.
Duffy Mountain in the western part of the coal field. The The sediments deposited in northwestern Colorado during
Twentymile Sandstone begins to lose its identity on the west- this time are now contained in the Mancos Shale and Iles and
ern edge of the coal field (Fig. 5), and pinches out toward the Williams Fork Formations of the Mesaverde Group (Fig. 3).
northwest in the subsurface of the Sand Wash Basin (Siepman, Variations in the sedimentation rate and drainage patterns, basin
1985) and to the southwest where it is not present in the Dan- subsidence, and eustatic changes in sea level contributed to the
forth Hills (Brownfield and others, in press). cyclic deposition of Upper Cretaceous lithofacies. Masters
Big White Sandstone. The term Big White sandstone was (1966) recognized three large-scale, regressive cycles in the
introduced by W. R. Grace and Company for a sandstone unit Mesaverde: (1) from the base of the Iles to the base of the marine
that lies about 200 ft above the Twentymile Sandstone in the shale underlying the Trout Creek Sandstone, (2) from the base of
central part of the coal field. In the western part of the coal the Trout Creek Sandstone to the base of the marine shale under-
field, a discontinuous sandstone exists at the same strati- lying the Twentymile Sandstone, and (3) from the base of the
graphic level as the Big White and is referred to by USGS Twentymile Sandstone to the top of the Williams Fork (Fig. 4).
geologists as the Fuhr Gulch sandstone of the Williams Fork In each cycle, the top of the coal bearing package and the base of
Formation (Johnson, 1987). Johnson (1987) describes the the overlying marine shale are separated by a transgressive dis-
unit at the mouth of Fuhr Gulch as a light-gray, very fine to conformity. In general, the Iles represents net shoreline regres-
fine grained sandstone with a thickness of 40 ft. In the eastern sion, and the Williams Fork represents a net shoreline
part of the coal field, Campbell (1923) named a sandstone transgression. However in detail, the Williams Fork shows evi-
unit at the same stratigraphic level of Big White (in Eckman dence of several shoreline fluctuations, as evidenced by the ver-
Park, Figs. 1 and 8) the Fish Creek sandstone. With little tical juxtaposition of the formation’s three major depositional
doubt, all three sandstone units are equivalent, and represent settings: offshore marine, nearshore marine, and fluvial.
a poorly exposed, regional unit. The westward-thinning tongues of mudrock that directly
Three White Sandstones. In the western part of the coal underlie the Trout Creek and Twentymile Sandstones were
field, the upper part of the Williams Fork Formation is domi- deposited on an open-marine shelf, as evidenced by the pres-
nated by three, vertically stacked, thick sandstone units that are ence of marine body fossils and deep-marine trace fossils. Most
referred to informally by USGS geologists as the Three White likely, this same environment is represented in the strata that
sandstones (Fig. 4). Lithologically, the three sandstones resem- underlie the sub-Twentymile sandstone in the central part of the
ble the Twentymile Sandstone. In addition, each sandstone unit coal field. A shallower marine environment is represented in the
overlies a thin sequence of marine rocks and is overlain by non- strata that underlie the Three White sandstones in the western
marine rocks. Each of the three sandstone units is 50 to 60 ft part of the coal field.
thick. Where best exposed along the Yampa River southwest of The thick sandstones, such as the Trout Creek, Twentymile
Craig, the base of the lower sandstone lies about 310 ft above Sandstone Members, and the sub Twentymile, Big White, and
the top of Big White sandstone. The stratigraphic distance Three White sandstones were deposited in a progradational
between the base of the lower sandstone and the base of the shoreface environment (Boyles and others, 1981). This conclu-
Coal mining in the 21st century: Yampa coal field, northwest Colorado 123
sion is supported by an upward increase in grain size and an lowed by an eastward shift of the shoreline that allowed
upward decrease in bed thickness, hummocky cross stratifica- nearshore marine sand to be deposited, now represented by the
tion followed upward by trough cross stratification, and the sub-Twentymile sandstone and its underlying marine shale.
occurrence of shallow-marine trace fossils. Relatively complete Next, nonmarine sediments prograded eastward over the sub-
stratigraphic sections of the Trout Creek and Twentymile reveal Twentymile followed by a westward shift of the shore line
that the units are composed of a basal transitional part deposited flooding the area, and marine mud was deposited over the non-
below wave base, and an overlying shoreface part deposited marine sediments represented by the marine shale underlying
above wave base with rare foreshore and backshore deposits. the Twentymile Sandstone.
Many workers, most recently Siepman (1985), believe that The last phase of the Williams Fork deposition occurred
these sandstone units were deposited along wave-dominated when the shoreline again moved eastward across the coal field
deltaic, strand plain, and barrier island systems, in a microtidal and nearshore marine sand and overlying nonmarine sediments
setting. now contained in the Twentymile Sandstone and uppermost part
The strata that overlie the nearshore marine sandstones dis- of the Williams Fork (upper coal group) were deposited. The
play characteristics suggesting fluvial, lagoonal, salt marsh, and depositional history at the close of Williams Fork time is repre-
freshwater swamp environments that accumulated on a coastal sented by a major westward transgression of the shoreline
plain that sloped gently seaward with little topographic relief. flooding the area with sea water, and the deposition of the
The fluvial lithofacies contain channel sandstones and associ- Lewis Shale. This signaled the end of Mesaverde deposition in
ated overbank deposits of sandstone, siltstone, shale, and lentic- northwestern Colorado. However, on the western side of the
ular coal beds. The lagoon and bay environments are coal field, the Lewis transgression was interrupted at least three
represented by low-energy deposits of shale, siltstone, and times by minor fluctuations in the shoreline during which time
minor fine-grained sandstone. The lagoonal sandstone deposits the Three White sandstones were deposited.
include washover fans, flood-tidal deltas, and crevasse splays.
The salt marsh lithofacies contain evenly laminated, fine- POST-CRETACEOUS DEFORMATIONAL HISTORY
grained deposits consisting of carbonaceous shale, siltstone, OF NORTHWESTERN COLORADO
very fine-grained sandstone, and impure coal. The freshwater
swamp lithofacies are represented by coal. Near the end of the Cretaceous, the Western Interior Sea-
way withdrew from what is now northwestern Colorado and
DEPOSITIONAL HISTORY OF THE MESAVERDE marine and coastal plain deposition ceased. This event was fol-
GROUP IN THE YAMPA COAL FIELD lowed closely by compressional tectonism associated with the
onset of the Laramide orogeny. In northwestern Colorado, the
Prior to the deposition of the Mesaverde Group, the shore- Lance and older rocks were folded, mildly uplifted, and eroded.
line of the Western Interior Seaway (Fig. 2) was positioned The peneplain that resulted is now represented by a regional
about 75 miles west of what is now northwestern Colorado unconformity between the Upper Cretaceous and Paleocene
(Zapp and Cobban, 1960), and in the area of the coal field, rocks. As the orogeny intensified in earliest Tertiary, the Park
marine mud (Mancos Shale), was accumulating. The lowest Range was uplifted and eroded, and a thin veneer of gravel
unit of the Mesaverde, the Tow Creek Sandstone Member of the spread westward as an alluvial fan over the erosional surface.
Iles Formation, was deposited when an eastward migration of This deposit now forms the basal conglomerate of the Pale-
the sea allowed nearshore sandstone to be deposited over the ocene Fort Union Formation. With the uplift of the Park Range
marine mud (Tow Creek regession). As this process continued, and other Laramide structures, the ancestral Sand Wash Basin
coal-bearing sediments of the Iles prograded eastward over the was defined, and all Tertiary formations younger than the Fort
Tow Creek. Minor fluctuations are represented by the double- Union are finer grained and show lateral facies changes consis-
ledge and Oak Creek sandstones in the eastern part of the coal tent with deposition in a slowly subsiding basin. During middle
field. Toward the end of Iles time, the shoreline transgressed and late Tertiary, extensional tectonism resulted in numerous
westward flooding the area with sea water, and marine mud was normal faults and volcanism in the region.
deposited over the nonmarine sediments. This event is repre-
sented by the marine rocks that underlie the Trout Creek Sand- Structural Features of the Yampa Coal Field
stone. At end of Iles time, the shoreline again moved eastward
across the coal field and the nearshore marine sand of the Trout The Yampa coal field occupies the southeastern corner of
Creek was deposited (Siepman, 1985). the Sand Wash Basin (Fig. 6). In the western and central parts
The lower Williams Fork sediments (middle coal group) of the coal field, the structural dip is toward the north, but far-
prograded eastward over the Trout Creek and accumulated just ther to the east, the regional structure swings counterclockwise
behind the strandline or low on the coastal plain. On the eastern until in the northeastern part of the coal field the dip is toward
side of the coal field, a minor westward shift of the shoreline the west. Another regional structure of significance in the area
caused an interval of marine sediments to be deposited, fol- of the coal field is the northwest-trending Axial Basin anticline
124 M. E. Brownfield et al.
R 93 W. R 92 W. R 91 W. R 90 W. R 89 W. R 88 W. R 87 W. R 86 W. R 85 W.
T. 8 N.
107 30'
40 37' 30''
7000
3500
107
4500 40 30'
5000
4000 6500
T. 6 N.
50 5000
00
7000
5500
6500
6000 6500
5500
6000 6000
6000 70
7000 00 5500 5500
0
600
8000 6000
00
6000 8000 7000 70
7000 T. 4 N.
8000 8000
7000
Surface outcrop of the 8000
Trout Creek Sandstone 8000
8500 40 15'
Fault, bar and ball on
downthrown side
0 5 mi T. 3 N.
3500 Structural contour
(contour interval 500 ft)
Figure 6. Structure contour map drawn on the top of the Trout Creek Sandstone Member of the Iles
formation Yampa coal field. Surface outcrop of the Trout Creek modified after Tweto (1976).
(Fig. 7). This structure defines the western boundary of the coal folds are the Round Bottom syncline, the Williams Fork anti-
field, and separates the Sand Wash Basin on the north from the cline, the Big Bottom syncline, and the Breeze Mountain-
Piceance Basin on the south. According to Stone (1986), this Buck Peak anticline.
fold is a minor part of a much larger tectonic structure that In the southernmost part of the coal field, folds trend and
extends from the Uinta Mountains in northeastern Utah to the plunge toward the northwest with their axial planes inclined
Eagle Basin in north-central Colorado. toward the northeast following the structural pattern in the west-
Folding in the Yampa coal field occurred after deposition ern part of the coal field. The dominant fold in this area is the
of the Lance Formation but before deposition of the Fort Hart syncline. This structure is bounded on the north by the
Union Formation (Tweto, 1976). In general, folds within the Beaver Creek anticline and on the south by the Seely anticline.
coal field can be grouped into a western group and an eastern The axes of these folds trend northwest, but the axis of the Hart
group (Fig. 7). On the east side of the coal field, southeast of syncline and the Seely anticline eventually veer westward.
Hayden, the fold axes trend north-northwest and north-north- Faulting in northwestern Colorado commenced during the
east, and plunge in a southerly direction. These folds are middle or late Tertiary, and in the Sand Wash Basin units as
asymmetrical with their axial planes inclined in a westerly young as the Miocene Browns Park Formation have been dis-
direction. Starting from the west, the more significant folds placed (Tweto, 1976). Undoubtedly, faulting continued into the
are the Sage Creek anticline, the Fish Creek anticline, the Quaternary and the region still experiences rare, mild earth-
Twentymile Park syncline, and the Tow Creek anticline. On quakes. Most of the faults in the coal field are high angle nor-
the western side of the coal field, south and west of Craig, fold mal faults that trend northwest (Fig. 7). Displacements are
axes trend and plunge toward the northwest. Most of the folds down to the northeast or southwest, and horst and grabben
are also asymmetrical with their axial planes inclined toward structures are common. Overall, faulting has not disrupted the
the northeast. Starting from the west, the more significant gross structure of the coal field to any significant degree.
R. 94 W. R. 93 W. R. 92 W. R. 91 W. R. 90 W. R. 89 W. R. 88 W. R. 87 W. R. 86 W. R. 85 W.
Explanation T. 9 N.
Lay
Craig
Hayden
1 Milner
3 Mount Harris T. 6 N.
4
8
12 7
2 T. 5 N.
10
T. 4 N.
9 5
6 Oak Creek
40 15'
T. 3 N.
11
0 5 mi T. 2 N.
Figure 7. Generalized geology of the Yampa coal field showing the Iles and Williams Fork Forma-
tions, major faults, and folds. Modified after Tweto (1976).
126 M. E. Brownfield et al.
COAL QUALITY cent and a sulfur content ranging between 0.5 to 0.6 percent.
The caloric value is reported to range between 10,380 and
Most Cretaceous coal in the Yampa coal field is noncoking, 11,570 Btu/lb.
high-volatile C bituminous, but some subbituminous A, B, and The upper coal group is presently being mined in the west-
C is reported (Speltz, 1976). Rare anthracite is also found adja- ern part of the coal field. The coal beds currently being
cent to Tertiary igneous intrusions. Coal analyses (as-received extracted from the Trapper Mine are a mixture of subbitumi-
basis) from the lower coal group have a moisture content rang- nous A, B, and C, and high-volatile bituminous C (Trapper
ing from 6.3 to 12.2 percent, an ash yield ranging from 4.3 to Mine, oral commun., 1999). Coal beds mined during September
11.3 percent, a sulfur content ranging from 0.3 to 0.9 percent, 1998 have an ash yield of 7.05 percent and a sulfur content of
and a caloric value ranging from 11,090 to 12,560 Btu/lb. Coals 0.40 percent, with a caloric value of 9,930 Btu/lb (reported on
analyzed (as-received basis) from the middle coal group have a an as-received basis).
moisture content ranging from 7.7 to 11.8 percent, an ash yield A comparison of the mean content of trace elements of
ranging from 3.4 to 11.5 percent, a sulfur content ranging from environmental concern (1990 Clean Air Act Amendment) is
0.3 to 0.6 percent, and a caloric value ranging from 10,740 to shown in Table 1. The Lennox, Wadge, Wolf Creek coal beds
12,260 Btu/lb (Speltz, 1976). (eastern part of the coal field) of the middle coal group and the
The 3 major coal beds (Wolf Creek, Wadge, and Lennox) Q bed (Trapper Mine) of the upper coal group are compared to
in the middle coal group were sample from the eastern part of mean Cretaceous coal values for the Colorado Plateau (Affolter,
the Yampa coal field and analyzed. The Wolf Creek bed has an in press).
average moisture content of 8.1 percent, an ash yield of 10.0
percent, a sulfur content of 0.72 percent, and a caloric value of COAL RESOURCES
10,860 Btu/lb. The Wadge bed has an average moisture content
of 9.3 percent, an ash yield of 7.8 percent, a sulfur content of The Yampa coal field contains an estimated remaining net
0.6 percent, and a caloric value of 11,180 Btu/lb. The Lennox coal resource of about 76 billion short tons in beds equal to or
bed has an average moisture content of 8.9 percent, an ash yield greater than 1.2 ft (Johnson and others, in press). Coal classified
of 6.7 percent, a sulfur content of 2.6 percent, and a caloric as identified makes up about 46 percent with the remaining 54
value of 11,420 Btu/lb. percent classified as hypothetical (Wood and others, 1983).
Little information is available on the quality of coal in Most of the hypothetical coal lies deep in the eastern Sand
the middle coal group in the western part of the coal field, Wash Basin (Fig. 7).
but as-received data provided by the Eagle No. 9 Mine (Zook Although a significant amount of coal remains in the Yampa
and Tremain, 1997) report the Eagle’s F bed as subbitumi- coal field, this figure must be regarded with caution because it
nous B, with ash yields ranging between 5.0 and 10.4 per- does not reflect geologic, land use, and environmental restric-
Coal mining in the 21st century: Yampa coal field, northwest Colorado 127
tions that might limit coal availability. For example, about 48 Silverthorne and then turn north on State Highway (SH) 9 to
percent of the coal is under more than 3,000 ft of overburden and Kremmling. Proceed west on U.S. 40 to SH 134 (about 7
thus is unavailable for underground mining. In addition, coal miles), then west over Gore Pass to Toponas. Turn right on SH
beds that dip more than 12 degrees are unavailable for under- 131 and drive north through Yampa and Phippsburg to Oak
ground mining at any depth, and such resources in the vicinity of Creek. Turn west on Routt County Road (RC) 27 just north of
the Sage Creek, Fish Creek, and Tow Creek anticlines are con- Oak Creek.
siderable. But even more significant, about 90 percent of the coal Stops 1 and 2. Oak Creek and Edna Mine. Oak Creek is
is under more than 500 ft of overburden, and is thus unavailable located on the southeastern margin of the Yampa coal field and
for surface mining. In addition, thin coal beds would be spoiled is the eastern limit of outcrops of the Mesaverde Group. Fenne-
during surface mining and the coal in beds greater than 14 ft man and Gale (1906) conducted the first detailed study of the
would be left behind in underground mines. To conclude, it is Oak Creek area. Development of coal in the Oak Creek area fol-
apparent that only a small percentage of the 76 billion tons of lowed the arrival of the railroad in 1906. Several large mines
remaining net coal resource could be recovered. were developed north and west of Oak Creek and continued
Over the next decade most of the coal produced in the into the 1940’s. The Moffat Coal Company (Argo or Oak Hills
Yampa coal field will come from areas presently under lease Mine) and the Victor-American Fuel Company (Pinnacle Mine,
and from coal mines already in existence. The middle coal Fig. 8) mined the Pinnacle bed (No. 2 bed of Bass and others,
group will continue to dominate production in the eastern part 1955, Fig. 9). Most of the coal mined in the Oak Creek district
of the field while the upper coal group will continue to domi- was from the lower coal group of the Iles Formation.
nate in the western part of the field. Coal mining in northwest- The Edna Mine is located about 3 miles west of Oak Creek
ern Colorado will remain an important social and economic and east of Trout Creek. The mine produced coal from the
factor in northwest Colorado well into the 21st century. Wadge bed (Fig. 9). The Edna Mine is now abandoned and the
mine area was reclaimed. About 1.5 miles south along Trout
ITINERARY Creek, the Apex Mine produced coal from the No. 2 bed until it
closed in the middle 1980’s (Fig. 9). Fenneman and Gale (1906)
The following is a brief description of the stops to be vis- named the Trout Creek Sandstone Member of the Iles for expo-
ited on the trip. Weather permitting, the field trip will visit all sure on Trout Creek. The Trout Creek forms the prominent
stops, but some may be modified because of road conditions white sandstone just east of Trout Creek.
and mining operations. Figure 1 is a location map of the Yampa Within the Oak Creek and Edna Mine area, the Upper Cre-
coal field with the trip stops. taceous rocks are dominated by marine facies including off-
shore marine, lower and upper shoreface, estuarine, and mire.
Day 1. Denver to Steamboat Springs Gaffke (1979) studied a portion of the section located along the
road to the old Oak Creek dump. Figure 10 shows a composite
The trip begins at the Colorado Convention Center. Driving section with depositional environments interpreted for rocks
west. Drive west on I-70, through the Eisenhower Tunnel, to exposed in the southern part of the Edna Mine. The section was
128 M. E. Brownfield et al.
constructed from USGS coal exploration drill-hole data and About 2 miles west of Stop 2, at Middle Creek, is an excel-
data collected by the authors. The middle coal group is about lent exposure of rippled sandstone and siltstone below the
300 ft thick in the vicinity of the Edna Mine. Estuarine deposits Lennox coal bed. The ripples are symmetrical indicating oscil-
associated with a bay or lagoon dominate the section. Estuarine latory flow caused by small waves. The ripple marks are flat-
deposits contain brackish-water pelecypods and brachiopods. topped, indicating that they were eroded by small waves that are
Coals in this section are overlain by splay deposits, or in the best explained by tidal action. Associated with the wave ripples
case of the Wadge coal, overlain by brackish water sediments. are Arenicolites (a marine to brackish-water trace fossil) and sea
The splays have sharp contacts with the underlying coal sug- urchin burrows. The rippled deposit, therefore, indicates an
gesting that 1) the deposits are distal splays occurring in a low- intertidal (lagoonal) setting.
energy environment, and 2) the peat is resistant to erosion. The After examining the intertidal deposit, time permitting,
rocks underlying the coals have sharp contacts and are not walk west along the road (RC 27) and observe the facies change
rooted suggesting that the initial peat formation may have been from coastal plain to offshore marine.
in the form of floating mats.
Sh. (part)
Lewis
Alluvial and
coastal plain Alluvial and
coastal plain
Upper shoreface
Iles Formation
Lower Shoreface
and offshore marine
500 ft
(part)
Mancos
Stop 3. Eckman Park and Foidel Creek Mine. Eckman crops out along Oak Creek near Haybro, about 3 miles north of
Park is about 7 miles west of Oak Creek on RC 27 and is devel- Oak Creek. The Oak Creek Sandstone can be mapped through
oped on gently northward dipping strata of the middle coal Trout and Middle Creeks to the Fish Creek Canyon. In the
group of the Williams Fork Formation. Eckman Park is the site canyon the unit is about 85 ft thick. The strike valley above the
of the old Energy 1 (strip) and the present day Foidel Creek Oak Creek Sandstone is formed in the marine sediments that
(longwall) mines. The Energy 1 Mine produced coal from the underlie the Trout Creek Sandstone. The Trout Creek is well
Lennox, Wadge, and Wolf Creek coal beds and the Foidel Creek exposed in the section. Directly above the Trout Creek is car-
Mine produces coal from the Wadge. Using longwall mining bonaceous shale that is interpreted as a salt marsh and is over-
methods the Foidel Creek Mine cut three one-mile long panels lain by a washover fan consisting of rippled sandstone and
setting a monthly coal production world record in 1996 of just siltstone. The Yampa bed (Brownfield and Johnson, 1986), an
over one million tons (Wynn and Coates, 1998; Fiscor, 1998). altered volcanic ash, is exposed above the washover fan. The
Within Eckman Park the Trout Creek Sandstone is con- Yampa was found within 5 feet of the Trout Creek Sandstone
formably overlain by a 330 ft thick section of the middle coal northeast of Eckman Park but was not preserved in the Edna
group. The Twentymile Sandstone Member of the Williams Mine section. The Wolf Creek coal bed is burned in the canyon
Fork Formation forms massive white sandstone cliffs on the but the Wadge bed (about 11.8 ft thick) was mined for a short
north and west sides of Eckman Park. The Twentymile overlies period while the Lennox bed was not found. The Twentymile
a 600 ft thick sequence of marine sediments containing Late Sandstone forms a massive white cliff and dip slope on the east-
Campanian Baculites reesidei (Izett and others, 1971). Underly- ern end of the canyon and is about 185 ft thick. This shoreface
ing the marine shale section is about 330 ft of the middle coal sandstone and the thick occurrences of the Trout Creek are
group sediments. The overlying coal-bearing rocks (upper coal much thicker than their modern counterparts. Siepman (1985)
group) are about 190 feet thick and contain the Fish Creek coal believes that these very thick sandstone units represent vertical
bed (Fig. 8). The Fish Creek coal bed ranges from 1 to 6 ft thick stacking of shoreface sand units. The overlying sediments of the
and was mined by Energy Fuels Corporation in the 1970’s. The Williams Fork Formation represent lagoon and tidal inlet.
Lewis Shale conformably overlies the Williams Fork. The After examining the Fish Creek Canyon section return to
Twentymile Sandstone, upper coal group, and Fish Creek sand- the Foidel School and drive north on RC 33 to Steamboat
stone are well exposed in the rail cut in Foidel Creek Canyon Springs. Between the Foidel Creek Canyon and Steamboat
(Fig. 1). Springs the road will cross both limbs of the Twentymile Park
Continue west on RC 27 to the junction with RC 33 and syncline (Fig. 7). Just north of the bridge crossing the Yampa
turn left passing the old Foidel Creek School. Driving south on River in Steamboat Springs is a faulted outcrop of the Lower
RC 27 the Twentymile Sandstone forms massive cliffs to the Cretaceous Dakota Sandstone and some associated hot springs.
right and the Foidel Creek Mine facilities are on the left. Normal
faults can be seen offsetting the top of the Twentymile. Pinnacle Day 2. Steamboat Springs to Craig
Peak to the south is capped by the Twentymile Sandstone and
bounded on the north and south by faults. RC 27 turns west Continue on west U.S. 40 from Steamboat Springs. The trip
passing through the Twentymile Sandstone. Drive west to RC will be traveling through the Lower and Upper Cretaceous Man-
37 and turn left following Fish Creek southwest to Fish Creek cos Shale and Upper Cretaceous Mesaverde Group. About 13
Canyon. miles west of Steamboat Springs is the town of Milner (Fig. 1).
Stop 4. Fish Creek Canyon. Fish Creek Canyon contains Milner is located in the northern extension of the Twentymile
one of the best stratigraphic sections (Fig. 11) of the uppermost Park syncline (Fig. 7). Several old mines were developed in the
Mancos Shale and the Mesaverde Group in the eastern part of Wadge bed along the flanks of the syncline. Just north of Milner
the coal field. On the west end of Fish Creek Canyon there are is the site of McGregor where one of the first coal fired electric
excellent exposures of the marine sandstone units in the upper plants was built in northwest Colorado. U.S. 40 crosses the
Mancos Shale. The canyon section is dipping about 40° to the asymmetrical Tow Creek anticline about 2.5 miles west of Mil-
east on the western end and increases to 55° at the eastern end. ner. Oil was discovered in 1924 and production is from the frac-
The Iles Formation within the canyon is dominated by 4 progra- tured Niobrara. Most of the wells are now abandoned north of
dational shoreface sandstone units: the Tow Creek Sandstone the highway but a few are still maintained to the south.
Member, the double-ledge sandstone (Bass and others, 1955), Stop 5. Mount Harris. Mount Harris was an important coal
the Oak Creek sandstone (Kucera, 1959), and the Trout Creek mining town from about 1915 to the early 1950’s (Fig. 12). Two
Sandstone Member. The Tow Creek Sandstone Member of the large mines were developed in the Wadge bed south of the
Iles Formation is the basal unit of the Iles Formation and is Yampa River: the Colorado and Utah Coal Company’s Harris
about 70 ft thick. About 340 ft above the Tow Creek is a massive Mine and the Victor-American Fuel Company’s Wadge Mine
ledge-forming sandstone unit called the double-ledge sandstone, (Campbell, 1923). These mines were closed after mining acci-
which is 250 ft thick. Kucera (1959) applied the name Oak dents killed several miners.
Creek Sandstone to a 100 ft thick cliff-forming sandstone that The Mount Harris section has been studied several times
130 M. E. Brownfield et al.
Sh. (part)
Lewis
consists of offshore marine shale and siltstone with minor sand-
stone about 550 ft thick. The Twentymile sandstone is about 160
ft thick and overlain by 230 ft of upper coal group rocks. The
Twentymile Ss. Mbr. depositional environments related to the Twentymile include
lower shoreface, middle shoreface, upper shoreface, tidal chan-
nel, foreshore, and backshore. The upper coal group consists of
the mine visit will be a tour of the mine property observing mine
Tow Creek Ss. Mbr
reclamation and coal deposits in the mine. Remember once
Sh. (part)
Figure 12. Tipple of the Colorado and Utah Coal Company’s Harris Mine and the town of Mount Harris from the south
side of the Yampa River. Photograph by L. C. McClure (Campbell, 1923). Twentymile Sandstone Member of the
Williams Fork Formation forms the massive sandstone cliff in the background.
Figure 13. Looking north from the Trapper Mine at the Craig power plant located near the center of
the photo. Cedar Mountain, to the left of the plant, consists of Miocene Browns Park Formation
capped by basalt.
during September 1998 are ash yield 7.05 percent and sulfur Twentymile Sandstone and upper coal group at the Trapper
content 0.4 percent, with a caloric value of 9,931 Btu/lb on an Mine area consist of shoreface, foreshore, lagoon, washover
as received basis. fan, splay, fluvial channel, marsh, and freshwater swamp
The upper coal group was deposited along wave-domi- deposits.
nated, progradational shoreface, and coastal plain setting. The Return to Denver and the Convention Center.
132 M. E. Brownfield et al.
REFERENCES CITED ________1906b, The Yampa coal field, Routt County, Colorado: U.S. Geolog-
ical Survey Bulletin 297, 96 p.
Affolter, R.H., in press, Quality characterization of the United States western ________1906c, The Yampa coal field, Routt County, Colorado: Mining
Cretaceous coal from the Colorado Plateau resource assessment area: reporter, v. 54, p. 251-252.
U.S. Geological Survey Professional Paper Fiscor, S., 1998, U.S. longwalls thrive: Coal Age, v. 103, no. 2, p. 22-27.
Bass, N.W., Eby, J.B., and Campbell, M.R., 1955, Geology and mineral fuels of Gaffke, T.M., 1979, Depositional environments of a coal-bearing section in the
parts of Routt and Moffat Counties, Colorado: U.S. Geological Survey Upper Cretaceous Mesaverde Group, Routt County, Colorado: U.S.
Bulletin 1027-D, p. 143-250. Geological Survey Open-File Report 79-1669, 15 p.
Boreck, D.L., and Murray, D.K., 1979, Colorado coal reserve depletion data Gale, H.S., 1909, Coal fields of northwestern Colorado and northeastern Utah:
and coal mine summaries: Colorado Geological Survey Open-File U.S. Geological Survey Bulletin 341-C, p. 283-315.
