LaGrange Et Al 2020
LaGrange Et Al 2020
Earth-Science Reviews
journal homepage: www.elsevier.com/locate/earscirev
A R T I C LE I N FO A B S T R A C T
Keywords: Sequence stratigraphy is commonly used to understand basin history and the distribution of conventional re-
Sequence stratigraphy servoir facies. Establishing a sequence stratigraphic framework in organic-rich mudstone successions is chal-
Chemostratigraphy lenging because macroscale sedimentological and petrophysical variations can be subtle, while biostratigraphic
Organic-rich mudstone and seismic data may be unavailable or of limited use. For these reasons, it is becoming increasingly common for
Black shale
chemostratigraphic profiles to be integrated with other datasets to facilitate sequence stratigraphic interpreta-
Geochemical proxies
tion. This paper summarizes the whole-rock inorganic geochemical proxies relevant to sequence stratigraphic
analysis in fine-grained, organic-rich marine units and reviews studies that have incorporated chemostrati-
graphic trends for sequence stratigraphy. This synthesis demonstrates that chemostratigraphic datasets are
useful in identification of transgressive-regressive cycles, allowing for a preliminary summary of the chemos-
tratigraphic characteristics of the maximum flooding surface, maximum regressive surface, transgressive systems
tract, and regressive systems tract to be established based on existing work. A preliminary synthesis of the
chemostratigraphic characteristics of the highstand systems tract is also possible for highstand systems tracts
recognized using other criteria. However, a chemostratigraphic means of identifying the correlative conformity
and basal surface of forced regression in order to subdivide the regressive systems tract into the lowstand systems
tract, falling-stage systems tract, and highstand systems has not yet been demonstrated. Further work is also
required in order to establish the differences in the chemostratigraphic signature of surfaces and systems tract
depending on the depositional setting. Chemostratigraphic proxies are an emergent and promising tool for the
identification of cyclicity in organic-rich mudstone intervals, which will become increasingly useful as further
research is conducted on the topic.
⁎
Corresponding author.
E-mail address: maya1@ualberta.ca (M.T. LaGrange).
https://doi.org/10.1016/j.earscirev.2020.103137
Received 4 July 2019; Received in revised form 13 February 2020; Accepted 14 February 2020
Available online 19 February 2020
0012-8252/ © 2020 Elsevier B.V. All rights reserved.
M.T. LaGrange, et al. Earth-Science Reviews 203 (2020) 103137
2
M.T. LaGrange, et al. Earth-Science Reviews 203 (2020) 103137
Table 1
Primary mineralogical controls on elements commonly used as chemostratigraphic proxies in organic-rich mudstone successions.
Element Typical mineralogical affinity Reference
Al Clay and feldspar. Ratcliffe et al. (2004), Brumsack (2006), Tribovillard et al. (2006),
Piper and Calvert (2009)
Cr Chromite, clay minerals, ferromagnesian minerals. Tribovillard et al. (2006)
K Clay and feldspar. Ratcliffe et al. (2004), Turner et al. (2016)
Nb Titanium oxides, silicates, or distinct mineral phase, clay minerals. Bonjour and Dabard (1991), Dinelli et al. (2007)
Rb Clay and feldspar. Ratcliffe et al. (2004), Dinelli et al. (2007)
Th Detrital. Associated with clay minerals, accessory minerals or adsorbed to mineral surfaces. Myers and Wignall (1987), Rowe et al. (2017)
Ti Titanium oxides, chlorite, illite/mica, biotite. Ratcliffe et al. (2004), Pearce et al. (2005), El Attar and Pranter
(2016)
U Detrital or authigenic. Accessory minerals or adsorbed to mineral surfaces. Can also be adsorbed to Myers and Wignall (1987), Wignall and Twitchett (1996),
organic matter or precipitated as uranium oxides. Tribovillard et al. (2006)
Y Heavy minerals (zircon, garnet, monzanite, apatite, hornblende). Dinelli et al. (2007)
Zr Zircon. Patchett et al. (1984), Ratcliffe et al. (2004), Mongelli et al.
(2006), El Attar and Pranter (2016)
Table 2
Elemental proxies useful for interpreting systems tracts, bounding surfaces, and sequences in marine organic-rich mudstone successions.
