You are on page 1of 12

Palaeogeography, Palaeoclimatology, Palaeoecology 574 (2021) 110440

Contents lists available at ScienceDirect

Palaeogeography, Palaeoclimatology, Palaeoecology


journal homepage: www.elsevier.com/locate/palaeo

Sequence stratigraphy and carbon isotopes from the Trenton and Black
River Groups near Union Furnace, PA: Constraining the role of land plants
in the Ordovician world
Page C. Quinton a, *, Michael C. Rygel a, Megan Heins a, b
a
Department of Geology, State University of New York, College at Potsdam, 44 Pierrepont Avenue, Potsdam, New York 13676, United States of America
b
Department of Earth and Atmospheric Sciences, University of Nebraska-Lincoln, 126 Bessey Hall, Lincoln, NE 68588, United States of America

A R T I C L E I N F O A B S T R A C T

Editor:Prof. Thomas Algeo Sea-level change influences carbon isotopic trends in both modern and ancient carbonate depositional envi­
ronments. Generally, this relationship is manifested as a positive carbon isotopic excursion where the rising limb
Keywords: of the excursion is associated with transgression. These excursions exist because sea level can influence 1) local/
Carbon cycle regional/global carbon cycling, 2) basin restriction and meteoric influence, and 3) carbonate sedimentation. The
Non-vascular plants
first two processes are influenced, in part, by the large carbon reservoir represented by the terrestrial biosphere.
Appalachian Basin
Our goal is to explore the relationship between carbon isotopes and sea level prior to the evolution of vascular
Sandbian
Sea level plants at a time when the terrestrial carbon reservoir was far smaller.
Epicontinental sea This study focuses on the Upper Ordovician Union Furnace section in Pennsylvania, where a previously
documented ~1‰ increase in δ13C values in the upper Sandbian (458.4–453.0 Ma; referred to as the “baseline
shift”) has been interpreted to reflect a perturbation to the global carbon cycle due to the proliferation of non-
vascular plants. By focusing on the interval deposited before the “baseline shift”, we were able to examine the
relationship between sea level, carbon isotopes, and some of the earliest terrestrial ecosystems. We present a
high-resolution δ13C record directly tied to the sedimentological and sequence stratigraphic framework. By
sampling genetically related packages of rock we were able to document a relationship between carbon isotopes
and sea level at the sequence and parasequence level. Our results show that carbon isotopic values decreased
during transgressions and increased during regressions. This seemingly inverse relationship between δ13C values
and sea level is interpreted to reflect increased terrestrial fluxes during regression in a world devoid of vascular
plants. Our results suggest that the early colonization of land by non-vascular plants impacted regional carbon
cycling in the Appalachian Basin and eventually the global carbon cycle in the Late Ordovician.

1. Introduction carbon, either in the water column or at the sediment-water interface,


decreases the δ13C DIC value of the ocean. These changes in ocean
Carbon isotopic trends from marine carbonates can record spatial chemistry are recorded in the carbon isotopic ratio of the carbonate
and temporal changes in the global carbon cycle (Saltzman and Thomas, sediments (δ13C carb) formed in many marine settings. This basic rela­
2012). They do so because significant fluctuations in the average δ13C tionship is the foundation for the use of δ13C records in paleoclimatic,
composition of the ocean (δ13C DIC) are a function of changes in the flux paleoredox, and chemostratigraphic studies. However, there are
of carbon between the atmosphere-ocean system and the sedimentary complicating factors that must be considered when using δ13C records
reservoir (Kump and Arthur, 1999; Saltzman and Thomas, 2012). This from epicontinental carbonates.
flux is primarily controlled by the burial and oxidation of organic car­ The δ13C trends recorded in carbonates of shallow epicontinental
bon, which is enriched in 12C (Kump and Arthur, 1999; Saltzman and settings can reflect a complicated set of global, regional, and local pro­
Thomas, 2012). The burial of organic carbon thus results in an increase cesses. These processes include, but are not limited to, carbonate
in the δ13CDIC value of the water. Whereas the oxidation of organic weathering, carbonate deposition, meteoric input, organic carbon

* Corresponding author.
E-mail address: quintopc@potsdam.edu (P.C. Quinton).

https://doi.org/10.1016/j.palaeo.2021.110440
Received 18 January 2021; Received in revised form 25 April 2021; Accepted 25 April 2021
Available online 29 April 2021
0031-0182/© 2021 Elsevier B.V. All rights reserved.
P.C. Quinton et al. Palaeogeography, Palaeoclimatology, Palaeoecology 574 (2021) 110440

burial/oxidation, carbon sources/flux, and local anoxia (Kump and been extended into New York and southern Ontario (Brett and Baird,
Arthur, 1999; Kump et al., 2000; Immenhauser et al., 2003; Panchuk 2002; Brett et al., 2004), but these studies and recent work in the Cin­
et al., 2005; Melchin and Holmden, 2006; Fanton and Holmden, 2007; cinnati Arch (Brett et al., 2020) has led to revision of the original
Schrag et al., 2013). All of these processes can be controlled directly or framework established by Holland and Patzkowsky (1996) and the
indirectly by relative sea-level change because it influences the burial/ addition of at least two sequences. Much of this work has focused on the
oxidation of organic carbon, proximity to terrestrial carbon sources, Katian Stage of the Upper Ordovician, with similar efforts not yet
basin restriction, carbonate production vs. weathering, carbonate completed for the earlier Sandbian Stage.
mineralogy, and/or meteoric fluids. This relationship is supported by Detailed sedimentological and sequence stratigraphic work on the
multiple studies which have documented a correlation between carbon Union Furnace section was done by Laughrey et al. (2017) as part of the
isotopic trends and sea level (e.g. Weissert, 1989; Jenkyns, 1996; Bug­ pre-conference field trip for the 82nd Annual Field Conference of
gisch et al., 2003; Immenhauser et al., 2003; Swart and Eberli, 2005; Pennsylvania Geologists and as part of a larger report (Patchen et al.,
Swart et al., 2009; Eltom et al., 2018). The general pattern that emerges 2006). They recognized four sequences in the Black River Group (BR
is one where δ13C values tend to increase during sea level rise and 1–4) and six sequences in the Trenton Group (TR 1–6). While the TR 3
decrease during sea level fall. The result is a positive carbon isotopic and TR4 sequences have been correlated to Holland and Patzkowsky’s
excursion with the rising limb during transgression. (1996) M4 and M5 sequences, the rest of the section has not been
While there are multiple processes that can potentially produce this correlated in detail and we use the nomenclature that Laughrey et al.
sea-level-induced positive carbon isotopic excursion, many of them are (2017) established at this location. Broad regional age equivalence of
linked to an abundance of land plants. For example, proximity to shore the Union Furnace section is possible with the age model developed by
and terrestrial carbon sources (typically enriched in 12C) can lead to Edwards et al., 2015a, placement of Midcontinent Conodont Biozones
decreasing δ13C values when relative sea level is low (Burdige, 2005; (Edwards et al., 2015a, Adiatma et al., 2019), and identification of the
Jarvis et al., 2006; Fanton and Holmden, 2007). At the same time, Deicke and Millbrig K-bentonites (altered volcanic ash beds; Laughrey
proximity to the shoreline and basin restriction might lead to an increase et al., 2017).
in meteoric fluids which are typically enriched in plant respired 12C
(Allan and Matthews, 1982; Immenhauser et al., 2003). The result is 2.2. Late Ordovician carbon isotopic record
decreased δ13C values when sea level is low. This influence of land plants
can be observed in modern carbonate settings such as the Great Bahama The Late Ordovician carbon isotopic record is well documented with
Banks and Florida Bay, where nearshore environments record δ13CDIC four globally recognized positive excursions and a relatively high-
values that are up to 2‰ lower than those close to the open ocean, in resolution composite curve spanning the Sandbian to the Hirnantian
part because of freshwater and terrestrial organic matter input (Patter­ Stages (Bergström et al., 2009; Saltzman and Thomas, 2012). Many of
son and Walter, 1994). The importance of land plants on the carbon these studies have focused on the Laurentian epicontinental region (e.g.
isotopic values of shallow epicontinental settings and the at least partial Young et al., 2005; Fanton and Holmden, 2007; Bergström et al., 2010;
dependence on land plants to produce the sea-level-induced positive Saltzman and Edwards, 2017; Adiatma et al., 2019). The large number
carbon isotopic excursion raises an important question: will the rela­ of carbon isotopic records and studies that attempt to constrain regional
tionship between sea level and carbon isotopes be different and/or more variability in δ13C values (Holmden et al., 1998; Panchuk et al., 2006)
variable prior to an abundance of terrestrial land plants? The aim of this have demonstrated that there was significant variation in carbon iso­
study is to examine carbon isotopic trends in a sequence stratigraphic topic values (~2‰) across the Laurentian epicontinental sea during the
framework in the absence of widespread terrestrial vegetation. Late Ordovician. These spatial gradients in carbonate δ13C values have
generally been attributed to basin restriction, differences in carbon
2. Geologic setting sources, and organic carbon burial (Panchuk et al., 2005; Panchuk et al.,
2006) but could also reflect miscorrelations between sections or certain
This study focused on Black River and Trenton Group carbonates mineralogical differences (Swart et al., 2009; Metzger and Fike, 2013;
(Late Ordovician) in a ~ 240 m thick exposure along PA Route 453 near Sell et al., 2015). Regardless of the cause, the presence of these spatial
Union Furnace, PA (40.6230◦ N, 78.1730◦ W). These rocks were gradients increases the chances that sea level could exert a resolvable
deposited in the shallow epicontinental sea that covered Laurentia influence on δ13C trends.
during the Sandbian–Katian Stages (458.4–445.2 Ma; Scotese, 2004; This prediction is supported by previous studies that documented a
Torsvik and Cocks, 2013; Walker et al., 2018). Our efforts were focused relationship between δ13C trends and sea level during the Late Ordovi­
on the Sandbian-aged Black River Group in the basal 105 meters of the cian. Fanton and Holmden (2007) used neodymium (Nd) isotopes to
section (Laughrey et al., 2017). The Union Furnace section is appro­ estimate sea level change and compared these inferred sea level cycles to
priate for this study because: 1) the sequence stratigraphic framework δ13C trends during the Katian. They suggested that the correlation be­
independently established by Laughrey et al. (2017) is easily identifiable tween positive carbon isotope excursions and Nd trends indicated that
and confirmable in the field, 2) the Late Ordovician carbon isotopic sea level was driving changes in basin restriction and upwelling (Fanton
record of Laurentia suggests that sea level might have influenced carbon and Holmden, 2007). Similarly, Young et al. (2008) argued that the
cycling in the Appalachian Basin, and 3) the Union Furnace section globally recognized Guttenberg Isotope Carbon Excursion (GICE) in the
spans an interval when non-vascular land plants were beginning to early Katian might have been enhanced in the Laurentian epicontinental
impact the carbon cycle. sea as sea level rose, driving upwelling of nutrient rich, oxygen poor
waters that fueled primary productivity and the burial of organic car­
2.1. Late Ordovician sequence stratigraphy bon. While these studies focused on Katian carbon isotopic record (and
this study is focused on the slightly older Sandbian record), these results
A detailed sequence stratigraphic framework for the Middle to Late suggest that the Union Furnace section is an appropriate interval for
Ordovician of eastern North America was established by Holland and exploring the relationship between sea level and δ13C.
Patzkowsky (1996). Their efforts focused on the central to southern
Appalachians and the Cincinnati Arch where they recognized and 2.3. The Late Ordovician & the evolution of land plants
named a total of 14 sequences. Those sequences are composed of
transgressive systems tracts (TST) and highstand systems tracts (HST), The colonization of the terrestrial realm by land plants during the
with lowstand systems tracts (LST) occasionally preserved (Holland and Ordovician–Devonian Periods would have had a significant and lasting
Patzkowsky, 1996). Parts of this sequence stratigraphic framework have impact on the carbon cycle (Figure 1). Increased biomass would have

