Sunteți pe pagina 1din 100

Abstract

On the Laurentian paleocontinent, repeated mass extinctions and radiations of trilobites and
associated fauna took place during the Upper Cambrian to Early Ordovician. This reoccurring pattern
is known as a “biomere” (Palmer, 1965), and the causes of their bounding extinctions are still poorly
understood. Using multidisciplinary approach, this study creates a high–resolution stratigraphic,
paleontological, and geochemical record for a biomere boundary, during which the extinction event
took place.

During the Sauk Sea Transgression, much of Laurentia was inundated. Mixed siliciclastic–
carbonate was deposited within the Inner Detrital Belt (IDB), landward of an extensive carbonate
platform. The boundary between the Pterocephaliid and Ptychaspid Biomeres is the focus of this
study. The extinction event occurred within the Sage Limestone–Conglomerate Member of the Snowy
Range Formation, deposited within the IDB. The Snowy Range Formation outcrops in southwest
Montana and northwest Wyoming, where the two sites studied are located: Nixon Gulch in the
Horseshoe Hills, Montana, and Wyoming Creek in the Beartooth Mountain Range, Wyoming. Detailed
stratigraphic columns of these two outcrops were drafted. Samples containing trilobites were
removed for identification and subsequent biostratigraphic correlation by Dr. John Taylor. Carbonate
samples were collected every 10–25 cm from our sites, and all 181 were packaged and shipped to
Harvard University. There they were prepared and analyzed using an Elemental Analysis Isotope
Ratio Mass Spectrometer (EA–IRMS) with help from Dr. David Johnston and his team.

Samples were analyzed for δ13Ccarb, δ34SCAS, and δ34Spyrite, in order to examine the paleoceanic
chemistry across the extinction event. There is extremely good correlation of isotopes between the
two sites, both isotopically and in relation to the assumed biomere boundaries. Carbonate carbon at
both sites presented a marked 0.3–0.7‰ negative shift of δ13Ccarb values at the extinction interval.
This correlates to a decrease in organically sequestered 12C due to the extinction of trilobites and
other fauna. δ34SCAS values also drop dramatically by 4–25‰ at the extinction boundary, due to the
upwelling of deep, cold ocean water, rich in 32S–enriched H2S. This hypothesis is supported by
biostratigraphic evidence that replacement fauna at the base of biomeres is composed of slope–
dwelling, cold–water trilobites (Westrop & Cuggy, 1999). δ34Spyrite values increase steadily by 15–
22‰ across the biomere boundary. This follows on an upwelling of 32S rich seawater, since bacteria
were less preferential about the use of light isotopes in bacterial sulfate reduction (BSR) if there is
greater 32S abundance in the environment. In summary, the biomere bounding extinction event was
brought about by cold, H2S rich waters brought onto the shelf during a sea–level highstand maximum
(Runkel et al., 2007).

1
Introduction

The goal of this research is to determine the cause or causes of the trilobite
extinction events that separate biomeres in the Upper Cambrian. There is a lot of
uncertainty regarding these extinction events, despite the large body of work by
Grant (1965), Palmer (1965, 1971, 1979, 1981, 1984, 1998), Perfetta et al. (1999),
Saltzman (1999) et al. (1995, 1998), Stitt (1971, 1975, 1983), Taylor (1997, 2006),
Thomas (1993), and Westrop (1995) et al. (1987, 1999). Although these studies
utilize such diverse techniques as sedimentology, biostratigraphy and geochemistry,
none of them are high–resolution in nature, and so this study involves a high–
resolution, detailed sedimentological, biostratigraphic, and stable isotope
geochemical analysis in order to create a detailed paleoenvironment record for the
biomere boundary.

Numerous hypotheses have been given to explain the closely spaced, repeating
extinction events found bounding biomeres. They range from the mundane to the
otherworldly. Öpik (1966) proposed a sudden transient increase in solar heat as the
possible cause of the extinctions. Cooling of ocean waters was suggested by both
Lochman and Duncan (1944) and Stitt (1975). Lochman and Duncan thought it was
associated with a regression, while Stitt postulated a coupled transgression and
rapid rise in thermocline, an idea supported by Taylor (1977). He also found
evidence for a highly stratified ocean in the Cambrian. Palmer (1984) hypothesizes
periodic bolide impacts as the mechanism of extinction. There is no other support
for this extraterrestrially forced trilobite extinction. Finally, chemical changes in the
ocean such as an oxygen minimums and euxinia are possible causes, although they
are under studied for the biomeres (Palmer, 1984).

The focus of this research is the Pterocephaliid–Ptychaspid Biomere boundary


that is located in the Sage Member of the Snowy Range Formation, of Wyoming and
Montana. During the Upper Cambrian this area was inundated under a shallow
inland sea, and the depositional environment of this formation consisted primarily
of mixed carbonate–siliciclastics. I undertook detailed stratigraphic analysis at two

2
sites: Nixon Gulch in Montana, and Wyoming Creek in Wyoming. Resulting
descriptions were used to create a depositional history, and to get an idea of sea–
level changes over the extinction interval. Lithologic data was then compared to
biostratigraphic data in order to correlate the extinction between sections. I also
integrated high–resolution carbon and sulfur isotope curves with stratigraphic and
sedimentological data in an effort to support or invalidate previous hypotheses and
help explain the perplexing trilobite extinctions of the Upper Cambrian.

3
Upper Cambrian Paleogeography:

During the Cambrian, Laurentia, made up of the present day North American
craton, sat astride the equator (Scotese et al., 1979). The paleocontinent, along with
Siberia, Baltica and the Gondwana supercontinent, had separated from the main
mass of the Rodinia supercontinent in the Neoproterozoic (McKerrow et al., 1992).
Laurentia was rotated 90° clockwise from its present day position (Figure 1). The
breakup of this supercontinent in the late Proterozoic to Lower Cambrian led to the
formation of the Cordilleran passive margin on the western edge of the North
American craton (Bond et al., 1984). The Cordilleran and Appalachian passive
margins were located at 15° N and S latitude, respectively (Saltzman et al., 2004).

By the Middle to Upper Cambrian, the drift stage of tectonic evolution on the
passive–margin was characterized by the development of an extensive carbonate
platform (Palmer, 1971; Levy and Christie–Blick, 1991). This carbonate platform
extended from near the cratonal hinge zone in Utah and Wyoming (Levy and
Christie–Blick, 1989) to the distally steepened ramp edge in Nevada (Osleger and
Read, 1993). The platform also extended south around Texas, and then north up
through Michigan and over into New England. The passive margins were
approximately 2 km thick prisms, and were characterized by a dominantly shallow–
water carbonate band referred to as the Middle Carbonate Belt. Oceanward, this
environment passed into a deeper–water siliciclastic facies association, the Outer
Detrital Belt. Landward, the middle–carbonate belt passed into interior cratonic
siliciclastic deposits called the Inner Detrital Belt (Saltzman et al., 2004). The study
area is within the Inner Detrital Belt, which is composed predominantly of mixed
siliciclastic and carbonate facies (Figure 2).

Sea–level at the time was generally high, near a Phanerozoic maximum due to
the greenhouse climate (Berner, 1990) and the effect that continental breakup had
on ocean–ridge volumes (Bond et al., 1984). From the Middle to Upper Cambrian,
there were a number of large–scale transgressions, known collectively as the Sauk

4
5
6
transgressions. This name stems from the shallow Sauk Sea, which covered most of
Laurentia at the time (Monroe and Wicander, 1997). The initial transgression was
marked in northern Wyoming by the deposition of the coarse siliciclastic Flathead
Formation. Subsequent deposition in the Inner Detrital Belt included three large–
scale siliciclastic–carbonate couplets, or grand cycles (Aitken, 1966). The
Pterocephaliid–Ptychaspid Biomere boundary, the focus of this study, falls in the
third of these, the Sauk III transgression (Figure 3). The basal carbonate–siliciclastic
transition of this transgression is stratigraphically tied to the contact between the
Pilgrim Limestone and the lowermost member of the Snowy Range Formation, the
Dry Creek Shale. The Snowy Range Formation consists of this shale, the overlying
Sage Limestone–Conglomerate Member, and the uppermost Grove Creek Dolomitic
Limestone, which is overlain in turn by the Middle Ordovician Bighorn Dolomite
(Grant, 1965).

According to Myrow et al. (in review), the Snowy Range Formation was
deposited within the Inner Detrital Belt, east and landward of the carbonate
platform (Figure 2). The water depth in this Inner Detrital Belt was several tens of
meters (Runkel et al., 2007; Myrow et al., in review). The depositional environment
of the Snowy Range Formation is in the deepest part of the wide Inner Detrital Belt
between the carbonate platform and the sandy shoreline. This epicratonic setting
was characterized by the deposition of mixed siliciclastic and carbonate sediment
because the warm, shallow tropical setting was subjected to runoff from adjacent
exposed rocks of the North American craton. The carbonate shelf was deposited in
part because of the equatorial location of Laurentia at this time. Being at the
equator, with a shallow sea along its shelf allowed for a high rate of precipitation of
carbonate rocks. The shoreline of the Inner Detrital Belt consisted of siliciclastic
sand, as deduced from sandstone deposits of the Upper Mississippi Valley region
(Figure 2) (Runkel et al., 2007). Even with the horizontal shifting of depositional
environments due to changes in relative sea–level, this part of the Inner Detrital Belt
remained a mixed–siliciclastic deposition zone since neither extreme of depositional

7
8
environment (either shoreface or carbonate shelf) extended far enough landward or
shoreward to turn the Snowy Range Formation into a predominantly carbonate or
siliciclastic deposit.

The formation records the “competition” between relatively in situ aggradation


of carbonate, and the input of siliciclastic sediment from rivers. The alternation
between siliciclastic and carbonate facies is due to changes in relative sea–level, the
causes and effects of which will be discussed later. According to the model of Myrow
et al. (in review), during relative sea–level rises the source of siliciclastics move
further landward, and the input of mud into the system was essentially cut off. They
interpret the carbonate facies as a product of relative sea–level highstand
conditions, in which the carbonate factory was active in the absence of siliciclastic
sediment input. The siliciclastic facies in this formation is shale, which was derived
from terrestrial sources (i.e., rivers), which during regressions were brought closer
to the depositional environment of the Snowy Range Formation. When siliciclastic
input was increased, the carbonate factory turned off for various reasons, including
increased turbidity.

The cause of the relative sea–level changes is uncertain, although Miller et al.
(2005) suggest that glacio–eustasy was a factor in the Upper Cambrian. An increase
in temperature would cause a rise in sea–level due to the melting of glaciers and
ice–caps. Myrow et al. (in review) speculate that increased oceanic surface area
during transgression was coupled with an increase in sea temperature due to the
increased absorption of solar radiation. This has a two–fold effect on deposition, an
increase in storm intensity (Emanuel, 2005) as well as an increase in carbonate
precipitation due to the loss of dissolved CO2 in warm, shallow waters. The latter
may have led to widespread cementation of the sea floor. If tied to an increased
distance of siliciclastic sources, carbonate deposition would prevail, and high–
energy storms would have produced flat–pebble conglomerate beds through
erosion of cemented sea floor crusts. The fine siliciclastics were deposited during

9
low stands, since the siliciclastic sources were closer to the depositional
environment, and storms were much less intense.

10
Biomeres

Palmer (1965) first defined the biomere as “a regional biostratigraphic unit


bounded by abrupt non–evolutionary changes in the dominant elements of a single
phylum.” Biomere means, literally, segment of life (Palmer, 1965). These faunal
changes aren’t necessarily related to any changes or discontinuities in the
sedimentary record. The trilobite taxa within a given biomere are not known or
abundant in older or younger faunas of the United States (Figure 4). Each biomere is
named for a family of trilobites that is particularly well represented in that interval,
although not necessarily restricted to it (Runkel et al., 2007). In a biomere, platform
trilobite diversity rises through several zones, and then drops rapidly through an
extinction horizon comprising one or two subzones (Stitt, 1971, 1983; Myrow et al.,
1999; Taylor, 2006).

The Upper Cambrian of North America is punctuated by three such mass


extinction horizons, known commonly as biomere boundaries (Lochman and
Duncan, 1944; Palmer, 1965, 1984; Thomas, 1993). These boundaries represent
quasi–isochronous surfaces (Palmer, 1979) and have been critical in the intra– and
inter–basinal correlation of different areas of the North American continent
(Osleger and Read, 1993). They may even be global in extent, as boundaries have
also been found in Australia (Öpik, 1966). These biomere boundaries are defined by
a large–scale diversity decline in trilobite populations (Runkel et al., 2007). Each
extinction event falls in the uppermost section of the underlying biomere. The next
biomere and its associated increase in faunal diversity starts above the critical
interval of the extinction. The start of the extinctions events defines the boundaries
of the Upper Cambrian Stages in North America (the Marjuman, Steptoean and
Sunwaptan Stages) (Runkel et al., 2007). The biomeres associated with these Stages
don’t correspond perfectly, as the critical intervals act as an overlap between the
Stages and their respective Biomeres: the Marjumiid, Pterocephaliid and Ptychaspid
biomeres (from oldest to youngest) from Ludvigsen and Westrop (1985). This study
focuses on the Pterocephaliid–Ptychaspid (Steptoean–Sunwaptan) biomere

11
12
boundary extinction event. The Pterocephaliid was the type biomere studied by
Palmer (1965) (Figure 5).

From the 1960’s through the 1980’s, there was considerable debate as to what
should define the exact boundaries of a biomere. Palmer (1965) and Stitt (1971)
suggested that the rise of new species should represent the start of a biomere.
Palmer (1979, 1981) and Ludvigsen and Westrop (1985) adopted a definition of the
lower boundary of the biomere as the horizon of faunal change that represents the
onset of the extinction event. However, Taylor (1997) suggested that a return to the
original definition be made. Upon further study of brachiopod and conodont data he
found that they reflected the extinction seen in the trilobites, although to a much
lesser extent (Palmer, 1998). There is rapid species–level replacement of
inarticulate and articulate brachiopods, first noted in the basal Sunwaptan strata of
Montana and Wyoming, which corresponds to the stratigraphic position examined
in this study. “Articulate brachiopods, in particular, appear to be characterized by
extensive immigration during the stage boundary interval” (Westrop, 1995). There
is a slight lag in turnover rates of other species compared to the trilobites in regard
to biomere boundary extinctions and repopulations, but it is thought that this may
simply be a consequence of differences in the speciation and extinction rates of
different taxa (Westrop 1995). Therefore, we place the biomere boundary after the
extinction interval.