Report 79-1, 65 p. ________1910, Coal fields of northwestern Colorado and northeastern Utah:
Boyles, M.J., Kauffman, E.G., Kiteley, L.W., and Scott, A.J., 1981, Deposi- U.S. Geological Survey Bulletin 415, 265 p.
tional systems Upper Cretaceous Mancos Shale and Mesaverde Group, Hail, W.J., Jr., 1974, Geologic map of the Rough Gulch quadrangle, Rio Blanco
northwestern Colorado: Society of Economic Paleontologists and Min- and Moffat Counties, Colorado: U.S. Geological Survey Geologic
eralogists, Rocky Mountain Section, Field Trip Guidebook, 146 p. Quadrangle Map GQ-1195, scale 1:24,000.
Brownfield, M.E., and Anderson, K., 1979, Geologic map and coal sections of Hancock, E.T., 1925, Geology and coal resources of the Axial and Monument
the Lay SE quadrangle, Moffat County, Colorado: U.S. Geological Sur- Butte quadrangles, Moffat County, Colorado: U.S. Geological Survey
vey Open-File Report 79-1680, scale 1:24,000. Bulletin 757, 134 p.
Brownfield, M.E., and Anderson, K., 1988, Geologic map and coal sections of Hancock, E.T., and Eby, J.B., 1930, Geology and coal resources of the Meeker
the Lay SE quadrangle, Moffat County, Colorado: U.S. Geological Sur- quadrangle, Moffat and Rio Blanco Counties, Colorado: U.S. Geologi-
vey Coal Investigation Map C-117, scale: 1:24,000. cal Survey Bulletin 812-C, p. 191-242.
Brownfield, M.E., and Prost, G.L., 1979, Geologic map and coal sections of the Haun, J.D., and Weimer, R.J., 1960, Cretaceous Stratigraphy of Colorado, in,
Lay quadrangle, Moffat County, Colorado: U.S. Geological Survey Weimer, R.J., and Haun, J.D., eds., Guide to the geology of Colorado:
Open-File Report 79-1679, scale: 1:24,000. Geological Society of America, Rocky Mountain Association of Geolo-
Brownfield, M.E., and Johnson, E.A., 1986, A regionally extensive altered air- gists, and Colorado Scientific Society Guidebook, p. 58-65.
fall ash for use in correlation of lithofacies in the Upper Cretaceous Hayden, F.V., 1877, Explorations made in Colorado under the direction of Pro-
Williams Fork Formation, northwestern Piceance Creek and southern fessor Ferdinand V. Hayden in 1876: American Naturalist, vol. XI, no. 2,
Sand Wash Basin, Colorado, in, Stone, D. S., ed., New interpretations of p. 73-86.
northwest Colorado geology: Rocky Mountain Association of Geolo- Hildebrand, R.T., Garrigues, R.S., Meyer, R.F., and Reheis, M.C., 1981, Geol-
gist, p. 165-169. ogy and chemical analyses of coal and coal-associated rock samples,
Brownfield, M.E., Roberts, L.N.R., Johnson, E.A., and Mercier, T.J., 1998, Williams Fork Formation (Upper Cretaceous), northwestern Colorado:
Assessment of the distribution and resources of coal in the Deserado U.S. Geological Survey Open-File Report 81-1348, 94 p.
coal area, Lower White River coal field, northwest Colorado: U.S. Geo- Holmes, W.H., 1877, Report (on the San Juan District, Colorado): U.S. Geo-
logical Survey Open-File Report 98-352, 28p. logical and Geographical Survey of the Territories, 9th Annual Report
Brownfield, M.E., Roberts, L.N.R., Johnson, E.A., and Mercier, T.J., in press, for 1875, p. 237-276.
Assessment of the distribution and resources of coal in the Deserado Horn, G.H., and Richardson, E.E., 1956, Geologic and structure map of the
coal area, Lower White River coal field, northwest Colorado: U.S. Geo- Williams Fork Mountains coal field, Moffat County, Colorado: U.S.
logical Survey Professional Paper. Geological Survey unnumbered map, scale 1:24,000.
Brownfield, M.E., Roberts, L.N.R., Johnson, E.A., and Mercier, T.J., in press, Izett, G.A., Cobban, W.A., and Gill, J.R., 1971, The Pierre Shale near Kremm-
Assessment of the Distribution and Resources of Coal in the Fairfield ling, Colorado, and its correlation to the east and west: U.S. Geological
Coal Group, Danforth Hills Coal Field, Northwest Colorado: U.S. Geo- Survey Professional Paper 684-A, 19 p.
logical Survey Professional Paper Johnson, E.A., 1987, Geologic map and coal sections of the Round Bottom
Campbell, M.R., 1906, Character and use of Yampa coals, in, Fenneman, N.M., quadrangle, Moffat County, Colorado: U.S. Geological Survey Coal
and Gale, H.S., The Yampa coal field, Routt County, Colorado: U.S. Investigations Map C-108, scale 1:24,000.
Geological Survey Bulletin 297, p. 82-91. Johnson, E.A., and Brownfield, M.E., 1984, Selected references on the geology
________1912, Miscellaneous analyses of coal samples from various fields of of the Yampa coal field and Sand Wash Basin, Moffat, Routt, and Rio
the United States: U.S. Geological Survey Bulletin 471-J, p. 629-655. Blanco Counties, Colorado: U.S. Geological Survey Open-File Report
________1923, The Twentymile Park district of the Yampa coal field, Routt 84-769, 42 p.
County, Colorado: U.S. Geological Survey Bulletin 748, 82 p. Johnson, E.A., Roberts, L.N.R., and Brownfield, M.E., in press, Geology and
Chisholm, F.F., 1887, The Elk Head anthracite coal field of Routt County, Col- resource assessment of the middle and upper coal groups in the Yampa
orado: Proceedings of the Colorado Scientific Society, v. 2, p. 147-149. coal field, northwestern Colorado: U.S. Geological Survey Professional
Collins, B.A., 1976, Coal deposits of the Carbondale, Grand Hogback, and Paper.
southern Danforth Hills coal field, eastern Piceance Basin, Colorado: Kerr, B.G., 1958, Geology of the Pagoda area, northwestern Colorado: Golden,
Quarterly of the Colorado School of Mines, v. 71, no. 1, 138 p. Colorado, Colorado School of Mines Masters thesis, 124 p.
Crawford, R.D., Willson, K.W., and Perini, V.C., 1920, Some anticlines of Routt Kitely, L.W., 1983, Paleogeography and eustatic-tectonic model of Late Cam-
County, Colorado: Colorado Geological Survey, Bulletin 23, 59 p. panian (Cretaceous) sedimentation, southwestern Wyoming and north-
Eakins, W., and Coates, M.M., 1998, Focus: Colorado coal: Colorado Geologi- western Colorado, in, Reynolds, M. W., and Dolly, E. D., eds., Mesozoic
cal Survey, Rock Talk, v. 1, no. 3, p. 1-4. paleogeography of the west-central United States: Rocky Mountain
Eby, J.B., 1924, Coal in Elkhead District of Yampa coal field, northwestern Col- Paleogeography Symposium 2, Society of Economic Paleontologists
orado: U.S. Geological Survey Press Notice 16653. and Mineralogist, Rocky Mountain Section, p. 273-302.
________1925, Contact metamorphism of some Colorado coals: American Konishi, K, 1959, Upper Cretaceous surface stratigraphy, Axial Basin and
Institute of Mining and Metallurgical Engineers Transactions, v. 71, p. Williams Fork area, Moffat and Routt Counties, Colorado, in, Haun,
250. J.D., and Weimer, R.J., eds, Symposium on Cretaceous rocks of Col-
Fenneman, N.M., and Gale, H.S., 1906a, The Yampa coal field, Routt County, orado and adjacent areas, 11th annual field conference: Rocky Mountain
Colorado: U.S. Geological Survey Bulletin 285-F, p. 226-239. Association of Geologists, p. 67-73.
Coal mining in the 21st century: Yampa coal field, northwest Colorado 133
Kucera, R.E., 1959, Cretaceous stratigraphy of the Yampa district, north- Paper 1561, 115 p.
western Colorado, in, Haun, J.D., and Weimer, R.J., eds., Sympo- Roehler, H.W., 1987, Surface -subsurface correlations of the Mesaverde Group
sium on Cretaceous rocks of Colorado and adjacent areas, 11th and associated Upper Cretaceous formations, Rock Springs, Wyoming,
annual field conference: Rocky Mountain Association of Geologists, to Mount Harris, Colorado: U.S. Geological Survey Miscellaneous Field
p. 37-45. Studies map MF-1937.
________1962, Geology of the Yampa District, northwestern Colorado: Boul- Roehler, H.W., and Hansen, D.E., 1989, Surface and subsurface correlations
der, Colorado, University of Colorado Ph.D. dissertation, 675 p. showing depositional environments of the Upper Cretaceous Mesaverde
Lakes, A, 1903, Coal and asphalt deposits along the Moffat railroad: Mines and Group and associated formations, Cow Creek in southwestern Wyoming
Minerals, v. 24, p. 134-136. to Mount Harris in northwest Colorado: U.S. Geological Survey Mis-
________1904, The Yampa coal field; a description of the anthracite, bitumi- cellaneous Field Studies map MF-2077.
nous, and lignite field transversed by the Moffat road in Routt County: Ryer, T.A., 1977, Geology and coal resources of the Foidel Creek EMRIA site
Mines and Minerals, v. 24, p. 249-251. and surrounding area, Routt County, Colorado: U.S. Geological Survey
________1905a, The Yampa coal field of Routt County: Mining Reporter, v. Open-File Report, 77-303, 31 p.
51, p. 404-405. Sears, J.D., 1925, Geology and oil and gas prospects of part of Moffat County,
Landis, E.R., 1959, Coal reserves of Colorado: U.S. Geological Survey Bulletin Colorado, and southern Sweetwater County, Wyoming: U.S. Geological
1072-C, p. 131-232. Survey Bulletin 751, p. 269-319.
Lauman, G.W., 1965, Geology of Iles Mountain area, Moffat County, north- Siepman, B.R., 1985, Stratigraphy and petroleum potential of trout Creek and
western Colorado: Golden, Colorado, Colorado School of Mines Mas- Twentymile Sandstones (Upper Cretaceous), Sand Wash Basin, Col-
ters thesis, 129 p. orado: Colorado School of Mines Quarterly, v. 80, no. 2, 59 p.
Massoth, T.W., 1982a, Depositional environments of a surface coal mine in Speltz, C.N., 1976, Strippable coal resources of Colorado - location, tonnage,
northwest Colorado, in, Gurgel, K.D., ed., Proceedings, 5th symposium and characteristics of coal and overburden: U.S. Bureau of Mines Infor-
on the geology of Rocky Mountain coal 1982: Utah Geological and mation Circular 8713, 70 p.
mineral Survey Bulletin 118, p. 115-120. Stone, D.S., 1986, Seismic and borehole evidence for important pre-Laramide
________1982b, Depositional environments of some Upper Cretaceous coal- faulting along the Axial arch in northwest Colorado, in, Stone, D. S.,
bearing strata at Trapper Mine, Craig, Colorado: University of Utah ed., New interpretations of northwest Colorado geology: Rocky Moun-
Masters thesis, 124 p. tain Association of Geologist, p. 19-36.
Masters, C.D., 1959, Correlation of the post-Mancos Upper Cretaceous sedi- Tremain, C.M., Hornbaker, A.L., Holt, R.D., Murray, D.K., and Ladwig, L.R.,
ments of the Sand Wash and Piceance Basins, in, Haun, J.D., and 1996, 1996 summary of coal resources in Colorado: Colorado Geologi-
Weimer, R.J., eds., Symposium on Cretaceous rocks of Colorado and cal Survey Special Publication 41, 19 p.
adjacent areas, 11th annual field conference: Rocky Mountain Associa- Tweto, O.,1976, Geologic map of the Craig 1o X 2o quadrangle, northwestern
tion of Geologists, p. 78-80. Colorado: U. S. Geological Survey Miscellaneous Investigations Series
________1966, Sedimentology of the Mesaverde Group and the upper part of I-972, scale 1:250,000.
the Mancos Formation, northwest Colorado: Yale University, Ph.D. dis- ______ 1979, Geologic map of Colorado: U. S. Geological Survey, scale
sertation, 88 p. 1:500,000.
________1967, Use of sedimentary structures in determination of depositional Wood, G.H., Kehn, T.M., Carter, M.D., and Culbertson, W.C., 1983, Coal
environments, Mesaverde Formation, Williams Fork Mountains, Col- resource classification system of the U.S. Geological Survey: U. S.
orado: American Association of Petroleum Geologists Bulletin, v. 51, Geological Survey Circular 891, 65 p.
no. 10, p. 2033-2043. Zook, J.M., and Tremain, C.M., 1997, Directory and statistics of Colorado coal
Resource Data International, 1998, COALdat data base: 1320 Pearl Street, Suite mines with distribution and electric generation map, 1995-96: Colorado
300, Boulder, Colorado 80302. Geological Survey Resource Series 32, 55 p.
Roberts, L.N., and Kirschbaum, M.A., 1995, Paleogeography of the late Creta- Zapp, A.D. and Cobban, W.A., 1960, Some Late Cretaceous strand lines in
ceous of the Western Interior of Middle North America - coal distribu- northwestern Colorado and northeastern Utah: U.S. Geological Survey
tion and sediment accumulation: U.S. Geological Survey Professional Professional Paper 450-D, p. 52-55.
Printed in U.S.A.
Geological Society of America
Field Guide 1
1999
This guide consists of three general sections: an introduc- globe have been reported where abundance anomalies of Ir
tion that includes discussions of Raton basin stratigraphy and occur in rock layers at the major extinction horizon of marine
the CretaceousTertiary (K-T) boundary; descriptions of the invertebrate and continental plant and animal fossils that
geology along the route from Denver, Colorado, to Raton, New defines the K-T boundary.
Mexico; and descriptions of several K-T sites in the Raton Significantly, anomalously high concentrations of Ir were
basin. Much of the information is from previous articles and first discovered at the K-T boundary in continental rocks in
field guides by the authors together with R. M. Flores and from 1981. A team consisting of chemists from Los Alamos National
road logs co-authored with Glenn R. Scott, both of the U.S. Laboratory aided by geologists from the U.S. Geological Sur-
Geological Survey. vey (USGS) reported an Ir abundance anomaly precisely at the
palynological K-T boundary in nonmarine rocks in the Raton
INTRODUCTION basin, New Mexico (Orth and others, 1981). The discovery was
made in core samples from a hole drilled specifically for paly-
Geologists have recognized for nearly a century that a nological and chemical analyses at the York Canyon coal mine
remarkable mass-extinction event occurred at the end of the about 35 miles west of Raton. Prior to this, all reports concern-
Cretaceous Period, about 65 million years ago. Thousands of ing the hypothesized asteroid impact at the K-T boundary were
species of plants and animals that lived during the Cretaceous from marine rocks, and some geologists had suggested that
no longer existed during the Tertiary Period immediately fol- marine processes concentrated the anomalous Ir; therefore, the
lowing. This disappearance has led to much speculation and Ir need not be extraterrestrial The discovery of an Ir abundance
controversy among scientists as to possible causes. Among the anomaly at the K-T boundary in nonmarine rocks in the Raton
causes suggested are changes in climate or sea level, neither of basin removed this objection and provided important supporting
which has been universally accepted. In 1980, researchers from evidence for the Berkeley team’s hypothesis. In later studies,
the University of California at Berkeley advanced the startling shock-metamorphosed quartz grains were found coincident
hypothesis that a large asteroid about 10 km in diameter struck with the Ir anomaly (Bohor and others, 1984; Pillmore and oth-
the Earth about 65 million years ago, causing a worldwide bio- ers, 1984; Izett and Pillmore, 1985a, b); these unusual grains
spheric catastrophe (Alvarez and others, 1980). are further evidence of a major impact event. Shock-metamor-
The Berkeley team found anomalously high concentrations phosed mineral grains are found only at known meterorite
of iridium (Ir) and other noble elements in a claystone layer that impact sites or nuclear bomb craters, never in volcanic rocks.
marks the K-T boundary in certain marine rocks in Italy, Den- Thorough and quite readable accounts of the K/T boundary
mark, and New Zealand. Ir is extremely rare in the Earth’s crust impact and extinction theory and the evidence in support of it
relative to its abundance in certain types of meteorites. Because can be found in two recently published, popular books: Alvarez
of this crustal Ir deficiency, the team proposed that the source of (1997) and Powell (1998); these books are highly recom-
the anomalously high Ir concentrations in the boundary layer mended.
was extraterrestrially derived and that the layer was formed Since the original Raton basin discovery, many more K-T
from fallout of ejecta after the asteroid impact occurred. Since boundary sites have been found in the eastcentral and southern
this original proposal, more than 100 sites scattered around the parts of the basin (Fig. 1), and other continental K-T sites have
Pillmore, C. L., Nichols, D. J., and Fleming, R. F., 1999, Field guide to the Continental Cretaceous-Tertiary boundary in the Raton basin, Colorado and
New Mexico, in Lageson, D. R., Lester, A. P., and Trudgill, B. D., eds., Colorado and Adjacent Areas: Boulder, Colorado, Geological Society of America Field
Guide 1.
135
Raton
Ca
na
Basin dian
Ve
rm
Vermejo
ejo
Park
Figure 1. Index map showing location of Raton basin and representative Cretaceous-Tertiary boundary sites in New
Mexico and Colorado. Columnar sections at or near top of lower zone of Raton Formation show lithology of the
boundary interval at selected sites: YCC, York Canyon; RAT, Raton; SUG, Sugarite; SVS, Starkville South; BER,
Berwind Canyon; and MAD, Madrid. Arrows indicate the K-T boundary. Measured Ir anomalies shown in ng/g
(10-9 g/g). Hachured line shows top of Trinidad Sandstone. Modified from Pillmore and Flores (1987).
Continental Cretaceous-Tertiary boundary in Raton basin, Colorado and New Mexico 137
been located in Wyoming, Montana, North Dakota, Alberta, and which overlies and intertongues with the Raton Formation.
Saskatchewan. On this trip we will visit several of the K-T Depositional and tectonic history. The formations present
boundary sites in the Raton basin to observe similarities and in the Raton basin are the Pierre Shale and Trinidad Sandstone,
differences in the character of the boundary sequence and the the Vermejo and Raton Formations, and the Poison Canyon For-
enclosing interval of rocks. mation as portrayed in the stratigraphic section (Fig. 2). Their
depositional and tectonic history is shown in Figure 3. The Cre-
RATON BASIN taceous epeiric sea covered the area during most of Late Creta-
ceous time. The sea was filled by a thick sequence of
The Raton basin (Fig. 1) is a large, asymmetric syncline calcareous, deep marine or basinal shales, overlain by the Pierre
(2,500 sq. mi. area) that extends from Huerfano Park, Colorado, Shale, 1,800 to 1,900 feet of shallow marine or shelf shale and
to Cimarron, New Mexico. The Cretaceous and Tertiary rocks siltstone. These beds range in age from Cenomanian to Maas-
dip steeply and form hogbacks along the western margin of the trichtian as determined from ammonite fossils studied by G.R.
basin on the east flank of the Sangre de Cristo Mountains. Scott and W.A. Cobban (G.R. Scott, written commun.). Most
These rocks dip more gently inward along the other margins of of the Pierre consists of gray, noncalcareous shale that coarsens
the Raton basin and are highly dissected. upward into silty shale and siltstone, reflecting the influx of silt
Stratigraphy. The sedimentary rocks of the Raton basin are as the paleoshoreline of the sea retreated to the east across Col-
shown in Figure 2. The marine Pierre Shale (Campanian to orado and New Mexico in late Campanian and early Maas-
Maastrichtian) and overlying marginal-marine Trinidad Sand- trichtian time. The upper part of the Pierre grades into the
stone (Maastrichtian) underlie the nonmarine Upper Cretaceous overlying Trinidad Sandstone
and Tertiary rocks in the basin. Nonmarine sedimentary rocks The Trinidad Sandstone is a tabular body, in most places
in the Raton basin, from oldest to youngest, include the coal- about 80-100 ft thick, composed mostly of fine to very fine-
bearing Vermejo Formation (Maastrichtian) and Raton Forma- grained sandstone that contains Ophiomorpha, Diplocraterion,
tion (Maastrichtian and Paleocene), and the non-coal-bearing and other trace fossils. It forms persistent, conspicuous, light-
Poison Canyon Formation (also Maastrichtian and Paleocene), colored cliffs at the east and south edges of the basin. The
Figure 2. Columnar section of rocks in the Raton basin. Thickness of zones in the Raton Formation
are: lower coal zone, 100-300 ft; barren series, 180-700 ft; and upper coal zone, 590-1100 ft.
138 C. L. Pillmore, D. J. Nichols, and R. F. Fleming
Figure 3. Diagrammatic block diagrams depicting paleoenvironments of Late Cretaceous and Pale-
ocene rocks of the Raton basin: A. The offshore environments of Pierre Shale; the contemporaneous
delta front and barrier environments of the Trinidad Sandstone; and the fluvio-deltaic plain of the
Vermejo Formation. Diagram shows oxbow lakes, crevasse splays, and coal-forming swamps related
to meandering streams of the lower delta plain. B. The floodplain and fine-grained meander-belt
environments on the alluvial plain that developed during the Late Cretaceous after deposition of the
basal conglomerate of the lower coal zone of the Raton Formation. Coal-forming swamps and oxbow
lakes characterized the floodplain. C. The depositional environments on the alluvial plain of the
lower coal zone of the Raton Formation at the end of the Cretaceous Period. The K-T boundary fall-
out material was deposited on a surface such as that shown here. The letters show representative
depositional sites: B, Berwind site (floodplain sequence, dominated by crevasse splays from nearby
stream channel); M, Madrid site (channel-floodplain-crevasse splay sequence); S, Starkville sites
(floodplain sequence developed on abandoned channel sequence); Y, York Canyon (floodplain-
crevasse splay sequence); R, Raton site (floodplain, marginal to major backswamp); SU, Sugarite
site (dominated by an extensive swamp). D. The braided stream and coarse meander-belt environ-
ments of the Paleocene barren series of the Raton Formation.
Trinidad was deposited in contemporaneous delta-front and steep, generally debris-covered slopes above the cliffs of the
interdeltaic barrier-bar environments as the sea continued to Trinidad Sandstone. The Vermejo varies in thickness from 370
regress eastward. The delta-front deposits include distributary to 390 along the western border of the Raton basin to 0 ft in the
mouthbar and distributary channel sandstone beds (Flores and eastern part of the basin, south and east of Raton, New Mexico.
Tur, 1982). The barrier-bar deposits consist of middle It contains coal beds as thick as 10-13 ft near the top and bot-
shoreface, river-estuarine-inlet, and beach sandstone beds tom of the formation (Pillmore, 1976). The sedimentary rocks
(Leighton, 1980) that are overlain by the fluvial-deltaic and and coal beds of the Vermejo were deposited in contemporane-
back-barrier deposits of the Vermejo Formation. ous fluvio-deltaic and back-barrier coastal plains fronted by
The Vermejo Formation consists of interbedded sandstone, barrier-bar and delta-front sandstone of the Trinidad (Flores and
siltstone, shale, carbonaceous shale, and coal that together form Tur, 1982). Lower alluvial plains dissected by meandering
Continental Cretaceous-Tertiary boundary in Raton basin, Colorado and New Mexico 139
streams separated by flood basins characterized the upper part sition and preservation of the impact ejecta and fallout resulting
of the Vermejo. Here, sand-filled stream channels and fine- from the K-T boundary asteroid impact.
grained sequences of silt and mud were laid down in floodplains Shortly after the close of the Cretaceous, tectonic condi-
associated with crevasse-splay and minor crevasse-channel sand- tions in the Raton basin changed as uplift of the source area was
stone. Coal beds in the lower part of the formation formed in reinitiated in the west (Fig. 3c). It is probable that episodic
poorly drained, back-barrier, coastal swamps and in swamps upthrusting along the fault belt to the west created extensive
adjacent to distributary channels of delta plains. These alluvial erosion and sediment input into the basin. Sediment load
deposits grade upward into the more landward deposits of the increased and a fluvial system characterized by braided streams
Raton Formation. merging basinward into meandering streams and well-drained
The Raton Formation contains the K-T boundary interval. flood basins once again characterized the depositional basin.
The formation consists of sandstone, siltstone, mudstone, coal, Streams aggraded broad belts across the alluvial plain, result-
carbonaceous shale, and conglomerate. It ranges in thickness ing in sheetlike to vertically stacked channel deposits. These
from more than 2,100 ft in the west-central part of the basin to deposits are locally interbedded with and laterally grade into
1,100 ft in the eastern part. A basal pebble-conglomerate bed floodplain mudstone and siltstone deposits. Associated car-
commonly rests unconformably directly on the Vermejo Forma- bonaceous shale and thin, lenticular coal beds formed in back-
tion, but in the vicinity of Raton and Trinidad, the conglomerate swamps. Along the eastern and southern margins of the basin,
is commonly absent and no unconformity is evident. This cliffs of the barren series commonly stand high above slopes of
scour-based sandstone forms a persistent cliff throughout much the lower coal zone and the Vermejo Formation. The upper coal
of the Raton basin, especially in the western and southern parts. zone of the Raton supports the hills and ridges above the barren
Lee (1917) originally divided the Raton into a basal conglom- zone in the interior part of the basin.
erate, a lower coal zone, a barren series, and an upper coal zone. The barren series was succeeded by the upper coal zone,
A similar subdivision that better fits our purposes, especially in which was deposited on a low-gradient alluvial plain, result-
the area around Raton and Trinidad where the basal conglomer- ing from a decrease in tectonic movement in the source area.
ate is lacking, includes the basal conglomerate in the lower coal This tectonic pause, perhaps combined with basin subsidence,
zone. The three subdivisions are mostly consistent and identifi- led to aggradation of the alluvial plain by a meandering fluvial
able throughout the Raton basin and thicken from east to west. system. The system was accompanied by floodplains that
The Raton Formation was deposited on an upper alluvial developed poorly drained backswamps in which coal beds as
plain (Flores, 1984) characterized by various modes of aggra- thick as 12 ft accumulated. The floodplains were locally filled
dation and erosion. Deposition on the alluvial plain was inter- by overbank and crevasse-splay detritus during episodes of
rupted by several styles of fluvial sedimentation. Following the floods. These deposits extended into the backswamps and
deposition of fluvio-delta plain and back-barrier sediments of caused splits of coal beds where the peat swamps reestab-
the Vermejo Formation, rapid uplift of the source area to the lished after detrital influx. Deposits of this setting grade
west (the present San Luis valley, termed the San Luis highland, upward into deposits of more landward environments of the
Tweto, 1987) caused widespread erosion of the highland and Poison Canyon Formation.
concurrent deposition of the basal conglomerate of the Raton. The Poison Canyon Formation overlies the Raton Forma-
The basal conglomerate, which consists of medium- to coarse- tion throughout most of the Raton basin, but to the west and
grained channel sandstone with lenses and stringers of con- southwest, it intertongues with the Raton. The Poison Canyon
glomerate, was probably deposited in basin-margin braided consists of thick to massive, lenticular, ledge-forming beds of
streams merging basinward into meandering streams. The coarse-grained to conglomeratic arkosic sandstone interca-
uplift of the highland during Laramide time may have moved lated with beds of nonresistant, yellow-weathering, sandy,
along thrust faults that bordered the western margin of the basin micaceous mudstone and siltstone. The contact with the
(Flores and Pillmore, 1987; Woodward and Snyder, 1976). underlying Raton generally is indefinite and gradational. The
After deposition of the basal conglomerate, more stable tectonic Poison Canyon was deposited on a high-gradient alluvial plain
conditions returned and sand was deposited in meandering (Strum, 1984), characterized by non-coal-bearing floodplain
streams at the same time that interbedded thin coal, carbona- and braided- to meandering-stream alluvial-fan deposits. The
ceous shale, mudstone, and sandstone accumulated in flood- high sediment input into the alluvial plain indicated by these
plains and backswamps. Crevasse splays periodically high-bedload streams and fans probably reflects intensive ero-
interrupted and infilled low-lying floodplains during river sion of a rapidly rising source area to the west. These tectonic
floods. Though some thick coal beds were deposited locally in and depositional conditions probably correspond to a pulse of
the lower coal zone, most coal swamps that formed in the flood upthrusting that exposed core rocks, as indicated by the abun-
basins were well-drained, small, and shallow, which limited dance of potassium feldspar grains in the channel sandstone.
peat accumulation and resulted in thin lenticular coal beds The Poison Canyon deposits coarsen to the west, forming the
mostly less than 8-12 in thick. At the close of the Cretaceous, alluvial fans of the piedmont environment marginal to the ris-
the extensive alluvial plain was an ideal environment for depo- ing ancestral San Luis highland.