2
P.C. Quinton et al. Palaeogeography, Palaeoclimatology, Palaeoecology 574 (2021) 110440

Fig. 1. Timescale with terrestrial plant stages as identified by Davies and Gibling (2010) and Dahl and Arens (2020). Major events in lant plant evolution and
influence on the carbon isotopic record are tied to the timescale. The Union Furnace section is placed within this framework using the age model from Edwards
(2015a). 1Stein et al., 2020, 2Glasspool and Scott, 2000, 3Gerrienne et al., 2011, 4Driese and Mora, 2001, 5Kennedy et al., 2013, 6Gensel et al., 2001, 7Edwards et al.,
1983, 8Tomescu and Rothwell, 2006, 9Tomescu et al., 2009, 10Kenrick et al., 2012, 11Adiatma et al., 2019, 12Jones et al., 2015, 13Malkowski and Racki, 2009.

resulted in significant terrestrial carbon storage (Berner, 1998). The components (Kleinhans et al., 2018; McMahon and Davies, 2018) and
advent of root systems would have led to an increase in silicate weath­ secreting organic acids through their rhizoids (Boucot and Gray, 2001;
ering that consumes CO2 (Algeo and Scheckler, 1998; Berner, 1998; Lenton et al., 2012; Quirk et al., 2015; Porada et al., 2016).
Quirk et al., 2015). The increase in silicate weathering would also supply A Late Ordovician increase in silicate weathering and nutrient supply
nutrients to the marine realm fueling primary productivity and causing could have caused a fundamental shift in the global cycle (Boucot and
an increase in marine organic carbon burial (Algeo et al., 1995; Algeo Gray, 2001; Lenton et al., 2016; Porada et al., 2016) marked by a ~ 2‰
and Scheckler, 1998). Traditionally, these changes were thought to increase in the background marine δ13C carbonate values in the late
mostly be confined to the Devonian Period because that is when complex Sandbian (Lenton et al., 2016). This geochemical support comes in part
root systems evolved and when significant organic carbon storage in the from the Union Furnace section where Adiatma et al. (2019) reported an
terrestrial realm began via the earliest coal (Algeo and Scheckler, 1998; ~1‰ increase in δ13C values from 60 to 120 m in their section. They
Glasspool and Scott, 2010; Dahl and Arens, 2020). However, recent refer to this increase as a “baseline shift” caused by the influence of land
studies looking at microfossil assemblages, molecular clock origination plants (Adiatma et al., 2019). By focusing on the interval below this
dates (Kenrick et al., 2012; Morris et al., 2018), and non-vascular land “baseline shift” in this study we were able to examine the relationship
plants’ ability to influence carbon storage and silicate weathering between sea level, carbon isotopes, and the earliest terrestrial
(Boucot and Gray, 2001; Edwards et al., 2015a, 2015b; Porada et al., ecosystems.
2016), suggest that land plants impacted the carbon cycle throughout
the Ordovician and Silurian and did so in a progressive, stepwise fashion 2.4. Previous carbon isotope work from the Union Furnace section
(Lenton et al., 2012, 2016; Dahl and Arens, 2020).
Recent studies suggest that land plants’ impact on the carbon cycle As noted above, Adiatma et al. (2019) reported δ13C results at a 0.5
may have started in the Late Ordovician (Lenton et al., 2012, 2016; to 3 m sampling interval from the Union Furnace section. That study
Adiatma et al., 2019; Dahl and Arens, 2020). The first unequivocal specifically focused on documenting large-scale trends related to
spores appear in the Middle Ordovician (Rubinstein et al., 2010; Kenrick changes in the global carbon reservoir (e.g. the GICE and the Sandbian
et al., 2012) and macroscopic plant material attributed to liverworts “baseline shift”). Here we focus on the generating a higher resolution
appears by the Late Ordovician (Tomescu et al., 2009). These observa­ record (0.1 to 1 m intervals) for the lower 113.2 m that is directly tied to
tions have led some to argue that significant terrestrial biomass existed detailed sedimentological observations and a sequence stratigraphic
by the Late Ordovician (Tomescu et al., 2009; Dahl and Arens, 2020). framework. This new paired geochemical and sedimentological dataset
These early terrestrial ecosystems would have been dominated by will allow us to test sea level’s role on the observed carbon isotopic
bryophytes, which have little to no subsurface component. Despite the patterns.
lack of roots, bryophytes could have exerted a significant influence on
silicate weathering rates by trapping sediments in their surface

3
P.C. Quinton et al. Palaeogeography, Palaeoclimatology, Palaeoecology 574 (2021) 110440

3. Methods deposits capping progradational parasequence sets. When preserved,


lowstand systems tract (LST) deposits record a basinward shift in facies
3.1. Sedimentology and sequence stratigraphy and a continuation of the progradational parasequence architecture.
Transgressive surfaces (TS) mark the start of progressive deepening;
We measured and described 220 m of the Union Furnace section where the LST is absent this surface is merged with the SB. The trans­
(Fig. 2 and supplementary materials file). We keyed our measurements gressive systems tract (TST) is composed of deposits that record pro­
to bedding units marked on the outcrop for the pre-conference field trip gressive deepening and a retrogradational parasequence architecture.
for the 82nd Annual Field Conference of Pennsylvania Geologists which Given the subtle lithologic differences in distal facies and the monoto­
also allowed us to directly compare our observations to the sedimento­ nous succession that result, individual parasequences were difficult to
logical and sequence stratigraphic work by Laughrey et al. (2017). In identify. The maximum flooding surface (MFS) is placed in the middle of
addition to our field-based observations, 15 samples were collected for the deepest water deposits; given the nature of the distal facies described
petrographic analysis. Using these sedimentological observations and above, its placement is approximate. The transgressive systems tract
building on the interpretations of Laughery et al. (2017), we developed a (TST) records progressive shallowing. In some intervals, marked dif­
simplified and standardized framework for a carbonate ramp informed ferences in facies result in an obvious progradational parasequence
by comprehensive reviews of carbonate systems (e.g. Read, 1985; Son­ architecture.
nenfeld, 1996; James and Jones, 2015). This depositional framework is In a few places we made slight adjustments (<2 m) to the position of
consistent with the barrier-bank type ramp that characterizes the Black sequence stratigraphic surfaces presented by Laughrey et al. (2017). In
River Group in this region (Laughrey et al., 2017). We were able to place sequence BR 2 (27.5–61.5 m) we were able to confidently identify 10
the Union Furnace section into 10 distinct depositional environments parasequences. This level of detail was not possible in the other se­
within that framework (Table 1, Figs. 2 and 3). quences and systems tracts because of a lack of lithologic contrast. Our
Our sequence stratigraphic framework is built on the work of expanded facies scheme and sequence stratigraphic framework is pre­
Laughrey et al. (2017), which is easily recognizable and verifiable in the sented in Figs. 2 and 3.
field. Sequence boundaries (SB) are placed atop the shallowest water

Fig. 2. Graphic log for 125 m of the Union Furnace section. Lithostratigraphy and bedding units are based on Laughrey et al. (2017). Graphic log, facies associations
(Table 1), and parasequence calls are based on work from this study. Sequence stratigraphic framework is based on Laughrey et al. (2017) and confirmed with our
sedimentological analysis. Grainsize scale abbreviations are mdst (mudstone), wkst (wackestone), pkst (packstone), grnst (grainstone), rxl (recrystallized). Detailed
field notes and full measured section are included in the supplementary materials file.