Although the main purpose of this study is to try to interpret the cause (or
causes) of the extinction at the Pterocephaliid–Ptychaspid biomere boundary, it is
important to understand the structure of the biomere as whole. According to Stitt
(1971), each biomere is defined by four successive trilobite evolution stages. The
first, stratigraphically lowest stage is characterized by species with high levels of
morphological variability. The second stage is characterized by species with longer
stratigraphic ranges and less individual morphological variability than those
trilobites in the first stage. The third stage generally contains trilobites of high
species diversity, all of which have a large stratigraphic range and increasingly low

13
14
morphological variation. Low species diversity and coquinas rich in at least one
species of trilobite characterize the fourth, final, and stratigraphically highest stage
(Stitt, 1971). All of these stages, when put together take no more than three to five
million years (Westrop and Ludvigsen, 1987), making biomeres some of the
shortest, repeating extinctions known.

The four stages described above define the nature and relative diversity of
trilobites following an extinction event. Westrop and Cuggy (1999) state that in the
aftermath of the extinction events a single biofacies consisting of immigrants from
off–shelf sites occupy all known lithofacies (in this study: Idahoiid Biofacies). The
initial burst of adaptive evolution (stage one) is this immigration of off–shelf
trilobites filling environmental niches (Figure 6). In fact, Westrop and Ludvigsen
(1987) showed that 86% of families extending into the upper–slope biofacies
survived, compared to just 36% of those confined to shelf biofacies. This study was
conducted on the biomere boundary above the one studied here, but the conclusions
drawn can be applied to all biomeres. Stage two represents the complete adjustment
to, and utilization of, their environments by a few genera. Stage three is similar to
stage two, except that it represents maximum adjustments to the environment. The
initial extinction defines the base of stage four, although some extinction of long–
ranging species takes place in stage three.

The factors responsible for the development of these patterns must have been
sufficiently strong and general enough to “overcome the perturbations caused by
local environmental differences” (Ashton and Rowell, 1975). Indeed, these patterns
are recorded in locations across Laurentia including a shelf bounding the western
margin of the craton (Great Basin), an earlier subsiding intracratonic basin
(Oklahoma) (Lochman–Balk, 1971), and a narrow carbonate platform of the
southern margin of the North American craton (Texas).

The focus of this thesis, the Pterocephaliid–Ptychaspid Biomere boundary, is the


least studied of the three major biomere boundaries of the Upper Cambrian,

15
16
although Palmer (1998) studied the faunal changes within the Steptoean and
Sunwaptan Stages, which occur proximal to the biomere boundary (Figure 4, 5). The
unifying faunal elements of the Pterocephaliid Biomere are trilobites of the
Pterocephaliidae and Elviniidae zones (Plate 1. A, B). At the top of this biomere the
extinction horizon is characterized specifically by the appearance of Comanchia
amplooculata (Frederickson), Dellea? Punctate (Palmer) and the proliferation of
Irvingella major (Plate 1. C. 1) (Ulrich and Resser). The overlying Sunwaptan Zone is
dominated by Dikelocephalidae, Ptychaspididae, Parabolinoididae, and
Ellipsocephaloididae (Plate 1. C) (Palmer, 1998).

17
18
Geochemical Background

Isotope Ratios:

Carbon, oxygen and sulfur are the most commonly used elements in
paleogeochemical reconstruction, and the analysis of secular variations in the
composition of seawater during the geologic past (Kampschulte and Strauss, 2004).
The natural isotopic abundance of these elements and others used in geochemistry
are expressed as a ratio of relative abundances. In this thesis I am presenting data
on the ratios of carbon and sulfur isotopes:
13
C
R carbon = ⁄12
C

and
34
S
R sulfur = ⁄32
S

Natural carbon and sulfur consist of several stable isotopes, and


chemostratigraphic studies of ancient rock focus on the ratios of 12C and 13C
(abundances of approximately 98.9 and 1.1% respectively), and 32S and 34S
(abundances of approximately 95.02 and 4.21% respectively).

Isotope abundances are given as a ratio between a sample and a standard. The
“δ–value” is used in reporting these values, defined by the Raleigh equation as:
R − R standard
δsample = [( sample ⁄R ) − 1] × 103
standard

where Rsample is the isotopic ratio of the sample and Rstandard is the isotopic ratio of a
corresponding standard. Higher values for δ represent a greater abundance of the
heavy isotope, while smaller values signify that the lighter isotope is present in
greater abundance in the sample than in the standard. The standards allow for
comparison of data from different laboratories worldwide. Additionally, since the
values for isotopic fractionation are so minute, they are measured as the relative
abundance of parts per thousand, or per mille (‰). Carbon isotopic abundance is

19
represented as δ13C, and is measured against the internationally accepted PeeDee
Belemnite standard. Sulfur isotopic abundance is given by δ34S and is standardized
against IAEA–SO–5, an inter–laboratory comparison standard distributed by the
International Atomic Energy Agency (IAEA) with an internationally accepted 34S
value.

Stable Isotope Geochemistry:

Stable isotope geochemistry allows for the use of physical and chemical
differences in stable isotopes of naturally occurring elements to look at geochemical
processes. These processes preferentially segregate various isotopes due to their
specific bond strength in molecules and certain mass–dependent reactions. All
isotopes of an element display the same chemical properties, but since they weigh
very slightly different amounts they are incorporated differently in reactions. In
general, the lighter isotopes are more mobile, and are thus more susceptible to
reactions such as isotope exchange reactions, condensation and evaporation,
crystallization and melting, adsorption and desorption, biologically mediated
reactions, and diffusion.

Preferential incorporation of isotopes can be due to biological, chemical or


physical processes. For isotopic fractionation, no single process can be linked
directly as a causal mechanism. Biological, physical and chemical fractionation all
occur simultaneously, and it is a combination of processes that interact to form a
distinct isotopic signature for any specific sedimentary deposit (Kendall and
Caldwell, 1998). Due to the complex interplay of factors that goes into creating an
isotopic signature for a deposit, there is difficulty in tying a positive or negative
variation in isotopic ratio to a specific causal mechanism. It is necessary to align
stratigraphic and paleontological data with chemostratigraphic data in order to
draw meaningful conclusions about the paleoenvironment and associated events.

20
Although samples are correlated using an international standard, δ–values for
specific geochemical events often vary worldwide, and even from location to
location within facies realms such as the Laurentian Inner Detrital Belt, the focus of
this thesis. The variation in values of isotopic excursions at different sites can be
significant in and of itself, as different sites may record events differently. Isotopic
differences within a particular section or basin may result from regional variations
(Beauchamp et al., 1987). The relative size of the excursions also relate to associated
geochemical events, even if the specific δ–values don’t necessarily agree with each
other from one outcrop to another over the same stratigraphic interval.

The global carbon and sulfur cycles include fluxes into and out of the
hydrosphere, atmosphere, biosphere and lithosphere. Since both of these elements
are involved in reactions in different oxidation states, they relate to nearly every
geological environment (Ripperdan, 2001) (Figure 7). This thesis focuses on the
carbon and sulfur isotope ratios in the hydrosphere during the Upper Cambrian. The
isotopic composition of the ocean is dependent on the isotopic ratio and total
amount of carbon and sulfur entering the ocean, as well as that fraction being
sequestered.

The main input of both carbon and sulfur into the ocean is from continental
weathering and erosion of sediment, and the subsequent transport of this
particulate sediment and dissolved ions to the ocean (Berner and Raiswell, 1983;
Hurtgen et al., 2009). A much smaller source of these elements in the ocean is due to
the adsorption of atmospheric CO2 and SO2 (and subordinate H2S). This carbon and
sulfur is brought into the ocean through wind and wave generated gas exchange
with the atmosphere, as well as subaqueous outgassing due to diagenetic alteration
of sediment, volcanism, and metamorphism (Halmer et al., 2002).

21
22
Carbon in the Ocean:

At present, most carbon in the ocean is dissolved inorganic carbon (DIC). DIC
makes up 98% of all oceanic carbon, and is composed primarily of bicarbonate
(HCO−
3 ) and carbonate (CO3 ). There are also small amounts of dissolved carbon
2−

dioxide present. As well as DIC, carbon derived from living or dead organic material
is present in the oceans in the form of dissolved organic carbon (DOC) and
particulate organic carbon (POC). All carbon, be it organic or inorganic, is removed
via sedimentary sequestration and subsequent subduction (Figure 8). Carbon has a
comparatively short residence time in the ocean, which contributes to the oceans
isotopic heterogeneity (Holser et al., 1986). The carbon–isotope composition of
modern oceans ranges between δ13CPDB values of –0.5 and 2.0‰ (Kroopnick, 1985).

The δ13C of ocean water is best reflected by the precipitation of DIC, which reacts
with Ca2+ to form CaCO3 under warm, low–pressure conditions. This
physicochemical precipitation of marine carbonate occurs in isotopic equilibrium
with seawater. Since there is almost no fractionation associated with this reaction
the precipitated carbonate is an extremely good proxy for the isotopic ratio of the
solution from which it precipitated. There are almost no effects on the δ13C of the
carbonate due to temperature in the near–surface temperature range (Anderson
and Arthur, 1983; Romanek et al., 1992). Thus, determining paleoceanic isotope
composition is best done using unaltered, abiotic (nonbiological) marine cements
precipitated as low–Mg calcite under oxic conditions (Lohmann and Walker, 1989;
Marshall, 1992). However, most studies, including this research, use bulk samples of
relatively pristine rock. Although this averages the isotopic composition of the
various components of the sample, it has proven to be a reliable method in studies
of strata from all ages.

Organic processes such as photosynthesis, respiration, and other metabolic


activities have a much larger effect on the δ13C of precipitated organic carbon (and
in reciprocal fashion, inorganic carbon) than inorganic precipitation has on

23
24
carbonate precipitation. Phytoplankton, algae, plants and some bacteria
preferentially incorporate 12C into their bodies, removing large volumes of 12C from
the ocean. These autotrophs photosynthesize CO2 and are consumed by larger
organisms that respire CO2. This isotopically light organic material is precipitated
and buried, leaving the ocean with a higher δ13C, which becomes inorganically
precipitated carbonate. According to Berger and Vincent (1986), the removal of 20
to 25% of the organic carbon in the ocean could lead to a δ13C shift of 4 to 5‰. This
is due to the fact that δ13C of marine organic matter is about 25‰ more negative
than inorganic bicarbonate (Marshall, 1992).

Sulfur in the Ocean:

The modern ocean has a δ34S value of 20‰, compared to the average modern
value of terrestrial run off of 0–10‰ (Holser et al., 1988), its main input. Obviously,
fractionation is occurring in the ocean on a large scale to create such a large
difference between terrestrial and oceanic δ34S. Sulfur present in the ocean is
composed primarily of sulfate and its reduced counterpart, sulfide (Figure 9).

The sulfate in the ocean (SO2−


4 ) is reduced biologically to sulfide, specifically

hydrogen sulfide (H2S). This process is known as bacterial sulfate reduction (BSR),
and is the single most important pathway for mineralization of organic sedimentary
matter in the ocean. Burial of organic–rich marine sediment depletes sediment pore
waters of oxygen due to aerobic respiration. This leads to anaerobic respiration and
the reduction of sulfate to sulfide (Bottrell, 2006). Also, due to the high solubility of
sulfate compared to dissolved oxygen, BSR is much more frequently the source of
mineralized organic matter in marine sediment (Mackin and Swider, 1989).

The bacteria that reduce sulfate to sulfide preferentially choose 32S to place in
the product, leaving the ocean enriched in 34S (Bottrell and Raiswell, 2000; Bottrell,
2006). Pyrite records this fractionation in the sediment via the following process.

25
26
Detrital iron–bearing minerals react with the hydrogen sulfide formed during BSR,
and thus pyrite (FeS2) is produced and sequestered away in ocean floor sediment.
This sequestration depends upon the local depositional environment, specifically
sedimentation rate, sulfate reduction rate, iron availability, and oceanic sulfate
concentration (Fike, 2007).

H2S and the pyrite formed from it can be depleted in 34S by up to 40 to 45‰
(Canfield, 2001; Detmers et al., 2001). Indeed, fractionations of up to 70‰ have
been reported between sulfate and pyrite in both modern and ancient
environments. This is commonly explained by redox recycling and associated
diproportionation reactions (Canfield and Thamdrup, 1994; Brunner and
Bernasconi, 2005), and illustrates the extreme isotopic variation that can be created
by bacterial sulfate reduction. Thus, the sulfur isotopic composition of seawater
sulfate is generally controlled by the large isotope fractionation by BSR (Bottrell,
2006), and although the majority of the sulfide (75 to 95%; Bottrell and Raiswell,
2000) is re–oxidized to sulfate and intermediate sulfur species (Jørgensen et al.,
1990) a large sulfide reservoir is produced, which is significant enough to change
the oceans δ34S appreciably.

The δ34S of oceanic sulfate is recorded in several minerals. Those with the most
sulfur contained within the mineral itself are anhydrites. These are formed by the
precipitation of sulfate minerals during evaporation of seawater, a process that has
little (0 to 3‰) associated fractionation (Raab and Spiro, 1991; Gill, 2007).
Therefore δ34S of anhydrites is a good proxy for oceanic δ34S at the time of
deposition. Unfortunately, these anhydrite deposits are extremely discontinuous in
the geological record and contain no fauna for stratigraphic correlation, and
therefore provide a poor source of high–resolution oceanic δ34S data.

Alternatively, sulfur can be removed from the ocean in trace amounts through
the incorporation of sulfate into the lattice structure of marine carbonate rocks. This
is a small flux, but it happens almost continuously and thus provides an excellent

27
and complete record of oceanic δ34S (Strauss, 1999). Referred to as both carbonate
associated sulfur (CAS) and structurally substituted sulfur in carbonate (SSS),
sulfate ions replace carbonate ions in marine carbonate (Bottrell, 2006).
Abundances vary from a few tens of ppm in inorganic precipitates (e.g., micrite) to
several thousand ppm in biogenic carbonates (Grossman et al., 1996; Kampschulte
and Strauss, 2004). The δ34S of CAS in modern marine sediment typically matches
that of the seawater to within 1‰ (Kampschulte and Strauss, 2004; Lyons et al.,
2004).