Parameter Potential proxies Details Reference
Detrital Sediment Elements predominantly associated with Vary depending on the interval but commonly include Al, K, Zr, and e.g., Pearce et al. (2005), Ratcliffe
detrital sediment Ti. et al. (2012b), Sano et al. (2013),
Nyhuis et al. (2016), Turner et al.
(2016)
Sums of detrital elements For example, Al2O3 + TiO2 + Fe2O3 + K2O used by Ratcliffe et al. e.g., Ratcliffe et al. (2012b), Sano
(2012b) or Al2O3 + K2O + TiO2 used by Sano et al. (2013). et al. (2013)
Ti/Al Higher Ti/Al as terrigenous input increases. Chen et al. (2013)
Grain Size Si/Al If silica is primarily detrital, can reflect changes in abundance of e.g., Ratcliffe et al. (2004), Dinelli
coarse sediment supply. et al. (2007), Ratcliffe et al. (2012c)
Zr/Al May reflect changes in abundance of coarse sediment supply. e.g., Ratcliffe et al. (2004), Dinelli
et al. (2007)
Zr/Nb May reflect changes in abundance of coarse sediment supply. e.g., Ratcliffe et al. (2006), Sano et al.
(2013)
Zr/Rb May reflect changes in abundance of coarse sediment supply. e.g., Dinelli et al. (2007)
Y/Al Reflects changes in abundance of coarse sediment supply. e.g., Dinelli et al. (2007)
Y/Rb May reflect changes in abundance of coarse sediment supply. e.g., Dinelli et al. (2007)
Paleoredox V Can become enriched in anoxic to euxinic environments. Wanty and Goldhaber (1992), Calvert
and Pedersen (1993), Tribovillard
et al. (2006)
Re Can become enriched in anoxic to euxinic environments. Colodner et al. (1993), Crusius et al.
(1996), Yamashita et al. (2007)
Cr Can become enriched in anoxic to euxinic environments. Calvert and Pedersen (1993), Algeo
and Maynard (2004), Tribovillard
et al. (2006)
U Can become enriched in anoxic to euxinic environments. Myers & Wignall (1987), Wignall and
Twitchett (1996), McManus et al.
(2005), Tribovillard et al. (2006)
Mo Can become enriched in euxinic environments. Erickson and Helz (2000)
Ni Can become enriched in euxinic environments Huerta-Diaz and Morse (1992),
Calvert and Pedersen (1993),
Tribovillard et al. (2006)
Th/U Th/U < 2 reflects anoxic conditions, Th/U 2–7 suggests oxic Wignall and Twitchett (1996)
environments, and Th/U > 7 is indicative of sediment deposited
under highly oxidizing conditions.
Basin Restriction Mo/TOC Mo/TOC > 35 × 10−4 in weakly restricted settings, Algeo & Lyons (2006)
~15–35 × 10−4 in environments characterized by moderate
restriction, and < 15 × 10−4 in strongly restricted settings.
Excess Silica Peaks in Si/Al Al typically reflects clay abundance in mudstones (Sageman and Davis et al. (1999), Sageman & Lyons
Lyons, 2004; Rowe et al., 2017). (2004), Turner et al. (2015, 2016),
Gambacorta et al. (2016), Zhang et al.
(2019)
Si-Al crossplot See Fig. 2. Tribovillard et al. (2006), Rowe et al.
(2012), El Attar and Pranter (2016)
Siexcess = Sisample(Si/Al)average shale Excess silica relative to average shale. Ross and Bustin (2009), Shen et al.