4
P.C. Quinton et al. Palaeogeography, Palaeoclimatology, Palaeoecology 574 (2021) 110440

Table 1
Facies associations and interpretated depositional environments for the Union Furnace section. The framework for depositional environments we deploy is informed by
Read (1985), Sonnenfeld (1996), and James and Jones (2015). The bioturbation index was developed by Droser and Bottjer (1986).
Facies Depositional Lithologies and sedimentary structures Fossils Processes
Associations Environment

A Anaerobic Laminated to faintly laminated dark gray mudstone Rare transported brachiopod Below storm wave base, limited carbonate
Outer Ramp (>90%) with discrete interbeds of fissile organic- shells. Rare Chondrites. production, low oxygen limits bioturbation.
rich(?) mudstone (<10%). Rare fossiliferous Dominated by suspension deposition.
wackstone beds. Sparse bioturbation (B.I. = 1–2).
Fossils and burrows are locally replaced by chert.

B Aerobic Outer Faintly laminated to internally massive mudstone Brachiopods with scattered Below storm wave base, limited carbonate
Ramp interbedded with dark gray nodular limestone and gastropods and rugose coral. production but closer to carbonate sources on the
rare sharp-based packstone/grainstone shelf, aerobic conditions allow for more intense
(fossiliferous). Abundant bioturbation (B.I. = 3–4); bioturbation and somewhat higher faunal
burrows commonly filled with coarser grained diversity. Combination of suspension deposition
material and downslope transportation from C

C Middle Ramp Faintly laminated mudstone and wackestone with Abundant fossils include Above storm wave base and below fairweather
subordinate sharp-based fossiliferous grainstones, brachiopods, bryozoans, trilobites, wave base, abundant oxygen and fully marine
1–2 cm thick laminated zones locally. Abundant rugose corals, echinoderms, conditions results in intense bioturbation and
bioturbation (B.I. = 3–4). Scattered hardgrounds mollusks, and gastropods. abundant/diverse fauna. Grainstones and lag
with bored surfaces and intraformational deposits represent storm beds.
conglomerates.

D Shoreface Grainstone with interbedded packstone; both Abundant abraded shell Above fairweather wave base. Although fully
contain abraded fossils, intraclasts, and peloids. fragments. marine, high energy conditions and shifting
Faintly laminated to internally massive. Minor substrate results in abraded fossils and modest
mudstone/wackestone interbeds locally. bioturbation.
Bioturbation is difficult to distinguish in coarse-
grained facies; rare grainstone-filled burrows in
mudstone/wackestone (B.I. = 1).

E Upper Nearly identical to D, except with abundant cross- Similar to D, except that energy levels are even
Shoreface beds and no mudstone/wackestone. higher and cross-beds become abundant.

F Restricted Interbedded wackestone/packstone (fossiliferous) Abundant abraded shell Topographic low between the upper shoreface
Upper with subordinate mudstone. Faintly laminated to fragments. and the backshoal. Wave activity still occurs, but
Shoreface internally massive. Minor bioturbation (B.I. = 1). mud can accumulate in sheltered areas.
Commonly grades into G.

G Back-shoal Interbedded grainstones and packstones with minor Abundant abraded shell Highest energy (shallowest water) environment
wackestone. Grainstones may be laminated, ripple fragments; rare bryozoans. on the ramp; represents the transition between
cross-laminated, or cross-bedded. Grainstones the wave-dominated shoal and the tidally-
appear sandy, and have abraded fossils and influenced lagoon.
intraclasts. Burrows are difficult to distinguish (B.I.
= 1). Commonly grades into F.

H Lagoon Mudstone to wackestone with minor packstone/ Abundant abraded shell Shoreward of shoal and subaqueous even at low
grainstone lags (fossils and intraclasts). Mudstone fragments; bryozoans (some in tide. Occasional washover events bring in
and wackstone varies from internally massive to situ) grainstones; intraclasts record storm events.
faintly laminated and bioturbation ranges from Variable conditions result in restricted fauna.
scattered to abundant (B.I. = 2–5). Pyrite abundant
locally.

I Intertidal Cyclically interbedded successions of a) faintly Abundant abraded shell Subaerially exposed at low tide and covered in
laminated mudstone/wackestone, b) laminated to fragments; bryozoans and shallow water at high tide. Generally a sheltered,
internally massive fossiliferous packstone, and c) scattered brachiopods. low energy environment behind the back-shoal/
fossiliferous to intraclastic grainstones. Individual lagoon complex; occasional high energy events.
cycles are a few tens of centimeters and not well
represented given the scale of our measured
section. Scattered bioturbation (B.I. = 1–2).
Commonly grades into J.
J Supratidal Mudstone to wackestone interbedded with Abraded shell fragments. Subaerially exposed, even at high tide. High
laminated to ripple cross-laminated grainstone. energy events can flood the area.
Scattered fossils in mudstone/wackestone;
grainstones appear sandy and may be capped with
fossiliferous horizons. Scattered bioturbation (B.I.
= 1–2). Commonly grades into I.

3.2. Stable carbon isotopic samples and analyses samples through the parasequence. This approach allowed us to
generate a δ13C record that captured changes within parasequences.
Samples for carbonate carbon isotopic analyses were collected along Samples were given a general lithological identification (mudstone,
with the generation of our detailed measured section. This sampling wackestone.. etc) in the field and this was confirmed when sampling for
protocol allowed us to directly tie generated carbon isotopic trends to carbon isotopic values in the laboratory.
the sedimentological information and interpretations. We used a sliding Sample powders were generated from freshly exposed surfaces using
sampling resolution based on sedimentological context. We did not take a low speed drill. When possible, we targeted the micritic zones because
samples at some regular stratigraphic interval. Instead, when possible muddy lithologies have a lower susceptibility to post-depositional
parasequence boundaries were identified, we took a minimum of three alteration (due to their low permeability) than coeval coarser-grained

5
P.C. Quinton et al. Palaeogeography, Palaeoclimatology, Palaeoecology 574 (2021) 110440

Fig. 3. Bulk carbonate carbon isotopic results from


this study are plotted with a 3-point moving average.
Lithostratigraphy (column 1) and bedding units
(column 2) are based on Laughrey et al. (2017).
Facies associations and inferred depositional envi­
ronments are based on the standardized framework
developed in this study (column 3). Systems tracts
and sequence stratigraphic surfaces (column 4) are
based on Laughrey et al. (2017) and confirmed based
on the work from this study. Black River and Trenton
Group cycles as identified by Laughrey et al. (2017)
are given in column 5. The late Sandbian “baseline
shift” previously identified by Adiatma et al. (2019)
is highlighted in green and the interval of proposed
sea level influence on carbon isotopic trends in
highlighted in blue. (For interpretation of the refer­
ences to colour in this figure legend, the reader is
referred to the web version of this article.)

lithologies (e.g. Hayes et al., 1989). However, that was not always supplementary data file. These values, with an average of − 0.28‰ are
possible because some intervals lacked carbonate mud. Sample lithology similar to those reported elsewhere for this interval (Young et al., 2010;
for all carbon isotopic values are reported in supplementary data file. Bergström et al., 2010; Quinton et al., 2016) and broad trends match
Sample powders were analyzed for bulk carbonate δ13C and δ18O values those reported previously by Adiatma et al. (2019). In particular, there is
on a ThermoFinnigan Delta Plus Dual Inlet isotope ratio mass spec­ a general trend of increasing values from ~0.5‰ at 41 m to 1.5‰ at 113
trometer connected to a Kiel III Carbonate Interface at the University of m. We did not sample the upper 113–220 m of the section (i.e. the GICE
Missouri Stable Isotope and Biogeochemistry Laboratory. Analytical interval) because a) the outer ramp deposits of those upper 107 m do not
precision is ±0.04‰ (one standard deviation) for δ13C and ± 0.06‰ contain the sedimentological contrast needed to confidently identify
(one standard deviation) for δ18O and is calculated from multiple ana­ parasequences and b) this study focused on Sandbian events.
lyses of NBS-19. In the lower 61.5 m, below the interval previously identified by
Adiatma et al. (2019) as the “baseline shift” in the late Sandbian, δ13C
4. Results values are correlated with the sequence stratigraphic framework. We see
decreasing values in the TST, increasing values in the HST, and stable
We report 170 bulk carbonate δ13C values from the lower 113 m of values in the single occurrence of a LST. Inflection points in these trends
the Union Furnace section. All values are plotted on Fig. 3 with a three- occur at or near sequence stratigraphic surfaces. In sequence BR 1,
point moving average; a complete data table is included in the values decrease from –1‰ to –1.5‰ in the TST and increase from − 1.5‰

6
P.C. Quinton et al. Palaeogeography, Palaeoclimatology, Palaeoecology 574 (2021) 110440

to 0‰ in the HST. In sequence BR 2, values decrease from 0 to –1.5‰ in


the TST and increase from − 1.5‰ to 0‰ in the HST. We used Pearson’s
regression analyses to test the direction and significance of these visually
identified trends. Those results are reported in Table 2 and indicate that
there is a moderate to strong correlation in the negative direction be­
tween δ13C values and transgressive system tracts. For highstand sys­
tems tracts, there is a strong correlation in the positive direction. To
demonstrate that this correlation was truly unique to the identified
systems tracts, we also tested the correlation between δ13C values and
the sequences (BR 1 and BR 2) and found a weak correlation.
In sequence BR 2 we were able to confidently identify parasequences
and we report corresponding δ13C trends in Fig. 4. Averages from those
10 parasequences display a stepwise pattern consistent with their place
in the sequence stratigraphic framework (parasequence averages and
standard deviations are reported in the supplementary data file). Para­
sequence 1 (P1) represents the LST (P1 average 0.24‰). Parasequences
2 through 4 (P2 – P4) comprise the TST and show progressively lower
averages (P2 average 0.02‰, P3 average − 1.10‰, P4 average −
1.51‰). Parasequences 5 through 10 (P5 – P10) comprise the HST and
progressively higher averages (P5 average − 1.23‰, P6 and P7 average
− 0.56‰, P8 average − 0.57‰, P9 average − 0.47‰, P10 average −
0.41‰).