CAS has several distinct advantages over anhydrite deposits. First, it is much
more widely distributed in the geological record. Second, the carbonate used for CAS
generally contains fossil faunas that allow for precise chronostratigraphic
correlation. More importantly, carbonate successions are often relatively
continuous, with rapid rates of deposition, which allows for extremely high
temporal resolution of changes in marine δ34S. Indeed, other methods that utilize
marine barite and trace sulfate in phosphate work well, but don’t provide the spatial
or temporal coverage afforded by CAS (Gill, 2007). Studies of δ34S in Precambrian
interbedded gypsum and carbonate has established that the CAS signal can be
preserved over long periods of geological time (Kah et al., 2001, 2004). Bottrell
(2006) showed that high temporal resolution allows for preservation of a record of
marked rapid changes in marine δ34S associated with major Earth history events,
such as extinctions and the oxidation state of the ocean. Furthermore, his detailed
study of CAS records showed that CAS δ34S did not fluctuate significantly during
periods with no major or rapid environmental changes.

SPICE and Related Isotopic Excursions:

Secular variation in δ13C in seawater arises from fractionation driven by


continental weathering, global sedimentation rates, primary productivity, organic
carbon burial, and ocean circulation (Montañez et al., 2000). Sulfur isotopic studies

28
can be used to detect these types of paleoenvironmental conditions. Specifically,
δ34SCAS and δ34Spyrite can be used to interpret the level of biological turnover of
sulfate and changes in continental weathering of sulfur–bearing minerals
(Kampschulte and Strauss, 2004).

Of special interest and import in stable–isotope studies are stratigraphically and


regionally extensive positive and negative excursions. These excursions generally
take place over major boundaries throughout the stratigraphic record and are
indicators of tectonic, climatic, oceanographic, and evolutionary changes that
influence the cycling of the elements involved (Magaritz, 1991). Major biological
events and stratigraphic boundaries often coincide with strong δ13C and δ34S
excursions (Veizer et al., 1999).

One such excursion is recorded in Upper Cambrian successions globally, from


North America to China, Siberia, Baltica, Australia and Kazakhstan (Saltzman et al.,
2000; Ahlberg, 2009). This is a large positive δ13C excursion (nearly 4‰; Saltzman
et al., 1998) called the Steptoean Positive Carbon Isotope Excursion, or SPICE
(Figure 10). As stated, this excursion occurs in the Steptoean zone, during the
Pterocephaliid Biomere, and lasted for about 3.5 m.y. (Saltzman et al., 2004). Its
peak coincides with the maximum regression of the Sauk II–Sauk III hiatus (Osleger
and Read, 1993; Saltzman et al., 2000), and the rise of this excursion coincides with
a time of maximum faunal diversity in the Pterocephaliid Biomere of Laurentia
(Palmer, 1984). The shift back to background values at the end of the SPICE occurs
prior to the end of the Pterocephaliid Biomere, during the Elvinia Zone (Saltzman et
al., 1998), so it is unrelated to the biomere boundary studied in this thesis. However,
it is instructional in understanding Upper Cambrian isotopic excursions and the
implications they have for the paleoenvironment of the time.

Such positive isotope excursions are generally interpreted as the result of one or
more of the following possible factors. An increase in organic productivity and
associated burial of organic material will draw 12C preferentially into the sediment,
resulting in an increase of δ13Ccarb. This generally occurs during explosive

29
30
evolutionary events. An increase in sedimentation rate in the oceans, along with its
increase in burial of organic material will again preferentially sequester 12C and thus
increase δ13Ccarb. An enhanced preservation of organic material caused by marine
anoxia or an expanded oxygen minimum zone can result in an increase in preserved
organic material, thus increasing δ13Ccarb (Weissert, 1989; Brasier et al., 1994).

A complete analysis of isotopic records requires integration with a wide variety


of additional stratigraphic data. For instance, it is known that the boundary between
the Sauk II and Sauk III transgressions falls almost directly over the maximum
δ13Ccarb of the SPICE excursion (Saltzman, 1999). Thus, one could extrapolate from
the evidence at hand that the increase in δ13Ccarb values could have been a tracer for
the increase of shallow, productive seas created during regression. These
productive seas would have sequestered light carbon as organically deposited
materials, thus increasing δ13C of the ocean (Saltzman et al., 1998).

Further integration with stable sulfur isotope records is also of great use in
analysis. In this study, sulfur and carbon were the two elements used to test
paleooceanic chemistry. In regards to the SPICE, δ34SCAS increased nearly 20‰
(Saltzman et al., 2011), and strongly paralleled the carbon isotope excursion (Gill et
al., 2011). The positive excursion in δ34SCAS was most likely due to enhanced
turnover of sulfate via BSR (Kampschulte and Strauss, 2004). The 32S enhanced
organic pyrite sulfur was preserved beneath water columns that were both anoxic
and euxinic (containing free hydrogen sulfide). These conditions led to a parallel
sequestration of carbon and sulfur in organic matter and pyrite, since the lack of
oxygen prevented the oxidation and removal from the sediment of this organic
material (Gill et al. 2011).

A negative δ13Ccarb event can be associated with changes in the amount carbon
sequestered as organic or inorganic precipitate. Specifically, a sharp drop in primary
productivity and/or eustatic changes (which tend to be expressed over longer
intervals) can lead to a negative excursion in δ13Ccarb (Holser, 1997; Kump and

31
Arthur, 1999). Alternatively, negative δ13C excursions have been interpreted as
resulting from a pCO2 pulse or an episode of methane release from gas hydrate
reserves (Álvaro, 2008), which can lead to anoxic ocean events. This is supported by
evidence from Australia and New Zealand that points towards a large degassing of
CO2 directly before the onset of a large negative isotopic excursion known
commonly as the Drumian Carbon Isotope Excursion, or DICE (Álvaro, 2008).

Another negative δ13Ccarb (approximately 7‰) excursion takes place over the
Ediacaran–Cambrian boundary, and stable sulfur isotope studies have been
conducted in order to help elucidate the paleoenvironmental causes of this
excursion. The excursion is known commonly as the Ara anomaly, and contains
evidence for ocean anoxia (Figure 11). Over this excursion both δ34SCAS and δ34Spyrite
are enriched by approximately 20‰ (Fike, 2007). This enrichment is thought have
been brought about by an increase in the sequestration of pyrite, due to ocean
anoxia and the subsequent lack of reoxidation (Schröder and Grotzinger, 2007), as
well as possible high sedimentation rate (Fike, 2007). Primary production also
increased over this interval, as supported by carbon isotope evidence for enhanced
organic carbon burial over the latest Ediacaran to earliest Cambrian (Saylor et al.,
1998).

Sea–level variation has been suggested as a significant control on the rate of the
oxidation of organic matter in the ocean, which in turn controls both δ13C and δ34S
variation in the ocean. Increased organic matter oxidation, due to a drop in sea–
level, leads to a drop in both these values as the result of increased exposure of
continental shelves and the development of a shallower ocean (Gao and Land,
1991). A sea–level rise should lead to an increase in these values as organic
oxidation drops, and 12C and 32S are preferentially sequestered in the ocean floor by
organisms and BSR respectively.

Unfortunately, sea–level change can also be associated with the exact opposite
effects on stable isotope ratios in the ocean. A sea–level drop will increase the
weathering of continental material, increasing the influx of isotopically heavy

32
33
material to the sea (the fractionation associated with weathering). This will increase
δ13C and δ34S of the ocean (Saltzman et al., 2004). This is an example of the difficulty
found in drawing conclusions from stable isotope data.

34
35
Study Area

The study area is located in northern Wyoming and southwestern Montana. The
westernmost exposure of the Snowy Range Formation is in Three Forks, Montana.
Further west and north, age–equivalent strata consist of a dark shale and reddish
purple limestone, mapped as the Red Lion Formation. In northern Wyoming, the
easternmost exposures of the formation are in the Powder River Basin, beyond
which rocks of this age are mapped as the Deadwood Formation (Grant, 1965). My
two study areas are in the Horseshoe Hills of southwestern Montana, and the
Beartooth Mountain Range, which spans the border between northwestern
Wyoming and southwestern Montana (Figure 12).

Facies Descriptions:

Shale/Marl Facies:

This facies consists of grey and grey–green shale, and grey, grey–green, and
grey–orange marl. Bed thickness ranges from millimeter scale upwards to 42 cm to
65 cm, or potentially larger under covered sections, with average bed thickness ~12
cm. This facies is predominantly shale at the Wyoming Creek locality, whereas at
Nixon Gulch it is dominantly marl (~75%). This facies is frequently associated with
thin interbeds of the grainstone and flat–pebble conglomerate facies. Locally it
contains abundant oval to spherical micritic nodules, which are most likely
diagenetic (Plate 2. A).

Grainstone Facies:

This facies is light to dark grey, white–weathering carbonate grainstone. Grain


sizes range from very fine to coarse, with thicker beds containing generally coarser
grains, whereas thinner beds, which are often interbedded with shale, are finer
grained (Plate 3. A). Bed thicknesses range from millimeter scale to 28 cm, with an
average of 4 cm. In general, this facies is interbedded with the shale facies, and the
beds are tabular and laterally continuous. Where there is no shale interbedding,

36
37
38
grainstone beds are vertically amalgamated, with scoured basal contacts. These
beds show pinch–and–swell geometries, wavy lower contacts, are truncated
laterally in places, and a few beds in the upper Nixon Gulch section have stylolitic
contacts (Plate 3. A).

Flat–pebble Conglomerate Facies:

This facies is made up of light to dark grey flat–pebble conglomerate with


grainstone matrix. The clasts are almost identical in lithology to the grainstone
facies, and the matrix is generally coarse grained. The clasts themselves are tabular,
rounded, and composed of fine to medium grainstone. Clasts have a large range in
size, from pebbles (4 mm) to cobbles (150 mm) in length, but are only 4–20 mm
thick. The smaller clasts are generally present in beds with abundant grainstone
matrix, whereas the larger clasts are present in beds with obvious framework
texture and subordinate matrix of grainstone. The clasts are generally oriented sub–
horizontally, although the larger clasts in the thicker beds tend to be less well
oriented than the smaller clasts in thinner beds, which lie much closer to horizontal
(Plate 2. B). Bed thicknesses range from 5 to 46 cm, with an average of ~15 cm. Flat–
pebble conglomerate beds are generally tabular and laterally continuous for as far
as the outcrops extend. There are a number of flat–pebble beds that pinch and swell
dramatically. Irregular buildups of the thrombolite facies overlie some flat–pebble
beds (Plate 4, 5).

Thrombolite Facies:

The thrombolite facies is composed of medium gray micrite, with a slightly pink
to blue–white, rubbly weathering pattern. A conspicuous marker bed at both
localities consists of tall columns of large Epiphyton, or Collenia magna (Plate 6. B)
(Grant, 1965). The columns are elongate in paleodirection ~N17°E, with an average
minimum diameter of 26 cm, and an average maximum diameter of 56 cm. The
columns are spaced 8 to 16 cm apart, and are infilled locally with the flat–pebble
conglomerate and grainstone facies (Plate 6. C). The maximum measured height is

39
40
41
1.45 m for an individual column, while the maximum bed thickness is ~6.5 m. At
Nixon Gulch the thrombolite columns in the upper 1.5 m of the main bed are much
narrower than those below (Plate 6. D), averaging just 17 cm minimum diameter.

This facies also shows up as much smaller, laterally discontinuous beds, often
overlying and surrounded by the flat–pebble conglomerate facies (Plate 4, 5). These
thrombolite buildups range from 10 to 46 cm in height, with an average height of
~24 cm. This thrombolite forms clotted mounds, not well–defined, laterally
continuous beds like the marker bed at the base of the sections. The smaller isolated
mounds are typical of this facies in younger parts of the formation (Myrow et al.,
2004). In one case, thrombolite mounds appear to have grown on the side of a
paleochannel filled with flat–pebble conglomerate (Plate 5).

Facies Associations:

The two sections at Nixon Gulch and Wyoming Creek contain several facies
associations. The most abundant facies association at these two sites, a combination
of all five facies, makes up anywhere from 30% to 70% of the sections. It contains
fine to coarse grainstone and flat–pebble conglomerate often overlain by, or
inclusive of clasts of the thrombolite facies (Plate 4, 5), which in turn is generally
overlain by shale and/or marl interbedded with early–diagenetic micritic nodules
(Plate 2. A). About 50% of this facies association is made up of the shale and marl
facies. Generally this association makes up units 0.5 to 2 m thick. Myrow et al.
(2004) also noted that “most of the Snowy Range Formation (~90%) consists of this
mixed facies association.”

The second facies association consists of thinly bedded (millimeter to centimeter


scale) grainstone beds, slightly thicker (average bed thickness: ~12 cm) flat–pebble
conglomerate beds, and a minor component of shale (<10%). This facies association
makes up 10% to 25% of the sections. This association is present throughout the
sections in fairly thin units, generally no more than about 30 cm thick. At the top of
both sections, however, there is a thicker unit (2.5 m at Nixon Gulch, and 1 m at

42
43
Wyoming Creek, although it may be thicker, as we stopped our stratigraphic record
at an overhang above which the section continued.

The third facies association consists of a mix of the grainstone and shale/marl
facies, and it makes up no more than 20% of the sections. It is composed of thin beds
(millimeter to centimeter scale) of alternating mudstone to fine grainstone
carbonate beds, and shale/marl beds with micritic limestone nodules. These
alternating facies exist in generally thin beds at the base of the sections, and the
thickness of this association tends to increase up section at both localities. There is a
3 m thick section near the top of the Nixon Gulch section.

Outcrop Locations and Descriptions:

Nixon Gulch:

The Horseshoe Hills are located approximately 20 miles west of Bozeman,


Montana along Interstate 90, and 10 miles north of the highway. The Nixon Gulch
locality is found at the southernmost tip of the Horseshoe Hills, just outside of
Manhattan, MT, on the north side of the Gallatin River. The location is reached by
driving on Nixon Gulch Road north out of Manhattan. It crosses the Gallatin River,
and branches into Nixon Gulch road on the right and Horseshoe Gulch Road on the
left. The gulch with the outcrop is approximately 2.5 miles beyond the fork, on
Horseshoe Gulch Road (45°55’04”N, 111°21’25”W). Beds strike 280°NW and dip
43°NE.