140 C. L. Pillmore, D. J. Nichols, and R. F. Fleming
the terrestrial ecosystem was stressed by a significant, though claystone is texturally and chemically different from typical
geologically brief, event (Tschudy and others, 1984; Tschudy Raton basin tonsteins. As seen in ultrathin section, it is fine
and Tschudy, 1986). grained to amorphous but may exhibit an imbricate fabric and
Paleobotany. Early in the K-T boundary controversy, relics of small bubbles in a fine crystalline matrix of kaolin-
Hickey (1981) asserted that the megafloral record was inconsis- ite. Microspherules (40-120 microns in diameter) consisting
tent with the hypothesis that a catastrophe caused terrestrial of calcium, aluminum, strontium, cerium, rare earth ele-
extinctions. However, in their joint paper, Johnson and Hickey ments, and phosphorus (similar to goyazite, a hydrous stron-
(1990) reversed this position by presenting new evidence that tium alumino-phosphate in composition; microprobe analysis
the megafloral change is about 80 percent and that it coincides by Ralph Christian, USGS, 1984) have been observed in sam-
with a peak in palynofloral extinctions and the occurrence of Ir ples of the boundary claystone from the Raton site. These
and shock-metamorphosed mineral grains. They state that the phosphatic spherules are rarely seen at other sites but they
results of their analysis of the terrestrial plant record are “com- form discrete layers in the boundary claystone at the Dogie
patible with the hypothesis of a biotic crisis caused by extrater- Creek and Teapot Dome localities in the Powder River basin
restrial impact at the end of the Cretaceous” (Johnson and in Wyoming (Izett, 1990). Smit (1984) has referred to simi-
Hickey, 1990, p. 433). larly shaped grains in the boundary claystone from the Raton
Wolfe and Upchurch (1986) analyzed fossil leaves and basin and other areas as microtektite-like structures, imply-
dispersed fragments of leaf cuticles from K-T boundary ing an impact origin. The microtektite grains from Raton
sequences in the Raton Formation. Their results suggest a basin sites are mostly spheroidal to subspheroidal and resem-
brief low-temperature excursion (mean temperature near 0°C) ble dull to shiny resinous little balls under the microscope.
that caused a masskill and ecological disruption of terrestrial Some are hollow. Under the scanning electron microscope,
vegetation at the K-T boundary. Leaf size and shape indicate they have an uneven surface texture and commonly are pitted
that a major increase in precipitation occurred across the by irregularly shaped cavities. On the basis of textures and
boundary. Their conclusions are consistent with the bolide shapes observed in goyazite spherules in the Powder River
impact hypothesis. Basin, that are identical to those seen in microtektites, the
The K-T boundary claystone bed. The boundary clay- spherules are thought to result from the alteration of glassy
stone bed resembles a tonstein (tonsteins are kaolinitic clay- ejecta material (microtektites) blown out of the crater during
stone partings thought to result from alteration of volcanic the K-T impact event.
ash beds in acidic coal swamps), but, unlike typical tonsteins, High concentrations of Ir and shock-metamorphosed min-
it usually weathers to a lighter, pinkish, color and exhibits a eral grains, both compelling evidence of impact origin (Bohor
fine-grained to amorphous texture and a distinctive hackly to and others, 1984; Izett, 1990), are present in a discrete layer at
conchoidal fracture. The claystone is mostly gray and gray- the top of the boundary claystone. This layer was called the
ish pink to grayish yellow and commonly contains tiny flaky shale layer by Pillmore (Pillmore and others, 1984), the
specks and thin, contorted lenses or layers of organic matter, K-T boundary impact bed by Izett and Bohor (1986), and the
especially near the margins. Small spheroidal structures can fireball layer by Hildebrand and Boynton (1988). The shocked
be seen on fracture surfaces of some specimens and in thin grains consist mainly of quartz, with rare microcline and pla-
section. X-ray diffractograms show that, like many tonsteins, gioclase. The shock-metamorphosed quartz grains contain as
the boundary claystone is nearly pure, well-crystallized kaoli- many as nine intersecting sets of closely spaced planar features
nite with lesser amounts of randomly stratified illite-smectite per grain (Izett, 1990). Figure 5 is a photograph by Izett of one
clay and some quartz and feldspar (Pollastro and others, of the shocked grains of quartz from the Starkville South site,
1983; Pollastro and Pillmore, 1987). However, the boundary showing two sets of planar lamellae.
142 C. L. Pillmore, D. J. Nichols, and R. F. Fleming
east of Castle Rock and is situated approximately above the about 10,000 ft against Dawson Arkose by the Rampart Range
deepest part of the Denver basin. In March of this year (1999), reverse fault. This fault, the Ute Pass fault to the south, and
a 2,200-foot well was core-drilled at the Elbert County Fair- other faults along the Front Range may be low angle thrusts, on
grounds in Kiowa. This 2.5-inch-diameter core was obtained the basis of geophysics and the outcrop pattern (A.F. Jacob,
as part of the Denver Basin Project, which is a Denver Museum consulting geologist, written commun., 1985). The large scars
of Natural History project funded in part by the National Sci- visible on the mountain front of the Rampart Range north of
ence Foundation. The goals of the project are to reconstruct the Colorado Springs are quarries that were developed for concrete
sedimentary and tectonic history of the Denver basin during the aggregate in limestone and dolomite of the Williams Canyon
Late Cretaceous and early Tertiary. The hydrogeology of the Formation (Devonian) and the Manitou Limestone (Ordovi-
Denver basin is also a primary focus of the project and detailed cian).
studies of the aquifers cored in the Kiowa well will provide Just past the south entrance to the Air Force Academy, I-25
valuable information about water resources in the Denver basin. crosses Monument Creek and passes down section through out-
The multidisciplinary project includes hydrogeology, pale- crops of the coalbearing Laramie Formation and the Fox Hills
ontology (palynology and paleobotany), stratigraphy, sedimen- Sandstone (mostly covered), which is lithologically equivalent
tology, paleomagnetism, fission-track dating, and other to the Vermejo and Trinidad Sandstone Formations of the Raton
disciplines. Results from these various fields will be integrated basin, (Fig. 2) and into the Pierre Shale, all of Late Cretaceous
to provide a basis for interpreting the history of the Denver age. Much of the area immediately east of Monument Creek is
basin and implications of that history for topics ranging from the site of abandoned coal mines in the Laramie. Continuing
groundwater use to the extinction of the dinosaurs. south through Colorado Springs, good views are afforded of
In this project the K-T boundary in the Denver basin and Pikes Peak, which is underlain by the Precambrian Pikes Peak
extinction patterns across it will be examined. The study will Granite (about 1,000 Ma), and Cheyenne Mountain, underlain
integrate paleobotanical and palynological data within the gen- by granodiorite of the Routt Plutonic Suite (about 1,700 Ma;
eral sedimentological and stratigraphic framework in hopes of Tweto, 1987). Cheyenne Mountain houses the large under-
documenting the evolution of terrestrial vegetation from the ground installation of NORAD (North American Air Defense
Late Cretaceous to the early Tertiary in the Denver basin. Command). The Ute Pass fault, a large reverse and thrust fault
The K-T boundary claystone has been documented in that dips to the west as low as 30 degrees and trends along the
numerous sites to the south and north of the Denver basin. eastern base of the mountain, places Pikes Peak Granite against
However, attempts to locate the boundary claystone in the Den- Pierre Shale. To the west of Colorado Springs, the Fountain
ver basin have been unsuccessful. Preliminary palynological Formation is spectacularly displayed in erosional forms at the
results place the K-T boundary approximately 880 feet below Garden of the Gods. These arkosic sandstone and conglomer-
the town of Kiowa. No K-T boundary claystone similar to the ate units are about 4,000 ft thick and were deposited as alluvial
K-T boundary layer in the Raton basin was located in the core. fans on the east flank of the northwest-trending ancestral Rock-
However, the horizon of the palynological boundary will be ies during Pennsylvanian and Early Permian time.
projected to surface exposures east of Kiowa and suitable sites
will be examined in detail to locate the K-T boundary. Colorado Springs to Pueblo
Castle Rock to Colorado Springs Proceeding south from Colorado Springs, the route contin-
ues on the Pierre Shale to Pueblo. The immense training
The Castle Rock itself and many of the mesas south of the grounds of Fort Carson Army Base lie to the west of the Inter-
town of Castle Rock are capped by the Castle Rock Conglom- state for many miles south of Colorado Springs. About midway
erate. In this area, except at Castle Rock, the conglomerate is to Pueblo, the sharp eye will observe the Tepee Buttes in the
locally underlain by the silicic Wall Mountain Tuff. The Wall Pierre Shale outcrops on the horizon to the east across Fountain
Mountain is an ashflow tuff that occurs on the Rampart Range Creek. These conical-shaped buttes are capped by large resis-
to the west and is also present in the Thirtynine Mile volcanic tant concretions in the Pierre Shale.
field. The tuff was a widespread ash flow that erupted 34.8 ± The Canyon City Embayment, which is about 20 mi west
1.1 Ma (J.D. Obradovich, USGS, written commun., 1969) from of the Interstate, is a topographic and structural feature that
its presumed source near Mt. Aetna in the Sawatch Range, reflects the en-echelon arrangement of mountain ranges. The
about 100 mi. to the west. Front Range terminates south of Cheyenne Mountain and the
The route continues in the Dawson Arkose nearly to Col- Wet Mountains form the eastern front of the Rocky Mountains
orado Springs. In the vicinity of the Air Force Academy, out- farther south. The Canyon City Embayment occupies the inter-
standing examples of pediments of three ages can be seen west vening space.
of the highway, capped from highest to lowest by the Rocky The Florence oil field, which was discovered in 1862, was
Flats, Verdos, and Slocum Alluviums. Precambrian rocks of the the first oil field found west of the Mississippi River. Oil from
Rampart Range, which lies west of the pediments, are uplifted this field was produced from fractured Pierre Shale (Upper Cre-
144 C. L. Pillmore, D. J. Nichols, and R. F. Fleming
Table 2. Generalized stratigraphic section of Cretaceous and Tertiary rocks seen from Colorado Springs to Raton.
taceous) and was in great demand as a lubricant for wagon stone Member of the Niobrara and the Bridge Creek Limestone
wheels and sold for more than $5 per gallon. Member of the Greenhorn. Benches formed by these limestone
units can be seen in the breached anticline a few miles south of
Pueblo to Walsenburg Pueblo and across the drainage divide at the abandoned rest
stop on the left that is mostly on the pinon- and juniper-covered
In Pueblo we pass from the Pierre Shale into the underlying Bridge Creek Member beneath mesa crests formed by the Fort
Cretaceous Niobrara Formation (table 2) at about the Arkansas Hays Member.
River. South of the river the old Colorado Fuel and Iron (CF&I) At the Colorado City Exit from I-25, we cross the struc-
steel-making plant is on the left and we can see the new tural divide, the Apishapa arch, that separates the Denver and
Comanche power generating plant to the southeast. This power Raton basins. Just south of the excellent road-cut exposure of
plant was constructed in large part to provide power for electric the Fort Hays Limestone at the exit, the route descends the dip
steelmaking furnaces, which totally cut off the demand for slope of the Fort Hays Limestone, which crops out at the bot-
Raton basin coking coal that had been produced for many years tom of the hill along Greenhorn Creek, east of the bridge. After
at CF&I’s Allen and Maxwell mines (more recently called the crossing the creek, we go up section through the Smoky Hill
New Eagle and Aztec mines, now closed), west of Trinidad. Member of the Niobrara Formation until we cross the Apishapa
Coal for the steam plant is brought by unit train from the Belle fault near the top of the hill. The sandstone beds exposed in
Ayr mine near Gillette, Wyoming. roadcuts of the highway to the left are on the south side of the
South of Pueblo we will drive across gently rolling, dis- fault and consist of the Dakota Sandstone, the oldest Cretaceous
sected terrain underlain by marine Upper Cretaceous rocks unit in this vicinity; the large ridge at the crest of the hill is sup-
(table 2) of the Niobrara, Carlile, and Greenhorn Formations, ported by Dakota Sandstone. The Lower Dakota Sandstone
past ridges and mesas supported mainly by the Fort Hays Lime- (table 2) represents the transgression of the Cretaceous epeiric
Continental Cretaceous-Tertiary boundary in Raton basin, Colorado and New Mexico 145
sea, which extended across the North American continent from tion is approximately equivalent to the Raton Formation. The
the Arctic Ocean to the Gulf of Mexico in the Early Cretaceous scenery along this part of the route is dominated by the Spanish
and existed during most of the Late Cretaceous (Cenomanian to Peaks to the west. The Spanish Peaks, in part held up by intru-
Maastrichtian). The Dakota shows several hundred feet of reac- sives, are well known for many dikes that radiate from them for
tivated Laramide movement along the fault, which originated several miles. The Spanish Peaks plutons have been dated at
during the late Paleozoic (Ogden Tweto, USGS, oral commun, 21.7 ± 1.0 Ma (Stormer, 1972, East Peak Granite) and 22.9 ±
1980). The Apishapa fault trends about east across Las Animas 2.0 Ma (Smith, 1975, West Spanish Peak stock). Other dates
County and forms the northern boundary of the late Paleozoic ranging from 19.8 ± 1.6 Ma to 39.5 Ma have been reported
Apishapa arch. The fault is a major fracture zone and water from intrusive rocks in the Spanish Peaks area; the 39.5 Ma date
wells in the zone provide a large part of the water supply for appears questionable and most dates fall into the 20-25 Ma
Colorado City. range (Marvin and others, 1974, p. 3233). Ages obtained by
South from the top of the Dakota ridge we leave the Denver Penn (1994) indicate continued intrusive activity in the Spanish
Basin and proceed across the valley back up section through the Peaks region from 26.6 Ma to 21.3 Ma with the middle period
progressively younger, south-dipping marine shale, limestone, of activity (24.6-22.8 Ma) involving West and East Spanish
and limy mud of the Graneros, Greenhorn, and Carlile to the Peaks and the radial dikes. Most of the sills and dikes in the
next ridge, capped once again by Fort Hays Limestone (table Raton and Vermejo Park areas were intruded at about the same
2). At the top of this ridge we look south into the Raton basin, time (J.D. Obradovich, C.W. Naeser, and H.H. Mehnert, USGS,
a large arcuate structural basin that extends south into New written commun., 1976-1983).
Mexico (Fig. 1). The Fort Hays and Smoky Hill Members of West of the interstate and along the route to Trinidad, the
the Niobrara Formation form the dip slope for several miles to Pierre Shale is exposed beneath cliffs formed by the Trinidad
the south from the ridge crest. Farther along, the Pierre Shale Sandstone. Just past the turnoff to Ludlow, the view to the
overlies the Smoky Hill and underlies the route into Walsenburg south is dominated by what appears (to some) to be the silhou-
and on south to Trinidad. ette of an early 1920’s Franklin automobile, outlining the hood
The Wet Mountains, which dominate the skyline to the and trunk, which is formed by the basalt flows on Fisher’s Peak,
west of the Interstate highway, are composed mainly of Pre- the northern segment of Raton Mesa. Though not dated as yet,
cambrian crystalline rocks. The summit of Greenhorn Moun- the flows that form the crest are considered to be about 3.5 m.y.,
tain, which is the prominent peak at the southern end of the the same as Bartlett Mesa north of Raton, New Mexico
range (elevation 12,334 ft), is unusual in that a remnant of (Stormer, 1972; Stroud, 1998).
Oligocene volcanic rocks uncomformably overlies the Precam-
brian. Several prominent pediment surfaces are visible to the Trinidad
west along the route.
East of the highway and visible for several miles is Huer- The site of Trinidad was an ancient Indian ceremonial
fano Butte, a conspicuous conical peak that rises about 100 feet ground. The area was later visited and periodically inhabited
above the surrounding plain. According to Penn (1994), the by Onate and other Spanish explorer-colonizers, priests, and
butte is a biotite olivine alkali-gabbro plug bisected by two east- soldiers; French and American trappers and traders; and Kearny
trending dikes, rather than a volcanic edifice of some sort as and other subsequent U.S. military expeditions to New Mexico.
indicated in popular literature. The age of the intrusives is The first permanent structure, a sheepherder’s cabin on the
about 25.2 Ma, similar in age to the East and West Spanish south bank of the Purgatoire River (pronounced “pergatwar”),
Peaks intrusives. was erected in 1859, and the area soon attracted numerous
farmers and traders. The early years of Trinidad were turbulent;
Walsenburg to Trinidad conflicts between American and Mexican settlers and between
the settlers and the Ute Indians were common in the 1860’s.
Leaving Walsenburg, we continue on Pierre Shale for about Today the population of Trinidad is about 9,000. Coal min-
the next 37 mi, past the town of Aguilar to Trinidad. Aguilar is ing, which began in the late 1800’s, was the most important
located adjacent to coal-bearing Cretaceous and Tertiary rocks industry in the area until the 1950’s when coal-burning steam
of the Raton basin. These rocks (Fig. 2), which define the east- locomotives were superceded by diesel locomotives. Coal con-
ern limits of the Walsenburg and Trinidad coal fields, are tinued to be produced at the New Eagle and Maxwell mines
marked by abrupt cliffs formed by the Trinidad Sandstone several miles west of Trinidad to fire the coke ovens in Pueblo
(Upper Cretaceous). Overlying the Trinidad are the coalbearing until the late 1970’s and later as fuel for electric power genera-
Vermejo Formation (Upper Cretaceous) and the Raton Forma- tion, but mining activity ceased in 1994.
tion (Upper Cretaceous and Paleocene). The Raton is overlain In the Colorado part of the Raton Basin, more than 200
by the Poison Canyon Formation (Paleocene). The Dawson wells have been drilled in the past several years to develop the
Arkose of the Denver Basin is roughly equivalent to the Poison methane potential in coal beds of the Vermejo and Raton For-
Canyon Formation of the Raton basin, and the Denver Forma- mations. The center of activity is about 20 miles west and
146 C. L. Pillmore, D. J. Nichols, and R. F. Fleming
northwest of Trinidad. The main operators at present (1999) are ues to increase with recent activity (1999) southwest of
Evergreen Resources, GeoMet, Inc., and Chandler and Associ- Trinidad and in the New Mexico portion of the basin.
ates. Most wells initially flow less than 100 MCFGPD and a
few hundred barrels of water per day, but with all of the pro- DESCRIPTION OF STOPS AND ROUTES TO STOPS
ducing wells in the area the gas stream in the Colorado Inter-
state Gas pipeline is about 70 MM cu. ft./ day. Gas is also being Trinidad to the Long Canyon K-T site
produced from the abandoned workings of some of the old coal
mines in the area. Since the installation of a pipeline into the Note mileage and turn off I-25 to Colorado Highway 12
area in 1994, drilling activity has greatly increased and contin- and proceed west out of Trinidad. Trinidad Sandstone caps the
0.5 ft
5 ft
road level
Figure 6. Diagram showing the lithology of the K/T boundary sequence at the Long
Canyon K/T boundary site in the Watchable Wildlife area of Trinidad Lake State
Park. The inset shows the detail of the boundary interval and the variation in iridium
concentration. (Iridium analyses courtesy of F. Asaro, Lawrence Berkley Labora-
tory, written comm., 1999). Palynologic analyses of samples collected across the
boundary at the Madrid site (Fig. 1) show that more than 80 percent of the assem-
blage in the coal bed directly above the boundary claystone bed consists of fern
spores (R. H. Tschudy, USGS, written comm., 1982).
Continental Cretaceous-Tertiary boundary in Raton basin, Colorado and New Mexico 147
high ridge to the north of town and crops out along the high- ately beneath the thick sandstone ledge, is the K-T boundary
way. The entrance to the road across Trinidad Lake dam is near claystone bed overlain by a 1- to- 2-in-thick coal bed and under-
the top of the formation, and sandstone, mudstone, and car- lain by 7 to 8 in of brownish-orange-weathering carbonaceous
bonaceous shale and coal beds of the Vermejo Formation are shale and boney coal typical of the Madrid area. This coal
well exposed in high roadcuts on the right. At about 5.7 miles sequence suggests that the ejecta cloud from the impact
is the approximate position of the base of the Raton Formation. deposited a thin bed of glassy debris in a large coal-forming
Beds and stringers of pebble conglomerate generally occur at swamp mire where it eventually altered to kaolinitic clay. The
the base of the Raton, but they are commonly absent in this area thin coal bed indicates that plant material that accumulated fol-
and the base is represented by beds of sugary-textured, quart- lowing the impact was inundated by a splay deposit of sand-
zose sandstone. stone, as determined by its nearly flat base, which is overlain
At the town of Cokedale is the large Asarco coal mine by a splay-channel sandstone.
waste pile on the right, and rows of brick coke ovens on the left Return to the parking lot and retrace the route to the
that produced coke for the Asarco smelters for 40 years until the entrance of the Trinidad Lake dam. Turn right and proceed
mine closed in 1946. across the dam through Starkville across I-25 and turn right
About 1.5 miles past Cokedale watch for the sign to the along the service road to Stops 2a and b, the Starkville K-T
Watchable Wildlife Area and turn left across the new bridge boundary sites.
over the Purgatoire River, cross the railroad tracks, and bear left Proceed south on the service road past waste piles of sev-
down the gravel road. Upstream along the railroad the K-T eral abandoned coal mines in the Vermejo Formation about 2
boundary sequence crops out in railroad cuts for several hun- miles to the Starkville K-T boundary sites. The prominent chan-
dred yards beneath a tabular splay sandstone bed that is easily nel sandstone in the roadcut at about 1 mi rests directly on the
visible from the highway across the valley. upper Starkville coal bed, which is approximately 65 ft above
Where the road turns south up a small valley (about a 1/4 the Trinidad Sandstone.
mi), the K-T boundary claystone is exposed directly beneath a
prominent sandstone bed in a steep roadcut known as the Madrid Stop 2a: Starkville North K-T boundary site
East site. The boundary claystone (the thin, 1-in-thick, white
claystone bed at the top of the cut) lies beneath a 2-in-thick coal A short distance on down the road is the Starkville North
bed that is directly overlain by the sandstone. The claystone bed K-T boundary site. This first recognized Colorado location of
is underlain by a brownish-orange-weathering, boney coal bed the K-T boundary is now obscured by a landslide. The K-T
about 12 to 16 in thick. The claystone bed at this site and in this sequence was exposed in a carbonaceous shale bed 20 in thick,
area exhibits all of the characteristics of the boundary claystone 111 ft above pavement level at top of the high cut on the east
in the Raton basin: anomalously high concentrations of Ir, side of I25. The rocks exposed in the roadcut consists of
shocked quartz, the fern-spore spike, and the palynological interbedded channel sandstone and floodplain/back-swamp silt-
extinction horizon, all indications of asteroid impact.
Proceed on the road about 2 mi to Long Canyon and turn
left at the road junction about ½ mi to the parking area for the
Watchable Wildlife Area.
Figure 9. Diagram showing the lithology of the K-T boundary interval at the Starkville North K-T
boundary site, 3 mi south of Trinidad, Colorado. The large black dots show the variation in Ir con-
centration, the solid line and triangles show the fern-spore percentage, and the inset shows the detail
of the boundary interval. From Tschudy and others (1984).
Continental Cretaceous-Tertiary boundary in Raton basin, Colorado and New Mexico 149
about 200 ft above the Trinidad Sandstone in a sequence of The relative abundance of fern spores just below the
fine-grained deposits of the lower coal zone of the Raton For- kaolinitic claystone is about 30 percent and comprises several
mation. species. Just above the kaolinitic claystone is a 2-cm-thick
interval that to date has not yielded any palynomorphs and has
Stop 2b: Starkville South K-T boundary site been called the barren interval. Palynomorph assemblages
recovered from the coal just above the barren interval are dom-
Just a ¼ mile farther south is the Starkville South K-T inated by fern spores—this is the “fern-spore spike” interval
boundary site (Fig. 10). This site has probably been the most (see section on the Cretaceous-Tertiary Boundary for a discus-
heavily sampled site in the Raton basin. In 1984, a team from sion of the significance of the fern-spore spike). Fern spores
the Smithsonian Institution collected a 2½-ton sample of the constitute 80 percent of the assemblage from the lower half and
boundary interval and shipped it back to Washington, D.C. to 77 percent of the assemblage from the upper half of the coal; a
archive this significant rock sequence. The Starkville South single species of fern spore dominates the assemblages.
section shown in Figure 1 represents several sites in roadcuts Angiosperm (flowering plant) pollen in the fern spore spike
along I-25. At the Starkville South site, the strongest Ir anom- interval is greatly reduced in relative abundance and diversity.
aly ever measured in continental rocks (56 ng/g) was measured Above the coal, angiosperm pollen gradually increases in rela-
in the flaky shale at the top of the ejecta layer of the boundary tive abundance and diversity until it again dominates.
claystone (Pillmore and others, 1984). The Ir is concentrated at A 3-ft-thick, olive-gray mudstone just above the K-T
the top of the K-T boundary claystone in this thin bed now boundary coal bed contains fossil planktonic green algae (Pedi-
known as the fireball layer (Fig. 11). astrum and Scenedesmus), which also occur in modern fresh-
water lakes and ponds. The presence of these algae indicates
Palynology of the Starkville South K-T boundary site that the mudstone was deposited in a floodbasin lake or pond
(Fleming, 1986, 1987). This interpretation is corroborated by
The K-T boundary at Starkville South is placed at the level the presence of Pandaniidites radicus in the brown mudstone
of the disappearance of the Cretaceous palynomorphs of the immediately overlying the olivegray mudstone. The plants that
Proteacidites assemblage. These fossils are present in the Cre- produced P. radicus are attributed to modern Pandanus (screw-
taceous rocks below but disappear at the top of a kaolinitic pine), which grows in coastal and marshy areas of the tropics
claystone, which here is located about 6 ft above the top of the and subtropics. At Starkville South, the presence of Pandani-
point bar sandstone (Fig. 10). Cretaceous palynomorph assem- idites radicus indicates deposition in a marshy area, probably
blages in mudstone below the boundary (Fig. 11) are diverse on the margin of the floodbasin lake that contained planktonic
and include many species of fossil pollen and spores that also green algae. The palynological and sedimentological evidence
occur in the lower Paleocene (e.g., Pandaniidites radicus, suggests that the K-T boundary layers and associated sediments
Kurtzipites spp., and Ulmipollenites sp.), but more importantly were deposited in a floodbasin environment that included peat-
they include the Proteacidites assemblage, which does not
occur in the Paleocene.
forming swamps (represented by the coal), marshes (repre- late 1800’s, Morley became a coal-mining company town in
sented by brown, carbonaceous mudstone containing Pandani- 1906, when CF&I opened the Morley mine. The fine-grade
idites radicus), and floodbasin lakes or ponds (represented by coking coal mined here was utilized by the steel mills in Pueblo
olive-gray mudstone containing Pediastrum and Scenedesmus). and by Santa Fe Railroad locomotives. Peak coal production
The depositional framework at Starkville South is some- reached 500,000 tons per year in the late 1920’s, when the town
what different from that at Starkville North. At Starkville had a population of over 1,000. The Morley mine was never
North, the boundary lies near the top of a sequence of mudstone mechanized; the use of blasting powder and machinery was
beds 20 ft thick that rests on a channel sandstone 7 ft thick; at prohibited because of significant amounts of methane gas.
Starkville South, the mudstone sequence has thinned to 10 ft, Instead, the coal was extracted by hand labor and an under-
and the channel sandstone has thickened to 23 ft. ground herd of donkeys was maintained to haul the coal out of
Go back up the service road, cross the Interstate, and turn the mine; in fact, the Morley was reportedly the last mine in the
south. U.S. to use these beasts. Cutbacks in steel production at Pueblo
temporarily halted coal production in the early 1950’s, and the
Starkville sites to Raton, New Mexico mine was closed permanently on May 4, 1956, when all work-
able deposits had been exhausted. From 1907 until 1956 the
Leaving the Trinidad area, we will go south over Raton Morley mine produced 11 million short tons of coking coal.
Pass on I-25, through rocks of the Trinidad Sandstone (delta-
front and barrier-bar environments) and Vermejo and Raton Top of Raton Pass
Formations (deltaic-fluvial environments), to the contact of the
Raton and Poison Canyon Formations (fluvial environments), The view from the parking area includes the Trinidad coal
which is near the top of the pass. Several coal beds, dikes, and field, the Spanish Peaks, and the high peaks of the Sangre de
sills can be observed in roadcuts along the route. As we go Cristo Mountains. The Spanish Peaks form a landmark that is
south from Trinidad, coal dumps along the way are from early visible throughout hundreds of square miles. The area lying
mining in the area. between Raton Pass and the Spanish Peaks contains several
We pass gently dipping alluvial floodplain sequences of hundred million tons of coal resources and probably trillions of
sandstone, mudstone, shale, and coal beds of the Vermejo and cu. ft of related coalbed methane. The sign at rest area reads:
Raton Formations and, about six miles south of Trinidad, the
Morley dome becomes evident, marked by subtle changes in Raton Pass, named by the Spanish for the rock rats found
dip that bring the Trinidad Sandstone and Pierre Shale to the there, was crossed by the Mountain Branch (also called the
surface. The Morley dome has about 450 ft of closure and was Bent’s Fort Branch) of the Santa Fe Trail. Originally part of an
apparently formed by the intrusion of a Tertiary plug or laccol- old Indian trail, the pass was used by Spanish expeditions at
ith, which crops out about 1 mi northeast of the abandoned least as early as 1718 and probably much earlier. When the
town of Morley. The crest of the fold is between the town and Cimarron Branch of the Santa Fe Trail was abandoned because
the exposed plug, and the closing contours extend for some dis- of its dangerous desert stretches, pioneer traffic increased over
tance around the north and east sides of this igneous intrusion. Raton Pass. Kearny’s Army of the West came this way in 1846
Coal seams lacking evidence of natural coking or metamor- with some of the first vehicular supply wagons to cross the pass.
phism have been mined to the contact with the igneous plug. A “Uncle Dick” Wootton, frontier scout, built a road over the pass
minor oil seep was recognized in one mine and in 1948 Stano- in 1865 and collected tolls, often at the point of a gun until the
lind drilled a dry hole on the crest of the structure to a total coming of the railroads in 1878. For two years, a controversy
depth of 6,831 ft into “granite.” The K-T boundary is exposed raged between the Denver and Rio Grande and the Santa Fe
in several I-25 roadcuts between the Starkville exit and Morley. Railroads to determine which company had the right-of-way
Palynological examination of samples from a sequence of thin over Raton Pass. After several section crew fights and much
coal beds beneath a south-dipping sequence of sandstone beds legal maneuvering, the Rio Grande gave up its claim to Raton
of the barren zone of the Raton Formation just south of the Pass and the Santa Fe agreed not to contest another disputed
Morley coal waste pile indicates they are of Cretaceous age and right-of-way through the Royal Gorge.
the K-T boundary must lie above the coal beds, perhaps in the
basal sandstone. Proteacidites spp. is abundant in these sam- The contact between the Raton and Poison Canyon Forma-
ples. tions is gradational through several feet of section. It is mapped
approximately at the color change from grayish-yellow-weath-
Morley ering rocks to grayish-orange and grayish-red-and-brown-
weathering rocks just below the north exit from the Rest Area.