5. Discussion

5.1. Carbon isotopic trends at Union Furnace

Union Furnace carbon isotopic results record the same gradual in­
crease in δ13C values through the Erismodus quadridactylus to Phragmo­
dus undatus Midcontinent Conodont Biozones documented on Laurentia
(Quinton et al., 2016; Bergström et al., 2015) and Baltoscandia (Ainsaar
et al., 2010). This observation suggests that the Union Furnace carbon
isotopic record does reflect primary changes in the global carbon
reservoir as has been previously argued by Adiatma et al. (2019).
However, the correlation between δ13C trends and the sequence strati­
graphic framework observed in sequences BR 1 and BR 2 at Union
Furnace indicates that regional and/or local processes were also influ­
encing the record.
Carbon isotopic results from Union Furnace are interpreted to reflect
primary changes in the δ13C value of the local water mass (δ13CDIC)
rather than diagenetic processes. The main diagenetic concern for car­
bonate δ13C is recrystallization and exchange of carbon isotopes with
meteoric fluids (Allan and Matthews, 1982; Saltzman and Thomas,
2012). Carbon and oxygen isotopic results show no systemic relation­
ship (supplementary material file) as might be the case if the record was
altered by meteoric diagenesis. Traditionally, cross-plots between δ13C
and δ18O values of carbonates were used as a tool to test for systematic
alteration due to exchange with meteoric fluids (Marshall, 1992). The

Table 2
Results from Pearson’s regression analysis. Strength of correlation is determined
by using a scale informed from Akoglu (2018): 0.1 < |r| < 0.3 for weak corre­
lation, 0.3 < |r| < 0.5 for moderate correlation, and 0.5 < |r| for strong
correlation.
a b c
Sequence Systems Pearson’s Pearson’s n Statistical
Tract r r Significance

BR 1 TST − 0.284 − 0.512 16 Moderate


correlation
BR 2 TST − 0.820 − 0.906 14 Strong correlation
BR 1 HST 0.608 0.965 4 Strong correlation Fig. 4. Bulk carbonate carbon isotopic results for sequence BR 2 plotted by
BR 2 HST 0.539 0.687 30 Strong correlation parasequence. Each parasequence is numbered and the parasequence δ13C
BR 1 All − 0.045 − 0.062 20 Weak correlation average is indicated with a pink line. The carbon isotopic averages for para­
BR 2 All − 0.158 − 0.158 50 Weak correlation sequences in the transgressive systems tract get progressively lower and pro­
a
Calculated based on original data with no smoothing. gressively higher in the highstand systems tract. (For interpretation of the
b
Calculated based on 3-point moving average of carbon isotopic data. references to colour in this figure legend, the reader is referred to the web
c
Number of data points in analysis. version of this article.)

7
P.C. Quinton et al. Palaeogeography, Palaeoclimatology, Palaeoecology 574 (2021) 110440

positive correlation between δ13C and δ18O values is interpreted to carbon cycling, 2) meteoric influence, and 3) carbonate sedimentation.
reflect the varying contribution of light isotopes from meteoric fluids Carbon cycling is the most common explanation for the relationship
(Allan and Matthews, 1982; Marshall, 1992; Banner and Hanson, 1990). between sea level and carbon isotopic values (e.g. Jenkyns, 1996; Bur­
We should note that the utility of carbon and oxygen cross-plots as a test dige, 2005; Jarvis et al., 2006; Fanton and Holmden, 2007; Katz et al.,
for meteoric diagenesis has been challenged based on work in the Great 2007; Ainsaar et al., 2010). The two may be related because sea level
Bahama Banks. Swart and Oehlert (2018) demonstrated that positive controls the net surface area for primary productivity and upwelling of
correlation between carbon and oxygen isotopes might also result from nutrient rich (that can potentially fuel primary productivity) and/or
varying amounts of recrystallization and development of diagenetic oxygen poor waters (that can potentially promote the preservation of
cements and, more importantly, that meteoric diagenesis does not al­ organic carbon). The net result is that during transgressions, increased
ways produce a positive correlation. upwelling of nutrient rich waters and increased surface area for primary
Another way to constrain potential meteoric influence is to look at productivity results in increased organic carbon burial. Combined with
the relationship between δ13C and δ18O values and lithology. Lithologies increased terrestrial organic carbon input and basin restriction during
dominated by fine grained material (e.g. mudstone and wackestone) sea level falls, the result is a classic sea-level-induced positive carbon
have lower permeability and porosities, such that it is less likely that isotope excursion with the rising limb in the TST and the falling limb in
fluids will penetrate and alter these than their coarser grained coun­ the HST.
terparts (e.g. grainstone and packstone) (Hayes et al., 1989; Marshall, Meteoric fluids can also cause a relationship between sea level and
1992). We observe no systematic relationship between lithology and δ13C patterns. During sea level lows, penetration of meteoric fluids
isotopic values (Supplementary material file). Lastly, petrographic work (potentially enriched in respired 12C) can lead to alteration of recorded
did not show significant and/or obvious meteoric cements and we carbon isotopic values. The result is decreasing carbon isotopic values
avoided any such features when sampling for carbon isotopes. Overall, up to the sequence boundary. This model, described by Allan and Mat­
we see no evidence that the carbon isotopic values at Union Furnace thews (1982), is readily acknowledged and geochemical studies attempt
reflect significant alteration due to meteoric diagenesis. to avoid meteoric cements. However, Immenhauser et al. (2003) pro­
posed the idea that sea level’s control on meteoric fluids can potentially
5.2. Is sea level influencing carbon isotopes at Union Furnace? have a more active role in the generation of carbon isotopic trends. Sea
level can influence the degree of basin restriction and therefore the ratio
The statistically significant correlation between carbon isotopic of meteoric fluids to sea water. During sea level lows, basin restriction
trends and the sequence stratigraphic framework in sequences BR 1 and can create an isotopically evolved fluid (low carbon isotopic values due
BR 2 suggests that sea level was influencing the carbon isotopic record at to increased meteoric influence). As sea level rises, decreased basin re­
Union Furnace. There are multiple lines of evidence that support this striction and/or shifting shoreline leads to decreased meteoric influence
interpretation. The sequence stratigraphic framework was established and higher carbon isotopic values in any carbonate sediments produced
independently (though verified by us based on our facies identifications) at the time. The result is a positive carbon isotopic excursion that has the
of the δ13C record. For this reason, there is no chance that the rising limb associated with transgression (TST).
geochemical trends influenced our placement of sequence stratigraphic Sea level can also influence carbon isotopic patterns because it
surfaces. We see the same pattern in both sequences with similar start controls the production, export, mixing, and type of carbonate sedi­
and end values for each systems tract (TST trend ends at ~ − 1.5‰ and ments. Aragonite tends to form in the shallowest water settings and
HST trend ends at ~0‰). This observation suggests that there is some typically has carbon isotopic values 2–3‰ higher than the low magne­
specific process that is allowing sea level to influence δ13C trends (dis­ sium calcite (LMC) that forms on the flank of the carbonate banks (Swart
cussed below). Furthermore, the inflection points in the carbon isotopic and Eberli, 2005). Based on these relationships and observations from
trends occur at or near sequence stratigraphic surfaces. For example, in the Great Bahama Banks, Swart (2008) argued that sea level change
BR 2 when flooding begins at the transgressive surface, δ13C values start would result in positive carbon isotopic excursions that get preserved
to decrease and when progressive shallowing begins following the even after the aragonite is recrystallized to LMC after burial. In this
maximum flooding surface, δ13C values start to increase. model, as sea level rises, aragonite production increases and carbon
We also document a relationship between sea level and δ13C trends at isotopic values increase. Again, we get a positive carbon isotope
the parasequence level in sequence BR 2. The progressively deepening excursion with the rising limb in the transgressive systems tract.
parasequences of the TST (P2 – P4) show progressively lower δ13C In all of these scenarios, the result is the inverse of the one that we
values while the progressively shallowing parasequences of the HST document at Union Furnace. We propose two possible explanations for
show progressively higher δ13C values. While there is certainly some the observed relationship between carbon isotopes and sequence stra­
scatter in δ13C values, averages of each parasequence show the same tigraphy in BR 1 and BR 2 at Union Furnace: 1) increased weathering
relationship to δ13C that we observe at the systems tract level. Com­ fluxes during HST fueling marine primary productivity and/or 2) a
bined, these observations provide significant evidence that sea level is spatial gradient in marine δ13CDIC with higher values nearshore.
influencing δ13C values in the BR 1 and BR 2 sequences at Union Carbon isotopic values could be increasing in the HST due to
Furnace. The question is: what processes are allowing sea level to con­ increased marine primary productivity and a corresponding increase in
trol carbon isotopes and are they different from what is traditionally organic carbon burial. Normally, primary productivity and enhanced
deployed to explain positive carbon isotope excursions? organic carbon burial is thought to increase during transgression as
described above. However, it is equally as plausible that nutrients (e.g.
5.3. How is sea level controlling carbon isotopes at Union Furnace? phosphorus and nitrogen) can be sourced from the terrestrial realm.
Progressive shallowing during the HST and the basinward movement of
The pattern of decreasing δ13C values during the TST and increasing the shoreline could have led to increased weathering fluxes. Because
values during the HST in BR 1 and BR 2 is statistically significant adequate shallow water would still have existed for primary producers,
(Table 2) and the opposite of the pattern typically documented in most the nutrient supply led to an increase in primary productivity and
younger carbonates (Weissert, 1989; Jenkyns, 1996; Immenhauser eventually an increase in organic carbon burial in these shallow marine
et al., 2003; Swart and Eberli, 2005; Swart et al., 2009; Eltom et al., settings. Basin restriction would have enhanced the effect by stalling
2018). In these studies, a rise in sea level causes an increase in carbon mixing with the open ocean.
isotopic values followed by a return to pre-excursion values as sea level The second possible explanation is that spatial gradients in δ13CDIC
falls. The processes that are commonly thought to produce this sea level were the opposite of what we observe in a fully vegetated world. In this
induced positive carbon isotopic excursion fall into three categories 1) scenario, nearshore settings could have had higher δ13CDIC values than