The Sage Member of the Snowy Range Formation is well exposed on the east side
of Horseshoe Gulch Road (Plate 6. A). The basal biohermal microbialites are well
exposed at the bottom of the section, and the overlying Grove Creek Member, also of
the Snowy Range Formation is present towards the top of the section. The Devonian
Jefferson Formation forms a resistant–weathering dolomitic cliff that
disconformably rests above the Cambrian strata. Siliciclastic and carbonate
lithologies were described and detailed bed–by–bed measurements were taken

44
45
through 42.84 m of section. The contact between the Sage and the Grove Creek
members is hidden beneath the 14.36 m of cover that starts about halfway up this
section. I am only concerned with the Sage member at this locality.

Sage Member (Total thickness: 17.88 m – 32.24 m; Figure 13, 14)

The first 5.53 m of this section are made up of micritic limestone thrombolite
columns filled in with fine to coarse grainstone and local flat–pebble conglomerate
(Plate 6. B, C, D). This basal unit of biohermal microbialite is topped by a thin, 3 cm
thick micrite hardground (Plate 3. B, C). Thrombolites appear twice more in this
section, within and adjacent to flat–pebble conglomerate beds at 6.92 m and 7.77 m.
Individual mounds within these two beds have a maximum height of 30 cm and 25
cm, respectively.

Overlying, and cutting into the micritic hardground above the thick microbialite
bed is a 15 cm thick flat–pebble conglomerate bed with a scoured base (Plate 5. B,
C). From this bed up to 12.14 m, the section is composed of the first facies
association. There are thin units of the second and third associations, but nothing
very significant. Flat–pebble conglomerate beds are abundant in the first 1.5 m of
the section above the massive microbialite. Beds in this interval average 22 cm in
thickness, and they are interbedded with thin beds of alternating marl, grainstone
(mostly coarse), and a 30 cm thick thrombolite mound. At the top of this interval is a
16 cm thick, thinly bedded grainstone unit with one minor shale bed, which is
capped by a 4 cm thick brachiopod coquina containing abundant broken trilobite
spines (Plate 7. B, C).

Above this lower section of abundant thick flat–pebble conglomerate lies 65 cm


of fossiliferous, fissile grey shale with trilobite and brachiopod fossils. Thick beds of
thinly interbedded grainstone, shale, and nodular carbonate mudstone lie up to
10.15 m in the section. The grainstone beds are up to 84 cm thick and are coarser
towards the base. White–weathering carbonate mudstone beds exist at the top of
this interval, and form very thin (<0.5 cm thick) tabular and nodular beds. Micritic

46
47
48
limestone nodules are also present within a shale bed in this interval. The next
stratigraphic interval (up to 12.14 m) consists of thin grainstone beds and overlying
flat–pebble conglomerate beds, all separated by covered intervals (presumably
shale or carbonate mudstone). The flat–pebble conglomerate beds are 10 to 22 cm
thick.

A covered interval extends to 12.14 m, and is overlain by a 59 cm thick unit of


coarse grainstone (with a thin 4 cm shale bed). Above this, and extending to 15.53 m
is the third facies association; thinly bedded grainstone and marl with abundant
white micritic limestone nodules and minor thin–bedded tabular micritic limestone
units (Plate 7. A). The marl beds are ~25 cm thick on average, and the grainstone
consists of numerous thin beds (<0.5 cm to 3 cm).

From 15.53 m to the top of the section, there are very thin to thin beds (<4 cm)
of marl and nodular to tabular beds of carbonate mudstone, fine grainstone, and
slightly coarser grainstone with scattered flat pebbles. Stylolitic contacts are
present, as well as irregular and wavy top and bottom contacts (Plate 3. A). This unit
makes up the second facies association, as it is composed predominantly of
alternating thin beds of grainstone and flat–pebble conglomerate.

Wyoming Creek:

The Beartooth Mountain Range spans the Montana–Wyoming border, and is


found within the Custer, Gallatin and Shoshone National Forests. The Wyoming
Creek locality is in the middle of the Gallatin National Forest, just south of Silvergate
Montana. The locality is just several hundred meters south of the Wyoming–
Montana border. The outcrop is found by heading south out of Silver Gate (located
along Highway 212), and heading left across a bridge over Soda Butte Creek, parallel
to the cliff–forming Pilgrim Limestone Formation to the right. Approximately a
quarter mile beyond there is a turn off on the right into a small parking area just
before a second bridge over Soda Butte Creek. Head south towards the cliff, and go
east for approximately 150 m, crossing the river just under the waterfall cutting

49
50
through the Pilgrim Limestone Formation cliff. Head up the cliff through the dry
gully and turn right, back towards the waterfalls, now on top of the cliff. Soda Butte
Creek cuts through the Wyoming Creek section before it forms the waterfall. On the
east side of the creek, up a small, very steep tributary is the lower (approximately
14 m) section of our outcrop (45°00’11”N, 109°58’49”W). Approximately 15 m up
the stream from this tributary is the upper section of the outcrop, in another steep
tributary on the west side of the creek. This section continues from approximately
14 m up to the top of our measured section at about 21 m. Beds strike 170°SE and
dip 15°SW.

Wyoming Creek flows through the Snowy Range Formation (Plate 8. A), and
forms a waterfall that flows over and down the underlying, cliff–forming Pilgrim
Limestone. The basal biohermal microbialites of the Sage Member of the Snowy
Range Formation are well exposed at the bottom of this section, on the east side of
Wyoming Creek (Plate 8. A), overlying 1.46 m of the Dry Creek Member. The lower
part of this section is well exposed on the east side of the creek up to 14.4 m,
although there is some cover (Plate 8. B). The section continues on the west bank,
south and upstream (Plate 8. C, D). A Jacob’s Staff and Brunton were used to
measure 4.5 m of cover by matching up a marker bed at 8.9 m on the west bank
(Plate 8. C). The section was well exposed on the west bank from 14.4 m beyond the
top at 20.86 m. Overlying strata are exposed above the section, but further detailed
lithology descriptions and bed–by–bed measurements were unnecessary for this
investigation, as the biomere boundary is much lower in the section.

Sage Member (Total thickness: >20.86 m; Figure 13, 15)

The basal Collenia magna thrombolite at Wyoming Creek has its base at 1.46 m
above the base of the section. The first 1.46 m of this section is composed of the
uppermost unit of the underlying Dry Creek Member, which is comprised of beds of
intraclast conglomerate, lime mudstone and shale with micritic nodules. There is a
microbialite bed at 0.9 m that has a maximum thickness of 10 cm. It rests directly on

51
52
underlying flat–pebble conglomerate. There is a 20 cm thick unit of intraclast
conglomerate beds at 0.35 m. The clasts are not flat pebbles of grainstone, but
instead consist of micritic nodules that were presumably eroded from underlying
shale beds that contain abundant nodules. This lowermost unit resembles the first
and third facies associations. The Collenia magna thrombolite bed is 6.47 m thick,
white weathering, and extends to 7.93 m.

Directly overlying the thrombolite bed is a single, 15 cm thick, very coarse


grainstone, which is in turn overlain by a 21 cm thick flat–pebble conglomerate with
abundant grainstone matrix. Above this flat–pebble bed, are strata of the third facies
association, which consist of nodular micritic limestone and marl, with minor beds
of coarse grainstone (1–7 cm thick) and shale with micritic nodules. A covered unit,
presumably shale, leads up to a flat–pebble rich interval at 10.42, and the first
exposure of the first facies association.

The rest of section consists of the first facies association, which includes 10–20
cm thick flat–pebble conglomerate, shale/marl with micrite nodules, and minor
coarse grainstone beds. Thrombolite mounds (average maximum height: 30 cm)
overlie, and are within, several of the flat–pebble conglomerate beds. Flat–pebble
conglomerate beds rest directly on shale at 11.22 m, 12.18 m, 14.4 m (Plate 9), and
in some cases grainstone exists directly above flat–pebble conglomerate beds. Thus,
the relative position of these facies does not define obvious upward coarsening
cycles (shale, grainstone, flat–pebble conglomerate, thrombolite), as recorded in
other Cambrian Inner Detrital Belt strata (Myrow et al., 2004). Covered intervals
show up predominantly directly above flat–pebble conglomerate beds, and are
almost certainly shale, and thrombolites are in most cases associated with flat–
pebble beds. A particularly thick (29–46 cm) flat–pebble conglomerate bed at 17.9
m contains local thrombolite buildups, and has irregular contacts with the beds
above and below (Plate 5).

53
The second facies association is only present at in the last 82 cm of this section.
There are several examples of the third facies association at 8.29 m, 11.79 m and
17.16 m. These units are especially rich in shale, and contain micritic limestone beds
and a few thin (5 cm) medium grainstone beds.

54
Methods

Fieldwork:

During the week of July 12th, 2010, I worked in the Nixon Gulch and Wyoming
Creek areas with Dr. Paul Myrow (Colorado College), Dr. John Taylor (Indiana
University of Pennsylvania) and Jessica Creveling (PhD Student, Harvard University
and Colorado College graduate). Our goal was to measure a detailed stratigraphic
column of the section at the centimeter scale, collect samples for later geochemical
analysis, and collect trilobites for biostratigraphy and correlation.

We took bed–by–bed measurements of the outcrops at Nixon Gulch and


Wyoming Creek. Detailed stratigraphic columns and descriptions were made for
nearly 43 m at Nixon Gulch (although only 17.7 m of the section were analyzed for
stable isotopes due to the stratigraphic location of the biomere) and 20.9 m of
section at Wyoming Creek. Fist sized samples (approximately 300 g) were collected
at 10 to 25 cm intervals throughout the sections. Through several parts of the Nixon
Gulch section, we had to dig through cover to get bed measurements and collect
samples. There were more extensive covered sections at Wyoming Creek, but most
of the section consisted of relatively unweathered beds that yielded good
measurements for stratigraphy and samples.

A total of 181 samples were collected, 101 from Nixon Gulch and 80 from
Wyoming Creek. Samples were sent to Harvard University where I prepared them
for geochemical analysis. Trilobite bearing rocks were collected by Dr. John Taylor,
who cleaned, photographed, and described them.

55
Sample Preparation and Analysis:

Carbon and Oxygen Isotope Preparation and Analysis:

During August, 2010, and again for a week in September, 2010, I prepared and
analyzed samples in Dr. David Johnston’s geochemical laboratory at Harvard
University. Samples were cut, drilled and powdered (Plate 9) for carbon, oxygen and
sulfur isotope analysis. Cutting took place with a water–based diamond–bladed
masonry saw (Plate 9. A). Fresh surfaces were microdrilled using a press–drill with
a dentists drill tip (Plate 9. C). Care was taken to avoid veins, clasts and seams.
Approximately 10 mg of powder was collected, weighed and poured into
microcentrifuge tubes for carbon and oxygen isotope analysis. A VG Optima dual
inlet mass spectrometer attached to a VG isocarb preparation device (Plate 9. D) in
the Harvard University Laboratory for Geochemical Oceanography was used.

Approximately 1 mg of the powdered carbonate rock was reacted in a common


H3PO4 bath at 90° for 6–8 minutes, via an automated 60 carousel port. A total of 54
samples and 6 standards are loaded at a time. The CO2 evolved from dissolution was
purified and concentrated cryogenically with liquid nitrogen as the coolant, and
progressively distilled along a series of valve inlets to the appropriate volume. The
instrument then measured the abundances of the various isotopes. The in–house
standard, CM–2, is corrected to international standard PDB and the samples were
corrected using this correlation in order to find their ‘true’ isotopic values (Maloof,
2005).

Sulfur Isotope Preparation and Analysis:

Approximately 15 cm3 of rock was required for preparation of sulfur samples. In


cases, several pieces of a single sample were cut to avoid secondary veins, clasts and
seams as much as possible. Samples were dried for several hours after cutting.
Certain stratigraphically significant samples were finely powdered using a Spex
SamplePrep 8530 Mill Shatterbox (Plate 9. B). The samples were selected in order to

56
57
have a dense grouping of samples surrounding the biomere boundary at each site,
with a wide spread throughout the rest of each section. This powdered material was
used for carbonate–associated sulfate (CAS) and pyrite analysis, as well as total
organic carbon analysis.

Eighty–three samples were prepared for CAS analysis, although only 25 could
actually be sent off to IsoAnalytical for stable isotope analysis due to monetary
constraints. For CAS preparation, 20 to 60 g (approximately 30 g on average) of each
sample were treated with 10% NaCl solution for 12 hours in order to remove
soluble sulfate (gypsum, etc.) from the powdered rock. This was carefully decanted,
and the rock was treated with deionized water twice for a total of 24 hours. The
samples were then treated with 6.25% NaClO in order to remove any metastable
sulfides or organically bound sulfur. The samples were rinsed twice more with
deionized water before they were dissolved using 4N HCl and let sit for 8 hours. The
remaining sample and liquid was centrifuged, and filtered through a 0.45 µm filter
(Plate 10. C) to remove the insoluble portion of the sample. This solid was saved for
analysis of pyrite. Approximately 50 mL 1N BaCl2 solution was added to the
remaining solution in order to precipitate sulfate as BaSO4, and let sit for three days
to ensure complete precipitation. The BaSO4 powders were separated from the
remaining solution via filtration and allowed to dry before being accurately weighed
and placed in microcentrifuge tubes (Fike and Grotzinger, 2008).

The solid removed during the first filtration for CAS was extracted as chromium–
reducible sulfur (Plate 10. A, B) for stable isotope analysis of pyrite. Twenty–eight
samples were processed, although only 25 could be sent to IsoAnalytical due to
monetary constraints. Each sample was first diluted with deionized water and
decanted several times in order to increase its pH. The cleaned samples were then
placed in flasks (Plate 10. A. 2), and 20 mL each of 6N HCl and 0.4N chromium
chloride (Plate 10. A. 4) were added to this. The flasks were filled with nitrogen gas
(Plate 10. A. 3), which carried the gaseous SO42– to test tubes via a fractionation
column (Plate 10. A. 5). The test tubes were filled to about an inch from the top with

58
59
zinc acetate (Plate 10. A. 6), and the nitrogen gas was added to the flasks at a rate of
two bubbles per second. The flasks were let sit on heating pads (Plate 10. A. 1) for
2+ hours in order for all of the pyrite to volatize and reach the collecting test tube.
After completing the reaction, 500µL 0.3N AgNO3 were added to the solution to
precipitate AgSO4 (Canfield et al., 1986; Strauss, 1989). The precipitate was then
centrifuged and removed as a plug from the bottom of the tube and then stored in a
microcentrifuge tube.