Morley began as a railroad town on the A.T.&S.F. line in The top of the Trinidad Sandstone is about 1,250 ft below the
1879, and trains were commonly shortened here before begin- top of the pass. The thickness of the Vermejo Formation is esti-
ning the difficult climb over Raton Pass. Inactive during the mated to be about 70 ft in this vicinity, making the Raton about
Continental Cretaceous-Tertiary boundary in Raton basin, Colorado and New Mexico 151
1,180 ft thick. To the west, in the central part of the basin, will return to Raton and proceed south on Main Street through
thicknesses are greater than 2,000 ft for the Raton and about town to the Holiday Inn Express for the night.
300-400 ft for the Vermejo.
Proceeding on south from Raton Pass, we follow the route SECOND DAY: ROUTE TO SUGARITE, NEW MEXICO
of the Mountain Branch of the Santa Fe Trail (Scott, 1986) and
pass back down section through the three zones of the Raton This segment of the field trip depends on weather more
Formation. Thick coal beds are exposed in roadcuts in the than the others. It involves a hike along a good trail through
upper coal zone and stacked channel sandstone beds make up some rugged terrain and, in the event of inclement weather, we
the barren zone toward the bottom of the pass. The lower coal will scrub the climb and substitute a different K-T site. To
zone of the Raton and the Vermejo Formation are present in reach Sugarite (Fig. 1) we will go north on I-25 and turn east
roadcuts near the mouth of the valley. The Trinidad Sandstone from Raton via New Mexico highway 72. Our route crosses
crops out just before the Raton exit. The town of Raton is built outcrops in creek bottoms of the upper (Exiteloceras jenneyi)
on the Pierre Shale. Enter Raton. part (Campanian) of the Pierre Shale (Pillmore and Scott, 1994)
Turn right at the first intersection past the Melody Lane and gravelly alluvium on the broad flats formed by the Barilla
Motel to Moulton Avenue, which will take us up Goat Hill on and Beshoar pediment surfaces (Scott and Pillmore, 1993; Pill-
the old Raton Pass road to stop 3, the Raton Pass K-T bound- more and Scott, 1976). Along the route, to the north, large land-
ary site (Fig. 1). Proceed up Moulton past the Shady Moun- slides cover much of the slopes below the 3.5 Ma basalt caps of
tain Park trailer court through the Trinidad Sandstone. At the Bartlett Mesa. To the east, landslides also cover the slopes
junction, keep right. The top of the Trinidad Sandstone is beneath the 7.2 Ma basalt cap of Johnson Mesa (Stormer,
approximately at the next loop in the road. The upper and 1972). As a side observation, about a mile from the Interstate,
lower coal beds of the Raton coal zone are exposed in road- the view to your left reveals a broad valley containing a small
cuts above the Trinidad. The basal conglomerate of the Raton hill that contains a diatreme (breccia-filled volcanic pipe
is not present and its contact with the underlying Vermejo For- formed by a gaseous explosion). A fission-track date by C.
mation is indefinite. Continue on up section through the lower Naeser of the USGS on zircon grains in a quartzitic sandstone
coal zone of the Raton Formation to the saddle. Just below the (Trinidad Sandstone or possibly Dakota Sandstone) block
saddle the Sugarite coal zone (Upper Cretaceous) is exposed yielded 25 ± 1.4 m.y., younger than the sedimentary rocks in
in the roadcut. the area but older than the oldest basalt. This number dates the
last time the rock was heated past the annealing temperature of
Stop 3: The Raton Pass K-T boundary site zircon, 200 ± 25°C.
About 3.5 miles from the Interstate is the junction with
The Raton site is at the top of a saddle about 1 mi west of New Mexico 526. Bear left and proceed left up Chicorica
Raton on Southwell Mountain Road (the old Raton Pass Canyon on highway 526. On the left, cliffs formed by sand-
Road). The roadcut exposes the first outcrop discovery of the stone beds of the barren zone of the Raton Formation and the
Ir anomaly (about 1 ng/g) in the Raton basin. The anomaly is Trinidad Sandstone are visible in the upper slope. Basaltic lava
coincident with the disappearance of the Proteacidites assem- flows cap Horse Mesa on the right and Bartlett Mesa on the left.
blage. It is at the top of a 1-in-thick bed of rusty-weathering A short distance up the creek, large coal waste piles from
kaolinitic claystone about 7 to 8 in below the base of a thin the Sugarite mines become evident. We will leave the vans at
coal bed. The boundary here is about 150 ft above the the visitor’s center, follow the trail through the abandoned Sug-
Trinidad Sandstone. arite town site, and continue up the east side of the valley to
At the Raton Pass site three coal beds lie in an 11ftthick examine an exposure of the Sugarite coal bed that contains the
sequence of mudstone, siltstone, and carbonaceous shale. K-T boundary. The coal is about 6 ft thick and was mined for
These three coal beds can be quite definitely correlated with the several years from entries on both sides of the canyon.
Sugarite coal bed on the basis of the presence of the K-T
boundary in both sections. The boundary claystone occurs in a Sugarite
detrital claystone sequence and does not directly underlie a coal
bed, as is the usual case elsewhere in the basin—in fact, its Sugarite is an abandoned mining town. Coal from the Sug-
position beneath the coal bed ranges from 1.5 to 8 in. The arite bed was mined in this valley from about 1902 until the
boundary is below the base of the uppermost coal bed, which is mines were closed in 1941. Coal was produced from Wagon
6 in thick; a 16-in-thick zone of coal and carbonaceous shale Mine No. 2 until the main Sugarite mine (No. 1) was opened in
lies about 6 ft beneath the boundary, and a 28-in-thick coal bed 1912. The coal is high-resin, noncoking, and was prized for
is at the base of the sequence. This coal zone appears to be cor- domestic fuel (Lee, 1917). The Sugarite coal camp was estab-
relative with the coal bed mined at Sugarite, about 3 mi across lished in 1908 by the Chicorica Coal Co., and began full opera-
the high ridge to the northeast. tion the following year
After examining the K-T boundary at the Raton site, we The opening of the Sugarite mine was hailed as one of the
152 C. L. Pillmore, D. J. Nichols, and R. F. Fleming
important developments in northern New Mexico, and the mine base of the coal, and the kaolinite-rich K-T boundary claystone,
produced high-quality domestic coal for more than 30 years. 12 in thick, 6 in below the top (Fig. 12). As shown in the dia-
When the camp first opened it consisted of scattered tents, but gram (fig 13), the boundary claystone contains Ir concentrations
within a short time the construction of a full-fledged company of 2.7 ng/g and also marks the disappearance of the Pro-
town began. Managed by the St. Louis Rocky Mountain and teacidites assemblage and change in fern spore/angiosperm
Pacific Company, concrete and stone dwellings were built on pollen ratios (Pillmore and others, 1984; Gilmore and others,
slopes and terraces along the canyon sides, along with a mer- 1984; and Pillmore and Flores, 1987)
cantile store, schoolhouse, post office, and community center.
The population of the camp fluctuated between 400 and 1,000 Palynology of the Sugarite K-T boundary site
during its years of operation, and this mining community was
considered one of the best coal camps in the area because of its The palynology of the K-T boundary at the Sugarite site
beautiful setting along a running stream, its everyday amenities was first studied by R. H. Tschudy of USGS on the basis of
and community activities, and its first class equipment. samples collected by C. L. Pillmore. Tschudy located the
The Sugarite mines were located high on both sides of boundary within the coal using the palynological extinction
Chicorica Creek. Miners in the canyon relied on mules and bur- datum (Pillmore and others, 1984). This was an unexpected
ros to do the heavy work of pulling carts loaded with coal from result because, in prior investigations in the Raton basin, the
the underground mines to the surface. From there, the coal was boundary was found below a thin coal bed, never within coal.
initially hauled by wagon to Raton, where it was used for steam These results had important implications because they showed
generation and domestic purposes. A few years later, railroads clearly that the palynological K-T boundary datum is indepen-
served both the Sugarite and Yankee camps nearly every day, dent of lithofacies.
with coal runs to and from Raton and on to other markets. The original collection consisted of 15 samples from
Following closure of the mines, some of Sugarite’s houses within and above the Sugarite coal bed. The original samples,
and buildings were moved to Raton, while others were simply which were supplemented by 13 others collected later, became
torn down for salvage. Now all that remains of the camp are the basis for a more detailed study of the palynological and
rows of rock foundations, the old post office, a mule barn, and a
few residences and structures associated with the operation of
the Lake Alice and Lake Maloya water systems. Further
description of the mines and Sugarite Canyon State Park is
given by Virginia McLemore (McLemore, 1990).
The area is mostly covered with landslide debris, soil, and
vegetation. The Sugarite coal bed is exposed near the top of the
lower coal zone of the Raton Formation in a landslide scar at
about 7,300 ft elevation on the east wall of the canyon, about
450 ft above the valley floor. It has also been uncovered in a pit
dug beside the trail. The Trinidad Sandstone forms ledges
about mid-slope. The Sugarite coal bed is only about 90 ft
above the top of the Trinidad and about 10 ft below the base of
the cliff-forming sandstone of the barren series.
The Sugarite coal accumulated in a large, poorly drained
backswamp on a broad, low-gradient, alluvial plain near the
eastern edge of the Raton basin. The thickness of the coal indi-
cates a great accumulation of peat and a long period of stable
conditions in the area. The presence of the isochronous K-T
boundary clay allows us to correlate with confidence the coal
zone observed at the Raton site with the Sugarite coal bed at
Sugarite. The swamp or perhaps several smaller, coeval
swamps extended over an area of roughly 35 sq mi. The site
differs from all other sites located to date in that beds of coal
sandwich the K-T boundary claystone.
0.5 ft
T
Figure 13. Diagram showing the position of the K/T
K boundary claystone bed in the Sugarite coal bed. Inset
shows the variation in Ir values; the double peak in the Ir
curve suggests migration of Ir from the boundary clay-
stone into the adjacent coal.
0 10 20 30 40 50 60 70 80 90 100
1 ft Carbonaceous shale
Coal
paleobotanical changes across the boundary at this unique site the Raton basin. At Sugarite, Proteacidites pollen is present in
(Nichols and others, 1985). The total interval eventually stud- almost every sample below the boundary but in none above it.
ied encompassed 1.6 m of section, including the entire thick- Two important patterns are revealed by these observations: (1)
ness of the Sugarite coal bed and carbonaceous shale intervals as already noted, Proteacidites pollen represents species that
above and below it (Fig. 13). became extinct abruptly at the K-T boundary; and (2) pollen
Both the original study and the detailed study of the Sug- evidence shows that the vegetation varied in composition from
arite site utilized palynology to locate the boundary—specifi- south to north in latest Cretaceous time, yet all plant communi-
cally, the disappearance of the Proteacidites assemblage. At ties were similarly and simultaneously affected by the impact
Sugarite this assemblage was found to include not only the pre- event.
viously reported species Proteacidites retusus, “Tilia” wode- In addition to the extinction of the pollen species named,
housei, Trisectoris sp., and Trichopeltinites sp., but also the K-T boundary at Sugarite is marked by the fern-spore spike,
Aquilapollenites reticulatus, Libopollis jarzenii, Liliacidites the anomalous abundance (here up to 78 percent) of fern spores
complexus, and Tricolpites microreticulatus. These species are (mostly Cyathidites sp.) in samples just above the extinction
well known from the uppermost Maastrichtian at K-T boundary horizon. As interpreted by Tschudy and others (1984), the fern-
sites elsewhere in the Western Interior of the United States and spore spike is evidence of the temporary replacement of a “nor-
Canada (Nichols and Fleming, 1990). In all, 48 palynomorph mal” plant community by one dominated by an opportunistic
taxa were identified, and about 20 percent disappeared at the species—one that can quickly reoccupy a devastated landscape
K-T boundary at the Sugarite site. (see also Fleming and Nichols, 1990).
The occurrence of a species of Aquilapollenites pollen at The detailed study at Sugarite (Nichols and others, 1985)
Sugarite is especially interesting. Several other species of the involved analysis of dispersed fragments of leaf cuticles as well
genus Aquilapollenites disappear at the K-T boundary at local- as pollen and spores. Cuticle fragments, which represent leaf
ities north of those in the Raton basin (Nichols, 1996), but in fossils not otherwise preserved in coal, are recovered by labora-
the Raton basin only Aquilapollenites reticulatus is present, and tory techniques similar to those used in preparing palynologic
it is quite rare. In contrast, Proteacidites retusus and perhaps samples. The palynologic and dispersed-cuticle data suggest
other species of the genus Proteacidites are quite common in that in latest Cretaceous time an ecological transition—a
154 C. L. Pillmore, D. J. Nichols, and R. F. Fleming
hydrosere—was taking place in the mire that eventually formed Izett, G.A., 1990, The Cretaceous/Tertiary boundary interval, Raton Basin, Col-
orado and New Mexico, and its contents of shock-metamorphosed min-
the Sugarite coal. A mire community dominated by dicot
erals; evidence relevant to the K-T boundary impact-extinction theory:
angiosperms was changing to one dominated by monocots, Geological Society of America Special Paper, v. 249, 100 p.
probably palms. Sphagnum moss was also becoming more Izett, G.A., and Bohor, B.F., 1986, Microstratigraphy of continental sedimen-
common in the mire. This gradual change in plant communities tary rocks in the Cretaceous-Tertiary boundary interval in the Western
was abruptly terminated by the K-T extinction event. Ferns Interior of North America: Geological Society of America Abstracts
with Programs, v. 18, no. 6, p. 644.
dominated the mire immediately after the K-T event. Samples
Izett, G.A., and Pillmore, C.L., 1985a, Abrupt appearance of shocked quartz at
from above the coal show that, although palms survived extinc- the Cretaceous-Tertiary boundary, Raton basin, Colorado and New
tion, dicot angiosperms returned to dominance as dicots recolo- Mexico [abs.]: Geological Society of America Abstracts with Programs,
nized the area and the fern-dominated vegetation was replaced vol. 17, no 7, p. 617.
by the profoundly and permanently altered plant communities _____, 1985b, Shock-metamorphic minerals at the Cretaceous-Tertiary bound-
ary, Raton basin, Colorado and New Mexico, provide evidence of aster-
of earliest Tertiary time.
oid impact in continental crust: EOS Transactions of the American
The Sugarite K-T boundary site remains unique in many Geophysical Union, v. 66, p. 1149-1150.
ways. Further studies of the palynology and micro-paleobotany Johnson, K.R., and Hickey, J.J., 1990, Megafloral change across the Creta-
are continuing in conjunction with isotopic studies. ceous/Tertiary boundary in the northern Great Plains and Rocky Moun-
Return to Raton on highway 72. tains, U.S.A., in Sharpton, V.L. and Ward, P.D., eds., Global
catastrophies in Earth history: an interdisciplinary conference on
impacts, volcanism, and mass mortality: Geological Society of America
REFERENCES CITED Special Paper 247, p. 433-444.
Lee, W.T., 1917, Geology of the Raton Mesa and other regions in Colorado and
Alvarez, W., 1997, T. rex and the crater of doom: Princeton, N.J., Princeton New Mexico, p. 9221, in Lee, W.T., and Knowlton, F.H., Geology and
University Press, 185 p. paleontology of Raton Mesa and other regions in Colorado and New
Alvarez, L.W., Alvarez, W., Asaro, F., and Michel, H.V., 1980, Extraterrestrial Mexico: U.S. Geological Survey Professional Paper 101, 450 p. [1918].
cause of the Cretaceous-Tertiary extinction: Science, v. 208, p. 1095- Leffingwell, H.A., 1970, Palynology of the Lance (Late Cretaceous) and Fort
1108. Union (Paleocene) Formations of the type Lance area, Wyoming, in
Bohor, B.F., Foord, E.E., Moires, P.J., and Triplehorn, D.M., 1984, Mineralogic Kosanke, R.M., and Cross, A.T., eds., Symposium on palynology of the
evidence for an impact event at the Cretaceous-Tertiary boundary: Sci- Late Cretaceous and early Tertiary: Geological Society of America Spe-
ence, v. 224, no. 4651, p. 867-869. cial Paper 127, p. 1-64.
Brown, R.W., 1962, Paleocene Flora of the Rocky Mountains and Great Plains: Leighton, V.L., 1980, Depositional environments and petrography of the
U.S. Geological Survey Professional Paper 375, 199 p. Trinidad Sandstone and related formations, Raton area, New Mexico
Fleming, R.F., 1986, Fossil Scenedesmus (Chlorophyta) and its paleoecological [M.S. thesis]: Fort Collins, Colorado State University, 105 p.
significance: Nineteenth Annual Meeting, American Association of Marvin, R.F., Young, E.J., Mehnert, H.H., and Naeser, C.W., 1974, Summary of
Stratigraphic Palynologists, Program and Abstracts, p. 10. radiometric age determinations on Mesozoic and Cenozoic igneous
_____, 1987, Paleoenvironmental significance of fossil Chlorococcalean algae rocks and uranium and base metal deposits in Colorado: Isochron West,
from the Raton Formation, Colorado and New Mexico: Annual Meeting 1974, no. 11, p. 3233 p.
of the Rocky Mountain Section of the Geological Society of America. McLemore, V.T., 1990, Sugarite Canyon: New Mexico Geology, v. 12, no. 2, p.
Fleming, R.F., and Nichols, D.J., 1990, The fern-spore abundance anomaly at 38-42.
the Cretaceous-Tertiary boundary: A regional bioevent in western North Nichols, D.J., 1996, Vegetational history in Western Interior North America dur-
America, in, Kauffman, E.G., and Walliser, O.H., (editors), Extinction ing the Cretaceous-Tertiary transition, Chapter 29D in Jansonius, Jan,
events in Earth history: Lecture Notes in Earth Sciences, Springer-Ver- and McGregor, D.C., eds., Palynology; Principles and Applications:
lag, p. 347-349. Dallas, Texas, American Association of Stratigraphic Palynologists, v.
Flores, R.M., 1984, Comparative analysis of coal accumulation in Cretaceous 3, p. 1189-1195.
alluvial deposits, southern United States Rocky Mountain Basins, in Nichols, D.J., and Fleming, R.F., 1990, Plant microfossil record of the terminal
Stott, D.F., and Glass, D., eds., The Mesozoic of Middle North America: Cretaceous event in the western United States and Canada, in Sharpton,
Canadian Society of Petroleum Geologists Memoir 9, p. 373385. V.L., and Ward, P.D., eds., Global catastrophes in Earth history; an inter-
Flores, R.M., and Pillmore, C.L., 1987, Tectonic control on alluvial paleoarchi- disciplinary conference on impacts, volcanism, and mass mortality:
tecture of the Cretaceous and Tertiary Raton basin, Colorado and New Geological Society of America Special Paper 247, p. 445-455.
Mexico, in Ethridge, F.G., Flores, R.M., and Harvey, M.D., eds., Recent Nichols, D.J., Fleming, R.F., Upchurch, G.R., Tschudy, R.H., and Pillmore,
developments in fluvial sedimentology: Society of Economic Paleon- C.L., 1985, Paleobotanical changes across the Cretaceous-Tertiary
tologists and Mineralogists Special Publication 39, p. 311-320. boundary at Sugarite, New Mexico; new data and interpretations [abs.]:
Flores, R.M., and Tur, S.M., 1982, Characteristics of deltaic deposits in the Cre- Society of Economic Paleontologists and Mineralogists, 1985 Annual
taceous Pierre Shale, Trinidad Sandstone, and Vermejo Formation, Midyear Meeting Abstracts, v. 2, p. 68.
Raton Basin, Colorado: The Mountain Geologist, v. 19, p. 2540. Nichols, D.J., Jarzen, D.M, Orth, C.J., and Oliver, P.Q., 1986, Palynological
Gilmore, J.S., Knight, J.D., Orth, C.J., Pillmore, C.L., and Tschudy, R.H., 1984, and Ir anomalies at Cretaceous-Tertiary boundary, south-central Sask-
Trace element patterns at a nonmarine Cretaceous-Tertiary boundary, atchewan: Science, v. 231, no. 4739, p. 714-717 p.
Western Interior: Nature, v. 307, no. 5948, p. 224-228. Orth, C.J., Gilmore, J.S., Knight, J.D., Pillmore, C.L., Tschudy, R.H., and Fas-
Hickey, L.J., 1981, Land plant evidence compatible with gradual, not catastrophic, sett, J.E., 1981, An Ir abundance anomaly at the palynological Creta-
change at the end of the Cretaceous, Nature, vol. 292, p. 529-531. ceous-Tertiary boundary in northern New Mexico: Science, v. 214, no.
Hildebrand, A.R., and Boynton, W.V., 1988, Provenance of the K-T boundary 4527, p. 1341-1343.
layers, in Global catastrophes in Earth history - An interdisciplinary Penn, B.S, 1994, An Investigation of the temporal and geochemical characteristics,
conference on impacts, volcanism, and mass mortality [abs.]: Lunar and and the petrogenetic origins of the Spanish Peaks intrusive rocks of south-
Planetary Institute Contribution 673, p. 78-79. central Colorado [Ph.D. thesis]: Golden, Colorado School of Mines, 198 p.
Continental Cretaceous-Tertiary boundary in Raton basin, Colorado and New Mexico 155
Pillmore, C.L., 1969,Geologic map of the Casa Grande quadrangle, Colfax Scott, G.R., and Pillmore, C.L., 1993, Geologic and structure-contour map of
County, New Mexico, and Las Animas County, Colorado; U.S. Geolog- the Raton 30' x 60' quadrangle, Colfax and Union Counties, New Mex-
ical Survey Geologic Quadrangle Map GQ-823, scale 1:62,500. ico, and Las Animas County, Colorado: U.S. Geological Survey Mis-
Pillmore, C.L., 1976, Commercial coal beds of the Raton coal field, Colfax cellaneous Investigations Series Map, I-2266, Scale 1:100,000.
County, New Mexico, in Ewing, R.C., and Kues, B.S., eds., Guidebook Smit, Jan, 1984, Evidence for worldwide microtektite strewn field at the Creta-
of Vermejo Park, northeastern New Mexico: New Mexico Geological ceous-Tertiary boundary [abs.]: Geological Society of America
Society Guidebook, 27th Field Conference, p. 227-247. Abstracts with Programs, v. 16, no. 6, p. 659.
Pillmore, C.L., and Flores, R.M., 1987, Stratigraphy and depositional environ- Smith, R.P., 1975, Structure and petrology of Spanish Peaks dikes, south-central
ments of the Cretaceous-Tertiary boundary interval and associated Colorado [Ph.D. thesis]: Boulder, University of Colorado, 191 p.
rocks, Raton basin, New Mexico and Colorado, in Fassett, J.E., and Stormer, J.C., 1972, Ages and nature of volcanic activity on the southern high
Rigby, K.B., Jr., eds., Cretaceous-Tertiary boundary of the San Juan and plains, New Mexico and Colorado: Geological Society of America Bul-
Raton basins, northern New Mexico and southern Colorado: Geological letin, v. 83, p. 2443-2448.
Society of America Special Paper 209, p. 111-130. Stroud, J. R., 1998, Geochronology of the Raton-Clayton volcanic field, New
Pillmore, C.L. and Scott, G.R., 1976, Pediments of the Vermejo Park area, in Mexico, with implications for volcanic history and landscape evolution
Ewing, R.C., and Kues, B.S., eds., Guidebook of Vermejo Park, north- [M.S. thesis]: Socorro, New Mexico Institute of Mining and Technol-
eastern New Mexico: New Mexico Geological Society Guidebook, ogy, 51 p.
27th Field Conference, p. 111-120. Strum, S.R., 1984, Depositional environments and lithofacies of the Raton For-
Pillmore, C.L., and Scott, G.R., 1994, Map showing the geology of the Clifton mation, eastern Raton Basin, New Mexico [M.S. thesis]: Raleigh, North
House 7½' quadrangle, Colfax County, New Mexico, and fossil zones Carolina State University, 81 p.
in the Pierre Shale: U.S. Geological Survey Geologic Quadrangle Map Tschudy, R.H., 1970, Palynology of the Cretaceous-Tertiary boundary in the
GQ-1737, Scale 1:24,000. northern Rocky Mountain and Mississippi Embayment regions, in
Pillmore, C.L., Tschudy, R.H., Orth, C.J., Gilmore, J.S., and Knight, J.D., 1984, Kosanke, R.M., and Cross, A.T., eds., Symposium on palynology of the
Geologic framework of nonmarine Cretaceous-Tertiary boundary sites, Late Cretaceous and early Tertiary: Geological Society of America Spe-
Raton basin, New Mexico and Colorado: Science, v. 223, no. 4641, p. cial Paper 127, p. 65-111.
1180-1182. _____, 1973, The Gasbuggy Core—a palynological appraisal in Fassett, J.E.,
Pollastro, R.M., and Pillmore, C.L., 1987, Mineralogy and petrology of the Cre- ed., Cretaceous and Tertiary rocks of the southern Colorado Plateau:
taceous-Tertiary boundary clay bed and adjacent clayrich rocks, Raton Four Corners Geological Society Memoir, p. 131-143.
basin, New Mexico and Colorado: Journal of Sedimentary Petrology, v. Tschudy, R.H., Pillmore, C.L., Orth, C.J., Gilmore, J.S., and Knight, J.D., 1984,
57, no. 3, p. 456-466. Disruption of the terrestrial plant ecosystem at the Cretaceous-Tertiary
Pollastro, R.M., Pillmore, C.L., Tschudy, R.H., Orth, C.J., and Gilmore, J.S., boundary, Western Interior: Science, v. 225, no. 4666, p. 1030-1034.
1983, Clay petrology of the conformable Cretaceous/Tertiary boundary Tschudy, R.H., and Tschudy, B.D., 1986, Extinction and survival of plant life
interval, Raton Basin, New Mexico and Colorado [abs.]: Annual Clay following the Cretaceous/Tertiary boundary event, Western Interior,
Minerals Conference, 32nd, Buffalo, New York, 1983, Programs and North America: Geology, v. 14, no. 8, p. 667-670.
Abstracts, p. 83 (2 pages). Tweto, O.L., 1987, Rock units of the Precambrian basement in Colorado: U.S.
Powell, J.L., 1998, Night comes to the Cretaceous; dinosaur extinction and the Geological Survey Professional Paper 1321-A, 54 p.
transformation of modern geology: New York, N.Y., W.H. Freeman and Wolfe, J.A., and Upchurch, G.R., 1986, Vegetation, climatic and floral changes at
Company, 250 p. the Cretaceous-Tertiary boundary: Nature, v. 324, no. 6093, p. 148-152.
Scott, G.R., 1986, Historic trail maps of the Raton and Springer 30' x 60' quad- Woodward, L.A., and Snyder, D.O., 1976, Structural framework of the southern
rangles, New Mexico and Colorado: U.S. Geological Survey Miscella- Raton Basin, New Mexico: New Mexico Geological Society Guide-
neous Investigations Series Map, I-1641, Scale 1;100,00. book, 27th Field Conference, Vermejo Park, p. 125-127.
Printed in U.S.A.
Geological Society of America
Field Guide 1
1999
Paul M. Myrow
Department of Geology, Colorado College, Colorado Springs, Colorado 80903, United States; [email protected]
John F. Taylor
Department of Geoscience, Indiana University of Pennsylvania, Indiana, Pennsylvania 15705, United States
James F. Miller
Department of Geography, Geology, and Planning, Southwest Missouri State University, Springfield, Missouri 65804, United States
Ray L. Ethington
Department of Geological Sciences, University of Missouri, Columbia, Missouri 65211, United States
Robert L. Ripperdan
Department of Geology, University of Puerto Rico, Mayaguez, Puerto Rico 00681
Christina M. Brachle
Department of Geology, Colorado College, Colorado Springs, Colorado 80903, United States
INTRODUCTION facies and fauna in these settings are highly responsive to rela-
tive sea level changes and other environmental perturbations.