8
P.C. Quinton et al. Palaeogeography, Palaeoclimatology, Palaeoecology 574 (2021) 110440

distal settings. In modern settings, like the Great Bahama Banks or between sea level and carbon isotopic trends is documented elsewhere.
Florida Bay, nearshore settings have distinctly lower δ13CDIC values due The Late Ordovician carbon isotopic record has received significant
in part to the high influx of respired 12C from terrestrial vegetation attention, but much of it focuses on intervals with positive carbon iso­
(Patterson and Walter, 1994; Saltzman and Thomas, 2012). With smaller topic excursions (e.g. GICE and HICE) or on long term trends. There are a
amounts of biomass in the terrestrial realm, this might not have been few studies that cover the same interval (Pl. aculeata – E. quadridactylus
possible in the Ordovician world. This interpretation is supported by Midcontinent Conodont Biozones) at a high enough resolution to make
results from Panchuk et al. (2006) looking at broad spatial gradients comparison of δ13C curves possible (Fig. 5). The stratigraphic interval
across the Laurentian epicontinental sea during the Katian Stage. Their with the proposed relationship between sea level and carbon isotopes at
study documented that δ13C values in the Pennsylvania to Virginia re­ Union Furnace is approximately correlated with records from Oklahoma
gion generally increased toward shore and the Taconic Highlands. (Edwards et al., 2015a, 2015b), Alabama (Quinton et al., 2016), Ten­
There are a variety of reasons why terrestrial fluxes could have nessee (Kozik et al., 2019), and West Virginia (Adiatma et al., 2019)
contributed a greater percentage of 13C into the marine realm. Near­ based on the position of conodont biozones and the nature of the carbon
shore settings could have received a greater proportion of 13C from the isotopic curves. This correlation is compelling as it suggests that a
weathering of exposed carbonates (e.g. Ainsaar et al., 2010). Meteoric similar pattern is present across eastern North America. However, these
fluids, which traditionally have lots of respired 12C, would have had comparisons cannot be made without more chronostratigraphic con­
much higher δ13CDIC values during the Late Ordovician. With no root straints and detailed sequence stratigraphic frameworks. We hope that
systems to supply respired 12C, the main source of carbon in meteoric our work will motivate future studies to pair high resolution δ13C re­
fluids would have been atmospheric CO2. Meteoric δ13CDIC values would cords with detailed sedimentological and sequence stratigraphic
have on average reflected the atmospheric δ13C values at the time frameworks to test the relationship between early land plants, sea level,
(Driese and Mora, 2001), which have been estimated to be around − 6‰ and carbon isotopes during this dynamic interval in the Late Ordovician.
to − 8‰ (Tomescu et al., 2009). The presence of high meteoric δ13CDIC
values in the Late Ordovician is supported by geochemical studies 5.5. Support for the late Sandbian “baseline shift”
constraining the carbon isotopic composition of meteoric carbonates (e.
g. Tobin and Walker, 1994; Driese and Mora, 2001; Jones et al., 2015). Our results add additional support to the interpretation that the
“baseline shift” in δ13C values in the late Sandbian could reflect a
5.4. A role for early land plants at Union Furnace and more broadly? perturbation to the global carbon reservoir possibly due to early land
plants. The most compelling evidence this study can add is that the
Regardless of which proposed mechanism is generating the inverse relationship between δ13C values and sea level (in BR 1 and BR 2) ends at
relationship between sea level and carbon isotopic trends (and it could the start of the “baseline shift” identified by Adiatma et al. (2019). This
certainly be a combination of those processes), our results do suggest a observation suggests that the ~1‰ increase observed at Union Furnace
possible role for early land plants. The progressive increase in the does reflect a change in the global carbon reservoir that is overriding the
amount of vegetation in nearshore environments may have exerted a interpreted regional signal in BR 1 and BR 2.
significant influence on silicate weathering and supplied important However, there is an alternative explanation for the decoupling of
nutrients into the marine realm (Quirk et al., 2015; Lenton et al., 2016; sea level and δ13C trends in sequences BR 3 and BR 4. These cycles occur
Porada et al., 2016). The increasing δ13C values during the highstand in deeper water settings where changes in sea level might not be able to
systems tract in sequences BR 1 and BR2 could record intervals of exert as significant of a control on carbon isotopes. In these deep-water
enhanced marine primary productivity due to the influx of those plant settings, the shoreline (and accompanying terrestrial fluxes) is far
released nutrients. The early terrestrial biosphere could also have removed and development of restricted water masses is less likely. In
contributed to an inverse spatial gradient in δ13CDIC values by suppling a fact, identified facies and δ13C values are correlated, with inner shelf
higher percentage of 13C to nearshore marine settings. Terrestrial settings on average 0.55‰ (σ = 0.51‰) and mid-shelf to outer ramp
organic carbon in the Late Ordovician is estimated to have had higher settings on average at − 0.64‰ (σ = 0.58‰). A student t-test indicates
δ13C values than marine organic matter. The average δ13C value of that the difference in δ13C values between environments is statistically
marine organic matter during the Late Ordovician is estimated to be ~ significant (supplementary materials file). It should be noted that this
− 28‰, while estimated values for liverworts at the time range from increase in baseline δ13C values has been observed at the margin of the
− 26‰ to − 27‰ (Tomescu et al., 2009). Alternations between recycling epicontinental sea where records are thought to reflect the open ocean
marine and terrestrial organic matter in shallow water settings could signal (Quinton et al., 2016). This observation supports the interpreta­
explain some of the 1.5‰ shift we observe between transgressions and tion that the late Sandbian “baseline shift” does reflect a change in the
regression at Union Furnace. global carbon cycle.
If this relationship between early land plants, sea level, and carbon
isotopes exists, the Late Ordovician would reflect a unique interval in 6. Conclusions
Earth’s history where the limited size and extent of the terrestrial
biosphere allowed it to impact shallow carbonate settings in ways not By coupling a high resolution δ13C record to the detailed sedimen­
observed before or after. The temporary nature of the inverse relation­ tological and sequence stratigraphic framework, we demonstrate that
ship between sea level and carbon isotopes documented at Union sea level and local/regional processes can influence carbon isotopic
Furnace is supported by Cambrian-Middle Ordovician δ13C records that patterns. The observed decrease in δ13C values in the TST and increase in
display the predicted relationship with sea level (Buggisch et al., 2003; values in the HST suggests that sea level was controlling carbon isotopic
Schiffbauer et al., 2017). In both intervals, carbon isotopic values in­ values during the BR 1 and BR 2 sequences. This relationship likely re­
crease during transgressions and decrease during regression. Buggisch flects a combination of factors, but we argue that during regressions,
et al. (2003) interpreted this relationship to reflect changes in the marine organic carbon burial increased due to enhanced nutrient supply
relative proportions of marine organic carbon remineralization verses from nearshore terrestrial environments. The ratio of meteoric fluids
carbonate deposition. By the Late Ordovician, with the ever-increasing and terrestrial organic carbon was much higher when sea level was low
influx of terrestrial organic carbon enriched in 13C, land plants could and the shoreline closer. In the Late Ordovician both meteoric fluids and
have influenced nearshore settings and altered the relationship between organic carbon would have been enriched in 13C relative to the modern.
sea level and carbon isotopic trends. Consequently, nearshore environments would have had higher δ13CDIC
The proposed role of early land plants is uncertain and needs to be values than distal settings. This is the opposite of modern shallow car­
tested. The first step should be to determine if a similar relationship bonate settings where the abundance of vascular plants contributes large