Twenty–five CAS and 25 pyrite samples were sent to IsoAnalytical in Cheshire,


UK for analysis using Elemental Analysis Isotope Ratio Mass Spectrometry (EA–
IRMS). Tin capsules containing reference or sample material plus vanadium
pentoxide catalyst were loaded into an automatic sampler. They were then dropped,
in sequence, into a furnace held at 1080°C and combusted in the presence of oxygen.
Tin capsules flash combust, raising the temperature in the region of the sample to
about 1700°C. the combusted gases are then swept in a helium stream over
combustion catalysts (tungstic oxide/zirconium oxide) and through a reduction
stage of high purity copper wires to produce SO2, N2, CO2 and water. The water was
removed using a Naflon™ membrane. Sulfur dioxide was resolved from N2 and CO2
on a packed GC column at a temperature of 32°C. The resultant SO2 peak entered the
ion source of the IRMS whereupon it was ionized and accelerated. Gas species of
different mass were separated in a magnetic field then simultaneously measured on
a Faraday cup universal collector array. Analysis was based on monitoring of m/z
48, 49 and 50 of SO+ produced from SO2 in the ion source. IA–R061, IA–R025
(barium sulfate) and IA–R026 (silver sulfide) were used for calibration and
correction of the sulfur (IA–R061) and oxygen (all three) contributions to the SO+
ion beam. These are in–house standards, and the samples were then calibrated to
the international comparison standards NBS–127 (barium sulfate) and IAEA–S–1
(silver sulfide) (re. IsoAnalytical; Ian Begley, 2011).

60
Results

Biostratigraphy:

At the both of the outcrops studied in this thesis, the Irvingella major subzone is
either absent or very poorly represented. Although unfortunate, this is not unusual
for the Snowy Range Formation. Saltzman et al. (1999) did not recognize the
subzone at many of his sites, and thus used Grant’s (1965) estimation that it falls a
meter above the thrombolite reef facies described earlier at the outcrops in
northern Wyoming and southern Montana. Grant only found the subzone in three of
his 27 Snowy Range Formation sites, and in these cases it was only 1 to 20 cm thick
and existed in stratigraphic lenses. Dr. John Taylor (personal communication, 2011)
was able to find samples of pre– and post–extinction boundary trilobites at fairly
close stratigraphic levels, placing far greater constrain on the position of the
biomere boundary.

At Nixon Gulch, Dr. John Taylor identified the highest collection of pre–extinction
fauna at 6.87 m, consisting of the underlying Cliffia lataegenae subzone (Plate 1. A,
B). The base of the Eoorthis coquina that sits at the base of the Ptychaspid Biomere
was found at 7.04 m in the section (Plate 1. C. 2–6). Brachiopods comprise the
majority of the coquina and the replacement fauna directly above the critical
interval. Although no trilobites from the extinction event were found, the under–
and over–lying species constrain the extinction event to less than 17 cm of strata,
which corresponds perfectly to Grant’s (1965) estimation of the extent of the critical
interval.

At Wyoming Creek the Irvingella major Subzone, the critical interval, is present,
but its extent is hard to constrain. A species found in the Pterocephaliid Biomere
occurs at 7.95 m up section. A collection at 8.4 m contains a large number of Dellea
cranidia, another pre–extinction fauna. Diverse pre–extinction fauna of the Cliffia
lataegenae Subzone (Loch and Taylor) exist at 9.11 and 9.29 m, indicating that these
levels are below the critical interval. At 11.2 m, Dr. John Taylor believes he found a

61
Dellea cranidia, meaning that this level is also likely below the critical interval (Plate
1. A, B). This is inconsistent with Grant’s (1965) findings, which claimed that the
Irvingella major Subzone exists at 10.64 m. At 11.95, a small cranidia from the
critical interval was uncovered (Plate 1. C. 1), and at 12.01 m, Taylor found trilobites
of the Parabolinoides Subzone, which comprise the replacement fauna at the base of
the Ptychaspid Biomere (Plate 1. C. 2–6). Thus, the biomere boundary is no more
than 81 cm in extent; although it most likely falls in a much narrower range around
11.95 m, judging from Grant’s (1965) results, as well as my own from Nixon Gulch.

Chemostratigraphy:

δ13Ccarb:

There are minor fluctuations of ±0.1 to 0.2‰ in δ13C from both locales. These
oscillations are extremely short, rarely continuing for more than 20 to 30 cm. At
Nixon Gulch (Figure 16) δ13Ccarb is fairly uniform, dropping slightly from around
0.7‰ at the base of the outcrop to 0.1‰ at 18 m at the top. Over the biohermal
thrombolite, δ13Ccarb stays high, at around 0.7 to 0.8‰. At the top of the bioherm is a
sharp (less than 20 cm) drop to 0.4‰, followed by a further drop to the most
negative value of 0.1‰. Following the negative excursion there is a two–stage
increase up to 0.8‰, and eventually (at about 13 m) to around 1.0‰. This is
followed by a gradual decrease towards the top of the section.

The oscillations from one data point to the next at Wyoming Creek (Figure 17)
are slightly greater than those at Nixon Gulch. Also, the overall pattern for the
δ13Ccarb shows a fairly steady increase up section from about 0.1‰ at the base to 0.9
to 1.0‰ at 21 m at the top. A sharp, steady increase in δ13Ccarb occurs over the first
1.5 m of the section, up to the base of the biohermal thrombolite. Values stay stable
over the bioherm, with small, 0.1 to 0.2‰ oscillations around 0.7‰. At the top of
the bioherm, at 8 m, there is a very large drop to just under 0.4‰. This is followed
by a slight increase and a second drop in δ13Ccarb just below the biomere boundary.

62
63
64
This maximum negative excursion is represented by only one point, and it falls
anywhere from 30 cm to 1.2 m below the critical extinction interval, as the bounding
biostratigraphy at this site is less specific than at Nixon Gulch. The maximum
negative excursion drops to 0.2‰, and following this excursion there is a sharp
(over 2.5 m) increase to a maximum positive excursion at 1.0‰. The curve then
oscillates by 0.2 to 0.4‰ around 0.8‰ to the top of the section.

Despite the large amount of noise in both curves, the patterns in both are
strongly supported by the other (Figure 18). There is a stable high over the
bioherms, followed by a two–stage drop to a maximum negative excursion at, or
near the biomere boundary. The discrepancy here will be discussed below, although
it is unlikely that a strong conclusion can be drawn as there is too little data and not
enough sites were sampled. Above the negative excursion both sections have a large
increase in δ13Ccarb to the highest value. The fact that the curve at Nixon Gulch seems
to decrease up section, while that at Wyoming Creek appears to increase will also be
discussed below.

δ 34SCAS, δ 34Spyrite, and pyrite mass yield:

There isn’t as much sulfur data as there is carbon data, although sampling
frequency is still relatively high close to the boundary. The δ34SCAS values (Figures
16, 17) at both sites are constant around 25‰, with variations of up to 5‰. There
is a negative excursion in δ34S at both sites near the biomere boundary. At Nixon
Gulch, the excursion is 4‰. It is possible however that there is a much larger
excursion at the boundary and the data only captures the start of this decrease since
the next data point falls nearly a meter above the boundary. At Wyoming Creek, the
negative δ34S excursion is much larger, a nearly 20‰ drop to 5‰.

The δ34Spyrite curves are difficult to draw any patterns from, although a data point
from Nixon Gulch falls directly within the biomere boundary, and appears to be a
local maximum at about 12‰ (Figure 16). Similarly, there is an increase to a
maximum across the boundary at Wyoming Creek (Figure 17), although it appears

65
66
that a local minimum corresponds more closely with the δ13Ccarb negative excursion.
There are few patterns to draw from these two curves, other than an increase in
values near or across the critical interval. Additionally, the average values for these
two datasets are highly different. At Nixon Gulch, the δ 34Spyrite values fluctuate from
–20 to –4‰, while ay Wyoming Creek they sit between about –5 and 10‰.
Interestingly, the data for mass pyrite yields (Figure 19) shows a huge difference in
magnitude between the two sites. At Nixon Gulch, maximum yield is 3.25%, while
Wyoming Creek has a maximum pyrite yield of around 15%. However, both sites
show a prominent peak in mass yield at the biomere boundary, from 0.025 to
0.325% at Nixon Gulch, and ~1 to 4% at Wyoming Creek.

67
68
Discussion

Sedimentological Interpretation

The base of each section is marked by the thick biohermal Collenia magna
microbialite facies. These microbialites generally grew in clear water where there is
enough light to support them Myrow (2004). This would require low levels of
siliciclastic sediment input, and therefore imply a highstand. Furthermore, there is
no evidence of exposure in this or any other facies, and inferred vertical relief of
these microbial buildups restricts its environment to the subtidal. Finally, the
horizontal extent of this bed (50 km wide, 250 km along strike, approximately
parallel to shore; Saltzman, 1999) would require a very wide and uniform
depositional setting. Elrick and Snider (2002) and Brett et al. (2009) both support
an interpretation of widespread microbialites as having colonized the seafloor
during maximum flooding periods. The presence of flat–pebble conglomerate infills
between the columns, and the elongation of the columns ~N17E, both suggest the
presence of currents. The former suggest deposition above storm wave base, as the
flat pebbles are interpreted to result from storm reworking of cemented sea floor.
The fact that the columns are thinner up section would suggest that there was an
environmental forcing that caused the bioherm to die off. This could have been
brought about by a change in relative sea–level, ocean temperatures, siliciclastic
input and increased turbidity, or any of a number of other causes.

The shale facies, which often overlies the mixed thrombolite flat–pebble
conglomerate facies, was deposited in a subtidal environment. There is no evidence
of exposure, such as desiccation cracks. This shale is commonly calcareous and
grades into marl. This implies that although the siliciclastic input must have been
high during periods of mud deposition, carbonate deposition continued. Also, much
of this facies contains a high percentage of micritic limestone nodules. These
nodules, which locally grade into continuous pinch–and–swell beds, likely represent
precipitated features formed during early diagenesis in the surrounding mud, as

69
opposed to boudinage of lime mudstone beds during compaction. This precipitation
likely occurred in the upper ~10–20 cm of the sediment column, within
supersaturated pore water.

The grainstone facies is the most volumetrically abundant facies in these


sections. There is no evidence of cross–stratification, other than some possible
upper–flow–regime planar laminated beds. Ripple– and even dune–scale cross–
stratification likely exists, but is difficult to discern in this facies due to the
homogeneous grain size and weathering pattern. The grainstone is largely pelloidal,
but also contains bioclasts. An absence of ooids, and other facies typical of the
carbonate platform, means that this grainstone was probably produced within the
Inner Detrital Belt itself, as opposed to being transported from middle carbonate
belt. The third facies association was almost certainly deposited in a turbid
environment, given the relatively thick marl and shale beds. The grainstone facies
was deposited across a wide area, most likely during storms, which winnowed any
fine–grained siliciclastic sediment. The beds of alternating grainstone and
grainstone intraclast conglomerate were deposited in a higher energy setting than
would allow for the settling out of such fine siliciclastic material.

The flat–pebble conglomerate is made up of thin, tabular intraclasts of the


grainstone facies. This was caused by early lithification and reworking in 10s of
meters water depth (Runkel et al., 2008). However, these beds must have been
formed above storm weather wave base, since the energy required to rework the
partially lithified carbonate rocks must have been high. Grainstone beds, deposited
between large storm events, were lithified and reworked during subsequent storms.
These beds were likely formed relatively in situ, as opposed to being long–travelled
gravity flows, since the flat, tabular clasts would have created considerable
interparticle friction, and there was little evidence of any significant slope in the
inner–detrital belt, or imbrication of clasts in the flat–pebble–conglomerate facies.
Thus, the beds simply represent local, reworked areas of underlying, partially
lithified grainstone. This would explain the relative abundance of the second facies

70
association, which consists of flat–pebble conglomerate interbedded with
grainstone beds. The presence of thrombolite mound buildups on and within the
flat–pebble conglomerate is due to the microbialites preferential selection of hard
surfaces for colonization.

Depositional History:

We know that these two sections were deposited in predominantly deep, mostly
below fair weather wave base, depositional environments. Small relative sea–level
changes occur throughout the deposition of the Nixon Gulch and Wyoming Creek
sections, but these minor changes are hard to distinguish since the main changes
through out the section take place over a longer time scale than our stratigraphic
sections cover. The biomere boundary at both sites occurs over an interval poor in
the shale facies and rich in carbonate facies, indicating relatively high sea–levels.
However, it is hard to draw conclusions regarding large–scale changes in sea–level
and depositonal environment from only two, comparatively short sections.

Studies have been conducted on age equivalent, depositionally similar


environments around America that conclusively show that the Pterocephaliid–
Ptychaspid Biomere boundary falls near, if not on, the maximum flooding interval of
a transgressive systems tract. According to Runkel et al. (2007), the biomere
boundary studied in this thesis falls directly at a maximum flooding surface between
a transgressive system tract and a highstand system tract. Therefore, the boundary
falls at the end of a large sea–level rise, possibly one of the largest for several million
years on either side (Figure 20). This sea–level rise, with its associated rise in the
thermocline and various chemoclines, is one of the most widely regarded and
accepted hypotheses for the cause of extinction. The stratigraphy at both of the
outcrops supports a sea–level maximum at or near the biomere boundary. Both
outcrops were deposited under several tens of meters of water, but at the
boundaries there are comparatively thick grainstones present, suggesting a
relatively high sea–level, and thus a strong carbonate factory (Myrow et al., in
review).

71
72
Geochemical Interpretation:

Diagenesis:

In order to interpret stable isotope data, it is necessary to first determine if


primary information is preserved. The processes that sequester sediments, lithify
them, and allow us to study ancient isotope ratios can unfortunately also alter these
ratios. The outcrops that were examined in this study didn’t appear to have been put
under any great pressure or temperature, as the rocks were unaltered from their
original sedimentary forms. However, metamorphosis is not the only source of
alteration of the isotopic composition of sedimentary rocks. The main diagenetic
processes capable of causing changes in the isotope composition of marine
carbonates include neomorphism and recrystallization, and more importantly
reequilibration with diagenetic fluids of differing isotopic composition and the
precipitation of isotopically different diagenetic carbonate phases (Kaufman et al.,
1991; Marshall, 1992). Another facies specific option for post–depositional
alteration in the flat–pebble conglomerate beds is the transportation of the pebbles
out of their stratigraphic and geographic locale. The sample preparation methods
used in this study would not differentiate between the cement holding these pebbles
together and the pebbles themselves. However, the clasts are not thought to move
very far from their original depositional site, as they are simply reworked by storms.
Furthermore, these clasts would not be lain on top of any younger stratigraphy and
it is unlikely that they could be dug any significant amount into underlying, lithified
sediments.