The numerous extinctions that affected shallow marine However, the mixed carbonate and siliciclastic facies that dom-
faunas on the tropical shelves surrounding Laurentia in the inate inner shelf successions in the Cambrian-Ordovician
Cambrian and Early Ordovician have been the focus of many deposits of Laurentia are generally less fossiliferous than coeval
detailed biostratigraphic, evolutionary, and paleoecologic stud- carbonate platform facies, and in places contain numerous
ies. Data from carbonate platform and off-platform strata have stratigraphic gaps. As a result, they have received less attention
been used to propose process-response models that invoke sea than outer shelf facies. Data on trilobite biofacies is sparse for
level change as a forcing mechanism for extinctions and/or faunas that occupied proximal shelf and cratonic environments.
radiations within the Cambrian and Early Ordovician. Some In addition, sedimentologic information in most previous bios-
regressive features observed near horizons of faunal change tratigraphic studies of inner shelf assemblages is either lacking
within the Cambrian-Ordovician boundary interval on various or of insufficient precision to determine whether horizons and
continents have been used to propose a series of “eustatic intervals of faunal change coincide with sequence boundaries
events” (Nicholl et al., 1992). These include the “Lange Ranch or significant facies transitions. Chemostratigraphic techniques
Eustatic Event” and “Black Mountain Eustatic Event” of such as carbon isotope stratigraphy, when used in conjunction
Miller (1984, 1992) and the Acerocare Regressive Event and with high-resolution biostratigraphic information and detailed
Peltocare Regressive Event of Erdtmann (1986). There is much sedimentologic and sequence stratigraphic data, can signifi-
debate about the nature of these proposed events (Ludvigsen et cantly enhance temporal correlations between these deposits
al. 1986; Taylor et al. 1992; Landing 1993) based, at least in and the more heavily studied distal platform facies.
part, on the ambiguous nature of the sedimentological data and On this field trip we will examine a precise and integrated
insufficient precision of correlation. stratigraphic framework recently established for the Cambrian-
A rigorous test of the proposed linkage between extinctions Ordovician inner shelf deposits of Colorado. Within this frame-
and paleoceanographic events within the Cambrian-Ordovician work we will observe a complex record of relative sea level
boundary interval will require varied, high-resolution strati- changes (as recorded in the stratigraphic succession of lithofa-
graphic data from a complete onshore-offshore profile, includ- cies and sequence boundaries), paleoceanographic events
ing inner shelf, platform, and shelfbreak settings. Detailed data (reflected in the isotope stratigraphy), and bioevents (extinction
from inner shelf environments are particularly critical because horizons and intervals of adaptive radiation).
Myrow, P. M., Taylor, J. F., Miller, J. F., Ethington, R. L., Ripperdan, R. L., and Brachle, C. M., 1999, Stratigraphy, sedimentology, and paleontology of
the Cambrian-Ordovician of Colorado, in Lageson, D. R., Lester, A. P., and Trudgill, B. D., eds., Colorado and Adjacent Areas: Boulder, Colorado, Geological
Society of America Field Guide 1.
157
158 P. M. Myrow et al.
LITHOSTRATIGRAPHY
NORTHWEST SOUTHEAST
SUB-BASIN SUB-BASIN
Lower Paleozoic rocks of Colorado unconformably overlie
Proterozoic crystalline basement that include gneisses and Leavick
Tarn
Tie Gulch
FORMATION
Dolomite
FORMATION
Member
MANITOU
Member
MANITOU
? HM
Clinetop
FORMATION FORMATION
FORMATION FORMATION
LP MG
SAWATCH PEERLESS
DOTSERO
Crested
Butte Member
? Manitou WC
Springs Colorado
Springs
Glenwood
Canyon
SE Canyon
Member
Uncompaghre Sub-Basin City
Uplift Sierra
Grande
SAWATCH
N 25 Mi.
Uplift
Figure 1. Cambrian-Ordovician paleogeography of Colorado. Location
abbreviations: SF=South Fork, MEC=Main Elk Creek, GC=Glen-
wood Canyon, EBC=East Brush Creek, LC=Lime Creek, LP=Lam-
bertson’s Peak, HM=Horseshoe Mountain, MG=Missouri Gulch,
WC=William’s Canyon. Conodont biostratigraphy for these locations Figure 2. Lithostratigraphy of Cambrian and Ordovician deposits of
is given in Figure 5. Modified from Gerhard (1972). Colorado. Modified from Gerhard (1972).
Figure 3. Schematic lithostratigraphic correlation of units between the northwest and southeast sub-basins.
160 P. M. Myrow et al.
the Franconian Stage of the Cambrian, from these younger strata macrofossils and is responsible for the biostratigraphic and
and assigned them to the Peerless Formation. No fauna had ever paleogeographic interpretations based upon the trilobite data. A
been found in the type section (Horseshoe/Peerless Mountains) suite of macrofossils was also collected at Missouri Gulch (Stop
of the Peerless Formation until our recovery of conodonts of the 2) by James D. Loch of Central Missouri State University. Con-
Eoconodontus Zone from the upper Peerless (Myrow et al. odont samples were processed by Ethington and Miller. Carbon
1995) and Lower Ordovician Rossodus manitouensis Zone con- isotope study was undertaken by Colorado College students in
odonts from the base of the Manitou Formation. This indicates: collaboration with Ripperdan at the University of Puerto Rico-
(1) that the glauconitic unit along the Front Range is not the Mayaguez and Claudia Mora at the University of Tennessee.
Peerless Formation, but the middle member of the Sawatch For-
mation (Stop 1; Fig. 4), and (2) a substantial Cambrian–Ordovi- STOP 1A: FRONT RANGE
cian unconformity exists at Horseshoe Mountain. We now
recognize that this is the same sub-Ordovician unconformity that Take Highway 24 West from Colorado Springs to Manitou
Berg and Ross (1959) showed to progressively cut out underly- Springs (Fig. 6). Just past mile marker 298 there is a stoplight
ing Cambrian strata along the Front Range. The Manitou For- for Manitou Springs/Cave of the Winds. Go left at the light and
mation onlaps onto this erosion surface—from north to south proceed downhill along the Fountain Creek to a stop sign. Take
(Berg and Ross 1959; Fig. 4)—and rests successively on the a SHARP right up a hill leading to Hwy 24 for 0.4 miles.
middle Sawatch (Stops 1 and 2), lower Sawatch, and Precam- CAREFULLY cross the road into a pulloff on the left.
brian basement (Deadman Canyon, Colorado Springs) (Fig. 4).
The Manitou is of variable thickness and reaches up to 112 Lower Sawatch Description
m in the Sawatch Range (Stevens 1961). There are differences
in lithofacies within the Manitou between northwestern and At this stop we will examine the sedimentology of the
southeastern sub-basins, and therefore separate member lower and middle members of the Sawatch Formation, which is
nomenclatures exist for each area (Gerhard 1972, 1974; Fig. 2). summarized by Myrow (1998; Fig. 7). Along the Front Range
Everywhere in the southeast sub-basin, the entire Manitou is these Upper Cambrian rocks rest nonconformably on 1.75-Ga
Lower Ordovician (Rossodus manitouensis Zone or younger). metamorphic rocks of the Idaho Springs Formation and
In the more complete western Colorado exposures, the Cam- younger (1030 Ma) intrusive rocks of the Pikes Peak Granite.
brian–Ordovician boundary occurs within the lower part of the The lower member of the Sawatch Formation consists of 4.0 to
Manitou as it is defined in that area with its base at the top of 4.5 m of white-weathering, thin to medium bedded (5 to 20 cm
the Clinetop stromatolite bed. An interval of sandy dolomite thick), coarse, very coarse, and pebbly quartz arenite with < 5%
and dolomitic sandstone occurs immediately above the Dotsero feldspar (Lewis 1965). The nonconformity is remarkably flat at
in the western part of the White River Plateau and near the the outcrop scale, except where several of the large 0.5 to 3-m
Homestake Shear Zone in the northwestern part of the Sawatch diameter corestones in the underlying granite project up to 40
Range. Recovery of lower Ibexian conodonts from this unit cm above the nonconformity surface into the overlying
indicates that it is a proximal facies of the Manitou Formation. Sawatch. Stratification in the Sawatch clearly abuts against the
corestone, indicating that these are Precambrian-age weather-
PREVIOUS WORK ing features.
The lower 2.3 m of the lower Sawatch is parallel laminated
Most previous work on the Cambrian–Ordovician of Col- with widely-spaced, small-scale (<8 cm thick) trough cross-
orado (Anderson 1970; Bass and Northrop 1953; Brainerd et al. stratification, as well as thin pebble lags. Well-preserved polyg-
1933; Campbell 1972; Gerhard 1974; Johnson 1934, 1944, onal cracks filled with pebbly sandstone occur at the 0.95-m
1945; Ross and Tweto 1980; Tweto 1949) are invaluable mark. Above 2.3 m beds are dominantly bioturbated and sepa-
sources of basic information, but they contain very little rated by partially dispersed (bioturbated) pebble lags with sand-
process-oriented sedimentology or modern stratigraphic analy- stone intraclasts up to 10 cm in diameter. The upper 60 cm of
ses. In addition, little paleontological work had been done since the lower Sawatch is a distinctive white-weathering, bioturbated
the trilobite study of Berg and Ross (1959). In this guidebook pebbly coarse sandstone bed with abundant burrows, including
we present some of the results of our research on the Cam- Teichichnus and lined burrows of cf. Paleophycus. The upper-
brian–Ordovician of Colorado over the past seven years. We most 10 cm of the bed consists of a hematite-coated lag with
examined nearly 40 measured sections, including nearly all abundant quartz pebbles, large intraclasts of quartz sandstone,
those listed in Ross and Tweto (1980) and Bass and Northrop as well as pyritic steinkerns and pyrite-replaced and silicified
(1953). A few from each region were chosen for detailed cephalopod, gastropod, trilobite, and brachiopod shells. In thin
description and high-resolution sampling for biostratigraphic section, small, mm-diameter burrows are seen projecting
and chemostratigraphic analyses (Figs. 1, 3, 5). Myrow pro- inwards radially around rounded (< 1 cm diameter) intraclasts
vided the detailed lithologic descriptions and sedimentological of mudstone that must have been firm enough early on to with-
interpretations. Taylor systematically sampled the sections for stand transport and rounding. Directly overlying the lag surface
MOSQUITO RANGE — FRONT RANGE
CORRELATION
Horseshoe
Mountain Traditional View
Williams
Canyon Deadman's
Missouri
Gulch Canyon
Manitou
ce
Surfa
y
Unc onformit
C-O
Peerless
Sawatch
Figure 5. Conodont biostratigraphy for selected locations in western Colorado. Shaded areas denote intervals missing at unconformities. Note increase in number of zones/subzones
present in more distal sections to northwest.
Stratigraphy, sedimentology, and paleontology of the Cambrian-Ordovician of Colorado 163
volithes, and possible Thalassinoides horizontalis (Myrow over many tidal cycles. Because sediment transport increases
1995). There is a decrease in bioturbation and an increase in the nonlinearly with increasing shear stress (Middleton and
abundance of trough cross-stratified and intraclast-rich beds Southard 1984), small tidal asymmetry could have produced
towards the top of the member. locally consistent bedform migration directions.
A distinctive, 50–100-cm-thick, resistant-weathering, red, The greater scatter in the paleocurrent data of small-scale
coarse dolostone unit marks the contact between the middle cross-bedding (=superimposed dunes) is similar to modern tidal
Sawatch and overlying Manitou Formation. This highly recrys- dunes (Houbolt 1982; Fenster et al. 1990; Davis et al. 1993).
tallized coarse dolostone bed has a sharp erosional lower sur- This is a reflection of the highly unsteady and nonuniform
face (~40 cm of relief) and a complex internal nature of the currents at the base of tidal flows. A large percent-
microstratigraphy that includes karstic cavities with collapse age of the smaller dunes migrated directly up and down the
breccia and carbonate cement-filled vugs. Conodonts from this dune faces. This may indicate that the high threshold velocities
bed are from the Rossodus manitouensis Zone, similar to those required to move the coarse sediment of these dunes may have
from the base of the Manitou Formation at the Missouri Gulch only been exceeded during peak flow conditions, during which
section (Stop 2). The immediately overlying carbonate beds time secondary flow may have been oriented nearly perpendic-
begins with conodonts of the low-diversity interval (= base of ular to the dune crests.
Fauna D of Ethington and Clark 1981). Two closely spaced Interpretations of depositional environments are difficult
sequence boundaries therefore bracket this bed. The basal sur- for the rest of the upper member because it lacks diagnostic sed-
face is the Cambrian–Ordovician boundary unconformity, imentary structures (e.g., desiccation cracks). The lack of wave-
which along the Front Range of Colorado represents a consid- or storm-diagnostic features and the abundance of trough cross-
erable hiatus interpreted as a depositional sequence boundary stratification indicates that tidal currents may have continued to
within Palmer’s (1981) Sauk III subsequence (Landing 1993). influence deposition. Coarse quartz-rich sand was likely mixed
The stratigraphic shift to the glauconitic tidal dune deposits with locally derived grains of carbonate and glauconite during
of the lowermost middle member represents a significant pale- episodic flooding events. The upper member shoals to the Cam-
oenvironmental change across a tidal (?) ravinement surface. brian-Ordovician unconformity, as indicated by an upward
Large tidal dunes in many locations globally rest directly on increase in abundance of trough cross-stratification and flat-
ravinement surfaces produced during Holocene transgression pebble conglomerate and much less evidence of bioturbation.
(e.g., Davis et al. 1993). Marine transgression has generally
been implicated in the formation of dune complexes (e.g., Hine STOP 1B: FRONT RANGE
1977) and in deposition of glauconite (Brasier 1980; Odin and
Fullagar 1988). The dune deposits are considered to be con- Retrace route for Stop 1A by going downhill 0.4 mi, taking
densed deposits, for they are rich in glauconite and formed in a sharp left just past the gate to Manitou Springs at the first
response to deepening above a ravinement surface. The initia- intersection. Proceed to hairpin turn and park in pulloff.
tion of bedform development is likely to have occurred as the For those who wish to further examine the Upper Cambrian
result of amplification of tidal currents due to the interaction of deposits, exposure of the same section as Stop 1A occurs at the
rising relative sea level and the geometry of the transgressed sharp hairpin turn parking lot. Here the tidal dune deposits con-
landscape. The area around Manitou Springs must have been a tain much more trough cross-bedding, most of which is highly
large embayment in the Cambrian shoreline that had a funnel- oblique to the large-scale foresets. Several yards down hill from
ing effect on tidal currents and possibly also locally created the parking lot along the creek are extensive bedding planes in
geochemical conditions (i.e., slightly reducing) that favored strata from just above the dune deposits. These contain abundant
glauconite formation. Myrow (1998) calculated a minimum eocrinoid columnals and bioturbated dolomitic and glauconitic
water depth for the tidal dune deposits of 21 m based on pre- deposits. The main outcrop of Stop 1B is further downhill on the
served bedform height, but considers 35 m to be a more reason- east side of the road. Here there are exposures of the limestone
able estimate. Transgression was therefore potentially rapid, and dolostone of the Lower Ordovician Manitou Formation.
given that there is only 3.1 m of section between the surface At this stop, the lowermost Manitou contains well-devel-
with desiccation cracks and the base of the dune deposits.
The nearly symmetrical geometry and low dips of stoss and
lee sides of the Peerless deposits resemble modern tidal dunes Figure 7. Detailed measured section showing lithostratigraphy; strati-
(e.g., Fenster et al. 1990) that form under less extreme condi- graphic changes in quartz and glauconite; and sedimentological,
tions of tidal asymmetry. Deposits of such dunes (Allen’s sequence stratigraphic, and paleoenvironmental interpretations. The
[1980] Class V and VI) are poorly documented or understood left side of the grain size scale shows fine sand (FS) to cobble (Cob)
from the ancient. The uniform NNW dips of the large-scale for siliciclastic sediment. The right side shows mudstone, grainstone
and flat pebble conglomerate in carbonate facies. Measured section
foresets may represent the orientation of the dominant tidal cur- from Stops 1A and 1B, Manitou Springs (southeast part of Sec. 31,
rent (flood or ebb) as tidal dunes are generally flow-transverse. T. 13 S., R. 67 W., Manitou Springs 7.5' quadrangle, El Paso County,
Large-scale bedform migration resulted from net accumulation Colorado).
166 P. M. Myrow et al.
Grainstone
the Manitou at Missouri Gulch yield species of Symphysurina to 50% of the strata. These beds are tabular and generally range
and Bellefontia, confirming the presence of the Bellefontia- from 7-30 cm in thickness, although a few beds are > 1 m thick.
Xenostegium Zone, which is missing farther south in the Front The tabular clasts, which consist almost entirely of parallel-
Range because of delayed onlap of the basal Manitou. laminated limestone and dolostone, are usually less than a cen-
Conodont collections from the Manitou at this outcrop, the timeter thick and measure a few centimeters in their long and
type locality for the conodont Rossodus manitouensis, are intermediate axes. The matrix of the flat-pebble beds is fine to
impressive in size and quality. The Color Alteration Index coarse grainstone. The flat pebbles are dominantly oriented
(CAI) is very low. Conodonts from the lower part of the section with their long axes sub-parallel to horizontal.
are from the Rossodus manitouensis Zone and those from the The basal part of the overlying Manitou Formation is litho-
upper part of the section yield a less diverse assemblage char- logically identical to the Glenwood Canyon Member of the
acteristic of the low-diversity interval (lower Fauna D of Dotsero and varies in thickness from a few meters at Glenwood
Ethington and Clark 1981) (Figs. 5, 10). Canyon to nearly 10 m at Main Elk Creek (Stop 4). This facies
A karstic breccia horizon has been recently discovered at is typical of inner shelf lagoon deposits of the inner detrital belt
Missouri Gulch within the basal few meters of the Manitou For- (Lochman-Balk 1971) of the paleo-Pacific Ocean during the
mation. Conodont samples below and above the karst both yield Late Cambrian to Early Ordovician. The stratification in the
Rossodus manitouensis Zone conodonts, so the duration of the grainstone beds reflects deposition by wave and storm
hiatus associated with the paleokarst is not well resolved. A processes. Flat-pebble beds are also generally considered to
deeply channeled unconformity, present at the top of the Mani- result from storm processes. The clasts are similar in lithology,
tou Formation, is overlain by the Missisippian William’s texture, thickness, and sedimentary structures to the interbed-
Canyon Member of the Leadville Limestone. ded thin grainstone beds, which indicates that these beds were
broken up by storms and redeposited.
STOP 3: GLENWOOD CANYON Slump structures are abundant in the upper 15 m of the
Dotsero Formation and the lower 2 m of the Manitou Forma-
Heading west on I-70 past Vail and Eagle, take the Dotsero tion (Fig. 11). These features include enigmatic, isolated, coher-
Exit 133. Go right (north) off the exit ramp to the first intersec- ent blocks and contorted folds that appear to have originated as
tion and turn left (west) towards the Glenwood Canyon trail- locally derived sea-floor slumps and slides. They are particu-
head parking lot. Follow the service road parallel to the larly abundant in the stratigraphic interval (+/- 2 m) above and
highway for nearly 3 miles to reach the parking lot. From here, below the Clinetop bed. Some large slide blocks rest directly on
take the paved trail by foot under the highway and along the the upper surface of the stromatolitic biostrome. Well-preserved
Colorado River until it cuts back under the highway (about 0.9 thrust-faulted beds show classic fault-bend folds, and some in
miles). Emerging from the underpass turn right immediately cases these folds detached and came to rest well beyond the
onto a faint trail that leads to outcrops above the level of the thrust ramp as isolated bodies. Sparse preliminary orientation
highway. data indicate bipolar orientations for thrust directions, which
This is the first of two sections to be visited (see also Stop might argue against downslope, gravity-driven failures. Failure
4) in the White River Plateau region to examine the Dotsero and under shear stresses produced by wave oscillations would be
Manitou formations (Fig. 11). The Dotsero Formation is consistent with this data, particularly given the abundant evi-
divided into two members, a 25-35 m thick lower Glenwood dence for storm currents in this facies.
Canyon Member and a 0.5-1.5 m thick stromatolitic biostrome, The Clinetop Member is a widespread marker bed that pre-
the Clinetop Member (Fig. 2). The Glenwood Canyon Member viously was believed to occur only in the White River Plateau
consists of a complexly interbedded very thin- to thin-bedded area. Our discovery of the bed in the northwestern part of the
shale, very thin- to medium-bedded grainstone, and thin- to Sawatch Range to the south (East Brush Creek and Lime Creek,
thick-bedded flat-pebble conglomerate. The shaly intervals Figs. 1, 3) extends its range to approximately 3500 km2. Several
range from a few cm to 60 cm in thickness and consist of alter- extremely widespread stromatolite beds occur in the Upper
nating mm-scale thin laminae to very thin (< 3 cm) beds of Cambrian–Lower Ordovician of the Great Basin as well, and
shale and fine grainstone. The grainstone intervals are generally these appear to mark a worldwide resurgence of stromatolites
less than 25 cm in thickness, although some beds reach 90 cm. at this time (Shapiro 1998). This resurgence was likely due to
Thin, 1-3 cm thick, grainstone beds in some cases form amal- relatively high sea levels and an equatorial position of Lauren-
gamated units up to 1.5 m in thickness. The grainstone beds are tia. The Clinetop bed contains several irregular hardground sur-
dominantly parallel-laminated although many beds contain faces marked in part by truncation of stromatolitic lamination.
well-developed wave-ripple and small- to medium-scale hum- In nearly all Clinetop localities, the upper surface is a flat hard-
mocky cross-stratification. The bases of grainstone beds are ground that is directly overlain by a wave-rippled grainstone
covered with sole markings, including groove and prod casts. bed that is locally glauconitic.
Carbonate flat-pebble beds are more abundant in the upper Red silt-filled fractures interpreted as possible paleokarst
half of the Glenwood Canyon Member, locally making up close features occur within the Clinetop bed at Glenwood Canyon.
Figure 11. Generalized stratigraphic column of
Glenwood Canyon with conodont and trilobite
zonations.
170 P. M. Myrow et al.
Such paleokarst features also occur at several horizons within studied to date. All conodont zones are represented from Pro-
two to three meters of section directly below the Clinetop at this condontus tenuiserratus to Rossodus manitouensis zones,
location and in a number of beds both below and above the except for the Proconodontus muelleri Zone and the Hir-
Clinetop at Main Elk Creek (Stop 4). The Clinetop appears to sutodontus hirsutus Subzone of the Cordylodus proavus Zone.
be more strongly karsted at a remote location on top of the The section is highly condensed relative to those in the Great
White River Plateau, where the top of the bed is also cut by a Basin and elsewhere because of low subsidence rates in this
series of grainstone-filled channels. Conodonts recovered from cratonal setting. Nonetheless, it is one of the most complete
up to a meter above the bed are from the same conodont sub- known from the inner detrital belt in North America and was
zone as the bed itself (Eoconodontus notchpeakensis Subzone), deposited far enough into the inner shelf lagoon to record a pro-
so any break at the top of the bed may be minor. A larger break tracted history of relative sea level changes, carbon isotope fluc-
may be represented at the irregular base of the bed because the tuations, and conodont evolution.
highest conodont sample below the bed (~1 m below) is A detailed d13C profile from the Main Elk Creek section
assigned to the Proconodontus posterocostatus Zone. Trilobite (Figure 14) extends the depositional interpretations made from
range data also suggest the presence of a stratigraphic break or biostratigraphic and lithostratigraphic information. Carbon iso-
condensed interval within or just below the Clinetop Member tope data from outer platform sections show a strong cyclicity
(see Stop 4). Deposition of the Clinetop Member is considered in d13C values during the Cambrian-Ordovician boundary inter-
to be a three-stage process, as illustrated in Figure 12. val, followed by a sharp decline to more negative values near
The abundance and spacing of thin karst horizons within the base of the Rossodus manitouensis Zone (Ripperdan and
the Dotsero and lower Manitou formations reflect generally low Miller, 1995). The rapidity of d13C variation, coupled with rel-
accommodation space and the occurrence of high frequency, atively continuous deposition on the outer platform, facilitate
low amplitude relative sea level changes. These fluctuations did high-resolution correlation of the more fragmentary d13C record
not impart a cyclical stratigraphic facies pattern, in part because at Main Elk Creek. Examples of this are found at the 1st and
this was not a depositional system characterized by in situ car- 3rd karst horizons at Main Elk Creek. Overlying strata contain
bonate production (i.e., a system driven to keep up with sea d13C variations and conodont faunas that can be precisely cor-
level). Instead, the carbonate beds shows clear evidence of trac- related to the contemporaneous section at Lawson Cove, Wah
tion transport (with selective size sorting) from adjacent envi- Wah Mountains, Utah (d13C profile events Ibe6 and Ibe7, Figure
ronments during storms, which alternated with ambient 14). The range of d13C variation found within each section is
deposition of mud from suspension. approximately 2.0 permil, permitting temporal approximation
The bulk of the overlying Manitou Formation consists of of intermediate d13C values between known maxima and min-
parallel laminated and hummocky cross-stratified dolomitic ima. Equivalents to the Lawson Cove Ibe4 and Ibe5 d13C profile
grainstone and a few thin shale and siltstone beds. These strata events are virtually absent at Main Elk Creek, thus constraining
are the record of a long-standing, high-energy, storm-domi- the duration of non-deposition (exposure?). A similar analysis
nated, epicratonic carbonate setting. Some intervals particularly of profile events Ibe1 and Ibe2 suggest that the 1st karst horizon
rich in quartz sandstone represent potential sequence bound- also represents a depositional hiatus of substantial duration.
aries. The basal Manitou Formation at MEC is quite shaly for
about 10 m above the Clinetop bed before it gives way to more
STOP 4: MAIN ELK CREEK typical amalgamated grainstone of the upper Manitou (Fig. 13).
This is in striking contrast to the Glenwood Canyon section
Take Exit 105 off of I-70 west of Glenwood Springs. Go where the amalgamated grainstone facies begins less than a
right (north) to the first intersection. Turn left onto the service meter above the Clinetop. This may reflect the more distal
road leading west to New Castle. Drive 1.1 miles along this (=more western) position of this outcrop relative to the Homes-
road, which turns into Main Street. Before the end of town at 7th take Shear Zone.
Street, turn right onto 245 RD (Buford Road) and follow for 3 Extraordinary thin karsted beds occur in the section, one of
miles to FR 243. Turn right at the sign indicating “National For- which is 1 m below the Clinetop. Several more are scattered up
est Access, Main Elk Creek and Clinetop Road.” FR 243 will fol- to 12 m above the Clinetop (Fig. 13). Karsting features are
low Main Elk Creek for 5.8 miles (last 0.9 mi on dirt road). Just developed within flat pebble and thick grainstone beds. These
before small bridge and steep winding hill, look for Warren features include red silt-filled and cement-filled veins and cavi-
Stalts’ Elk Creek Ranch on the left. Park here and walk dirt road ties whose edges show sharp truncation of flat-pebble clasts and
to his house and ask for access to the outcrop on the west side of grainstone laminae. Large mud-crack features occur locally in
Main Elk Creek to the west of the rancher’s property. float, and in a few places on bedding planes. The karst surfaces
Sections in the northwest sub-basin are significantly more correspond with interruptions in isotopic curves that verify the
complete through the Cambrian-Ordovician boundary interval loss or partial preservation of some conodont subzones.
(Fig. 5) than those in the southeast sub-basin, and Main Elk Intensive sampling for trilobites in the Main Elk Creek sec-
Creek (MEC) (Fig. 13) is the most complete section of those tion has documented the presence of several zones and sub-
Stratigraphy, sedimentology, and paleontology of the Cambrian-Ordovician of Colorado 171
zones not previously reported from Colorado, including the two tions deposited on the eastern (Appalachian) side of Laurentia.
highest zones of the Cambrian (Illaenurus and Saukia Zones) W. bulbosa was recovered by A.R. Palmer from the thin grain-
and the two lowest zones of the Ordovician (Missisquoia and stone bed atop the Clinetop Member at Glenwood Canyon. C.
Symphysurina Zones). Several trilobite taxa from the MEC and typicalis is the commonest species present in the Illaenurus
GC sections strongly suggest that these strata were deposited on Zone collections recovered from MEC and from two sections
the western side of the Transcontinental Arch. For example, in the Sawatch Range to the south.
Wilcoxaspis bulbosa and Clelandia typicalis are common ele- The Cambro-Ordovician boundary at MEC occurs at a
ments of the Saukia and Illaenurus Zones (respectively) in cryptic unconformity whose hiatus includes the highest subzone
western North America, but have never been reported from sec- of the Saukia Zone and lowest subzone of the Missisquoia
200
Ibe7
Rossodus
Rossodus manitouensis manitouensis 180
Cordylodus
Cordylodus angulatus angulatus Ibe6 160
Iapetognathus Iapetagnathus n.sp.1
Cordylodus
Cordylodus lindstromi lindstromi s.l.