9
P.C. Quinton et al. Palaeogeography, Palaeoclimatology, Palaeoecology 574 (2021) 110440

Fig. 5. Regional comparison of bulk carbonate carbon isotopic trends for the late Sandbian. Correlation is approximate based on the position of the Pl. aculeata and
E. quadridactylus Midcontinent Conodont Biozone boundary and the correlation of that boundary to the B. gerdae North Atlantic Conodont Biozone based on Cooper
and Sadler (2012). Sections include the Interstate-35 section in Oklahoma (Edwards et al., 2015a, 2015b), the Tidwell Hollow section in Alabama (Quinton et al.,
2016), Evans Ferry section in Tennessee (Kozik et al., 2019), the Germany Valley section in West Virginia (Adiatma et al., 2019), and the Union Furnace section in
Pennsylvania (this study). The blue shaded region represents the stratigraphic interval in the Union Furnace section where sea level is influencing δ13C trends. The
green shaded region is the baseline shift identified by Adiatma et al. (2019). These intervals are approximately identified in the other sections based on the position of
conodont biozones and the nature of the carbon isotopic curves. North Atlantic Conodont Biozones for Oklahoma and Tennessee are as identified in Edwards et al.,
2015a, 2015b) and Kozik et al. (2019) respectively. Midcontinent Conodont Biozones for Alabama and West Virginia are as identified in Quinton et al. (2016) and
Adiatma et al. (2019) respectively. (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)

quantities of 12C. Declaration of Competing Interest


Our results also suggest that non-vascular plants could have played a
significant role in the Ordovician carbon cycle by supplying nutrients None.
and organic carbon enriched in 13C. This study adds support to the idea
that the late Sandbian increase in δ13C values reflects a change in the Acknowledgements
global carbon cycle. We demonstrate that the relationship between sea
level and carbon isotopes at Union Furnace ends when this “baseline This study was funded by the SUNY Potsdam Presidential Scholars
shift” begins. We argue that this observation could reflect the fact that program, the Kilmer Fund, and the Neil R. O’Brien & William T.
the global signal is overriding the local/regional signal observed in se­ Kirchgasser Undergraduate Research Fund at SUNY Potsdam. This
quences BR 1 and BR 2. Furthermore, the lack of relationship between manscript benefited from thoughtful reviews by Editor Thomas Algeo,
sea level and δ13C values in this part of the section (above 61.5 m) in­ William McMahon, and one anonymous reviwer. We thank Emily Downs
dicates that sea level is not driving the “baseline shift”. In short, there is for assistance with field work, Kenneth MacLeod for analyses at the
a perturbation to the global carbon cycle that falls at a time when land University of Missouri Stable Isotope and Biogeochemistry Laboratory,
plants were increasing in abundance and diversity. As suggested by and Sheetz Inc. for logistical support.
previous studies, there is likely a cause-and-effect relationship and we
are seeing land plants’ first impact on the global carbon cycle. Appendix A. Supplementary data
Finally, these results and interpretations are only possible because of
detailed sampling coupled with sedimentological observations. Sam­ Supplementary data to this article can be found online at https://doi.
pling at regular intervals that are not directly tied to sedimentological org/10.1016/j.palaeo.2021.110440.
observations can show general trends. Only by targeting genetically
related packages of rock and using a sliding sampling resolution can one References
capture δ13C patterns within those packages and constrain the rela­
tionship between carbon isotopic trends and environmental processes. Adiatma, Y.D., Saltzman, M.R., Young, S.A., Griffith, E.M., Kozik, N.P., Edwards, C.T.,
Leslie, S.A., Bancroft, A.M., 2019. Did early land plants produce a stepwise change in
atmospheric oxygen during the Late Ordovician (Sandbian ~485 Ma)? Palaeogeogr.
Palaeoclimatol. Palaeoecol. 534 https://doi.org/10.1016/j.palaeo.2019.109341.
Ainsaar, L., Kaljo, D., Martma, T., Meidla, T., Männik, P., Nõlvak, J., Tinn, O., 2010.
Middle and Upper Ordovician carbon isotope chronostratrigraphy in Baltoscandia: a