(2014), Arsairai et al. (2016)
2.2. Grain size clay size fraction (e.g., Ratcliffe et al., 2004; Dinelli et al., 2007;
Ratcliffe et al., 2012c). Similarly, changes in Zr/Nb and Zr/Al can in-
Chemostratigraphic proxies can also be used to infer grain size dicate changes in the abundance of coarse sediment supply if these
variations. For instance, if the Si present in an interval is predominantly ratios are primarily controlled by variations in the abundance of silt-
detrital, trends in Si/Al can reflect variations in the proportion of silt to size detrital zircon rather than variations in provenance (Ratcliffe et al.,
3
M.T. LaGrange, et al. Earth-Science Reviews 203 (2020) 103137
2004, 2006; Dinelli et al., 2007; Sano et al., 2013). In this case, Ratcliffe euxinic conditions where Re(VII) is reduced to Re(IV) (Colodner et al.,
et al. (2004) suggested that increasing Zr/Al results from an increased 1993; Crusius et al., 1996; Yamashita et al., 2007) and precipitated as
proportion of silt-size detrital zircon compared to clay minerals, sulfides or adsorbed to organic matter (Kendall et al., 2010). In anoxic
whereas Ratcliffe et al. (2006) and Sano et al. (2013) observed that Zr/ settings, the conversion of Cr(VI) to Cr(III) leads to the latter being
Nb followed grain size trends with higher Zr/Nb in coarser grained incorporated into organic compounds or Fe(III) and Mn(IV) oxyhydr-
intervals. Additionally, in a study of fine-grained Pleistocene and Ho- oxides (Calvert and Pedersen, 1993; Algeo and Maynard, 2004).
locene sediments, Dinelli et al. (2007) showed that ratios of Zr/Rb, Zr/ Chromium has been used to interpret paleoredox conditions in some
Al, Y/Al, and Y/Rb correlated well to the coarse silt-size sediment studies (e.g., El Attar and Pranter, 2016), although the possibility of Cr
fraction. Bloemsma et al. (2012) assessed geochemical grain size loss from remineralization of organic matter in euxinic conditions
proxies for three offshore fine-grained quaternary sediment cores. Two (Algeo and Maynard, 2004) and the potential for Cr enrichment from an
of these cores were retrieved from different locations off the coast of increased proportion of Cr in detrital sediment means that this proxy
West Africa, whereas the third core was collected off the coast of Chile. should be used with caution (Tribovillard et al., 2006).
In the two cores from West Africa, Si exhibited the strongest correla- Uranium is delivered to the oceans in detrital accessory minerals
tions with mean grain size, whereas Ti showed the strongest correla- and in solution in the form of U(VI) (Myers and Wignall, 1987). Under
tions with mean grain size for core collected offshore of Chile. Based on oxic conditions, U(VI) binds with bicarbonate to form a soluble anion,
these results, Bloemsma et al. (2012) concluded that geochemical grain while reducing conditions lead to the conversion of U(VI) to U(IV) and
size proxies vary depending on the particular setting. The above studies can lead to sedimentary enrichment of U in the form of uraninite, UO2
suggest that geochemical grain size proxies should be determined for (Myers and Wignall, 1987; Klinkhammer and Palmer, 1991; Wignall
each interval on a case by case basis through comparison with grain size and Twitchett, 1996; McManus et al., 2005; Partin et al., 2013).
data. Molybdenum accumulates in sediment under euxinic conditions
(Helz et al., 1996; Tribovillard et al., 2006; Scott and Lyons, 2012;
2.3. Paleoredox Robbins et al., 2016). In seawater, Mo is present as molybdate
(MoO42−) and in the presence of H2S reacts to form particle-reactive
Variations in redox conditions changes the solubility of many ele- thiomolybdates (e.g., MoS42−) (Helz et al., 1996), which is enriched in
ments, with decreased oxygen levels at the sediment water interface sediment through complexation with organic matter or reduced com-
resulting in the enrichment of certain metals in sediment; in this paper pounds (Helz et al., 1996; Erickson and Helz, 2000). Similarly, euxinic
we focus on mudstone successions. When discussing paleoredox con- conditions result in enrichment of Ni through incorporation of NiS into
ditions, the terms oxic, suboxic, anoxic, and euxinic are used. The sulfides (Huerta-Diaz and Morse, 1992; Calvert and Pedersen, 1993).