Since the only reasonably significant sources of diagenetic isotope ratio


alteration for the Nixon Gulch and Wyoming Creek sections come from pore water,
the focus here will be to examine and discredit sources of alteration tied to this.
Resetting δ13C and δ34S values in existing carbonate phases is hindered by the low
concentrations of carbon and sulfur in the diagenetic fluids. Furthermore, the
isotopic composition of porewater is determined by the composition of the

73
dissolving carbonate. Therefore, carbonate carbon and sulfur isotopic ratios can
only be modified in very open systems, or in the presence of brines with very
elevated levels of dissolved carbon and sulfur (Banner and Hanson, 1990).
Joachimski and Buggisch (1993) have also shown the preservation of trends in δ13C
from widely separated Upper Cambrian stratigraphic sections, supporting a low
amount of post–depositional alteration in our section.

Narbonne et al. (1994) suggests that the best way to determine whether
diagenetic alteration has had any significant effect on a section is to compare it to
other sections that have been shown by independent means to be of comparable
age. As we will see for our sections, they show good correlation, ruling out any
significant diagenesis. Furthermore, if there was diagenesis, it appears to have
altered the rock as a cohesive unit, not preferentially changing an isolated
stratigraphic section. Buggisch (2003) states that good continuation of isotope
ratios through a section containing varying lithologies is a good sign that primary
signals are being recorded. Nixon Gulch and Wyoming Creek are both comprised of
a large range of lithofacies, through which the isotopic data remains fairly stable.

The sharp drop above about 16 m at Nixon Gulch may be due to diagenetic
alteration and weathering, as the values for this section of Wyoming Creek are very
different, and the sedimentary rocks here appeared to have undergone intensive
weathering. This conclusion gains credibility when the good correlation in carbon
isotope curves between the sites is observed, especially across the bioherm, the
biomere boundary, and within the immediately surrounding stratigraphy.
Furthermore, there is good correlation between the carbonate carbon and CAS
curves, which would signify that the isotopic signature of the rock has not been
altered by post–depositional processes. Additionally, the coupled drop in carbonate

74
75
carbon and CAS isotope ratios at the biomere boundary adds confidence to the data
points.

Diaganetic alteration can also be determined by the strength of any sort of


covariance between carbonate δ13C and δ18O (Glumac, 2007). Covariance at Nixon
Gulch is 0.20, and 0.13 at Wyoming Creek. When the lack of covariance between
carbonate δ13C and δ18O (Figure 21) is coupled with the strong correlation between
the two sites, we can see that there has been little alteration of the isotope ratios of
the two outcrops.

Stable Isotope Geochemistry and Extinction Mechanisms:

There have been numerous proposed mechanisms for the biomere boundary
extinctions within the paleontological community. These range from changes in
solar heating to sudden cooling, coupled with marine regressions and
transgressions. The high resolution carbon and previously unused sulfur stable
isotope data that I have collected across the Pterocephaliid–Ptychaspid Biomere
boundary will help to elucidate which of the suggested mechanisms are probable,
and which are not.

The general trend of increasing δ13C up section at Wyoming Creek follows well
the regional and global trends for Upper Cambrian rocks (Figure 22) (Perfetta et al.,
1999). Saltzman et al. (1995) claims that this general increase represents a rise in
sea–level, oceanic overturn, and spread of anoxic waters across the craton. This
would lead to a gradual increase in δ13C values because of the increase in primary
productivity and the increased burial of organic material due to the expanded
oxygen minimum zone. This is supported by a similar increase across the end–
Cambrian extinctions, with their associated rises in sea–level (Westrop and
Ludvigsen, 1987; Ripperdan and Miller, 1995). The low resolution of these studies,
however, prevents a good look at what is occurring directly at the critical interval.

76
77
The high resolution of my research allows us to take a closer look, and see that there
is in fact a significant drop in δ13C and δ34SCAS values at, or very near to, the
extinction event. This drop occurs during the general trend of increasing isotope
ratios.

The rise in sea–level found by Runkel et al. (2007) is reflected in the


chemostratigraphic data across the critical interval for Nixon Gulch and Wyoming
Creek. When sea–level rises, it has been hypothesized that the thermocline and
chemocline rise up with it (Westrop and Ludvigsen, 1987). In this case, deep, cold
water floods onto the shelf, along with possibly anoxic (Saltzman et al., 1995), and
even euxinic waters (Gill et al., 2011). Indeed, Upper Cambrian oceans were more
susceptible to episodes of anoxia (Hurtgen et al., 2009) due to increased seawater
temperatures and thus decreased O2 solubility, as well as the extensive black shale
deposits found from the time (Berry and Wilde, 1978). These organic–rich, pyritic
black shales were common in the Upper Cambrian, evidence of ocean stratification
(Gill et al., 2011). The Upper Cambrian oceans were characterized by sluggish global
ocean–current circulation that favored stratification and anoxic deep waters
(Weissert, 1989). Sea–level was generally high, and salinity stratification was high,
while terrestrial runoff was low on Laurentia. This may have led to nutrient starved
waters over anoxic waters along the margins of Laurentia (Brasier, 1992).

Some authors suggest that an incursion of cold, deep, 12C enriched seawater
onto the carbonate shelf environment across the Marjumin–Pterocephaliid Biomere
boundary (the trilobite extinction event that occurred prior to the one in this study)
led to a decrease in δ13C of ocean water and an associated negative excursion
(Figure 23) (Patzkowsky et al., 1997; Perfetta et al., 1999; Fike, 2007; Gill et al.,
2011). The enrichment of light carbon in deep water is due to biological preferential
fractionation of light carbon and the subsequent sinking of this material. At depth,
the biological material is remineralized and released back into the water column.
Perfetta et al. (1999) showed that a decrease of 0.5 to 1.0 ‰ occurs across the
Pterocephaliid–Ptychaspid Biomere boundary. It has also been shown across the

78
79
Marjumin–Pterocephaliid Biomere boundary that the magnitude of this negative
shift varies as a function of distance from the paleoshoreline. More distal sections
display larger negative shifts (0.78‰), while landwards sections present smaller
changes (0.41 to 0.48‰).

We can see a strong negative excursion in the carbonate–carbon data (Figure


17), which is a proxy for DIC isotopic ratios in the ocean at the time of deposition.
The values from the Nixon Gulch and Wyoming Creek even correspond to those
found in the House Range across this very biomere boundary (Figure 24) (Saltsman,
1995). The interpretation that the cold, 12C enriched waters had a greater influence
where they first impinged on the shelf, and slowly lost their effect as they moved
progressively towards the shore (Perfetta et al., 1999) points towards a rise in the
thermocline (Stitt, 1975) due to sea–level change. However, the introduction of cold,
deep seawater could have also been due to the destratification of the Cambrian
ocean (Taylor, 1977), although this seems less likely.

The sulfur isotopic records reflect a stratified ocean. Short–term substantial


variations in δ34SCAS are inconsistent with the idea of vertical homogeneity of the
water column (Strauss ,1999). If an ocean is homogenous, then sulfur will change
gradually over time. However, in a stratified ocean, sea–level change can rapidly
alter the δ34SCAS values of the preserved bottom waters. Ocean stratification can also
cause isotopic sulfur gradients to develop between deep and shallow waters
(Bottrell, 2006). Deep water carbonate from the Neoproterozoic recorded positive
δ34SCAS values of 11 to 20‰, while their shallow water equivalents recorded values
of 25 to 45‰ (Chu et al., 2005). This disparity in δ34SCAS between deep and shallow
water is recorded in the negative excursion in stable sulfur isotope ratios across the
extinction boundary at both Nixon Gulch and Wyoming Creek. Especially at
Wyoming Creek, the negative excursion of δ34SCAS, is due to anoxic, isotopically light
water rising onto the shelf from offshore.

80
81
The disparity between the stable isotope excursions and the critical interval
at Wyoming Creek is strange, given that the excursions fall almost perfectly on the
interval at Nixon Gulch. Nonetheless, we can use this disparity as evidence for not
only a stratified ocean, but for the isotopically gradational ocean put forth by Chu et
al. (2005) as well. Myrow and Grotzinger (2000) illustrate the ability of an ocean to
have not only thermo– and chemoclines, but isotopic gradients as well. They work
with sea–level changes and the effect it can have on preserved stable isotope
signatures. Using steadily fluctuating stable isotope curves, they come to the
conclusion that sea–level change can lead to quite variable isotope curves at
different sites in a basin. Applied to the difference between Wyoming Creek and
Nixon Gulch, it is reasonable to assume that they were located far enough apart for
this disparity to be noticeable. As the isotope gradient moves up the shelf, along
with the rising sea–level, it would reach the sites at different times, stamping a
diachronous isotope signature on an isochronous biomere boundary (Öpik, 1966).

The δ34Spyrite curves don’t show very strong correlation, and are also fairly
low resolution. Since it is hard to draw conclusions from this sparse data, it is
extremely useful to couple it with the pyrite mass yield record. When put side by
side, bottom water anoxia is evident as a strong positive excursion in both pyrite
mass yield and δ34Spyrite at the biomere boundary. The lack of oxidation of bacterially
reduced sulfate that is coupled with anoxic water would lead to this strong peak in
pyrite mass yields, as well as the less distinct rise and peak in δ34Spyrite values
(Bottrell, 2006; Grotzinger, 2007). Additionally, Thomas (1993) found a positive
cerium anomaly in McGill, Nevada, coinciding with the Pterocephaliid–Ptychaspid
Biomere boundary. He interpreted this as support for the flooding of oxygen
depleted water onto the platform.

Saltzman et al. (1995) and Gill et al. (2011) further suggested that the cold,
anoxic waters flooding the shelf were coupled with toxic, metal–laden waters.
Although this study did not cover metal content of the ocean during the time of
deposition, the extremely high levels of pyrite, coupled with the low δ34SCAS, point

82
towards possibly euxinic, or sulfur rich (in the form of H2S), waters. These euxinic
waters would have been produced under a chemocline, in anoxic conditions where
BSR could not be counteracted by oxidation. This sulfur would be isotopically light
due to repeated fractionation through BSR (Chu et al., 2005). In these waters H2S
would build–up to toxic levels, and when the sea–level rose, these toxic waters
would flood the platform along with the cold, anoxic waters, heralding in the
extinction event. This flooding onto the platform of sulfur rich water would be
recorded in the pyrite mass yield curve as a spike, while the δ34SCAS records a low
(Bottrell, 2006).

Sea–level rise has other, smaller effects besides the rise in thermo– and
chemoclines. These are generally smaller contributing factors that should be
mentioned, as they may prove pertinent to the extinction event, and the following
expansion. When sea–level rises, it decreases the amount of weathering on land, and
therefore the amount of material that reaches our depositional environment. This
means that fewer nutrients are coming from onshore (Kampschulte, 2004), which
leads in turn to a decrease in biology on the platform. Furthermore, since there is
less weathered cratonic material being washed into the ocean, there is less
isotopically fresh carbon and sulfur, so any changes in water chemistry that occur
are exacerbated by the decreased reservoir (Saltzman, 2004).

There is also a massive upwelling of nutrients with the deep, cold water that
comes onto the platform. This allows a massive expansion of biology on the shelf
following the extinction event (Saltzman et al., 1995; Perfetta et al., 1999). The
expansion of primary producers is recorded in the δ13C curve as a positive excursion
between 3 and 5 m above the extinction associated negative excursion. This positive
increase is due to a large amount of light carbon being sequestered by the recovery
of primary producers, leaving the ocean isotopically heavy. This post extinction
expansion is known as the “Strangelove ocean” effect (Hsü et al., 1985), and is also
partially due to the decreased requirement of these primary producers for 12C
(Saltzman et al., 1995).

83
Although there were numerous sea–level fluctuations during the biomeres, it
is only at some highstands that these mass extinctions occur. Runkel et al. (2007)
points out that there are numerous highstands and transgressions between biomere
boundaries (Figure 17). Biomeres occur at an interval much greater (~5 Ma) than
could be explained by sea–level changes driven by Milankovitch cycles (100 ka, 40
ka, 20 ka; Imbrie et al., 1984). It is possible that the biomeres are controlled by some
sort of periodic sea–level rise that brings water up higher than normal onto the
platform, and the thermo– and chemoclines with it, while at other times it simply
does not get high enough. However, the cyclicity would suggest otherwise. I propose
that the cyclic buildup of anoxic (Gill et al., 2011) and euxinic conditions off shore,
coincides with sea–level rise on a periodic basis, leading to the biomeres and their
bounding extinction events.

84
Conclusion
This research produced a detailed sedimentological, biostratigraphic, and
geochemical analysis of the Pterocephaliid-Ptychaspid Biomere boundary in the
Snowy Range Formation. This formation was deposited in the Inner Detrital Belt of
the carbonate platform on the passive margin of Late Cambrian Laurentia. The
depositional environment consisted of mixed siliciclastic–carbonate facies, as there
was an alternation between input of siliciclastics from the craton and carbonates on
site and from the Middle Carbonate Belt. This interchange of deposition is due to the
small scale fluctuation of sea–level. The critical interval of the biomere boundary fell
directly on a highstand of sea–level (Runkel et al., 2007), and the stratigraphy
around the critical interval shows a predominance of thick grainstone facies,
signifying high sea-levels and thus a distant shore and weak siliciclastic input.
Biostratigraphy for this research was tightly constrained, giving critical intervals
of, at most, 17 cm at Nixon Gulch and 80 cm at Wyoming Creek. This is much higher
resolution biostratigraphy than has been used previously to correlate
chemostratigraphy. Biomeres or composed of a rapid evolution and expansion of
deep water, cryophilic trilobites onto the platform, which eventually adapt to the
warm, shallow water on the platform. After 3 to 5 million years, these trilobites
undergo a massive extinction event, in which all but a few species go extinct.
Following this extinction event, offshore trilobites again invade the platform and
expand in the new environment.
The isotope research conducted for this thesis was composed of carbonate
carbon, carbonate associated sulfur, and pyritic sulfur stable isotope analysis. At
both sites there was a drop of about 0.3 to 0.7‰ in δ13C at or just below the critical
interval. Following the extinction, δ13C showed a marked rise to a maximum value of
about 1.0‰. δ34SCAS paralleled the negative excursion with drops of 4 to 15‰ at
Nixon Gulch and Wyoming Creek respectively. There is also a large spike in pyrite
mass yield from the collected samples at both sites, from 0 to 0.3 at Nixon Gulch, and
5‰ at Wyoming Creek. Finally, the δ34Spyrite data shows a rise and maximum across
the critical interval of about 25‰.