10
P.C. Quinton et al. Palaeogeography, Palaeoclimatology, Palaeoecology 574 (2021) 110440

correlation standard and clues to environmental history. Palaeogeogr. Eltom, H.A., Gonzalez, L.A., Hasiotis, S.T., Rankey, E.C., Cantrell, D.L., 2018.
Palaeoclimatol. Palaeoecol. 294, 189–201. https://doi.org/10.1016/j. Paleogeographic and paleo-oceanographic influences on carbon isotope signatures:
palaeo.2010.01.003. Implications for global and regional correlation, Middle-Upper Jurassic of Saudi
Akoglu, H., 2018. User’s guide to correlation coefficients. Turkish J. Emerg. Med. 18 (3), Arabia. Sediment. Geol. 364, 89–102. https://doi.org/10.1016/j.
91–93. https://www.sciencedirect.com/science/article/pii/S2452247318302164. sedgeo.2017.12.011.
Algeo, T.J., Scheckler, S.E., 1998. Terrestrial-marine teleconnections in the Devonian: Fanton, K.C., Holmden, C., 2007. Sea-level forcing of carbon isotope excursions in epeiric
links between the evolution of land plants, weathering processes, and marine anoxic seas: implications for chemostratigraphy. Can. J. Earth Sci. 44, 807–818. https://doi.
events. Philos. Trans. R. Soc. Lond. 353, 113–130. https://doi.org/10.1098/ org/10.1139/e06-122.
rstb.1998.0195. Gensel, P.G., Kotyk, M.E., Basinger, J.F., 2001. Morphology of above- and below-ground
Algeo, T.J., Berner, R.A., Maynard, J.B., Scheckler, S.E., 1995. Late Devonian oceanic structures in Early Devonian (Pragian–Emsian) Plants. In: Gensel, P.G., Edwards, D.
anoxic events and biotic crises: “rooted” in the evolution of vascular land plants. GSA (Eds.), Plants Invade the Land: Evolutionary and Environmental Perspectives.
Today 5 (3), 45–66. https://doi.org/10.7312/gens11160.
Allan, J.R., Matthews, R.K., 1982. Isotope signatures associated with early meteoric Gerrienne, P., Gensel, P.G., Strullu-Derrien, C., Lardeux, H., Steemans, P., Prestianni, C.,
diagenesis. Sedimentology 29, 797–817. https://doi.org/10.1111/j.1365- 2011. A simple type of wood in two Early Devonian plants. Science 333 (6044), 837.
3091.1982.tb00085.x. https://doi.org/10.1126/science.1208882.
Banner, J.L., Hanson, G.N., 1990. Calculation of simultaneous isotopic and trace element Glasspool, I.J., Scott, A.C., 2010. Phanerozoic concentrations of atmospheric oxygen
variations during water–rock interaction with applications to carbonate diagenesis. reconstructed from sedimentary charcoal. Nat. Geosci. 3 (9), 627–630. https://doi.
Geochim. Cosmochim. Acta 54, 3123–3137. https://doi.org/10.1016/0016-7037 org/10.1038/ngeo923.
(90)90128-8. Hayes, J.M., Popp, B.N., Takigiku, R., Johnson, M.W., 1989. An isotopic study of
Bergström, S.M., Chen, X., Gutierrez-Marco, J.C., Dronov, A., 2009. The new biogeochemical relationships between carbonates and organic carbon in the
chronostratigraphic classification of the Ordovician System and its relations to major Greenhorn Formation. Geochim. Cosmochim. Acta 53, 2961–2972. https://doi.org/
regional series and stages and the δ13C chemostratigraphy. Lethaia 42, 97–107. 10.1016/0016-7037(89)90172-5.
https://doi.org/10.1111/j.1502-3931.2008.00136.x. Holland, S.M., Patzkowsky, M.E., 1996. Sequence stratigraphy and long-term
Bergström, S.M., Schmitz, B., Saltzman, M.R., Huff, W.D., 2010. The Upper Ordovician paleoceanographic changes in the Middle and Upper Ordovician of the eastern
Guttenberg δ13C excursion (GICE) in North American and Baltoscandia: occurrence, United States. Geol. Soc. Am. Spec. Pap. 306, 117–128.
chronostratigraphic significance, and paleoenvironmental relationships. In: Holmden, C., Creaser, R.A., Muehlenbachs, K., Leslie, S.A., Bergström, S.M., 1998.
Finney, S.C., Berry, W.B.N. (Eds.), The Ordovician System, The Geological Society of Isotopic evidence for geochemical decoupling between ancient epeiric seas and
America Special Paper 466, pp. 37–67. bordering oceans: Implications for secular curves. Geology 26 (6), 567–570. https://
Bergström, S.M., Saltzman, M.R., Leslie, S.A., Ferretti, A., Young, S.A., 2015. Trans- doi.org/10.1130/0091-7613(1998)026<0567:IEFGDB>2.3.CO;2.
Atlantic application of the Baltic Middle and Upper Ordovician carbon isotope Immenhauser, A., Della Prorta, G., Kenter, J.A.M., Bahamonde, J.R., 2003. An alternative
zonation. Est. J. Earth Sci. 64, 8–12. model for positive shifts in shallow-marines carbonate δ13C and δ18O. Sedimentology
Berner, R.A., 1998. The carbon cycle and carbon dioxide over Phanerozoic time: the role 50, 953–959. https://doi.org/10.1046/j.1365-3091.2003.00590.x.
of land plants. Philos. Trans. R. Soc. Lond. B Biol. Sci. 353 (1365), 75–82. https:// James, N.P., Jones, B., 2015. Origin of carbonate sedimentary rocks. Am. Geophys.
doi.org/10.1098/rstb.1998.0192. Union 464. ISBN: 978-1-118-65270-1.
Boucot, A.J., Gray, J., 2001. A critique of Phanerozoic climatic models involving changes Jarvis, I., Gale, A.S., Jenkyns, H.C., Pearce, M.A., 2006. Secular variation in Late
in the CO2 content of the atmosphere. Earth Sci. Rev. 56 (1–4), 1–159. https://doi. Cretaceous carbon isotopes: a new δ13C carbonate reference curve for the
org/10.1016/S0012-8252(01)00066-6. Cenomanian-Campanian (99.6-70.6 Ma). Geol. Mag. 143 (5), 561–608. https://doi.
Brett, C.E., Baird, G.C., 2002. Revised stratigraphy of the Trenton Group in its type area, org/10.1017/S0016756806002421.
central New York State: sedimentology and tectonics of a Middle Ordovician shelf-to- Jenkyns, H.C., 1996. Relative sea-level change and carbon isotopes: data from the Upper
basin succession. Phys. Chem. Earth, Parts A/B/C 27 (1–3), 231–263. https://doi. Jurassic (Oxfordian) of central and Southern Europe. Terra Nova 8, 75–85. https://
org/10.1016/S1474-7065(01)00007-9. doi.org/10.1111/j.1365-3121.1996.tb00727.x.
Brett, C.E., McLaughlin, P.I., Cornell, S.R., Baird, G.C., 2004. Comparative sequence Jones, D.S., Creel, R.C., Rios, B., Santiago Ramos, D.P., 2015. Chemostratigraphy of an
stratigraphy of two classic Upper Ordovician successions, Trenton Shelf (New Ordovician–Silurian carbonate platform: δ13C records below glacioeustatic exposure
York–Ontario) and Lexington Platform (Kentucky–Ohio): implications for eustasy surfaces. Geol 43 (1), 59–62.
and local tectonism in eastern Laurentia. Palaeogeogr. Palaeoclimatol. Palaeoecol. Katz, D.A., Buoniconti, M.R., Montañez, I.P., Swart, P.K., Eberli, G.P., Smith, L.B., 2007.
210 (2–4), 295–329. https://doi.org/10.1016/j.palaeo.2004.02.038. Timing and local perturbations to the carbon pool in the lower Mississippian
Brett, C.E., Aucoin, C.D., Dattilo, B.F., Freeman, R.L., Hartshorn, K.R., McLaughlin, P.I., Madison Limestone, Montana and Wyoming. Palaeogeogr. Palaeoclimatol.
Schwalbach, C.E., 2020. Revised sequence stratigraphy of the upper Katian Stage Palaeoecol. 256, 231–253. https://doi.org/10.1016/j.palaeo.2007.02.048.
(Cincinnatian) strata in the Cincinnati Arch reference area: Geological and Kennedy, K.L., Gibling, M.R., Eble, C.F., Gastaldo, R.A., Gensel, P.G., Werner-
paleontological implications. Palaeogeogr. Palaeoclimatol. Palaeoecol. 540, 109483. Zwanziger, U., Wilson, R.A., 2013. Lower Devonian coaly shales of northern New
https://doi.org/10.1016/j.palaeo.2019.109483. Brunswick, Canada: plant accumulations in the early stages of Terrestrial
Buggisch, W., Keller, M., Lehnert, O., 2003. Carbon isotope record of Late Cambrian to colonization. J. Sediment. Res. 83 (12), 1202–1215. https://doi.org/10.2110/
Early Ordovician carbonates of the Argentine Precordillera. Palaeogeogr. jsr.2013.86.
Palaeoclimatol. Palaeoecol. 195 (3–4), 357–373. Kenrick, P., Wellman, C.H., Schneider, H., Edgecombe, G.D., 2012. A timeline for
Burdige, D.J., 2005. Burial of terrestrial organic matter in marine sediments: A re- terrestrialization: consequences for the carbon cycle in the Palaeozoic. Philosoph.
assessment. Glob. Biogeochem. Cycles 19 (4). https://doi.org/10.1029/ Transact. Royal Society B 367 (1588), 519–536. https://doi.org/10.1098/
2004GB002368. rstb.2011.0271.
Cooper, R.A., Sadler, P.M., 2012. The Ordovician Period. In: Gradstein, F.M., Ogg, J.G., Kleinhans, M.G., de Vries, B., Braat, L., van Oorschot, M., 2018. Living landscapes:
Schimtz, M.D., Ogg, G.M. (Eds.), The Geologic Time Scale 2012. Elsevier, muddy and vegetated floodplain effects on fluvial pattern in an incised river. Earth
Amsterdam, pp. 489–523. Surf. Process. Landf. 43 (14), 2948–2963. https://onlinelibrary.wiley.com/doi/full/
Dahl, T.W., Arens, S.K., 2020. The impacts of land plant evolution on Earth’s climate and 10.1002/esp.4437.
oxygenation state–An interdisciplinary review. Chem. Geol. 547 https://doi.org/ Kozik, N.P., Young, S.A., Bowman, C.N., Saltzman, M.R., Them II, T.R., 2019.
10.1016/j.chemgeo.2020.119665. Middle–Upper Ordovician (Darriwilian–Sandbian) paired carbon and sulfur isotope
Davies, N.S., Gibling, M.R., 2010. Cambrian to Devonian evolution of alluvial systems: stratigraphy from the Appalachian Basin, USA: Implications for dynamic redox
The sedimentological impact of earliest land plants. Earth Sci. Rev. 98 (3–4), conditions spanning the peak of the Great Ordovician Biodiversification Event.
171–200. https://doi.org/10.1016/j.earscirev.2009.11.002. Palaeogeogr. Palaeoclimatol. Palaeoecol. 520, 188–202.
Driese, S.G., Mora, C.I., 2001. Diversification of Siluro-Devonian plant traces in paleosols Kump, L.R., Arthur, M.A., 1999. Interpreting carbon-isotope excursions: carbonates and
and influence on estimates of paleoatmospheric CO2 levels. In: Gensel, P.G., organic matter. Chem. Geol. 161, 181–198. https://doi.org/10.1016/S0009-2541
Edwards, D. (Eds.), Plants Invade the Land: Evolutionary and Environmental (99)00086-8.
Perspectives, pp. 237–253. https://doi.org/10.7312/gens11160. Kump, L.R., Brantley, S.L., Arthur, M.A., 2000. Chemical weathering, atmospheric CO2,
Droser, M.L., Bottjer, D.J., 1986. A semiquantitative field classification of ichnofabric. and climate. Annu. Rev. Earth Planet. Sci. 28 (1), 611–667. https://doi.org/
J. Sediment. Res. 4 (1986), 558–559. https://pubs.geoscienceworld.org/sepm/ 10.1146/annurev.earth.28.1.611.
jsedres/article-abstract/56/4/558/113776/A-semiquantitative-field-classification- Laughrey, C.D., Kostelnik, J., Doden, A.G., Gold, D.P., Harper, J.A., Morrow, L.,
of?redirectedFrom=PDF. Pedlow, G., 2017. Trenton and black river carbonates in the union furnace roadcut,
Edwards, D., Feehan, J., Smith, D.G., 1983. A late Wenlock flora from Co. Tipperary, state route 453, Huntingdon County, Pennsylvania. In: 82nd Annual Field Conference
Ireland. Bot. J. Linn. Soc. 86 (1–2), 19–36. https://doi.org/10.1111/j.1095- of Pennsylvania Geologists Pre-conference Field Trip Guidebook, pp. 1–56.
8339.1983.tb00715.x. Lenton, T.M., Crouch, M., Johnson, M., Pires, N., Dolan, 2012. First plants cooled the
Edwards, C.T., Saltzman, M.R., Leslie, S.A., Bergström, S.M., Sedlacek, A.R., Howard, A., Ordovician. Nat. Geosci. 5, 86–89. https://doi.org/10.1038/ngeo1390.
Bauer, J.A., Sweet, W.C., Young, S.A., 2015a. Strontium isotope (87Sr/86Sr) Lenton, T.M., Dahl, T.W., Daines, S.J., Mills, B.J., Ozaki, K., Saltzman, M.R., Porada, P.,
stratigraphy of Ordovician bulk carbonate: implications for preservation of primary 2016. Earliest land plants created modern levels of atmospheric oxygen. Proc. Natl.
seawater values. Bulletin 127 (9–10), 1275–1289. https://doi.org/10.1130/ Acad. Sci. 113 (35), 9704–9709. https://doi.org/10.1073/pnas.1604787113.
B31149.1. Malkowski, K., Racki, G., 2009. A global biogeochemical perturbation across the
Edwards, D., Cherns, L., Raven, J.A., 2015b. Could land-based early photosynthesizing Silurian–Devonian boundary: Ocean–continent–biosphere feedbacks. Palaeogeogr.
ecosystems have bioengineered the planet in mid-Palaeozoic times? Palaeontology Palaeoclimatol. Palaeoecol. 276, 244–254. https://doi.org/10.1016/j.
58 (5), 803–837. https://doi.org/10.1111/pala.12187. palaeo.2009.03.010.