distinction between oxic, suboxic, and anoxic has commonly been used The Th/U ratio can be used as an indicator of paleoredox conditions
to distinguish the amount of dissolved oxygen available, with different (Myers and Wignall, 1987; Wignall and Twitchett, 1996; Kimura and
authors using different concentrations (e.g., oxic, > 4.5 uM; suboxic, Watanabe, 2001; Sano et al., 2013). As previously discussed, U can
4.5 uM – 10 nM; anoxic, < 10 nM; Morrison et al., 1999; Revsbech become enriched in sediment under anoxic conditions (Myers and
et al., 2009; Tyson and Pearson, 1991). The terms are similarly used to Wignall, 1987; Wignall and Twitchett, 1996; Partin et al., 2013). In
refer to specific microbial metabolisms and the terminal electron ac- contrast, the abundance of Th in sediment is controlled by detrital input
ceptor (TEA) coupled to the oxidation of organic carbon (e.g., Froelich as it is insoluble in seawater, and consequently it does not become
et al., 1979; Canfield and Thamdrup, 2009). Thus, oxic refers to aerobic authigenically enriched (Myers and Wignall, 1987; Wignall and
respiration (where cells use O2), suboxic refers to using nitrate (NO3−), Twitchett, 1996). Owing to their differing mobility, Th/U < 2 is in-
while anoxic refers to using Mn(IV), Fe(III) or sulfate (SO42−) terpreted to reflect deposition under anoxic conditions, Th/U of 2–7
(Konhauser, 2007). Additionally, the term euxinic will be used herein to suggests oxic environments, and Th/U > 7 are indicative of sediment
describe anoxic conditions in which sulfate reduction results in the deposited under highly oxidizing conditions (Wignall and Twitchett,
presence of hydrogen sulfide in the water column. 1996).
Trace element enrichment is commonly compared to an average
shale with the enrichment factor for a given element (X) calculated as 2.4. Basin restriction
follows (Brumsack, 2006; Tribovillard et al., 2006):
In basins characterized by a degree of restriction from the global
EFX = (X /Al)sample /(X /Al)average shale (1)
oceans, Algeo & Lyons (2006) proposed that Mo/total organic carbon
An element is considered to be enriched if this ratio exceeds 1 and (TOC) is an indicator of the extent of basin isolation. In that study,
depleted if the ratio is below 1 (Brumsack, 2006; Tribovillard et al., Mo–TOC relationships of sediment were demonstrated to vary as a
2006). Alternatively, Brumsack (2006) argued that for intervals with function of deep-water dissolved Mo and deep-water renewal time in
low detrital input and, therefore low Al abundance, it is more suitable several restricted modern anoxic environments. Algeo & Lyons (2006)
to evaluate enrichment of an element (X) by considering its non-detrital observed sedimentary Mo/TOC values of > 35 × 10−4 in weakly re-
fraction, which can be calculated using the following formula: stricted settings, ~15–35 × 10−4 in environments characterized by
moderate restriction, and < 15 × 10−4 in strongly restricted settings.
Xnon − detrital = Xsample –Al sample (X /Al)average shale (2)
This trend was interpreted to reflect a growing drawdown of the aqu-
Elemental concentrations from Average Shale (Wedepohl, 1971) or eous Mo reservoir in increasingly restricted settings caused by Mo re-
Post-Archean Average Australian Shale (Taylor and McLennan, 1985) moval to sediment outpacing Mo re-supply. These relationships have
are commonly used as comparators (e.g., Ross and Bustin, 2009; Rowe been used to interpret the level of basin isolation in several studies of
et al., 2009; Zhou and Jiang, 2009; Gambacorta et al., 2016). organic-rich mudstone intervals including Harris et al. (2013), Turner
Several trace elements, including V, Re, Cr, and U, may become and Slatt (2016), and Hines et al. (2019).
enriched in anoxic to euxinic sediment. Vanadium is fixed in sediment
under mildly anoxic conditions if the insoluble hydroxide VO(OH)2 is 2.5. Biogenic sediment
formed once V(V) is reduced to V(IV) (Wanty and Goldhaber, 1992;
Calvert and Pedersen, 1993). Under euxinic conditions, sedimentary V Certain elements, elemental ratios, and cross-plots are used to es-
enrichment occurs through precipitation of V2O3 or V(OH)3 upon re- timate the proportion of biogenic silica (i.e., silica derived from or-
duction of V(IV) to V(III) (Wanty and Goldhaber, 1992; Calvert and ganisms such as diatoms, radiolarians, or siliceous sponges) relative to
Pedersen, 1993). Sedimentary enrichment of Re occurs under anoxic to other sources of silica. Understanding variations in the abundance of
4
M.T. LaGrange, et al. Earth-Science Reviews 203 (2020) 103137
these studies (e.g., Dean and Arthur, 1998; Hart, 2015) suggesting that
Ca could potentially be used as a proxy for pelagic carbonate if this
relationship is first confirmed through comparison to XRD and sedi-
mentological datasets.