140
Clavohamulus hintzei Clavohamulus Ibe5
hintzei * **
120
HOUSE LIMESTONE
Ibe4
Hirsutodontus
Hirsutodontus simplex simplex *
Clavohamulus elongatus Clavohamulus 100
Fryxellodontus inornatus elongatus
Ibe3
Fryxellodontus
** Ibe2 80
inornatus
Cordylodus
*
proavus
Ibe1
* 60
*
Cambrooistodus minutus Cambrooistodus 40
LAVA DAM MEMBER
minutus
20
Eoconodontus
NOTCH PEAK FORMATION
notchpeakensis
RED
TOPS
Figure 15. Diagram contrasting positions of biomere and stadial boundaries in Laurentian platform successions. Black arrows in stage column
indicate levels (bases of biomeres) dominated by olenimorphic trilobites illustrated to right of column. Cranidia illustrated (bottom to top) are
Aphelaspis, Parabolinoides, and Apoplanias. Brachiopods to the right of Parabolinoides and Apoplanias are Eoorthis and Apheoorthis, respec-
tively. Jagged lines and question marks used to represent biomere boundaries indicate potential diachroneity.
Zone. This relationship is similar to that reported by Miller Subzone (marked by the appearance and dominance of Apopla-
(1984, 1992) in his description of his Lange Ranch Eustatic nias) is a better horizon for designation as the top of the
Event. Recovery of the trilobite Apoplanias rejectus from strata youngest Cambrian biomere (the Ptychaspid Biomere) than
immediately above the unconformity assigns these beds to the other zonal boundaries utilized for that purpose in previous
Missisquoia typicalis Subzone and establishes a precise corre- studies (Longacre 1970; Stitt 1975; Palmer 1984; among oth-
lation with boundary sections throughout North America. The ers). Figure 15 shows the relationship of the biomere bound-
collections of Apoplanias from MEC are also worthy of men- aries to the extinction horizons that currently serve as stadial
tion because they contributed to a recent breakthrough in the boundaries for the Upper Cambrian of Laurentia. With this
study of biomeres, the stage-level biostratigraphic units origi- repositioning, the top of the Ptychaspid Biomere is marked by
nally defined by Palmer (1965) in Upper Cambrian platform the appearance of an olenimorph-dominated fauna of minimum
successions in North America. Each biomere (a few million taxonomic and morphologic diversity similar to those at the
years in duration) records an adaptive radiation of Laurentian bases of the Aphelaspis and Taenicephalus Zones, which define
platform trilobite faunas, followed by a step-wise decline in the bases of the underlying Marjumiid and Pterocephaliid Bio-
species diversity through an extinction interval, culminating in a meres, respectively.
return to minimum diversity which marks the base of the next Interestingly, at MEC, the horizon of major turnover in
biomere. conodont and trilobite faunas (the disconformity whose hiatus
Analysis of the MEC collections led to the realization includes both the base of the Ibexian Series and top of the Pty-
(Taylor 1997; Taylor, in prep.) that the base of the M. typicalis chaspid Biomere) lies a short distance below a surface of karst-
Stratigraphy, sedimentology, and paleontology of the Cambrian-Ordovician of Colorado 175
ing (Fig. 13). This same pattern, i.e., mass extinction preceding odonts from the Ibex area, western Millard County, Utah: Brigham
evidence for exposure, characterizes other biomere boundaries Young University Studies, v. 28, pt. 2, 155 p.
in inner platform facies in western North America (Robert Fenster, M.S., Fitzgerald, D.M., Bohlen, W.F., Lewis, R.S., and Baldwin, C.T.,
1990, Stability of giant sand waves in eastern Long Island Sound,
Thomas, pers. comm., 1997) and in the Appalachians (Loch and U.S.A.: Marine Geology, v. 91, p. 207-225.
Taylor 1995; Taylor et al. 1999). Collections from the 6-7 Frey, R.W. and Seilacher, A., 1980, Uniformity in marine invertebrate ichnol-
meters of strata below the Clinetop Member at MEC and in the ogy: Lethaia, v. 13, p. 183-207.
Glenwood Canyon section confirm the presence of the Illaenu- Gerhard, L.C., 1972, Canadian depositional environments and paleotectonics,
rus Zone in the Dotsero Formation. Strata above the Clinetop central Colorado: in De Voto, R.H., ed., Paleozoic stratigraphy and
structural evolution of Colorado: Quarterly of the Colorado School of
in those sections yielded species characteristic of the overlying Mines, v. 67, no. 4, p. 1-36.
Saukia Zone, specifically, the Saukeilla serotina Subzone. At Gerhard, L.C., 1974, Redescription and new nomenclature of Manitou Forma-
present, the trilobite data indicate that the underlying Saukiella tion, Colorado: American Association of Petroleum Geologists Bulletin,
junia Subzone is very thin or absent. v. 58, p. 1397-1406.
Hine, A.C., 1977, Lily Bank, Bahamas: History of an active oolite sand shoal:
Journal of Sedimentary Petrology, v. 47, p. 1554-1581.
REFERENCES CITED Houbolt, J.J.H.C., 1982, A comparison of recent shallow marine tidal sand
ridges with Miocene sand ridges in Belgium, in Scrutton, R.A. and Tal-
Allen, G.P., 1992, Sedimentary processes and facies in the Gironde estuary: a wani, M., eds., The Ocean Floor: New York, John Wiley and Sons, Ltd.,
recent model for macrotidal estuarine systems, in Smith, D.G., Reinson, p. 69-80.
G.E., Zaitlin, B.A., and Rahmani, R.A., eds., Canadian Society of Petro- Johnson, J.H., 1934, Paleozoic formations of the Mosquito Range, Colorado:
leum Geologists Memoir 16, p. 29-40. U.S. Geological Survey Professional Paper 185-B, 43 p.
Allen, G.P. and Posamentier, H.W., 1993, Sequence stratigraphy and facies Johnson, J.H., 1944, Paleozoic stratigraphy of the Sawatch Range, Colorado:
model of an incised valley fill: the Gironde Estuary, France: Journal of Geological Society of America Bulletin, v. 55, p. 303-378.
Sedimentary Petrology, v. 63, p. 378-391. Johnson, J.H., 1945, A resumé of the Paleozoic stratigraphy of Colorado: Col-
Allen, J.R.L., 1980, Sand waves: A model of origin and internal structure: Sed- orado School of Mines Quarterly, v. 40, 109 p.
imentary Geology, v. 26, p. 281-328. Landing, E., 1993, Cambrian–Ordovician boundary in the Taconic Allochthon,
Anderson, T.B., 1970, Cambrian and Ordovician stratigraphy of the southern eastern New York, and its interregional correlation: Journal of Paleon-
Mosquito Range, Colorado: The Mountain Geologist, v. 7, p. 51-64. tology, v. 67, p. 1-19.
Bass, N.W. and Northrop, S.A., 1953, Dotsero and Manitou formations, White Lewis, J.H., 1965, Petrology and diagenesis of Upper Cambrian rocks of central
River Plateau, Colorado, with special reference to Clinetop Algal Lime- and western Colorado: Unpublished PhD thesis, University of Colorado,
stone Member of Dotsero Formation: American Association of Petro- Boulder, Colorado, 184 p.
leum Geologists Bulletin, v. 37, p. 889-912. Loch, J.D., and Taylor, J.F., 1995, High-resolution biostratigraphy in the Upper
Belknap, D.F. and Kraft, J.C., 1981, Preservation potential of transgressive Cambrian Ore Hill Member of the Gatesburg Formation, south-central
coastal lithosomes on the U.S. Atlantic shelf: Marine Geology, v. 42, p. Pennsylvania, in Mann, K.O. and Lane, R.L. (eds.), Graphic Correla-
429-442. tion, Society for Sedimentary Geology Special Publication 53, p. 131-
Belknap, D.F. and Kraft, J.C., 1985, Influence of antecedent geology on strati- 137.
graphic preservation potential and evolution of Delaware’s barrier sys- Lochman-Balk, C., 1956, The Cambrian of the Rocky Mountains and south-
tem: Marine Geology, v. 63, p. 235-262. west deserts of the United States and adjoining Senora Province, Mex-
Berg R.R., and Ross, R.J., Jr., 1959, Trilobites from the Peerless and Manitou ico, in J. Rodgers, ed., El Sistemo Camrico, su paleogeografia y el
formations, Colorado: Journal of Paleontology, v. 33, p. 106-119. problema de su base: Internat. Geological Congress, 20th Mexico, v. 2,
Brainerd, A.E., Baldwin, H.L., Jr., and Keyte, I.A., 1933, Pre-Pennsylvanian p. 529-661.
stratigraphy of Front Range in Colorado: American Association of Lochman-Balk, C., 1971, The Cambrian of the craton of the United States, in C.
Petroleum Geologists Bulletin, v. 17, p. 375-396. H. Holland, ed., Cambrian of the New World, Lower Paleozoic Rocks of
Brasier, M.D., 1980, The Lower Cambrian transgression and glauconite-phos- the World: New York, Wiley Interscience, 1, p. 79-167.
phate facies in western Europe: Journal Geological Society of London, Longacre, S.A., 1970, Trilobites of the Upper Cambrian Ptychaspid Biomere,
v. 137, p. 695-703. Wilberns Formation, central Texas: Paleontological Society Memoir 4,
Campbell, J.A., 1972, Lower Paleozoic systems, White River Plateau: in De Journal of Paleontology, v. 44, Supplement, 70 p.
Voto, R.H., ed., Paleozoic stratigraphy and structural evolution of Col- Ludvigson, R., Pratt, B.R. and Westrop, S.R., 1986, The myth of a eustatic sea
orado: Quarterly of the Colorado School of Mines, v. 67, no. 4, p. 37-62. level drop near the base of the Ibexian Series: New York State Museum
Campbell, J.A., 1976, Upper Cambrian stromatolitic biostrome, Clinetop Mem- Bulletin, v. 462, p. 65-70.
ber of the Dotsero Formation, western Colorado: Geological Society of Middleton, G.V. and Southard, J.B., 1984, Mechanics of sediment movement
America Bulletin, v. 87, p. 1331-1335. (2nd ed): Society of Economic Paleontologists and Mineralogists, Short
Davis, R.A., Jr, Klay, J., and Jewell, P., IV, 1993, Sedimentology and stratigra- Course No. 3, Providence, Rhode Island, 401 p.
phy of tidal sand ridges southwest Florida inner shelf: Journal of Sedi- Miller, J.F., 1984, Cambrian and earliest Ordovician conodont evolution, biofa-
mentary Petrology, v. 63, p. 91-104. cies and provincialism, in D. L. Clark, ed., Conodont Biofacies and
Erdtmann, B.-D., Early Ordovician eustatic cycles and their bearing on punctu- Provincialism: Geological Society of America Special Paper 196, p. 43-
ations in early nematophorid (planctic) graptolite evolution: Lecture 68.
Notes in Earth Sciences, v. 8, p. 139-152. Miller, J.F., 1992, The Lange Ranch Eustatic Event: A regressive-transgressive
Ethington, R.L. and Clark, D.L., 1971, Lower Ordovician conodonts in North couplet near the base of the Ordovician System, in Webby, D. B. and
America, in W. C. Sweet and S. M. Bergström, ed., Symposium on Con- Laurie, J. R., eds., Global Perspectives on Ordovician Geology: Rotter-
odont Biostratigraphy: Geological Society of America Memoir, 127, p. dam, Netherlands, A.A. Balkema Publishers, Proceedings of the Sixth
63-82. International Symposium on the Ordovician System, p. 395-407.
Ethington, R.L. and Clark, D.L., 1981, Lower and Middle Ordovician con- Myrow, P., 1998, Transgressive Stratigraphy and Depositional Framework of
176 P. M. Myrow et al.
Cambrian Tidal Sandwave Deposits, Peerless Formation, Central Col- Colorado, in, Kent, H.C. and Porter, K.W., eds., Colorado Geology: Rocky
orado, in Alexander, C., Davis, R., and Henry, J., eds., Clastic Tidal Mountain Association of Geologists — 1980 Symposium, p. 47-56.
Deposition, SEPM Special Publication. Sepkoski, J.J., 1982, Flat-pebble conglomerates, storm deposits, and the Cam-
Myrow, P.M., 1995, Thalassinoides and the enigma of early Paleozoic open- brian bottom fauna, in Einsele, G. and Seilacher, A., ed., Cyclic and
framework burrow systems: Palaios, v. 10, P. 58-74. Event Stratigraphy: Springer-Verlag, p. 375-385.
Myrow, P.M., Ethington, R.L., and Miller, J.F., 1995, Cambro-Ordovician prox- Shapiro, R.S., 1998, Upper Cambrian–lowermost Ordovician stratigraphy and
imal shelf deposits of Colorado: Short Papers for the Seventh Interna- microbialites of the Great Basin, U.S.A.: Unpublished Ph.D. thesis,
tional Symposium on the Ordovician System, Ordovician Odyssey, p. University of California, Santa Barbara, 435 p.
375-379. Stevens, D.N., 1961, Cambrian and Lower Ordovician stratigraphy of central
Nicoll, R.S., Laurie, J.R., and Shergold, J.H., 1992, Preliminary correlation of Colorado, in Rocky Mountain Association Geologists Guidebook, 12th
latest Cambrian to Early Ordovician sea level events in Australia and Annual Field Conference, p. 7-15.
Scandinavia, in B. D. Webby and J. R. Laurie, (eds.), Global Perspec- Stitt, J.H., 1975, Adaptive radiation, trilobite paleoecology, and extinction, Pty-
tives on Ordovician Geology: Rotterdam, Netherlands, Balkema Pub- chaspid Biomere, Late Cambrian of Oklahoma: Fossils and Strata 4, p.
lishers, p. 381-394. 381-390.
Odin, G.S. and Fullagar, P.D., 1988, Geological significance of the glaucony Taylor, J.F., in review, Biomeres and stages: distinctly different but equally valid
facies, in Odin, G.S., ed., Green Marine Clays: Developments in Sedi- units in the Upper Cambrian of Laurentia: manuscript in review for pub-
mentology, Elsevier, Amsterdam, pp. 295-332. lication in Geology, 12 ms. pages, 4 figures.
Palmer, A.R., 1965, Biomere–a new kind of biostratigraphic unit: Journal of Taylor, J.F., 1997, Upper Cambrian biomeres and stages, two distinctly different
Paleontology, v. 39, p. 149-153. and equally vital stratigraphic units: 2nd International Trilobite Confer-
Palmer, A.R., 1981, Subdivision of the Sauk Sequence, in M.E. Taylor, ed., ence, St. Catherines, Ontario, Abstracts Volume, p. 47.
Short Papers for the Second International Symposium on the Cambrian Taylor, J.F., Loch, J.D., and Perfetta, P.R., 1999, Trilobite faunas from Upper
System: United States Geological Survey Open-File Report 81-743, p. Cambrian reefs in the central Appalachians: Journal of Paleontology, v.
160-162. 32, p. 326-336.
Palmer, A.R., 1984, The biomere problem: evolution of an idea: Journal of Pale- Taylor, J.F., Repetski, J.E. and Orndorff, R.C., 1992, The Stonehenge Trans-
ontology, v. 58, p. 599-611. gression: A rapid submergence of the central Appalachian platform in
Pemberton, S. G. and Frey , R. W., 1985, The Glossifungites ichnofacies: mod- the Early Ordovician, in B. D. Webby and J. R. Laurie, eds., Global Per-
ern examples from the Georgia coast, U. S. A., in Curran, H.A., ed., spectives on Ordovician Geology: Rotterdam, Netherlands, Balkema
Biogenic Structures: Their Use in Interpreting Depositional Environ- Publishers, Proceedings of the Sixth International Symposium on the
ments: Society of Economic Paleontologists and Mineralogists, Special Ordovician System, p. 409-418.
Publication No. 35, p. 237-259. Tweto, O., 1949, Stratigraphy of the Pando area, Eagle County, Colorado: Col-
Ross, J.R., Jr., and Tweto, O., 1980, Lower Paleozoic sediments and tectonics in orado Science Society Proceedings, v. 15, no. 4, p. 147-235.
Printed in U.S.A.
Geological Society of America
Field Guide 1
1999
David H. Malone
Department of Geography-Geology, Illinois State University, Normal, Illinois 61790-4400, United States
Thomas A. Hauge
Exxon Production Research Company, P.O. Box 2189, Houston, Texas 77252-2189, United States
Edward C. Beutner
Department of Geosciences, Franklin and Marshall College, Lancaster, Pennsylvania 17604, United States
ABSTRACT
For more than a century, the Heart Mountain Detachment has been an important
natural laboratory that has contributed to the education of thousands of students rep-
resenting most of the colleges and universities of the nation. The purpose of this field
trip is to provide a forum in which the important features of the Heart Mountain
Detachment (HMD) can be observed and the various explanations for the origin of
the structure can be discussed. The foremost questions to be addressed will likely
include: What factors (e.g., paleotopographic slope; direction and magnitude of bed-
ding dip; seismicity; eruptive processes, presence and pressure of fluids) triggered the
formation of the HMD? What factors facilitated movement across the gently dipping
detachment surface? Did the allochthon consist of numerous detached blocks of Pale-
ozoic rocks or was the allochthon continuous and consist of predominantly Eocene
volcanic rocks? Were the detached rocks emplaced gradually or catastrophically?
What data can be used to constrain the rates of emplacement?
This field guide begins with an overview of the regional setting of the HMD.
Detailed site descriptions of the various detachments and their associated structures
are integrated into the road log (Figure 1). Most outcrops that are viewed from a dis-
tance on this trip are accessible for detailed examination via hikes of 1 to 2 hours. Time
constraints preclude visiting more than a few of these exposures for detailed examina-
tion during this field trip, but others meriting such examination are identified for
future reference. The staging area for this trip is the Double Diamond X Guest Ranch
in the upper South Fork Shoshone River valley. The ranch is owned by Russ and Patsy
Frazier. We warmly thank Russ and Patsy for their friendship and hospitality.
THE HEART MOUNTAIN DETACHMENT preserved over an area of at least 3400 km2. The upper plate of the
detachment was emplaced during the middle Eocene, during the
The HMD is a rootless, low-angle normal fault that accom- late stages of the Laramide orogeny. Heart Mountain faulting
modated transport of upper-plate rocks for distances of as much as involved rocks ranging in age from Ordovician to middle Eocene
50 km or more (Figure 1). Transport was largely southeastward, (Figure 2), mostly Paleozoic cratonic strata and andesitic Eocene
from the northeast flank of the northern Absaroka Mount-ains, a volcanic rocks. The reader is referred to Pierce (1973) and Hauge
Laramide volcanic center and basement uplift, toward and into the (1993) for descriptions of the general features of the HMD. Other
western margin of the Laramide Bighorn Basin. The detachment is field guides to the area include Tucker (1982) and Hauge (1992).
Malone, D. H., Hauge, T. A., and Beutner, E. C., 1999, Field guide for the Heart Mountain detachment and associated structures, northeast Absaroka
Range, Wyoming, in Lageson, D. R., Lester, A. P., and Trudgill, B. D., eds., Colorado and Adjacent Areas: Boulder, Colorado, Geological Society of America
Field Guide 1.
177
178 D. H. Malone, T. A. Hauge, and E. C. Beutner
Based on its relationship to Eocene sedimentary and vol- this detachment faulting, a tectonically denuded surface was
canic rocks, the Heart Mountain allochthon was emplaced during formed. Immediately after faulting had ceased, massive out-
the late stages of the Laramide orogeny. Allochthonous Paleozoic pourings of Wapiti Formation volcanic rocks were deposited
rocks at Heart Mountain and McCulloch Peak overlie nonmarine on the detached blocks as well as on the tectonically denuded
lower Eocene (Wasatchian) Willwood strata, indicating that surface (Pierce 1973, 1987). The most compelling line of evi-
emplacement of the allochthon postdated most of the Laramide dence for this interpretation is the complete lack of erosion on
fill of the Bighorn Basin. Thus, most of the offset across the base- the exposed fault plane, indicating that the time interval
ment-involved fault zone that defines the boundary between the between slide block emplacement and the deposition of the
Bighorn Basin and adjacent uplifts to the west had taken place by Wapiti Formation must have been very short.
the time of Heart Mountain faulting. Wise (1983) has argued that During the 1980s, a different model for the emplacement of
the north-trending portion of this fault zone postdated northwest- the HMD allochthons was advanced. In this view, the upper plate is
trending structures along the Bear-tooth range front, and he interpreted to have been a continuous allochthon rather than a
inferred that the northwest-trending boundary between the series of individual slide blocks (Figure 3b; Hauge 1985, 1990,
Absaroka Mountains and the Beartooth Mountains was also 1993). Volcanic rocks overlying the detachment, originally viewed
probably an earlier Laramide structure. Thus, development of the as in depositional contact, were reinterpreted as allochthonous, and
presently observed basement framework of Laramide structures as comprising much of the upper plate (Hauge 1990). Thus, the
in the detachment area was essentially complete when Heart continuous allochthon model requires no tectonic denudation or
Mountain faulting occurred, and little subsequent tectonic defor- catastrophic emplacement of numerous slide blocks, and the model
mation has affected the region. Further constraints on the age of eliminates the mechanical enigma that tectonic denudation poses.
Heart Mountain faulting are provided by relationships west of Other continuous allochthon models that do not require tectonic
Buffalo Bill Reservoir, where allochthonous Paleozoic rocks denudation, but do require catastrophic emplacement rates have
overlie middle Eocene strata (Bridgerian Aycross Formation), been advanced by Sales (1983) and Beutner and Craven (1996).
indicating emplacement of the allochthonous rocks in this area as
middle Eocene or younger (Torres and Gingerich, 1983). ABSAROKA VOLCANISM
A middle Eocene upper age limit for Heart Mountain
faulting is also indicated by the Bridgerian age of volcanic The Eocene volcanic succession within the Absaroka
rocks overlying the allochthonous Paleozoic rocks at this local- Range has been formally named the Absaroka Volcanic Super-
ity (Torres and Gingerich, 1983). These Bridgerian rocks are group (AVS) (Smedes and Prostka, 1972). The rocks of the AVS
assigned to the Wapiti Formation and Trout Peak Trachyan- extend over an area of approximately 18,000 km2 most of
desite (Pierce and Nelson, 1968). The Wapiti Formation and which is underlain by a Laramide structural basin (Absaroka
Trout Peak Trachyandesite have been interpreted as postdating Basin); the volcanic rocks overlie rock units which range in age
Heart Mountain faulting (Pierce, 1987a), which would indicate from Archean to Eocene. Deeply incised valleys provide excel-
that faulting was wholly Bridgerian (pre-Wapiti) in age. Alter- lent natural cross sections, and they display a volcanic strati-
natively, rocks assigned to the Wapiti Formation have been graphic succession in excess of 1875 m in thickness. The rocks
interpreted as allochthonous (Hauge, 1985), and the Trout Peak of the AVS are unconformably overlain to the west by Quater-
Trachyandesite, which locally comprises the hanging wall of nary volcanic rocks of the Yellowstone Volcanic Plateau. The
the breakaway fault, has been interpreted as involved in the inferred depositional setting of the AVS is a series of high stra-
final phases of Heart Mountain faulting (Hauge, 1990). This tovolcanoes with coalescing alluvial aprons (Sundell, 1993),
alternative interpretation suggests that Heart Mountain faulting rather like the Cascade Range or Andes Mountains of today.
was complete slightly later in the Bridgerian (after Trout Peak Volcanic centers define two northwest-trending belts, the loca-
Trachyandesite time). By either interpretation, Heart Mountain tions of which are probably controlled by weakness zones
faulting may have been wholly Bridgerian (47.5 to 49.5 Ma; within the Precambrian basement (Chadwick, 1970). Volcanism
Torres and Gingerich, 1983), though earlier minor movements was initiated in southern Montana during late early Eocene time
(Pierce, 1973) are not precluded by the data. Days One and (53 Ma) (Chadwick, 1970), and continued throughout middle
Two (morning) of this field trip provide an introduction to the Eocene time, with the locus migrating southeastward until cul-
general features of the HMD, with emphasis on the relation- minating in the southeastern Absaroka Range during late
ships between Eocene volcanic rocks and Paleozoic sedimen- Eocene time (38 Ma; Sundell, 1985).
tary rocks above and along the HMD. Reworked epiclastic volcanic rocks, including volcanic
Two fundamentally different models describing the breccias, sandstones, conglomerates, siltstones, and claystones
geometry, kinematic pattern, and emplacement of the upper are the predominant rock types. Primary volcanic rocks (lava
plate of the HMD have been proposed. For many years, the flows, flow breccias, pyroclastic breccias, and tuffs) increase in
upper plate was viewed to have been emplaced catastrophi- abundance near the intrusive centers. Three groups comprise the
cally as numerous independent slide blocks (Figure 3a; AVS: (1) the lower(?) and middle Eocene Washburn Group, (2)
Bucher 1947; Pierce 1957, 1973, 1987a), and as a result of the middle Eocene Sunlight Group, and (3) the middle and
Figure 1. Geologic map of the HMD area (modified from Malone, 1995). The field trip stops discussed in the text are indicated.
180 D. H. Malone, T. A. Hauge, and E. C. Beutner
desite in the north, and the Aycross Formation in the south. The
Thorofare Creek Group consists of the Langford, Two Ocean,
Wiggins, and Teepee Trail formations. The Langford and Two
Ocean formations crop out in the northern Absaroka Range, and
the Wiggins and Teepee formations occur through the southerly
reaches of the range.
The Deer Creek Member of the Wapiti Formation consists
of blocks (individually as large as several km2 in area) of vent-
medial-facies lava flows, breccias, and sandstones within a thin,
heterogeneous matrix of boulder- to sand-sized volcaniclastic
material. It is interpreted as the deposit of a large debris aval-
anche, formed by the collapse of a large stratovolcano within the
Absaroka Range during the early middle Eocene. The areal
extent and volume of the proximal facies of the Deer Creek
Member are ~450 km2 and ~100 km3, respectively. The unit was
first described by Malone (1995) and it was later described in
detail by Malone (1996). The Deer Creek member was assigned
to the Wapiti Formation because it is within the type section of
that unit as defined by Nelson and Pierce (1968). Further work
by Malone (1997) indicates that a distal facies of the unit also
probably occurs throughout the upper South Fork Shoshone
River valley. In its proximal area, the unit extends from the
southern end of Sheep Mountain 32 km northward to Dead
Indian Creek. The easternmost extent tops Rattlesnake and Pat
O’Hara Mountains; to the west, the unit is buried by younger
volcanic rocks. A transport distance of 40 km from the center of
the inferred source area is indicated. The average thickness of
the debris-avalanche deposit is estimated to be ~220 m. Its thick-
ness ranges from zero at its stratigraphic pinch-out to >470 m at
the west of Sheep Mountain. The Deer Creek Member is
bounded below by an erosional surface that displays both large-
and small-scale paleotopography. Overall, the basal surface
slopes an average of 3°-4° to the southeast, with >1000 m of
total relief on this surface demonstrated. The top of the unit is
an erosional surface as well. Overall, the upper surface is rela-
tively flat and slopes gently 1°-2° to the southeast. The age of
the Deer Creek Member is closely determined. It falls into the
same 2 million year window as Heart Mountain faulting.
breakaway fault, which forms the western boundary of the bedding-parallel component of the detachment near the base of
detachment area and is located near the eastern boundary of the Ordovician up to the middle Eocene. Whereas the footwall of
Yellowstone National Park. The breakaway fault extends the bedding-parallel component of the detachment is a struc-
upward from the detachment and ends upward within Eocene turally simple homocline, the footwall of the detachment ramp is
volcanic rock, demonstrating that the HMD is a rootless struc- the structurally complex transition from the Absaroka Mountains
ture. The breakaway fault is the subject of Stop 1-2 of this field to the Bighorn Basin, and the configuration of the detachment
trip, and it is described in detail in the text for that stop. The ramp reflects this footwall structure. Pierce (1957, 1960, 1985)
second component of the HMD, which is visible from our pre- subdivided the detachment ramp into (1) a “shear thrust” (1957)
sent vantage point and is the subject of Stops 1-3 through 1-7 or “transgressive fault” (1960), where it climbs abruptly in pre-
and Stop 2-1 of this field trip, is the portion of the basal detach- sent-day elevation to the top of Dead Indian Hill, and (2) an “ero-
ment that parallels bedding in the homoclinal Paleozoic section sion thrust” (1957; after Hewett, 1920) or “fault on former land
of the northeastern Absaroka Mountains. surface” (1960) from Dead Indian Hill eastward. The 3-to-5-km
The bedding-parallel component of the detachment lies wide “transgressive fault” presently dips about 10 degrees west-
along a footwall bedding plane located about 2 m above the ward; the “fault on former land surface” as much as 48 km wide,
base of the Ordovician section. This stratigraphic position is presently dips an average of 1 degree (and locally up to 4 degrees)
remarkable from the perspective of rock mechanics, as dis- eastward.
cussed by Pierce (1973) and Hauge (1993), because the detach-
ment lies along a bedding plane within “strong” dolomite rather Return to vehicles. Turn around, retrace route west along US 212.
than within the thick, “weak” Cambrian shales that underlie it 7.0 78.0 Junction with WYO 296. Continue west along
by only a few meters. The detachment does not cut below this US 212.
stratigraphic horizon (see Hauge, 1983, for minor exceptions to 14.2 92.2 = 0.0 Cooke City, Montana. Continue west along
this), and upper-plate units are no older than Ordovician, indi- US 212.
cating that the detachment is rootless. The bedding-parallel 3.0 3.0 Silver Gate, Montana.
component of the detachment is bounded to the west by the 1.1 4.1 Northeast Entrance to Yellowstone National
breakaway, to the northeast by erosion, to the southeast by a Park.
footwall ramp, and to the southwest, speculatively, by the Black 1.2 5.3 Turn left into Soda Butte Creek picnic area.