11
P.C. Quinton et al. Palaeogeography, Palaeoclimatology, Palaeoecology 574 (2021) 110440

Marshall, J.D., 1992. Climatic and oceanographic isotopic signals from the carbonate Schrag, D.P., Higgins, J.A., Macdonald, F.A., Johnston, D.T., 2013. Authigenic carbonate
rock record and their preservation. Geol. Mag. 129 (2), 143–160. https://doi.org/ and the history of the global carbon cycle. Science 339 (February), 540–543. https://
10.1017/S0016756800008244. doi.org/10.1126/science.1229578.
McMahon, W.J., Davies, N.S., 2018. Evolution of alluvial mudrock forced by early land Scotese, C.R., 2004. A continental drift flipbook. J. Geol. 112, 729–742. https://www.
plants. Science 359 (6379), 1022–1024. https://science.sciencemag.org/content journals.uchicago.edu/doi/abs/10.1086/424867.
/359/6379/1022.abstract. Sell, B.K., Samson, S.D., Mitchell, C.E., MacLaughlin, P.J., Koenig, A.E., Leslie, S.A.,
Melchin, M.J., Holmden, C., 2006. Carbon isotope chemostratigraphy of the llandovery 2015. Stratigraphic correlations using trace elements in apatite from Late Ordovician
in arctic canada: implications for global correlation and sea-level change. gff 128, (Sandbian-Katian) K-bentonites of eastern North American. Geol. Soc. Am. Bull. 127,
173–180. https://doi.org/10.1080/11035890601282173. 1259–1274. https://doi.org/10.1130/B31194.1.
metzger, g.j., fike, d.a., 2013. Techniques for assessing spatial heterogeneity of carbonate Sonnenfeld, M.D., 1996. Sequence evolution and hierarchy within the lower
δ13c values: implications for craton-wide isotope gradients. Sedimentology 60, Mississippian Madison Limestone of Wyoming. In: Longman, M.W., Sonnenfeld, M.D.
1405–1431. https://doi.org/10.1111/sed.12033. (Eds.), Paleozoic systems of the Rocky Mountain region. Society for Economic
Morris, J.L., Puttick, M.N., Clark, J.W., Edwards, D., Kenrick, P., Pressel, S., Wellman, C. Paleontologist and Mineralogist (Society for Sedimentary Geology) Rocky Mountain
H., Yang, Z., Schneider, H., Donoghue, P.C., 2018. The timescale of early land plant Section, pp. 165–192.
evolution. Proc. Natl. Acad. Sci. 115 (10), E2274–E2283. https://doi.org/10.1073/ Stein, W.E., Berry, C.M., Morris, J.L., Hernick, L.V., Mannolini, F., Ver Straeten, C.,
pnas.1719588115. Landing, E., Marshall, J.E.A., Wellman, C.H., Beerling, D., Leake, J.R., 2020. Mid-
Panchuk, K.M., Holmden, C., Kump, L.R., 2005. Sensitivity of the epeiric sea carbon Devonian Archaeopteris roots signal revolutionary change in earliest fossil forests.
isotope record to local-scale carbon cycle processes: tales from the Mohawkian Sea. Curr. Biol. 30, 1–11. https://doi.org/10.1016/j.cub.2019.11.067.
Palaeogeography, Paleaclimatology, Palaeoecology 228, 320–337. https://doi.org/ Swart, P.K., 2008. Global synchronous changes in the carbon isotopic composition of
10.1016/j.palaeo.2005.06.019. carbonate sediments unrelated to changes in the global carbon cycle. Proc. Nation.
Panchuk, K.M., Holmden, C.E., Leslie, S.A., 2006. Local controls on carbon cycling in the Acad. Sci. 105 (37), 13741–13745.
Ordovician Midcontinent region of North America, with implications for carbon Swart, P.K., Eberli, G., 2005. The nature of the δ13C of periplatform sediments:
isotope secular curves. J. Sediment. Res. 76, 200–211. https://doi.org/10.2110/ implications for stratigraphy and the global carbon cycle. Sediment. Geol. 175,
jsr.2006.017. 115–129. https://doi.org/10.1016/j.sedgeo.2004.12.029.
Patchen, D.G., Hickman, J.B., Harris, D.C., Drahovzal, J.A., Lake, P.D., Smith, L.B., Swart, P.K., Oehlert, A.M., 2018. Revised interpretations of stable C and O patterns in
Nyahay, R., Schulze, R., Riley, R.A., Baranoski, M.T., Wickstrom, L.H., Laughrey, C. carbonate rocks resulting from meteoric diagenesis. Sediment. Geol. 364, 14–23.
D., Kostelnik, J., Harper, J.A., Avary, K.L., Bocan, J., Hohn, M.E., McDowell, R., https://doi.org/10.1016/j.sedgeo.2017.12.005.
2006. A geologic play book for Trenton-Black River Appalachian Basin exploration. Swart, P.K., Reijmer, J.J.G., Otto, T., 2009. A reevaluation of facies on Great Bahama
Department of Energy (DOE Award Number DE-FC26-03NT41856) Report. https Bank II: variations in the δ13C, δ18O and mineralogy of surface sediments. Int. Assoc.
://www.netl.doe.gov/sites/default/files/2018-04/NT41856-Final-Rpt.pdf. Sedimentol. Spec. Publ. 41, 47–59.
Patterson, W.P., Walter, L.M., 1994. Depletion of 13C in seawater ΣCO2 on modern Tobin, J., Walker, K.R., 1994. Meteoric diagenesis below a submerged platform:
carbonate platforms: Significance for the carbon isotopic record of carbonates. implications for δ13C compositions prior to pre-vascular plant evolution, Middle
Geology 22, 885–888. https://doi.org/10.1130/0091-7613(1994)022%3C0885: Ordovician, Alabama, USA. Sediment. Geol. 90 (1–2), 95–111. https://doi.org/
DOCISC%3E2.3.CO;2. 10.1016/0037-0738(94)90019-1.
Porada, P., Lenton, T.M., Pohl, A., Weber, B., Mander, L., Donnadieu, Y., Beer, C., Tomescu, A.M.F., Rothwell, G.W., 2006. Wetlands before tracheophytes: Thalloid
Pöschl, U., Kleidon, A., 2016. High potential for weathering and climate effects of terrestrial communities of the Early Silurian Passage Creek biota (Virginia). In:
non-vascular vegetation in the Late Ordovician. Nat. Commun. 7 (1), 1–13. https:// Greb, S.F., Dimichele, W.A. (Eds.), Wetlands through time, Special Papers-Geological
doi.org/10.1038/ncomms12113. Society of American, 399, p. 41.
Quinton, P.C., Herrmann, A.D., Leslie, S.A., MacLeod, K.G., 2016. Carbon cycling across Tomescu, A.M., Pratt, L.M., Rothwell, G.W., Strother, P.K., Nadon, G.C., 2009. Carbon
the southern margin of Laurentia during the Late Ordovician. Palaeogeogr. isotopes support the presence of extensive land floras pre-dating the origin of
Palaeoclimatol. Palaeoecol. 458, 63–76. https://doi.org/10.1016/j. vascular plants. Palaeogeogr. Palaeoclimatol. Palaeoecol. 283 (1–2), 46–59. https://
palaeo.2015.08.020. doi.org/10.1016/j.palaeo.2009.09.002.
Quirk, J., Leake, J.R., Johnson, D.A., Taylor, L.L., Saccone, L., Beerling, D.J., 2015. Torsvik, T.H., Cocks, R.M., 2013. New global palaeogeographical reconstructions for the
Constraining the role of early land plants in Palaeozoic weathering and global Early Palaeozoic and their generation. In: Harper, D.A.T., Servais, T. (Eds.), Early
cooling. Proc. R. Soc. B Biol. Sci. 282 (1813), 20151115. https://doi.org/10.1098/ Palaeozoic Biogeography and Palaeogeography. Geological Society of London,
rspb.2015.1115. Memoirs 38, pp. 5–24. https://mem.lyellcollection.org/content/38/1/5.short.
Read, J.F., 1985. Carbonate platform facies models. Am. Assoc. Pet. Geol. Bull. 66, Walker, J.D., Geissman, J.W., Bowring, S.A., Babcock, L.E., 2018. Geologic Time Scale v.
860–878. https://doi.org/10.1306/AD461B79-16F7-11D7-8645000102C1865D. 5.0. Geological Society of America. https://doi.org/10.1130/2018.CTS005R3C.
Rubinstein, C.V., Gerrienne, P., de la Puente, G.S., Astini, R.A., Steemans, P., 2010. Early Weissert, H., 1989. C-isotope stratigraphy, a monitor of paleoenvironmental change: A
Middle Ordovician evidence for land plants in Argentina (eastern Gondwana). New case study from the Early Cretaceous. Surv. Geophys. 10, 1–61. https://link.springer.
Phytol. 188 (2), 365–369. https://nph.onlinelibrary.wiley.com/doi/full/10.1111/ com/article/10.1007/BF01901664.
j.1469-8137.2010.03433.x. Young, S.A., Saltzman, M.R., Ausich, W.I., Desrochers, A., Kaljo, D., 2010. Did changes in
Saltzman, M.R., Edwards, C.T., 2017. Gradients in the carbon isotopic composition of atmospheric CO2 coincide with latest Ordovician glacial–interglacial cycles?
Ordovician shallow water carbonates: A potential pitfall in estimates of ancient CO2. Palaeogeogr. Palaeoclimatol. Palaeoecol. 296 (3–4), 376–388.
Earth Planet. Sci. Lett. 464, 46–54. https://doi.org/10.1016/j.epsl.2017.02.011. Young, S.A., Saltzman, M.R., Bergström, S.M., 2005. Upper Ordovician (Mohakian)
Saltzman, M.R., Thomas, E., 2012. Carbon isotope stratigraphy. In: Gradstein, F.M., carbon isotope (δ13C) stratigraphy in eastern and central North America: regional
Ogg, J.G., Schmitz, M., Ogg, G. (Eds.), The Geologic Time Scale, pp. 207–232. expression of a perturbation of the global carbon cycle. Palaeogeogr. Palaeoclimatol.
https://doi.org/10.1016/B978-0-444-59425-9.00011-1. Palaeoecol. 222, 53–76. https://doi.org/10.1016/j.palaeo.2005.03.008.
Schiffbauer, J.D., Huntley, J.W., Fike, D.A., Jeffrey, M.J., Gregg, J.M., Shelton, K.L., Young, S.A., Saltzman, M.R., Bergström, S.M., Leslie, S.A., Xu, C., 2008. Paired δ13Ccarb
2017. Decoupling biogeochemical records, extinction, and environmental change and δ13Corg records of Upper Ordovician (Sandbian-Katian) carbonates in North
during the Cambrian SPICE event. Sci. Advanc. 3 (3), e1602158. America and China: implications for paleoceanographic change. Palaeogeogr.
Palaeoclimatol. Palaeoecol. 270, 166–178.

12

You might also like