5
M.T. LaGrange, et al. Earth-Science Reviews 203 (2020) 103137
type II kerogens (Lewan and Maynard, 1982; Lewan, 1984). This may Shale of the Arkoma Basin, Oklahoma. The third-order maximum
result in lower than expected enrichment of certain trace metals in flooding surface identified in this interval is characterized by a max-
organic-rich mudstone units that have undergone oil generation and imum gamma ray response, relatively high Si/Al, local minimum con-
expulsion (Harris et al., 2013). Chemical weathering also affects com- centration of terrigenous elements (Al, K, Ti, Zr). This surface also se-
position (Nesbitt and Young, 1982). For example, in a study of two parates underlying strata with lower Mo and V abundance from
organic-rich mudstone intervals from the Guizhou province in China, overlying strata displaying higher Mo and V concentrations (Fig. 3).
Liu et al. (2016) observed the loss of major elements, such as Na, Ca, Higher-frequency surfaces that mark the shift from regression to
Mg, Fe, and trace elements, including V and Ni, in outcrop samples transgression were recognized chemostratigraphically and are marked
compared to core samples. Weathering of organic matter in mudstone by local minima in terrigenous elements (e.g., Al, K, Ti, Zr) and com-
intervals has been observed to cause a pronounced breakdown of Ni and monly coincide with local maxima in Si/Al and Ca. Turner et al. (2016)
V metallo-organic complexes (Grosjean et al., 2004). In a study of the refer to these as chemostratigraphic flooding surfaces, herein, these
effects of weathering on proxies used for paleoenvironmental inter- surfaces are interpreted as higher frequency maximum flooding sur-
pretation in mudstone successions, Marynowski et al. (2017) observed faces. Ratcliffe et al. (2012c) and Sano et al. (2013) identified max-
significant decreases in the concentrations of trace metals including, imum flooding surfaces in the Upper Jurassic Haynesville Shale, an
Mo, Ni, and U, which were attributed to degradation of organic matter organic-rich mudstone deposited on the slope of a carbonate platform
and oxidation of pyrite. The depletion of Re from weathering of or- (Frébourg et al., 2013) in a restricted basin setting (Hammes et al.,
ganic-rich mudstones has also been observed (e.g., Peucker-Ehrenbrink 2011). These studies placed maximum flooding surfaces in the Hay-
and Hannigan, 2000; Jaffe et al., 2002). nesville Shale at minimums in Zr/Nb, while chemostratigraphic data
Due to complexities inherent in element and mineral distributions, presented in Sano et al. (2013) also illustrates that maximum flooding
good practice confirms element affinities using cross plots, compar- surfaces in the Haynesville Shale are characterized by minimums in V
isons, and statistical analyses. When possible, integrating chemostrati- abundance (Fig. 4). Data presented in Ratcliffe et al. (2012c) shows that
graphic interpretations with other datasets is also recommended (e.g., maximum flooding surfaces in this interval typical fall at local lows in
Algeo et al., 2004; Ver Straeten et al., 2011; Hammes and Frébourg, terrigenous content. Harris et al. (2018) presented geochemical data for
2012). cores from the Duvernay Formation — an organic-rich mudstone de-
posited in inter-reef settings during the Late Devonian in Alberta
2.7. Normalization to aluminium and the average shale (Knapp et al., 2019) — that are included in a previously established
sequence stratigraphic framework based on sedimentological and pet-
In many studies, elemental concentrations have been normalized to rophysical attributes. Harris et al. (2018) use the formula for excess Si
Al with the aim of accounting for dilution by other components and (Eq. (3)) and interpret all excess Si to be biogenic in origin, potentially
facilitating comparison between different locations and intervals because of the presence of siliceous radiolaria in several of the facies
(Pearce et al., 1999; Van der Weijden, 2002; Algeo et al., 2004; Algeo observed by Knapp et al. (2017). Maximum flooding surfaces in cores
and Maynard, 2004; Harris et al., 2013). Moreover, elements are nor- from the West Shale Basin are often near lows in Al and typically cor-
malized to Al in order to compare with average shale values, which respond to highs in excess Si and Mo/Al. In the East Shale Basin,
then provides an estimate of enrichment of those elements (Brumsack, maximum flooding surfaces are typically near minima in Al, and se-
2006; Tribovillard et al., 2006). However, Van der Weijden (2002) parate intervals of lower Al with overlying intervals characterized by
demonstrated that normalization to Al can result in apparent correla- higher Al abundance. Patterns of excess Si enrichment and Mo/Al do
tions between unrelated variables or vice versa, especially if the coef- not match those observed in the West Shale Basin, which Harris et al.