85
The conclusions I drew from the above data are multifaceted, playing into and
supporting each other. The main hypothesis for the cause of these trilobite
extinction events comes from a rise in sea–level coupled with a rise in the
thermocline, and a pulse onto the platform of anoxic and euxinic water. The drop in
stable carbon isotope ratios at the boundary comes from the influx of cold, 12C
enriched water from offshore with the rise in sea-level. The disparity at Wyoming
Creek between the negative excursion and the extinction event is most likely due to
isotopic gradients in the ocean during the Camrbian. However, there were not
enough studied sites in my research to confidently state whether this is the case or
not.
The sulfur data also supports a strongly stratified ocean, since a non–stratified
ocean would not give stable isotope ratio curves with strong, fast fluctuations, as we
see in the data. Also, the negative excursion supports an influx of anoxic, H2S
enriched water. This sulfur would be isotopically lightened by repeated BSR, under
oxygen depleted water, since the sulfide needed to produce pyrite would otherwise
be oxidized back to sulfate. The peak in pyrite mass yield values at the critical
interval along with the rise in δ34Spyrite values signifies an influx of sulfur–enriched
water, with a lack of oxidation to release the bacterially reduced sulfur back into the
water column.
Finally, the peak in δ13C values above the extinction event are representative of
the “Strangelove ocean” effect. This occurs when there is a massive expansion of
primary producers following an extinction event. In this case, nutrients were
brought in from offshore with the upwelling of cold, anoxic, euxinic water. Following
the extinction, these nutrients allowed primary producers to flourish, increasing the
amount of 12C enriched organic carbon being sequestered, thereby increasing the
δ13C value of ocean water at the time.
Further research should be conducted on this and the other biomere boundaries
using the high–resolution techniques presented here. Although my research showed
some extremely interesting and convincing data, there is simply not enough of it to
confidently draw conclusions regarding the Pterocephaliid0Ptychaspid Biomere
boundary. More sites should be studied for better regional correlation, especially

86
regarding an isotopic gradient in the ocean and the subsequent disparity between
isotope excursions and the extinction event. Higher resolution in all of the stable
sulfur isotope data would also provide a better image of what is going on, as the
resolution of this data is not high enough to claim that there are in fact any real
increases or decreases at the biomere boundary. Finally, organic carbon isotope
data would be invaluable to give us a better view into what is going on with the
biology at the time. Despite the success of this study, there is still much more work
to be done on this, and all biomeres in the Upper Cambrian.

87
Works Cited

Ahlberg, P., Axheimer, N., Babcock, L.E., Eriksson, M.E., Schmitz, B., Terfelt, F., 2009.
Cambrian high–resolution biostratigraphy and carbon isotope
Chemostratigraphy in Scania, Sweden: first record of the SPICE and DICE
excursions in Scandinavia: Lethaia 42, 2–16.

Aitken, J. D., 1966. Middle Cambrian to Middle Ordovician cyclic sedimentation,


southern Rocky Mountains of Alberta: Canadian Petroleum Geologists Bulletin, v.
14, p. 405–441.

Álvaro, J. J., Subías, B. B. I., Pierre, C., Vizcaïno, D., 2008. Carbon chemostratigraphy of
the Cambrian–Ordovician transition in a midlatitude mixed platform, Montagne
Noire, France: GSA Bulletin, v. 120, p. 962–975.

Anderson, T. F., Arthur, M. A., 1983. Stable isotopes of oxygen and carbon and their
application to sedimentologic and paleoenvironmental problems: in Arthur, M.A.,
Anderson, T. F., Kaplan, I. R., Veizer, J., Land, L. S. (eds.), Stable Isotopes in
Sedimentary Geology: SEPM, Short Course 10, p. 1–1–1–151.

Ashton, J. H., Rowell, A. J., 1975. Environmental Stability and Species Proliferation in
Late Cambrian Trilobite Faunas: A Test of the Niche–Variation Hypothesis:
Paleobiology, v. 1, no. 2, p. 161–174.

Banner, J. L., Hanson, G. N., 1990. Calculation of simultaneous isotopic and trace
element variations during water–rock interaction with applications to carbonate
diagenesis: Geochimica et Cosmochimica Acta, v. 54, p. 3123–3127.

Beauchamp, B., Oldershaw, A. E., Krouse, H. R., 1987. Upper Carboniferous to Upper
Permian 13C–enriched primary carbonates in the Sverdrup Basin, Canadian
Arctic: Comparison to coeval western North American ocean margins: Chemical
Geology, Isotope Geoscience Section, v. 65, p. 391–413.

88
Berger, W. H., Vincent, E., 1986. Deep–sea carbonates: Reading the carbon–isotope
signal: Geologische Rundschau, v. 75, p. 249–269.

Berner 1990

Berner, R.A., Raiswell, R., 1983. Burial of organic–carbon and pyrite sulfur in
sediments over Phanerozoic time—a new theory: Geochimica et Cosmochimica
Acta, v. 47, 855–862.

Berry, W. B. N., Wilde, P., 1978. Progressive ventilation of the oceans—an


explanation for the distribution of the lower Paleozoic black shales: American
Journal of Science, v. 278, p. 257–275.

Bond, G. C., Nickeson, P. A., Kominz, M. A., 1984. Breakup of a supercontinent


between 625 Ma and 555 Ma: New evidence and implications for continental
histories: Earth and Planetary Science Letters, v. 70, p. 325–345.

Bottrell, S. H., Newton, R. J., 2006. Reconstruction of changes in global sulfur cycling
from marine sulfate isotopes: Earth–Science Reviews, v. 75, p. 59–83.

Bottrell, S. H., Raiswell, R., 2000. Sulphur isotopes and microbial sulphur cycling in
sediments: in Riding, R. E., Awramik, S. M. (eds.), Microbial Sediments, p. 96–104.

Brasier, M. D., 1992. Nutrient–enriched waters and the early skeletal fossil record:
Geological Society of London, v. 149, p. 621–629.

Brasier, M. D., Corfield, R. M., Derry, L. A., Rozanov, A. YU., Zhuravlev, A. YU., 1994.
Multiple δ13C excursions spanning the Cambrian explosion to the Botomian crisis
in Siberia: Geology, v. 22, p. 455–458.

Brett, C.E., Allison, P.A., DeSantis, M.K., Liddell, W.D., Kramer, A., 2009. Sequence
stratigraphy, cyclic facies, and lagerstatten in the Middle Cambrian Wheeler and
Marjum Formations, Great Basin, Utah: Palaeogeography, Palaeoclimatology,
Palaeoecology, v. 277, p. 9–33.

89
Brunner, B., Bernasconi, S. M., 2005. A revised isotope fractionation model for
dissimilatory sulfate reduction in sulfate reducing bacteria: Geochimica et
Cosmochimica Acta, v. 69, p. 4759–4771.

Buggisch, W., 2003. Carbon isotope record of Middle Cambrian to Late Silurian
carbonate and shale of Northeastern Ellsemere Island: Palaeogeography,
Palaeoclimatology, Palaeoecology, v. 195, v. p. 357–373.

Canfield D. E., 2001. Biogeochemistry of sulfur isotopes. Reviews in Mineralogy and


Geochemistry, v. 43, p. 607–636.

Canfield D. E., Raiswell R., Westrich J. T., Reaves C. M. Berner R. A., 1986. The use of
chromium reduction in the analysis of reduced inorganic sulfur in sediments and
shales: Chemical Geology, v. 54, p. 149–155.

Canfield, D. E., Thamdrup, B., 1994. The production of 34S–depleted sulfide during
bacterial disproportionation of elemental sulfur: Science, v. 266, p. 1973–1975.

Chu, X., Zhang, T., Strauss, H., Zhang, Q., Feng, J., 2005. Dynamic ocean chemistry
around the Marinoan glaciation—isotopic evidence from cap carbonates:
Geochimica et Cosmochimica, v. 549.

Detmers, J., Brüchert, V., Habicht, K. S., Kuever, J., 2001. Diversity of sulfur isotope
fractionations by sulfate–reducing prokaryotes: Applied and Environmental
Microbiology, v. 67, p. 888–894.

Elrick, M., and Snider, A.C., 2002. Deep–water stratigraphic cyclicity and carbonate
mud mound development in the Middle Cambrian Marjum Formation, House
Range, Utah, USA: Sedimentology, v. 49, p. 1021–1047.

Emanuel, K., 2005. Increasing destructiveness of tropical cyclones over the past 30
years: Nature, v. 436, p. 686–688.

90
Fike D. A., 2007. Carbon and sulfur isotopic constraints on Ediacaran
biogeochemical processes, Huqf Supergroup, Sultanate of Oman [Ph.D. thesis]:
Boston, Massachusetts Institute of Technology.

Fike, D. A., Grotzinger, J. P., 2008. A paired sulfate–pyrite δ34S approach to


understanding the evolution of the Ediacaran–Cambrian sulfur cycle: Geochimica
et Cosmochimica Acta, v. 72, p. 2636–2648.

Gao, G., and Land, L. S., 1991. Geochemistry of Cambrian–Ordovician Arbuckle


Limestone, Oklahoma: Implications for diagenetic δ18O alteration and secular
δ13C and 87Sr/86Sr variation: Geochimica et Cosmochimica Acta, v. 55, p. 2911–
2920.

Gill, B. C., Lyons, T. W., Saltzman, M. R., 2007. Parallel, high–resolution carbon and
sulfur isotope records of the evolving Paleozoic marine sulfur reservoir,
Palaeogeography, Palaeoclimatology, Palaeoecology, v. 256, p. 156–173.

Gill, B. C., Lyons, T. W., Young, S. A., Kump, L. R., Knoll, A. H., Saltzman, M. R., 2011.
Geochemical evidence for widespread euxinia in the Later Cambrian ocean:
Nature, v. 469, p. 80–83.

Glumac, B., Mutti, L. E., 2007. Late Cambrian (Steptoean) sedimentation and
responses to sea–level change along the northeastern Laurentian margin:
Insights from carbon isotope stratigraphy: GSA Bulletin, v. 119, p. 623–636.

Grant, R. E., 1965. Faunas and stratigraphy of the Snowy Range Formation (Upper
Cambrian) in southwestern Montana and northwestern Wyoming: GSA Memoir,
v. 96, 171 p.

Grossman, E. L., Mii, H. –S., Zhang, C., Yancey, T. E., 1996. Chemical variation in
Pennsylvanian brachiopod shells—diagenetic, taxonomic, microstructural, and
seasonal effects: Journal of Sedimentary Research, v. 66, p. 1011–1022.

Halmer, M. M., Schmincke, H. U., Graf, H. F., 2002. The annual volcanic gas input into

91
the atmosphere, in particular into the stratosphere: a global data set for the past
100 years: Journal of Volcanology and Geothermal Research, v. 115, p. 511–528.

Holser, W. T., Mackenzie, F. T., Maynard, J. B., Schidlowski, M., 1988. Geochemical
cycles of carbon and sulfur: in Gregor, C.B. (ed.), Chemical Cycles in the Evolution
of the Earth, p. 105–173.

Holser, W. T., Magaritz, M., and Wright, J., 1986. Chemical and isotopic variations in
the world ocean during Phanerozoic time: in Walliser, O.H. (ed.), Global Bio–
Events: a critical approach: Lecture Notes in Earth Sciences, v. 8, p. 63–74.

Holser, W.T., 1977. Catastrophic chemical events in earth history of the ocean:
Nature, v. 267, 403–408.

Hsü, K. J., Oberhansli, H., Gao, J. Y., Shu, S., Haihong, C., Krahenbuhl, U., 1985.
“Strange– love ocean” before the Cambrian explosion: Nature, v. 316, p. 809–811.

Hurtgen, M. T., Pruss, S. B., Knoll, A. H., 2009. Evaluating the relationship between
the carbon and sulfur cycles in the later Cambrian ocean: An example from the
Port au Port group, western Newfoundland, Canada: Earth Planet Science
Letters, v. 281, p. 288–297.

Imbrie, J., Hays, J. D., Martinson, D., McIntyre, A., Mix, A., Morley, J., Pisias, N., Prell,
W., Shackleton, N. J., 1984. The orbital theory of Pleistocene climate: support
from a revised chronology of the marine δ18O record: in Berger, A. L. (ed.),
Milankovitch and Climate, Part I, p. 269–305.

Joachimski, M. M., Buggisch, W., 1993. Anoxic events in the late Frasnian—Causes of
the Frasnian–Famennian faunal crisis: Geology, v. 21, p. 675–678.

Jørgensen, B. B., Bang, M., Blackburn, T. H., 1990. Anaerobic mineralization in marine
sediments from the Baltic Sea–North Sea transition: Marine Ecology, v. 59, p. 39–
54.

92
Kah, L. C., Lyons, T. W., Chesley, J. T., 2001. Geochemistry of a 1.2 Ga carbonate–
evaporite succession, northern Baffin and Bylot Islands: implications for
Mesoproterozoic marine evolution. Pre– cambrian Research 111 (1–4), 203–
234.

Kah, L. C., Lyons, T. W., Frank, T. D., 2004. Low marine sulphate and protracted
oxygenation of the Proterozoic biosphere: Nature, v. 431, p. 834–838.

Kampschulte, A., Strauss, H., 2004. The sulfur isotopic evolution of Phanerozoic
seawater based on the analysis of structurally sub– stituted sulfate in
carbonates. Chemical Geology 204 (3–4), 255 – 286.

Kaufman, A. J., Hayes, J. M., Knoll, A. H., Germs, G. J. B., 1991. Isotopic compositions of
carbonates and organic carbon from upper Proterozoic successions in Namibia:
stratigraphic variation and the effects of diagenesis and metamorphism:
Precambrian Research, v. 49, p. 301–327.

Kendall, C., Caldwell, E. A., 1998. Fundamental of Isotope Geochemistry: in Kendall,


C., Mcdonnell, J. J. (eds.), Isotope Tracers in Catchment Hydrology, p. 51–86.