Mountain fault. The bedding-parallel component of the detach- 0.3 5.6 Park vehicles at last picnic area. Walk up hill
ment presently dips an average of 3 to 5 degrees to the south- to east for a view of the breakaway fault.
southwest (Pierce, 1985). The autochthon in the area of the
bedding-parallel component of the detachment exhibits little STOP 1-2: VIEW OF THE BREAKAWAY FAULT
deformation either related to or subsequent to faulting. In con-
trast, the allochthon was strongly extended during Heart Moun- From our vantage point the breakaway fault, which forms the
tain faulting. Allochthonous Paleozoic rocks include 400 m western boundary of the detachment area, is exposed across 650 m
thick untilted sections with little section missing along the of relief (Figure 4). At this locality it dips 70 degrees to the east and
detachment; sections tilted up to 30 degrees or more with strata cuts steeply down through Eocene volcanic rocks and Mississip-
truncating downward at the detachment; and local exposures pian, Devonian, and Ordovician sedimentary rocks in its footwall,
where Ordovician and Devonian strata are omitted along the ending near the base of the Ordovician at the detachment. The hang-
detachment, without significant angular discordance between ing wall consists of Eocene volcanic rocks. In other areas, due to
allochthon and autochthon. Volcanic rocks lie between and limited vertical exposure and difficult access, the trace and cross-
upon the masses of allochthonous Paleozoic strata, but dis- sectional geometry of the breakaway are less well known. Pierce
agreements exist as to the involvement of these rocks in fault- (1960, 1980, 1987b) interpreted the breakaway as having been tec-
ing. Some workers have argued that in most areas the contact tonically denuded and interpreted the volcanic rocks overlying it as
between volcanic rocks and the detachment is a “half fault” Wapiti Formation that was deposited on the denuded surface of the
(Pierce, 1980), the volcanic rocks having been deposited upon breakaway. He described the breakaway fault as a “half-fault,” for
the detachment (Pierce, 1987a), or emplaced upon it as a debris- only one side [the footwall side] is a fault surface; the other side [the
avalanche (Malone, 1994), after it had been tectonically hanging-wall side] is a surface of deposition (Pierce, 1980, p. 276).
denuded. Others inferred the contact is in many areas (Prostka, In contrast, the hanging-wall volcanic rocks have been interpreted
1978) or everywhere tectonic (Hauge, 1982,1985, 1990; Tem- as in part (Prostka and others, 1975) or wholly (Hauge, 1982, 1985,
pleton and others, 1995; Beutner and Craven, 1996). This, 1990) allochthonous.
along with the issue of emplacement rate of of rocks overlying
the detachment, is the essence of the current debate about the Turn vehicles around and retrace route to Silver Gate.
nature of Heart Mountain faulting. 0.3 5.9 Turn right from picnic area onto US 212.
The third component of the detachment, is a footwall ramp 1.2 7.1 Exit Yellowstone National Park.
that, in general terms, cuts up section to the southeast, from the 1.1 8.2 Pull off road to left (north) near gas station.
Heart Mountain detachment and associated structures, northeast Absaroka Range, Wyoming 183
STOP 1-3: VIEW OF THE DETACHMENT FROM west-southwest of the exposure of allochthonous Paleozoic
SILVER GATE, MONTANA (OPTIONAL) rocks, recognized by Hauge (1990); volcanic flow units dipping
30-45 degrees southeast, recognized by Pierce and others (1973)
For a distance of 5 km east of the breakaway exposure just and Hauge (1990); 4) volcanic rocks underlain along the detach-
described, the detachment is commonly well exposed and is typi- ment by striated microbreccia a few hundred meters farther west,
cally overlain by volcanic rocks with a preserved vertical thickness recognized by Hauge (1985); 5) clastic dikes of fault breccia,
on the order of 500 m. Locally, masses of allochthonous Missis- interpreted as having been emplaced after Heart Mountain fault-
sippian and Devonian rocks up to a few tens of meters thick over- ing (Pierce, 1987a) or during faulting (Hauge, 1990). This local-
lie the detachment and underlie the volcanic rocks. The exposure ity is recommended for detailed examination, with the conflicting
visible from this stop, 2 km east of the breakaway and south of the interpretations of the relationships in mind. Access, described by
town of Silver Gate, Montana, was singled out in several publica- Pierce (1987b), is via a climb beginning on the south side of Sil-
tions (Pierce, 1979, 1980, 1987a, 1987b; Pierce and Nelson, 1986) ver Gate.
as a showcase example of relationships demonstrating that vol-
canic rocks lie in situ on the detachment (as well as on allochtho- Proceed east along US 212.
nous carbonate rocks) and, hence, that tectonic denudation of the 3.0 11.2 Cooke City
detachment is proven. However, according to Hauge (1990), the 0.7 11.9 Park at junction of US 212 and forest service
relationships cited by these authors are also compatible with tec- road to Daisy Pass. Walk across US 212 to
tonic emplacement of the volcanic rocks, and other relationships in south side of road.
this area require that at least the basal 200 m of volcanic rocks in
this area are allochthonous. The critical features described at this STOP 1-4: VIEW OF REPUBLIC MOUNTAIN
locality are: 1) the contact between the volcanic rocks and the
allochthonous Paleozoic rocks, which is interpreted as depositional This stop affords a view of allochthonous Paleozoic rocks at
(Pierce, 1987a) and as a faulted unconformity (Hauge, 1990); 2) Republic Mountain and the volcanic rocks that overlie them
faults within the upper-plate Paleozoic rocks, which either do not (Figure 5). The Paleozoic rocks are Ordovician, Devonian, and
(Pierce, 1987a, 1987b) or do (Hauge, 1990) offset the contact Mississippian strata that are internally faulted and tilted to small
between the Paleozoic upper plate and the overlying volcanic angular discordance with the detachment (Elliott, 1979). These
rocks; 3) faults within volcanic rocks in the area immediately rocks were first recognized as allochthonous by Pierce (1960).
The volcanic rocks are dominantly massive andesitic volcani- rather than numerous detached blocks. Upper plate rocks that
clastic rocks, in which primary stratification is difficult to dis- underlie Pilot Peak include Paleozoic rocks visible in the
cern. These volcanic rocks have been variably mapped as Wapiti drainages of Fox and Pilot Creeks, volcanic rocks variably
Formation that postdates Heart Mountain faulting (Pierce, 1978) mapped as Wapiti Formation (in situ) or Lamar River and Cathe-
and as Lamar River and Cathedral Cliffs formations (Nelson and dral Cliffs Formations (allochthonous), and Trout Peak Tra-
others, 1980; Elliott, 1979). Hauge (1983) described a fault con- chyandesite (Pierce and others, 1973). A fault contact between
tact between the Eocene volcanic and Paleozoic sedimentary upper-plate Paleozoic and Eocene rocks well exposed in the
rocks of the upper plate. From our vantage point 30 degree drainage of Pilot Creek is described in Hauge (1985).
southward dips of the volcanic rocks can be discerned. We will
discuss whether these dips are consistent with the volcanic rocks Proceed eastward on US 212.
being in situ Wapiti Formation (Wapiti vents are south of this 3.4 22.2 Turn right into Crazy Creek Campground. Park
area), or are better explained as tectonic dips. in the loop at the end of the campground road
and walk a few meters south.
Proceed eastward on US 212.
6.9 18.8 Park at entrance to Fox Creek campground. STOP 1-6: VIEW OF ROCKS OVERLYING THE HMD,
FROM PILOT AND INDEX PEAKS TO THE NORTH-
STOP 1-5: VIEW TO WEST OF PILOT PEAK, INDEX WEST, TO THE ONEMILE CREEK AREA TO THE
PEAK, AND THE UNDERLYING DETACHMENT SOUTHEAST
(OPTIONAL)
Several features and localities of interest are visible from
This vantage point provides a distant view of a superb expo- this vantage point. The first (N60-75W) is many of the features
sure of the detachment, Cambrian rocks of its footwall (intruded seen at Stop 1-5, seen here in a more distant view. The second is
at this locality by an Eocene latite porphyry sill—Pierce and oth- a view up Pilot Creek (N75-80W). On the north side of Pilot
ers, 1973), and volcanic rocks immediately overlying the detach- Creek, a major unconformity, best viewed in early morning sun-
ment. The detachment follows a bedding horizon 2 m above the light, is visible. This unconformity separates undeformed, sub-
Cambrian-Ordovician contact (Pierce, 1968), and from this view horizontal volcanic strata from underlying volcanic rocks that,
the parallelism of the detachment and footwall strata is evident. from this view, show no apparent stratification. The volcanic
The 13-16 degrees southwest dip of the detachment in this area rocks beneath this unconformity, which are mapped as Lamar
(Pierce and others, 1973) is atypical of most of the bedding-par- River and Cathedral Cliffs Formations by Pierce and others
allel component of the detachment, dips of less than 1 degree, to (1973) and as mostly Wapiti Formation by Pierce (1978), are
the south or southwest, being more typical. The volcanic rocks characterized by dips as steep as 36 degrees (Pierce and others,
overlying the exposure of the detachment visible from this stop 1973; Hauge, 1983), and are variably interpreted as in situ
are interpreted as Wapiti Formation on most published maps (Pierce and others, 1973: Pierce and Nelson’s interpretation) or
(Pierce and others, 1973; Pierce, 1978; Nelson and others, allochtonous (Pierce and others, 1973; Prostka’s interpretation;
1980). The Wapiti Formation is defined as postdating Heart Hauge, 1983). The volcanic rocks above this unconformity
Mountain faulting (Nelson and Pierce, 1968). Pierce (1968) postdate faulting. From the perspective of the continuous
interpreted the detachment at this locality as having been tecton- allochthon model of Heart Mountain faulting, this view sug-
ically denuded and interpreted the volcanic rocks to be in depo- gests the thickness of the continuous allochthon that was pre-
sitional contact with the detachment. Based on observations of served when faulting ceased and the allochthon was overlain by
the volcanic rocks along the detachment immediately to the younger volcanic rocks. This preserved thickness, which is less
north and south of this exposure (the exposure seen from this than the thickness of the active allochthon by some unknown
stop is too steep to be safely accessible), Hauge (1983, 1985) amount, is roughly 600 m.
interpreted these volcanic rocks as allochthonous. Also visible from this stop are Jim Smith Peak (S70W) and
From this perspective the characteristic sharp, planar nature the spot where Jim Smith Creek crosses the detachment
of the detachment is particularly impressive. This and other expo- (S75W). Like the area south of Silver Gate viewed at Stop 1-3,
sures of the detachment show no direct evidence of subaerial the excellent detachment exposure at Jim Smith Creek has been
exposure; nowhere is it incised by erosion that postdated faulting a focus of disagreement in the literature. Pierce and others
but predated the overlying volcanic rocks, and nowhere is it over- (1991) interpreted the volcanic rocks overlying the detachment
lain by fluvial deposits. These relationships led Pierce (1957, on the east wall of Jim Smith Creek as in depositional contact
1973) to infer that the period of time after tectonic denudation of with the detachment, and Pierce and others (1991) interpreted
the detachment and deposition of the Wapiti volcanic rocks was the volcanic rocks on the west wall of the creek as allochtho-
very brief. These relationships were cited by Hauge (1985) as nous (see also Pierce and others, 1973). Hauge (1985) inter-
supportive of the argument that tectonic denudation did not occur, preted the volcanic rocks immediately above the detachment on
and the Heart Mountain upper plate was a continuous allochthon both sides of the creek as allochthonous. Hauge’s (1985) inter-
Heart Mountain detachment and associated structures, northeast Absaroka Range, Wyoming 185
pretation was based on mesoscopic features (Figures 6 and 7) created during slow (~typical geologic strain rates) collapse and
supplemented by thin-section examination. Pierce and others extension of the continuous allochthon. Beutner agrees with the
(1991) provide descriptions of thin sections of the detachment kinematics of Hauge but maintains that movement was cata-
horizon that they interpret as incompatible with tectonic strophic and aided by a gas-fluidized cushion (now represented
emplacement of the volcanic rocks. I (Hauge) view the features by microbreccia) at the base of the allochthon. To observe most
described by Pierce and others as compatible with tectonic exposures of volcanics on the fault requires a climb of 1000' to
emplacement of the volcanic rocks. This locality is accessible 1500' through brush. The outcrops we will visit are not large but
via a Forest Service road: turn south off of US 212 just west of demonstrate the relations well and are easily reached by trail.
Crazy Creek, cross the bridge over the Clarks Fork, drive past (NOTE: THIS IS GRIZZLY COUNTRY AND WE HAVE SEEN
the B Four Ranch 1.5 mi (this is a Forest Service road; public THEM ALONG THIS TRAIL. WE STRONGLY RECOMMEND
access is permitted), and hike south about 1.5 km. Time and CARRYING BEAR PEPPER SPRAY AND AGAINST GOING
weather permitting, we will visit this locality today. Throughout INTO THIS AREA ALONE.)
the area of exposure of the bedding-plane component of the Go 6.4 miles east on the Chief Joseph Highway from its
detachment, numerous normal faults, most probably with small intersection with U.S. 212 (0.1 mile E. of the bridge over the
offset, are present in volcanic rocks overlying the detachment. Clark’s Fork) and go up the Squaw Cr. trail across a clearcut
Hauge (1983, 1985) measured the orientations of these faults and into the woods. Massive outcrops of Crandall Conglomer-
and their slickenlines, and Buetner and Craven (1996) interpret ate are uphill to the south and can be examined with a short,
a vertical contraction axis and an approximately horizontal brisk climb or in float. In the first small drainage crossed, Cran-
extension axis oriented N59W-S59E. This is compatible with dall Conglomerate composed of angular fragments (channel
the interpretation that they were emplaced as part of a continu- margin talus?) can be seen in the bed of the ravine. When the
ous expanding allochthon (Beutner and Craven, 1996). Squaw Cr. trail turns south at a sign just after crossing a small
stream, leave it and continue west on the trail along the bench
Proceed eastward on US 212. on top of the Pilgrim Limestone. Cross another small stram and
2.8 25.0 = 0.0 Turn right onto WY 296. then Squaw Cr., where the trail is washed out. Immediately
6.4 6.4 Turn west on Park Co. Road XUX. Go 3.5 mi. downstream from the washout is an outcrop of Pilgrim with an
and park near the bridge over Squaw Cr., which attitude of N60°W; 31°NE. This outcrop is probably on the NE
is just past the Squaw Cr. trailhead. flank of the Blacktail anticline, which must plunge out between
here and the Pilgrim cliffs 1 km to the NW. Immediately after
STOP 1-7: SQUAW CREEK AREA crossing Squaw Cr., turn left (W) where the trail forks and go
upstream <1/8 mile, crossing an open slope, to the first drainage
This stop will examine the critical relationship, alluded to from the right (NW). Go up this small stream. The fault is well
earlier, between the bedding plane portion of the Heart Mountain exposed in the first ravine from the right (NE), where it is
detachment fault and overlying volcanogenic rocks. Malone marked by reddish microbreccia ~20 cm thick underlying shat-
argues that the volcanics were, at least in large part, emplaced by tered volcaniclastics and overlying comminuted Cambrian
megaslandsliding onto the denuded fault surface immediately fol- shale. If you want samples, please do not sample the micro-
lowing dispersal of the carbonate blocks. Hauge argues that the breccia in place—there are sufficient float blocks available. The
volcanics originally overlay the Paleozoic carbonates and moved fault is poorly exposed (grey microbreccia up to 0.5 m thick) in
down upon normal faults to fill gaps between carbonate blocks the next small drainage; it appears to be offset by a small fault
186 D. H. Malone, T. A. Hauge, and E. C. Beutner
between these locations. It is well exposed further up in the Hurricane Mesa 11 km due west. Windy Mountain is the highest
main stream, with pale grey-green microbreccia 3-5 cm thick peak visible to the south. Both Hurricane Mesa and Windy
containing superb tool marks (asymmetric grooves) (Figure 8) Mountain are underlain by Absaroka volcanic and intrusive
on its base indicating movement toward S50°-55°E. The foot- rocks. The carbonate exposure immediately to the south is Cathe-
wall in this area is sheared shale and fractured limestone of the dral Cliffs. At this stop we will discuss the relationship between
Snowy Range Formation. The stratigraphic level of the fault is Absaroka volcanic rocks and upper-plate Paleozoic rocks in two
lower and the footwall is more intensely deformed than usual areas: at the west end of Cathedral Cliffs (near Corral Creek), and
here, presumably because the fault is cutting through the preex- along the length of the face of Cathedral Cliffs.
isting Blacktail anticline. The microbreccia at the base of the The first topic of this stop is the poorly exposed contact
volcanics is, as usual, a massive, cohesive rock with no internal between the upper-plate Paleozoic sedimentary rocks and Eocene
shear planes or macroscopic orientation to its fabric. It contains volcanic rocks that abut them to the west along Corral Creek. This
volcanic and carbonate clasts in a microcrystalline carbonate- area is best viewed from the road that leads to the K-Z Ranch from
rich matrix. the west; we will stop at this better vantage point if time permits.
The volcanic rocks of this area were mapped as Wapiti Formation,
Return to Highway 212 and proceed eastward. in depositional contact with the detachment and Paleozoic upper
1.4 7.8 Cross Crandall Creek plate, by Pierce and Nelson (1971) and Pierce (1978). There has
1.8 9.6 Pull off highway to the right into Scenic View been disagreement in the literature, however, as to whether vol-
pull-out. canic rocks in this area were deposited on the detachment and
against the allochthonous Paleozoic rocks (Pierce, 1987a) or were
STOP 1-8: CATHEDRAL CLIFFS AREA tectonically emplaced (Hauge, 1985, 1990) (Figure 9).
Cathedral Cliffs (Figure 10) is a 5 km wide exposure of
Several features of interest are visible in this panoramic allochthonous Paleozoic sedimentary rocks, with Eocene volcanic
view. Hunter Peak 5 km to the northwest (N60W), is underlain by rocks along much of the skyline. The Paleozoic rocks appear
allochthonous Paleozoic rocks that traveled perhaps 6 to 15 km remarkably little deformed, despite a probable transport distance
across the detachment. South of Hunter Peak are the drainage of of 5 to 15 km. In most of this exposure, the base of the allochtho-
Crandall Creek (bearing due W to S80W) and the high country of nous Paleozoic rocks consists of shattered Bighorn Dolomite, with
Figure 6. Volcanic rocks along the bedding-plane detachment immediately west of Jim Smith Creek, showing basal
shatter zone up to 10 m thick, steeply dipping truncated strata (right of center), and internal faulting of volcanic rocks,
The detachment and underlying autochthonous strata are also visible. View is westward. From Hauge, 1985.
Heart Mountain detachment and associated structures, northeast Absaroka Range, Wyoming 187
little section omitted along the detachment. The allochthonous at this stop is indicated by the fault contacts between Eocene and
Paleozoic rocks are bounded to the west by volcanic rocks, as was Paleozoic rocks (Pierce and Nelson, 1971; Hauge, 1983) and
described immediately above. They are bounded to the east by vol- within volcanic rocks (Hauge, 1983) at this locality. From this per-
canic rocks that were mapped by Pierce and Nelson (1971) and spective, the volcanic rocks on the skyline seem to overlie the
Pierce (1978) as Wapiti Formation (in depositional contact with the allochthonous Paleozoic rocks. Instead, the volcanic rocks are in
detachment) and were interpreted by Hauge (1983, 1985) as fault contact with the Paleozoic rocks across a steeply south-dip-
allochthonous. An asymmetric graben, about 200 m wide at its ping fault. From this vantage, this south-dipping contact can be
exposed base, is visible near the center of the exposure of seen best in the area of the graben and, farther west, where Paleo-
allochthonous Paleozoic rocks (Pierce and Nelson, 1971). Hauge zoic rocks form the highest cliffs. Numerous Eocene latitic to
(1990) interpreted this small graben as representative of an early basaltic dikes intrude the upper plate rocks (both Paleozoic and
phase of extension of the allochthon. He envisioned downfaulting Eocene) at this and numerous other localities. These dikes, which
of Paleozoic and Eocene rocks within grabens such as these, with die out downward at or within a few m of detachment, are
continued extension leading to the downfaulting of broad expanses restricted to the upper plate. Although many hundreds have been
of Eocene volcanic rocks to the detachment horizon, such as are mapped within the allochthon (e.g., Pierce and Nelson, 1971; Nel-
seen east and west of Cathedral Cliffs (Hauge, 1990). The involve- son and others, 1980), only a few have been identified within the
ment of volcanic rocks in the formation of the small graben visible autochthon (Hauge, 1983, 1985). Pierce (1987a) interpreted these
188 D. H. Malone, T. A. Hauge, and E. C. Beutner
Mileage description
Figure 10. View of Cathedral Cliffs. Dash-C is Heart Mountain footwall rocks, mostly Cambrian; MO is allochthonous Ordovician-
Mississippian cratonic strata; Tv is Tertiary volcanic rocks; d is HMD.
These observations and inferences, in conjunction with their 0.0 White Mountain.
inferences regarding the kinematics of emplacement of the 6.7 6.7 Junction with WYO 296 (Chief Joseph Scenic
volcanic rocks associated with Heart Mountain faulting, led Highway). Proceed east on WYO 296.
them to conclude that Heart Mountain faulting comprised 10.0 16.7 Pull off Wyoming Highway 296 to right (west)
catastrophic emplacement of a continuous allochthon. Pierce side of highway at the overlook at Dead Indian
(1965, 1978) and Pierce and Nelson (1971) interpreted these Pass.
volcanic rocks as younger than Heart Mountain faulting, but
Hauge (1983, 1985) and Pierce (1985) interpreted them as STOP 2-2: DISCUSSION OF DETACHMENT RAMP
allochthonous. Hauge (1985) described a normal fault within
the volcanic rocks that is exposed across 200 m of topo- Dead Indian Pass provides a spectacular view of the
graphic relief. This is one of the best accessible exposures of Beartooth Plateau to the north, the Clarks Fork fault that bounds
a fault within volcanic rocks in the upper plate. its southern flank in this area, and the Absaroka Mountains to
the west and southwest. The prominent valley immediately to
the west in the Absaroka Mountains is Sunlight Basin, visited at
Retrace route through Sunlight Basin to WYO 296; turn right Stop 2-1; the distinctive peak at N80W along the north side of
onto 296 and proceed eastward up the switch backs to Dead Sunlight Basin is White Mountain. The bedding-parallel com-
Indian Pass. ponent of the HMD underlies much of the country to the west,
190 D. H. Malone, T. A. Hauge, and E. C. Beutner
divide between the Clarks Fork and Shoshone River drainage is Key marker beds within blocks have not been identified, but
the ridge to the north. To the east is the Rattlesnake Mountain anti- a crude coarsening upward stratigraphy has been recognized in
cline. In the lower elevations to the north and west is the Deer some of the larger blocks with vent-facies rocks increasing in
Creek Member. The high peaks to the west are Trout and Dead abundance toward the tops. The basal parts of blocks are com-
Indian peaks. Deer Creek Member blocks are commonly elongate posed primarily of well-stratified, medial-facies epiclastic brec-
or lenticular in shape and they make up about 80-90% of the total cias and sandstones. These units are overlain by massive laharic
volume and area of the Deer Creek Member. The largest block breccias, flow breccias and trachyandesite lava flows. This coars-
identified thus far is about 12 km long and 5 km wide (Figure 15). ening upward relationship is indicative of the pre-collapse strati-
It occupies most of the area between Rattlesnake and Trout creeks graphic configuration of the volcanic edifice, and it reflects a
(lower elevations to the west). Blocks decrease in size and rather mature stage of development. The proportion of vent-
increase in number and structural complexity to the southeast. The facies rocks within blocks is higher in this area and they are the
rock type, facies, and depositional environments of the individual dominant rocks exposed between Robber’s Roost and Rat-
units within blocks are similar to other vent-medial-facies units tlesnake creeks. This prevalence of vent-facies rocks indicates a
exposed throughout the northeastern Absaroka Range. proximity to the source area. Throughout this area, the interior
structures of individual blocks are relatively simple with only a Member in the Sheep Mountain area. The afternoon stops will
slight backward tilt toward the inferred source area. Minor local focus on the stratigraphy of the upper South Fork Shoshone
changes in attitude within the larger blocks are due either to local River valley and the distal facies of the Deer Creek Member.
variations in initial dip or to broad folds and unmappable faults. The mileage between stops is approximate.
Bedding within blocks becomes highly deformed toward the base
and lateral margins, with folds, faults, and fractures common. Mileage description
Some of these larger blocks are intruded by swarms of trachyan-
desite dikes, most of which trend roughly north-south. These 0.0 Double Diamond X Ranch on the South Fork
dikes are typically perpendicular to bedding. Some dikes are Road.
truncated along bedding-parallel shear planes, and all are trun- 3.5 3.5 Bridge over the South Fork Shoshone River.
cated along block-matrix contacts. This relation indicates that the 3.3 6.8 TE Ranch Road
intrusion of the dike swarm almost surely predates the emplace- 4.4 11.2 Bridge over Rock Creek
ment of the debris avalanche. In this area, both the upper plate 4.7 15.9 Carter Mountain Road
allochthons and the Deer Creek Member occur within a broad N- 2.2 18.1 Park at sign on the north side of the highway
S trending paleovalley. At the time of emplacement, Rattlesnake that discusses the origin of Castle Rock.
Mountain stood as a prominent topographic barrier to the disper-
sal of the allochthons. Exposures of the Deer Creek Member do STOP 3-1: OVERVIEW OF THE TYPE AREA OF THE
not occur east of this locality. Time permitting, the contact DEER CREEK MEMBER
between the Logan Mountain Block and the Deer Creek Mem-
ber can be examined in detail. Turn caravan around and proceed This stop provides an excellent vantage of the type area of
back to US 14-16-20. Return to the Double Diamond X Ranch the Deer Creek Member (about 5 km to the north). At this local-
via Cody and the South Fork Road. ity the relief on the Eocene land surface, two mountain-size
blocks, and the matrix beneath and between these blocks can be
DAY 3: DEER CREEK MEMBER observed (Figures 16 and 17). The steep dark brown exposures
of the eastern block consists of proximal-medial facies breccias,
Introduction sandstones and conglomerates that dip about 30° to the north.
Beneath this block the Deer Creek Member matrix forms a 10
Day 3, like Day 2 begins at the Double Diamond X Ranch. m thick veneer above the Willwood Formation. The matrix at
We will spend today observing the volcanic stratigraphy and this locality has an imbricate fabric. Petrified wood is locally
detached extensional structures in the South Fork Shoshone common between these imbrications and indicate a forest was
River valley. The morning stops will include overviews and probably overrun during emplacement. The basal contact of the
detailed examinations of the proximal facies of the Deer Creek Deer Creek Member here preserves the early middle Eocene
topography. The grassy covered area between the large blocks STOP 3-2: BEAR CREEK AREA
consists of a chaotic assortment of matrix, small volcanic
blocks, small blocks of the Willwood Formation, and small This locality provides an excellent opportunity to view the
blocks of Mississippian limestone. Pierce and Nelson (1969) structural relationship between the Deer Creek Member and
map a fault contact between the Wapiti Formation and adjacent HMD and to observe the small-scale structure of the Deer
Willwood Formation along Deer Creek. This interpretation is Creek Member Matrix. Traverse on horse trail along the south
based on an abrupt change in stratigraphic level of the underly- side of Bear Creek for about 2.5 km. The trail slowly climbs up
ing Willwood Formation west of the inferred fault zone. Malone hill away from the creek. Several smaller volcanic and lime-
(1994) interprets the contact to be depositional, and attributes stone blocks (~10 m in diameter) occur along the trail about
the change in stratigraphic position to the deposition of the ~120 m above the creek. View to the north of the contact
younger Deer Creek Member of the Wapiti Formation on the between a Deer Creek Member block and an upper plate
east side of a steep middle Eocene hill that is composed of the allochthon of the HMD (Figure 18). At this locality, the dip of
Willwood Formation. A detailed examination of this locality is volcanic strata is gently to the northwest. The Paleozoic strata
highly recommended if time permits. The exposure can be of the upper plate allochthon also dip gently to the northwest. A
reached by several horse trails from the Castle Rock Ranch. 10-20 m wide zone of gray matrix occurs between these two
blocks. The high-angle truncation of volcanic strata along the
Continue northeast on the South Fork Road. contact indicates that these volcanic rocks are allochthonous.
2.8 20.9 Junction with the lower South Fork road. Proceed northward down to Bear Creek and up the other side
Turn left (north). along the contact to view the Deer Creek Member matrix (the
0.2 21.1 Cross the South Fork Shoshone River. climb is about 150 m).
1.5 22.6 Junction with Castle Rock Ranch road. Turn The debris-avalanche matrix is a highly sheared and perva-
left (west). Sheep Mountain allochthon is to sively brecciated zone that is found beneath and between blocks.
the north. The recognition of matrix is the key to mapping the distribution
1.4 24.0 Junction with Hidden Valley Ranch road. Turn of the unit and to understanding its origin. The matrix is typically
right (northwest). Proceed northwest to the Hid- light gray and bears a strong resemblance to ready-mix concrete.
den Valley Ranch at the end of the lane. Most of the matrix consists of sand-sized particles, which consist
1.1 25.1 Park cars and ask permission for access. The in decreasing abundance, of lithic fragments, hornblende, pyrox-
Sheep Mountain allochthon is to the east, two ene, glass, feldspar, and quartz. Pebble- to boulder-sized frag-
smaller allochthons occur to the west. To the ments are composed dominantly of light gray and pink
north is a large volcanic block of the Deer Creek hornblende andesite, and lesser dark gray trachyandesite. Matrix
Member. usually breaks through, rather than around, enclosed clasts, yield-
Figure 16. Panoramic view to the north of the type area of the Deer Creek Member of the Wapiti Formation from the South Fork Shoshone River Val-
ley, about 3 km away (from Malone, 1995). The light-colored, grassy foothills are underlain by the Willwood Formation (Twl) and Cody Shale (Kc).