ficient of variation of Al is large compared to the coefficient of variation (2018) attributed to higher carbonate dilution, lower productivity, and
of the elements being normalized. Furthermore, Algeo & Lyons (2006) higher dissolved oxygenation in the East compared to West Shale Basin.
argued that given trace elements and TOC are both affected by dilution, Results from these studies suggest that maximum flooding surfaces
trace metals should not be normalized to Al for comparison with trends are typically characterized by relatively low levels of terrigenous ele-
in TOC. Finally, understanding the relative proportion and abundance ments (e.g., Al, K, Ti, Zr), minima in grain size proxies (e.g. Zr/Nb), and
of different elements or minerals is useful for sequence stratigraphy maxima in Si/Al (Table 3). The enrichment of redox sensitive trace
(e.g., identifying condensed sections) and normalizing to Al conflates elements (e.g., Mo, V) at these surfaces appears to be variable. Seeing as
the dataset. Consequently, normalization to Al should only be per- maximum flooding surface records a time when continental sediment is
formed if the coefficient of variation of Al is similar to that of the ele- stored in landward settings, which is accompanied by the formation of
ment of interest and it is also recommended to consider both normal- condensed sections in the marine environment (Loutit et al., 1988), a
ized profiles and raw profiles. Given its lower solubility, some trace minimum abundance of detrital elements is expected (Turner et al.,
element studies of ancient shales have instead relied on metal/Ti ratios 2016). In shallow-marine systems, a maximum flooding surface also
as a proxy for authigenic enrichments (e.g., Reinhard et al., 2013; Fru records the change from fining upwards to coarsening upwards (Embry,
et al., 2016). 2010). Peaks in Si/Al or excess Si observed by Turner et al. (2015,
2016) and Harris et al. (2018), respectively, were interpreted to reflect
3. Observed chemostratigraphic expression of surfaces a higher proportion of biogenic silica. Both studies observed increases
in biogenic silica proxies at maximum flooding surfaces, suggesting that
The geochemical expressions of sequence stratigraphic surfaces higher biogenic silica content is associated with maximum flooding
observed in fine-grained organic-rich intervals are summarized in surfaces, which is also expected in condensed sections (e.g., Bohacs,
Table 3. 1998; Gutierrez Parades et al., 2017). This has been interpreted as the
result of increased primary productivity (e.g., Harris et al., 2018), de-
3.1. Maximum flooding surface creased clastic dilution, or a combination thereof (Turner et al., 2016).
Differences in redox sensitive trace metal enrichment associated
The maximum flooding surface marks the switch from transgression with maximum flooding surfaces from study to study likely occur be-
to regression and the time of the maximum landward position of the cause of differences in the paleohydrographic characteristics of the
shoreline (Posamentier and Allen, 1999). A few studies provide ex- setting. As water depth at a point on the open shelf increases, dissolved
amples of the chemostratigraphic expression of this surface. Turner oxygen at this location typically declines (Wilde et al., 1996). In
et al. (2016) produced a sequence stratigraphic framework for two modern oceans, consumption of oxygen by degrading organic matter
outcrops and three cores in the organic-rich Upper Devonian Woodford results in the decrease of dissolved oxygen from the surface to a
6
M.T. LaGrange, et al. Earth-Science Reviews 203 (2020) 103137
Table 3
The observed chemostratigraphic expression of sequence stratigraphic surfaces in organic-rich mudstone units.