Kroopnick, P. M., 1985. The distribution of 13C of CO2 in the world oceans: Deep–Sea
Research, v. 32, p. 57–84.

Kump L. R. and Arthur M. A., 1999. Interpreting carbon–isotope excursions:


carbonates and organic matter: Chemical Geology, v. 161, p. 181–198.

Levy, M., Christie–Blick, N., 1989. Pre–Mesozoic palinspastic reconstruction of the


eastern Great Basin (western United States): Science, v. 245, p. 1454–1462.

Levy, M., Christie–Blick, N., 1991. Tectonic subsidence of the early Paleozoic passive
continental margin in eastern California and southern Nevada: GSA Bulletin, v.
103, p. 1590–1606.

93
Lochman–Balk, C., 1971. The Cambrian of the craton of the United States: Holland,
Cambrian of the New World, p. 79–168.

Lochman, C., Duncan, D,. 1944. Early Upper Cambrian faunas of central Montana:
GSA Special Papers, v. 54, 181 p.

Lohmann, K. C., Walker, J. C. G., 1989. The δ18O record of Phanerozoic abiotic marine
calcite cements: Geophysical Research Letters, v. 16, p. 319–322.

Ludvigsen, R., Westrop, S. R., 1985. Three new Upper Cambrian stages for North
America: Geology, v. 13, p.139–143.

Lyons, T. W., Kah, L. C., Gellatly, A. M., 2004. The Precambrian sulphur isotope record
of evolving atmospheric oxygen: in Ericksson, P. G., Altermann, W., Nelson, D. R.,
Mueller, W. U., Catuneanu, O. (eds.), The Precambrian Earth: Tempos and Events,
Developments in Precambrian Geology, v. 12, p. 421–440.

Lyons, T. W., Walter, L. M., Gellatly, A. M., Martini, A. M., Blake, R. E., 2004. Sites of
anomalous organic remineralization in the carbonate sediments of South
Florida, USA: the sulfur cycle and carbonate–associated sulfate: in Amend, J. P.,
Edwards, K. J., Lyons, T. W. (eds.), Sulfur Biogeochemistry—Past and Present,
GSA Special Paper, v. 379. p. 161–176.

Mackin, J. E., Swider, K. T., 1989. Organic–matter decomposition pathways and


oxygen–consumption in coastal marine–sediments: Journal of Marine Research,
v. 47, p. 681–716.

Magaritz, M., 1991. Carbon isotopes, time boundaries and evolution: Terra Nova, v.
3, p. 251–256.

Maloof, A. C., Schrag, D. P., Crowley, J. L., Bowring, S. A., 2005. An expanded record of
Early Cambrian carbon cycling from the Anti–Atlas Margin, Morocco: Canadian
Journal of Earth Science, v. 42, p. 2195–2216.

94
Marshall, J. D., 1992. Climatic and oceanographic isotope signals from the carbonate
rock record and their preservation: Geological Magazine, v. 2, p. 143–160.

Mckerrow, W. S., Scotese, C. R., Brasier, M. D., 1992. Early Cambrian continental
reconstructions: Journal of the Geological Society, v. 149, p. 593–599.

Miller, K. G., Kominz, M. A., Browning, J. V., Wright, J. D., Mountain, G. S., Katz, M. E.,
Sugarman, P. J., Cramer, B. S., Christie–Blick, N., Pekar, S. F., 2005. The
Phanerozoic record of Global Sea–Level Change: Science, v. 310, p. 1293–1298.

Monroe, J. S., Wicander, R., 1997. The Changing Earth: Exploring Geology and
Evolution, 2nd ed., p. 533.

Montañez, I. P., Osleger, D. A., Banner, J. L., Mack, L. E., Musgrove, M., 2000. Evolution
of the Sr and C isotope composition of Cambrian Oceans. GSA Today, v. 10, p. 1–
7.

Myrow, P. M., Grotzinger, J. P., 2000. Chemostratigraphic proxy records: Forward


modeling the effects of unconformities, variable sediment accumulation rates,
and sampling–interval bias: Society for Sedimentary Geology, v. 67, p. 43–55.

Myrow, P. M., Taylor, J. F., Miller, J. F., Ethington, R. L., Ripperdan, R. L., Brachle, C. M.,
1999. Stratigraphy, sedimentology, and paleontology of the Cambrian–
Ordovician of Colorado: in Lageson, D.R., et al. (eds.), Colorado and adjacent
areas: GSA Field Guide 1, p. 157–176.

Myrow, P. M., Tice, L., Archuleta, B., Clark, B., Taylor, J. F., Ripperdan, R. L., 2004.
Flat–pebble conglomerate: its multiple origins and relationship to metre–scale
depositional cycles: International Association of Sedimentologists, v. 51, p. 973–
996.

Narbonne, G. M., Kaufman, A. J., Knoll, A. H., 1994. Integrated chemostratigraphy and
biostratigraphy of the upper Windermere Supergroup (Neoproterozoic),
Mackenzie Mountains, northwestern Canada: GSA Bulletin, v. 106, p. 1281–1291.

95
Öpik, A. A., 1966. The early Upper Cambrian crisis and its correlations: Journal and
Proceedings of the Royal Society of New South Wales, v. 100, p. 9–14.

Osleger, D. A., Read, J. F., 1993. Comparative analysis of methods used to define
eustatic variations in outcrop: Late Cambrian interbasinal sequence
development: American Journal of Science, v. 293, p. 157–216.

Palmer, A. R., 1965, Biomere—A new kind of biostratigraphic unit: Journal of


Paleontology, v. 39, p. 149–153.

Palmer, A. R., 1965. Trilobites of the Late Cambrian Pterocephaliid Biomere in the
Great Basin, United States: U.S. Geological Survey Professional Paper, v. 493, p.
1–105.

Palmer, A. R., 1971. The Cambrian of the Great Basin and adjacent areas, western
United States: in Holland, C. H. (ed.), Cambrian of the New World: Interscience, p.
1–78.

Palmer, A. R., 1979. Biomere boundaries reexamined: Alcheringa, v. 3, p. 33–41.

Palmer, A. R., 1981. Subdivision of the Sauk sequence: in Taylor, M. E. (ed.), Short
papers for the Second International Symposium on the Cambrian Sys– tem: U.S.
Geological Survey Open–File Report 81–743, p. 160–162.

Palmer, A. R., 1984. The biomere problem: Evolution of an idea: Journal of


Paleontology, v. 58, p. 599–611.

Palmer, A. R., 1998. A proposed nomenclature for stages and series of the Cambrian
of Laurentia: Canadian Journal of Earth Sciences, v. 35, p. 323–328.

Patzkowsky, M. E., Slupik, L. M., Arthur, M. A., Pancost, R. D., Freeman, K. H., 1997.
Late Middle Ordovician environmental change and extinction: Harbinger of the
Late Ordovician or continuation of Cambrian patterns: Geology, v. 25, p. 911–
914.

96
Perfetta, P. J., Shelton, K. L., Stitt, J. H., 1999. Carbon isotope evidence for deep–water
invasion at the Marjumiid–Pterocephaliid biomere boundary, Black Hills, USA: A
common origin for biotic crises on Late Cambrian shelves: Geology, v. 27, p. 403,
406.

Raab, M., Spiro, B., 1991. Sulfur isotopic variations during seawater evaporation
with fractional crystallization: Chemical Geology, v. 86, p. 323–333.

Ripperdan, R. L., 2001. Stratigraphic variation in marine carbonate carbon isotope


ratios: Reviews in Mineralogy and Geochemistry, v. 43, p. 637–662.

Ripperdan, R. L., Miller, J. F., 1995. Carbon isotope ratios from the Cambrian–
Ordovician boundary section at Lawson Cove, Ibex area, Utah: in Cooper, J. D.,
Droser, M. L., Finney, S. C. (eds.), Ordovician Odyssey, Pacific section v. 77, p.
129–132.

Romanek, C. S., Grossman, E. L., Morse, J. W., 1992. Carbon isotopic fractionation in
synthetic aragonite and calcite: Effects of temperature and precipitation rate:
Geochimica et Cosmochimica Acta, v. 56, p. 419–430.

Runkel, A. C., Miller, J. F., McKay, R. M., Palmer, A. R., Taylor, J. F., 2007. High–
resolution sequence stratigraphy of lower Paleozoic sheet sandstones in central
North America: The role of special conditions of cratonic interiors in
development of stratal architecture: GSA Bulletin, v. 119, p. 860–881.

Runkel, A. C., Miller, J. F., McKay, R. M., Palmer, A. R., Taylor, J.F., 2008. The record of
time in cratonic interior strata: does exceptionally slow subsidence necessarily
result in exceptionally poor stratigraphic completeness: in Pratt, B.L., Holmden,
C. (eds.), Dynamics of Epeiric Seas: Geological Association of Canada Special
Paper, v. 48, p. 341–362.

Saltzman, M. R., 1998. Trilobite mass extinction event at the boundary of the Elvinia
and Tsenicephalus bio– zones, Yellowstone National Park: National Park Service
Paleontological Research, Technical Report v. 98, p. 154–157.

97
Saltzman, M. R., 1999. Upper Cambrian carbonate platform evolution, Elvinia and
Taenicephalus zones (pterocephaliid–ptychaspid biomere boundary),
northwestern Wyoming: Journal of Sedimentary Research, v. 69, p. 926–938.

Saltzman, M. R., Cowan, C. A., Runkel, A. C., Runnegar, B., Steward, M. C., and Palmer,
A. R., 2004. The Late Cambrian SPICE 13C event and the Sauk II–III regression:
New evidence from Laurentian basins in Utah, Iowa, and Newfoundland: Journal
of Sedimentary Research, v. 74, p. 366–377.

Saltzman, M. R., Davidson, J. P., Holden, P., Runnegar, B., Lohmann, K. C., 1995. Sea–
level–driven changes in ocean chemistry at an Upper Cambrian extinction
horizon: Geology, v. 23, p. 893–896.

Saltzman, M. R., Ripperdan, R. L., Brasier, M. D., Lohmann, K. C., Robison, R. A., Chang,
W. T., Peng, S., Ergaliev, E. K., and Runnegar, B., 2000. A global carbon isotope
excursion (SPICE) during the Late Cambrian: Relation to trilobite extinctions,
organic–matter burial and sea–level: Palaeogeography, Palaeoclimatology,
Palaeoecology, v. 162, p. 211–223.

Saltzman, M. R., Runnegar, B., Lohmann, K. C, 1998. Carbon isotope stratigraphy of


Upper Cambrian (Steptoean Stage) sequences of the eastern Great Basin: Record
of a global oceanographic event: GSA Bulletin, v. 110, p. 285–297.

Saltzman, M. R., Young, S. A., Kump, L. R., Gill, B. C., Lyons, T. W., Runnegar, B., 2011.
Pulse of atmospheric oxygen during the late Cambrian: PNAS Early Edition, 6 p.

Saylor B. Z., Kaufman A. J., Grotzinger J. P., Urban F., 1998. A composite reference
section for terminal Proterozoic strata of southern Namibia: Journal of
Sedimentary Research, v. 68, p. 1223–1235.

Schroöder S., Grotzinger J. P., 2007. Evidence for anoxia at the Ediacaran–Cambrian
boundary: the record of redox–sensitive trace elements and rare earth elements
in Oman: Journal of Geological Society, v. 164, p. 175–187.

98
Scotese, C. R., Bambach, R. K., Barton, C., Van Der Voo, R., and Zeigler, A. M., 1979.
Paleozoic base maps: Journal of Geology, v. 87, p. 217–278.

Stitt, J. H., 1971. Repeating evolutionary pattern in Late Cambrian biomeres: Journal
of Paleontology, v. 45, p. 178–181.

Stitt, J. H., 1975. Adaptive radiation, trilobite paleoecology, and extinction,


Ptychaspid Biomere, Late Cambrian of Oklahoma: Fossils and Strata, v. 4, p. 381–
390.

Stitt, J. H., 1983. Trilobite biostratigraphy and lithostratigraphy of the McKenzie Hill
Limestone (Lower Ordovician), Wichita and Arbuckle Mountains, Oklahoma:
Oklahoma Geological Survey Bulletin, v. 134, 54 p.

Strauss, H., 1999. Geological evolution from isotope proxy signals— sulfur: Chemical
Geology, v. 161, p. 89–101.

Strauss, H., Schieber, J., 1989. A sulfur isotope study of pyrite genesis: The Mid-
Proterozoic Newland Formation, Belt Supergroup, Montana: Geochimica et
Cosmochimica Acta, v. 54, p. 197–204.

Taylor, J. F. 1997. Upper Cambrian biomeres and stages, two distinctly different and
equally vital stratigraphic units: GSA Field Guides, v. 1, p. 157–176.

Taylor, J. F., 2006. History and status of the biomere concept: in Paterson, J. R.,
Laurie, J. R. (eds.), Cambro–Ordovician Studies II: Association of Australasian
Palaeontologists Memoir, v. 32, p. 247–265.

Thomas, R. C., 1993. The Marjumiid–Pterocephaliid (Upper Cambrian) mass


extinction event in the western United States [Ph.D. thesis]: Seattle, University of
Washington, 352 p.

Veizer, J., Ala, D., Azmy, K., Brukschein, P., Buhl, D., Bruhn, F., Carden, G. A. F., Diener,
A., Abneth, S., Godderis, Y., Japer, T., Korte, C., Pawellek, F., Podlaha, O. F., Strauss,

99
H., 1999. 87Sr/86Sr, δ13C, and δ18O evolution of Phanerozoic seawater: Chemical
Geology, v. 161, p. 59–88.

Weissert, H., 1989. C–isotope stratigraphy, a monitor of paleoenvironmental change:


A case study from the early Cretaceous: Surveys in Geophysics, v. 10, p. 1–61.

Westrop, R., 1995. Sunwaptan (Upper Cambrian) trilobites of the Rabbitkettle


Formation, Mountain River region, northern Mackenzie Mountains, northwest
Canada: Paleontographica Canadaiana, v. 12, p. 1–75.

Westrop, S. R., and Cuggy, M. B., 1999. Comparative paleoecology of Cambrian


trilobite extinctions: Journal of Paleontology, v. 73, p. 337–354.

Westrop, S. R., and Ludvigsen, R., 1987. Biogeographic control of trilobite mass
extinction at an Upper Cambrian “biomere” boundary: Paleobiology, v. 13, p. 84–
99.

100

S-ar putea să vă placă și