The heavy dashed line is the early middle Eocene land surface with more than 320 m of relief. In this scene, two blocks (>1 km in diameter) are visi-
ble (Twdb). The block to the right (east) consists of about 250 m of interbedded breccias, sandstones, and conglomerates, and dips about 25° to the
north. The two blocks are bounded by a poorly exposed, lighter-colored interval of matrix. Matrix (Twdm) also occurs beneath each block but is too
thin to be resolved from this distance.
196 D. H. Malone, T. A. Hauge, and E. C. Beutner
Figure 17. View to the east of the type area of the Deer Creek Member of the Wapiti Formation (from Malone, 1996). Shown here are two blocks
separated by a zone of matrix up to 25 m wide. The block to the left (north) dips about 35° to the northwest, and the block to the right (south) dips
about 10° to the north. This locality displays an imbricate relationship between Deer Creek Member blocks. The width of photo is about 320 m.
ing smooth and rounded outcrops. Matrix was formed by the pro- intrude overlying blocks. These clastic dikes were likely injected
gressive disaggregation of block material during transport. Thus, late, and they indicate that some blocks were strong enough to
matrix is entirely tectonic in origin, and no primary (depositional) support the development of extension fractures during the late
layering is present. Matrix as defined here also includes all vol- stages of emplacement. Return to vehicles via the horse trail
canic fragments <25 m in diameter. Broken and shattered clasts along Bear Creek. Retrace route back along the South Fork Road
(ranging from pebble to boulder size) are the most common to the TE Ranch road.
structural feature within the matrix. Clasts of competent, dark- 41.4 0.0 Junction of the South Fork Road with the TE
colored trachyandesite lava display a wide variety of fracture pat- Ranch Road.
terns, ranging from well-developed conjugate shear sets to 1.6 1.6 Park at “Entering Shoshone National Forest”
pervasive shattering. Most of the shattering and fracturing of indi- Sign.
vidual clasts does not penetrate the surrounding finer grained
material, and the orientations of fractures within adjacent clasts STOP 3-3: OVERVIEW OF THE UPPER SOUTH FORK
are variable. In some places, minor displacement along the frac- SHOSHONE RIVER VALLEY
tures has produced visible offset. In extreme cases, fractures
within clasts were intruded by a slurry-like injection of sand- More than 1550 m of layered volcanic rocks are exposed
sized material. Less common, but also significant, are strained in the upper South Fork Shoshone River valley. In this area,
clasts. Most strained clasts are composed of incompetent, light- rocks from the Wapiti Formation, the Trout Peak Trachyan-
colored, hornblende andesite. As seen in thin section, the bound- desite, and the Wiggins Formation have all been identified
aries of strained clasts are difficult to discern, and (Malone, 1997; Figures 19 and 20). The Willwood Formation is
intrusive-appearing contacts between these clasts and the sur- locally exposed at the base of the volcanic succession in the
rounding material are common. The progressive deformation of Ishawooa Hills and Slide Mountain areas. The following dis-
these incompetent clasts led eventually to their complete destruc- cussion is modified from Malone (1997). A geologic map and
tion and homogenization into the surrounding material. Planar manuscript describing the stratigraphy and structure of this area
and curviplanar clastic dikes are found throughout the matrix. is in preparation. The Deer Creek Member in this area is
Individual dikes range in width from a few centimeters to a meter, extremely poorly sorted and consists of particles ranging from
and are as much as ten meters in length. Most clastic dikes are silt- to boulder-size. No internal stratification within the unit
light gray and composed dominantly of fine sand- to small-peb- has been recognized, indicating that the entire unit was likely
ble-sized material. At a few localities, multiple episodes of clastic emplaced during a single depositional event. The unit is matrix-
dikes are present and complex cross-cutting relations among supported and resembles a very large laharic breccia. Most
dikes are common, indicating that the matrix was dynamic, and clasts consist of well- to sub-rounded, dark gray trachyandesite,
that high pore-fluid pressures existed within the matrix through- red pyroxene andesite porphyry; and light gray hornblende
out the emplacement of the unit. Some of these clastic dikes andesite; the dark gray trachyandesite is the most abundant.
Heart Mountain detachment and associated structures, northeast Absaroka Range, Wyoming 197
Several large, irregular inclusions of trachyandesite lava averages about 600 ft in thickness, and ranges from a low of 90 m
and epiclastic breccia occur throughout the deposit, but they are between Sheephead and Hunter creeks to a high of 450 m
concentrated in the middle. Some of these inclusions are com- between Legg and Schoolhouse creeks. The top of the unit is
prised of light colored, well-stratified epiclastic sandstones and mostly flat, and the variation in thickness is mainly a function of
breccias, but this variety is rare. These inclusions range in size relief on the lower surface. In general, the unit is sheet-like in
from about 3 m to more than 155 m in diameter. Layering geometry, but it thins gradually to the south and west. In most
within these inclusions is often apparent from a distance, but cases, the relief on the basal surface is less dramatic. Throughout
upon closer inspection it is difficult to discern. Where present, most of the area, the lower contact occurs between 2240 m and
this layering appears to be highly contorted, and it is truncated 2305 m in elevation. These values are similar to the higher eleva-
along the margins of the inclusions. Contacts of these inclusions tions observed along the base of the Deer Creek Member in the
range from sharp to gradational. Where contacts are gradational lower North and South Fork Shoshone River valleys.
and the volcanic material within and adjacent to the inclusion It is likely that a distal-facies of the Deer Creek Member
are texturally similar, the contact is only recognized by a occurs throughout the upper South Fork Shoshone River valley.
change in color. Where contacts between the inclusions and Evidence for this assertion include: 1) the presently known south-
adjacent material are sharp, the contacts are locally striated— eastern extent of the Deer Creek Member is only 18 km from the
most commonly where the inclusions are comprised of tra- Ishawooa Hills area; the only stratigraphic pinch-out of the prox-
chyandesite lava. These striations are typically on the inclusion imal facies of the Deer Creek Member occurs near Jordan Creek
rather than on the surrounding material. In most cases (a total of at an elevation of about 2460 m; this pinch-out is attributed to a
five striated inclusions were observed), the striations are unidi- paleotopographic high, and the unit could well occur to the south-
rectional and subhorizontal and indicate laminar rather than tur- west of this paleohill at lower elevations; 2) in both areas, the
bulent flow. rocks display a similar color, texture, scale, and degree of defor-
Decker (1990) interpreted these inclusions as the result of mation; 3) in both areas, the units overlie a major early middle
the liquefaction and dismemberment of the overlying lava flows, Eocene unconformity; 4) kinematic data indicate a southeastward
flow breccias, and epiclastic breccias. In the present study, no dis- transport direction. This is consistent with a collapse directed out-
ruption of the overlying lava flows was observed, and all lique- ward from the inferred source area. The structural differences
faction domains are probably confined to the subjacent between the two areas can be explained by a difference in the
well-stratified succession. More likely, most of these inclusions emplacement mechanisms. In the lower North and South Fork
are probably disaggregated debris-avalanche blocks that were Shoshone River valleys, sliding was the dominant emplacement
suspended and supported buoyantly by the strength of the sur- mechanism, and blocks remained relatively intact and unde-
rounding matrix material when the predominant transport mech- formed. With further transport, blocks disaggregated, and the
anism evolved from slide to flow. The origin of the lighter dominant emplacement mechanism evolved from slide to flow,
colored inclusions is less obvious, as no light-colored strata with the resultant deposit having characteristics resembling a
within Deer Creek Member has been recognized elsewhere. It is very large, laharic debris flow deposit. The total distance from the
possible that these inclusions were incorporated from the subja- inferred source area at Sunlight Peak to the extreme southern
cent units during emplacement. The Deer Creek Member here exposure of the unit in this area is at least 58 km. This increases
198 D. H. Malone, T. A. Hauge, and E. C. Beutner
Figure 19. Geologic map of the upper South Fork Shoshone River valley.
the previously known transport distance by some 18 km. The are areally distinct and is another reason to suggest that the Deer
actual maximum transport distance is probably greater, as the Creek Member and Heart Mountain allochthon were emplaced
southernmost part is concealed by younger volcanic rocks. sequentially rather than coevally. The emplacement of the Deer
Another implication of the assignment of these rocks to the Deer Creek Member could have contributed to the liquefaction-related
Creek Member is that a clear distinction can be made between deformation in the underlying well-stratified epiclastic units, and
the respective distribution areas of the latter and the upper plate several domains of chaotically dismembered bedding have been
of the HMD. The closest upper plate remnant is in the Sheep identified (Decker, 1990) at Deer Creek Canyon, between Deer
Mountain area which is more than 32 km to the northeast, and no and Cabin Creeks, between Sheepeater and Hunter Creeks, and
Paleozoic limestone fragments of any size were identified within Southwest of Legg Creek. The deformation in these domains has
the Deer Creek Member in this area. This demonstrates that they been attributed to in-situ liquefaction caused by excess pore fluid
Heart Mountain detachment and associated structures, northeast Absaroka Range, Wyoming 199
pressures resulting from cyclic loading during large earthquakes. position rapidly. The underlying well-stratified succession is
It is possible that the liquefaction-related deformation may have commonly extensively fractured just beneath the contact. Sev-
been induced by impulsive loading caused by the emplacement eral large, irregular inclusions of trachyandesite lava can be
of the debris avalanche rather than repetitive loading during observed here within the Deer Creek Member. The contacts of
earthquakes. Unfortunately, there is no way to evaluate or quan- these inclusions are striated. These striations are subhorizontal
tify the relative importance of either stimulus. or gently inclined and trend nearly north-south.
Proceed southwest on the South Fork Road. Return to the vehicles and proceed south along the South Fork
1.9 3.5 Bridge over the South Fork Shoshone River. Road.
1.8 5.3 Park at the red U.S. Forest Service gate and pro- 1.5 6.8 Double Diamond X Ranch on right. Slide
ceed north across the sage covered foot hills to Mountain to the east.
the base of the cliff. 0.8 7.6 Ishawooa Mesa Trail Head
1.9 9.5 Bridge over Legg Creek.
STOP 3-4: DISTAL FACIES OF THE DEER CREEK
MEMBER IN THE ISHAWOOA HILLS AREA STOP 3-5: OVERVIEW OF DRAMATIC PALEO-
TOPOGRAPHY NORTHEAST OF LEGG CREEK
This location provides an opportunity to closely observe an
outcrop of the distal facies of the Deer Creek Member. Access Pull vehicles off the road near the bridge over Legg Creek.
to the Deer Creek Member is easiest here because the lower View to the north of a spectacular paleovalley filled by the Deer
contact is a relatively low elevation (paleovalley). The lower 30 Creek Member. During emplacement, the Deer Creek Member
m of the exposure consists of northward-dipping epiclastic filled and topped steep gorges that were carved in the underly-
sandstone, breccia, and conglomerate. The base of the Deer ing well-stratified succession (Figure 20). The west contact is
Creek Member is well-exposed here and changes stratigraphic nearly vertical for about 300 m of paleorelief. The east contact
200 D. H. Malone, T. A. Hauge, and E. C. Beutner
is less steep; the paleoslope on this side of the ancient canyon is logical Association Field Conference for 1947, Bighorn Basin, p. 189-
about 20 degrees. The origin and preservation of such a feature 197.
Chadwick, R.A., 1970, Belts of eruptive centers in the Absaroka-Gallatin Vol-
is enigmatic; it probably would not have been preserved if it had canic Province, Wyoming and Montana: Geological Society of America
not been immediately and fully buried by the deposition of the Bulletin, v. 81, p. 267-273.
Deer Creek Member. This relationship is not recognized on the Decker, P.L., 1990, Style and mechanics of liquefaction-related deformation,
lower Absaroka Volcanic Supergroup (Eocene), Wyoming: Geological
opposite cliff on the south side of the South Fork valley. Society of America Special Paper #240, 71 pp.
Along the west side of this paleovalley, there is evidence of Elliott, J. E., 1979, Geologic map of the southwestern part of the Cooke City
drag of the well-stratified units near the contact. In addition, sub- quadrangle, Montana and Wyoming: United States Geological Survey
Miscellaneous Investigations Series, Map I-1084, 1:24,000.
horizontal striations were observed in two places along the con- Erskine, D. W., and Kudo, A. M., 1991, Evidence from a Shoshonite dike in
tact. Both of these structures indicate that the feature is not merely support of the continuous allochthon model, Heart Mountain fault,
an unconformity but rather both an unconformity and a fault at the northwestern Wyoming [Abstract]: EOS., v. 72, no. 44, p. 490.
Hague, A., Iddings, J.P., Weed, W.H., Walcott, C.D., Girty, G.H., Stanton, T.W.,
same time. This relationship is remarkably similar to what is and Knowleton, F.G., 1899, Descriptive geology, petrography, and pale-
observed along the breakaway of the HMD, and a brief compari- ontology, Part II of Geology of the Yellowstone National Park; U. S.
son of the features of each is warranted. 1) Although at its south- Geological Survey Monograph #32, 893 pp.
Hauge, T. A., 1982, The HMD fault, northwest Wyoming: Involvement of
ernmost mapped point, the breakaway is buried by younger Wapiti Absaroka volcanic rock: Wyoming Geological Association, 33rd Annual
Formation rocks and is more than (48 km) away, the Legg Creek Field Conference, Guidebook, p. 175-179.
locality is along strike with the breakaway fault. 2) Both the break- Hauge, T. A., 1983, Geometry and kinematics of the HMD fault, northwestern
Wyoming and Montana [Ph. D. thesis]: Los Angeles, California, Uni-
away and this structure are half-faults (Pierce, 1979), the lower versity of Southern California, 265 p.
surface is a fault surface while the upper surface is a site of depo- Hauge, T. A., 1985, Gravity-spreading origin of the Heart Mountain allochthon,
sition. 3) Both the breakaway and this structure display a similar northwestern Wyoming: Geological Society of America Bulletin, v. 96,
p. 1440-1456.
relief across a short lateral distance. 4) Massive, dark-colored Hauge, T. A., 1990, Continuous-allochthon model of Heart Mountain faulting:
breccias immediately buried both the breakaway and this feature. Geological Society of America Bulletin, v. 102, p. 1174-1188.
It is likely that the Deer Creek Member is present within the hang- Hauge, T. A., 1992, HMD, Northwestern Wyoming, in Elliott, J. E., ed, Guide-
book for the Red Lodge - Beartooth Mountains - Stillwater Area: North-
ing wall of both localities. 5) The footwall of each is the western west Geology v. 20-21 (Tobacco Root Geological Society, Seventeenth
side and consists of light colored, well-stratified medial-distal- Annual Field Conference), p. 21-46.
facies volcanic rocks. 6) Both structures are similar (probably Hauge, T. A., 1993, The HMD, northwestern Wyoming: 100 years of contro-
versy: in Snoke, A. W., Steadtman, J. R., and Roberts, S. M., eds., Geol-
identical) in age. Despite these similarities, it is unlikely that this ogy of Wyoming (Blackstone - Love Volume), Wyoming Geological
structure in the upper South Fork valley is a southern extension of Survey Memoir No. 5, p. 530-571.
the breakaway of the HMD. Other large paleovalleys exist in the Hewett, D. F., 1920, The Heart Mountain overthrust, Wyoming: Journal of
Geology, v. 28, p. 536-557.
Ishawooa Hills area and along Little Boulder Creek. The paleore- Hughes, C. J., 1970a, Role of cohesive strength in the mechanics of overthrust
lief in these areas is less spectacular, but it is still in excess of 250 faulting and of landsliding: Discussion: Geological Society of America
m. Unfortunately, since exposures are limited to the canyon walls, Bulletin v. 81, p. 607-608.
Hughes, C. J., 1970b, The HMD fault - a volcanic phenomenon?: Journal of
it is impossible to trace these paleovalleys laterally. Geology, v. 78, no. 1, p. 107-116.
Hughes, C. J., 1970c, The HMD fault - a volcanic phenomenon? A Reply: Jour-
Proceed southwest along the South Fork Road. nal of Geology, v. 78, p. 629-630.
Hughes, C. J., 1973, Igneous activity, metamorphism, and Heart Mountain
4.5 14.0 End of the South Fork Road. faulting at White Mountain, northwestern Wyoming: Discussion: Geo-
logical Society of America Bulletin, v. 84, p. 3109-3110.
STOP 3.6: OVERVIEW OF DISTAL FACIES OF THE Malone, D.H., 1994, A Debris-Avalanche Origin for Absaroka Volcanic Rocks
Overlying the HMD, Northwest Wyoming. Unpublished Ph.D. disserta-
DEER CREEK MEMBER AT CABIN CREEK tion, The University of Wisconsin, 292 pp.
Malone, D.H., 1995, A very large debris-avalanche deposit within the Eocene
The Deer Creek member is beautifully exposed in the steep volcanic succession of the Northeasten Absaroka Range, Wyoming:
Geology, v. 23, no.7, p.661-664.
cliff face just north of and 170 m above Cabin Creek. At this local- Malone, D. H., 1996, Revised Stratigraphy of Eocene Volcanic Rocks in the
ity, the unit is approximately 100 m in thickness. Several large Lower North and South Fork Shoshone River Valleys, Wyoming:
inclusions of bedded distal facies rocks occur near the base of the Wyoming Geological Association Annual Field Conference Guidebook,
vol. 47, p. 109-138.
exposure. Two light-colored dacite dikes intrude the units. The dis- Malone, D.H., 1997, Recognition of a Distal Facies Greatly Extends the
continuous geometry of these dikes indicate that the Deer Creek Domain of the Deer Creek Debris-Avalanche Deposit (Eocene),
Member was not fully lithified as the dikes intruded. These dikes Absaroka Range, Wyoming: Wyoming Geological Association Annual
Field Conference Guidebook, vol. 48, p. 1-9.
are related to a series of small plutons in the area that intruded dur- Nelson, W. H., and Pierce, W. G., 1968, Wapiti Formation and Trout Peak tra-
ing the deposition of the overlying Wiggins Formation. chyandesite, northwestern Wyoming: United States Geological Survey
Bulletin, 1254-H, p. Hl-Hll.
Nelson, W. H., Pierce, W. G., Parsons, W. H., and Brophy, G. P., 1972, Igneous
REFERENCES CITED activity, metamorphism, and Heart Mountain faulting at White Moun-
tain, northwestern Wyoming: Geological Society of America Bulletin,
Beutner, E.C. and Craven, A.E., 1996, Volcanic fluidization and the Heart v. 83, no. 9, p. 2607-2620.
Mountain detachment, WY: Geology, v. 24, p. 595-598. Nelson, W. H., Pierce, W. G., and Brophy, G. P., 1973, Igneous activity, meta-
Bucher, W. H., 1947, Heart Mountain problem, in Guidebook, Wyoming Geo- morphism, and Heart Mountain faulting at White Mountain, northwest-
Heart Mountain detachment and associated structures, northeast Absaroka Range, Wyoming 201
ern Wyoming: Reply: Geological Society of America Bulletin, v. 84, p. Conference, p. 155-164.
3111-3112. Pierce, W. G., Nelson, W. H., and Prostka, H. J., 1973, Geologic Map of the
Nelson, W. H., Prostka, H. J., and Williams, F. E., 1980, Geology and mineral Pilot Peak Quadrangle, Park County, Wyoming, and Park County, Mon-
resources of the North Absaroka Wilderness and vicinity, Park County, tana: U.S. Geological Survey Miscellaneous Geologic Investigations
Wyoming: United States Geological Survey Bulletin 1447, with sections Map I-816, 1:62,500.
on Mineralization of the Cooke City mining district by James E. Elliott Pierce, W. G., Nelson, W. H., Tokarski, A. K., and Piekarska, E., 1991, Heart
and Aeromagnetic survey by Donald L. Peterson, 101 p. Mountain, Wyoming, detachment lineations -- are they in microbreccia
Pierce, W. G., 1957, Heart Mountain and South Fork detachment thrusts of or in volcanic tuff?; Geological Society of America Bulletin, v. 103, p.
Wyoming: American Association of Petroleum Geologists Bulletin, v. 1133-1145.
41, no. 4, p. 591-626. Prostka, H. J., 1978, Heart Mountain fault and Absaroka volcanism, Wyoming
Pierce, W. G., 1960, The "break-away" point of the HMD fault in northwestern and Montana, U.S.A., in Voight, B., editor, Rockslides and Avalanches,
Wyoming: in Geological Survey Research 1960: United States Geolog- I, Natural Phenomena: Amsterdam, Oxford, New York, Elsevier Scien-
ical Survey Professional Paper 400-B, p. B236-B237. tific Publishing Company, p. 423-437.
Pierce, W. G., 1965, Geologic map of the Deep Lake Quadrangle, Park County, Prostka, H. J., Ruppel, E. T., and Christiansen, R. L., 1975, Geologic map of the
Wyoming: U.S. Geological Survey Geologic Quadrangle Map GQ-478, Abiathar Peak Quadrangle, Yellowstone National Park, Wyoming and
1:62,500. Montana: U.S. Geological Survey Geologic Quadrangle Map GQ-1244,
Pierce, W. G., 1968, Tectonic denudation as exemplified by the Heart Mountain 1:62,500.
fault, Wyoming, in Orogenic Belts: 23rd International Geological Con- Sales, J.K., 1983, Heart Mountain; blocks in a giant volcanic rock glacier:
gress, Prague, Czechoslovakia, 1968, Report, Section 3, Proceedings: p. Wyoming Geological Association, 34th Annual Field Conference
191-197. Guidebook, p. 117-165.
Pierce, W. G., 1973, Principal features of the Heart Mountain fault, and the Smedes, H. W., and Prostka, H. J., 1972, Stratigraphic framework of the
mechanism problem, in DeJong, K. A., and Scholten, R., editors, Grav- Absaroka volcanic supergroup in the Yellowstone National Park region:
ity and tectonics: New York, John Wiley and Sons, p. 457-471.Pierce, U. S. Geol. Survey Prof. Paper 729-C, 33 p.
W. G., 1978, Geologic map of the Cody 1° x 2° Quadrangle, northwest- Sundell, K.A., 1985, The Castle Rocks Chaos; a gigantic Eocene landslide-
ern Wyoming: United States Geological Survey Miscellaneous Field debris flow within the southeastern Absaroka Range, Wyoming: Unpub-
Studies Map MF-963, 1:62,500. lished Ph.D. Dissertation, University of California, 283 pp.
Pierce, W. G., 1979, Clastic dikes of Heart Mountain fault breccia, northwestern Sundell, K.A., 1993, The Absaroka volcanic province, in Snoke A.W., Steidt-
Wyoming, and their significance: United States Geological Survey Pro- man, J.R., and Roberts, S.M., eds, Geology of Wyoming: Geological
fessional Paper 1133, p. 1-25. Survey Memoir #5, p. 572-603.
Pierce, W. G., 1980, The Heart Mountain break-away fault, northwestern Templeton, A. S., Sweeny, J. Jr., Manske, H., Tilghman, J. F., Calhoun, S. C.,
Wyoming: Geological Society of America Bulletin, Part 1, v. 91, p. 272- Violich, A., and Chamberlain, C. Paige, 1995, Fluids and the Heart
281. Mountain fault revisited: Geology, v. 23, p. 929-932.
Pierce, W. G., 1985, Map showing present configuration of Heart Mountain Torres, V., and Gingerich, P. D., 1983, Summary of Eocene stratigraphy at the
fault and related features, Wyoming and Montana: Geological Survey of base of Jim Mountain, North Fork of the Shoshone River, northwestern
Wyoming, Map Series MS-15, 1:125,000. Wyoming: Wyoming Geological Association 34th Annual Field Confer-
Pierce, W. G., 1987a, The case for tectonic denudation by the Heart Mountain fault ence Guidebook, p. 205-208.
- a response: Geological Society of America Bulletin, v. 99, p. 552-568. Tucker, T. E., 1982, Dead Indian Hill to Yellowstone National Park, Northeast
Pierce, W. G., 1987b, HMD fault and clastic dikes of fault breccia, and Heart Entrance: Wyoming Geological Association, 33rd Annual Field Confer-
Mountain break-away fault, Wyoming and Montana, in Beus, S. S., edi- ence, Guidebook, p. 380-386.
tor, Geological Society of America Centennial Field Guide, Rocky Voight, B., 1974a, Architecture and mechanics of the Heart Mountain and South
Mountain Section, p. 147-154. Fork rockslides, in Voight, B. and Voight, M. A., editors, Rock Mechan-
Pierce, W. G., and Nelson, W. H., 1968, Geologic map of the Pat O'Hara Moun- ics: The American Northwest, 3rd Congress International Society of
tain Quadrangle, Park County, Wyoming: U.S. Geol. Survey Geological Rock Mechanics Expedition Guidebook: Special Publication, Experi-
Quadrangle Map GQ-755, 1:62,500. ment Station, College of Earth and Mineral Sciences, The Pennsylvania
Pierce, W. G., and Nelson, W. H., 1969, Geologic Map of the Wapiti Quadran- State University, p. 26-36.
gle, Park County, Wyoming: U.S. Geol. Survey Geological Quadrangle Voight, B., 1974b, Roadlog: Wapiti-Heart Mountain Area - Canyon, in Voight,
Map GQ-778, 1:62,500. B. and Voight, M. A., editors, Rock Mechanics: The American North-
Pierce, W. G., and Nelson, W. H., 1970, The HMD fault -a volcanic phenome- west, 3rd Congress International Society of Rock Mechanics Expedi-
non? A discussion: Journal of Geology, v. 78, p. 116-122. tion Guidebook: Special Publication, Experiment Station, College of
Pierce, W. G., and Nelson, W. H., 1971, Geologic map of the Beartooth Butte Earth and Mineral Sciences, The Pennsylvania State University,
quadrangle, Park County, Wyoming: United States Geological Survey p. 112-124.
Geological Quadrangle Map GQ-935, 1:62,500. Wise, D. U., 1983, Overprinting of Laramide structural grains in the Clarks
Pierce, W. G., and Nelson, W. H., 1986, Some features indicating tectonic Fork canyon area and eastern Beartooth Mountains of Wyoming:
denudation by the Heart Mountain fault: Guidebook, 1986 Montana Wyoming Geological Association 34th Annual Field Conference Guide-
Geological Society -- Yellowstone-Bighorn Research Association Field book, p. 77-87.
Printed in U.S.A.
GSA Field Guide 1
Colorado and Adjacent Areas
David R. Lageson, Alan P. Lester, and Bruce D. Trudgill, eds.
Contents
1. Bouncing boulders, rising rivers, and sneaky soils: A primer of geologic hazards and
engineering geology along Colorado’s Front Range
D. C. Noe, J. M. Soule, J. L. Hynes, and K. A. Berry
2. Laramide to Holocene structural development of the northern Colorado Front Range
E. A. Erslev, K. S. Kellogg, B. Bryant, T. K. Ehrlich, S. M. Holdaway, and C. W. Naeser
3. Laramide faulting and tectonics of the northeastern Front Range of Colorado
E. A. Erslev and S. M. Holdaway
4. Hydrogeology and wetlands of the mountains and foothills near Denver, Colorado
K. E. Kolm and J. C. Emerick
5. Field trip to Manitou Springs, Colorado, with specific emphasis on the sediments of Cave
of the Winds and their relationship to nearby alluvial deposits and spring sediments
F. G. Luiszer
6. 200,000 years of climate change recorded in eolian sediments of the High Plains of
eastern Colorado and western Nebraska
D. R. Muhs, J. B. Swinehart, D. B. Loope, J. N. Aleinikoff, and J. Been
7. Walking tour of paleontologist George G. Simpson’s boyhood neighborhood
L. Laporte
8. Active evaporite tectonics and collapse in the Eagle River valley and the southwestern
flank of the White River uplift, Colorado
R. B. Scott, D. J. Lidke, M. R. Hudson, W. J. Perry, Jr., B. Bryant, M. J. Kunk, J. R. Budahn,
and F. M. Byers, Jr.
9. Coal mining in the 21st century: Yampa coal field, northwest Colorado
M. E. Brownfield, E. A. Johnson, R. H. Affolter, and C. E. Barker
10. Field guide to the continental Cretaceous-Tertiary boundary in the Raton basin, Colorado
and New Mexico
C. L. Pillmore, D. J. Nichols, and R. F. Fleming
11. Stratigraphy, sedimentology, and paleontology of the Cambrian-Ordovician of Colorado
P. M. Myrow, J. F. Taylor, J. F. Miller, R. L. Ethington, R. L. Ripperdan, and C. M. Brachle
12. Field guide for the Heart Mountain detachment and associated structures, northeast
Absaroka Range, Wyoming
D. H. Malone, T. A. Hauge, and E. C. Beutner
ISBN 0-8137-0001-9
,!7IA8B3-ha abe!