Sequence stratigraphic surface Observed chemostratigraphic signature Examples
Maximum flooding surface Low abundance of terrigenous proxies Ratcliffe et al. (2012c), Turner et al. (2015, 2016), Harris et al. (2018)
Minima in grain size proxies Ratcliffe et al. (2012c), Sano et al. (2013)
Increase in proxies for biogenic silica Turner et al. (2015, 2016), Harris et al. (2018)
Variable levels of redox proxies depending on paleohydrography Sano et al. (2013), Turner et al. (2015, 2016), Harris et al. (2018)
Maximum regressive surface Peak in terrigenous proxies Ratcliffe et al. (2012c), Harris et al. (2018)
Peak in grain size proxies Ratcliffe et al. (2012c), Sano et al. (2013)
Minima in proxies for biogenic silica Harris et al. (2018)
Variable levels of redox proxies depending on paleohydrography Sano et al. (2013), Harris et al. (2018)
Fig. 3. Chemostratigraphic profiles from the Woodford Formation in the Wyche Farm Quarry Core–1 in Pontotoc County, Oklahoma. Gamma ray log data and
chemostratigraphic profiles are presented with the interpreted transgressive-regressive cycles shown to the left of each dataset. The shoreline trajectory shown
presents the inferred local shoreline trajectory. The solid grey lines at the change from transgression to regression are herein interpreted as higher frequency
maximum flooding surfaces. Acronyms: GR–gamma ray; HST–highstand systems tract; MFS–maximum flooding surface; TST–transgressive systems tract. Reproduced
with permission from Turner et al. (2016).
minimum in the zone of lowest circulation (Wyrtki, 1962; Berry and In restricted basin settings, oxygen becomes depleted through decay
Wilde, 1978). Below this depth, dissolved oxygen increases as a result of organic matter and is not replenished because these basins lack cir-
of ventilation by cold, dense, oxygenated waters produced at high la- culation and ventilation (Berry and Wilde, 1978; Schönfeld et al.,
titudes during the formation of sea ice (Wyrtki, 1962; Berry and Wilde, 2015). If transgression results in increased connectivity to the open
1978). The depth of the oxygen minimum can intersect the seafloor ocean, oxygenation can increase from input of ventilated water and
from the shelf to the slope (Reichart et al., 1998; Helly and Levin, improved circulation (Savrda and Bottjer, 1989; Bohacs, 1998). Ad-
2004). At depths near the oxygen minimum, transgression could result ditionally, Hines et al. (2019) attributed increased bottom water oxy-
in a shift to more oxygenated conditions, whereas for positions well genation during times of transgression to lower productivity, with
above the oxygen minimum, transgression should lead to a decline in higher productivity during regression resulting from an increased
dissolved oxygen. Additionally, transgression can result in rising wave supply of terrigenous sediment and nutrients. For example, in a study of
base, causing decreased water column mixing and oxygenation, with the Woodford Shale, Turner and Slatt (2016) used Mo/TOC relation-
the opposite occurring during falling relative sea level (Harris et al., ships and a decreasing upwards trend in redox-sensitive trace metal
2013). This change in oxygen is reflected by the ichnofacies model enrichment to interpret renewed circulation and decreased basin iso-
(MacEachern et al., 2009a) with trace fossil assemblages transitioning lation resulting from transgression. Similarly, Sano et al. (2013) con-
to those made by organisms more tolerant of suboxic to anoxic condi- cluded that anoxic conditions during deposition of the Haynesville
tions as water depth increases and mixing from waves becomes less Shale, which occurred in a restricted intra-shelf basin, prevailed at the
common resulting in lower dissolved oxygen at the sediment-water onset of transgression. This conclusion has also been made using other
interface (MacEachern et al., 2009b; Dashtgard and MacEachern, datasets. For example, by comparing trace fossil assemblages to orga-
2016). nic‑carbon and carbonate content in the Upper Cretaceous Niobrara
7
M.T. LaGrange, et al. Earth-Science Reviews 203 (2020) 103137
Fig. 4. Chemostratigraphic profiles for the Upper Jurassic strata of three wells in eastern Louisiana and western Texas and associated sequence stratigraphic
interpretation. Package 1 corresponds to the Smackover Formation, package 2 comprises the Gilmer Lime Formation, and package 3 and 4 are the Haynesville
Formation. The solid Zr/Nb profile represents a moving average with n = 2 and the associated squares are the raw data. Acronyms: T-R–transgressive-regressive;
EFV–enrichment factor of Vanadium. Reproduced from Sano et al. (2013). Copyright ©2013 by The American Association of Petroleum Geologists. Reprinted by
permission of the AAPG whose permission is required for further use.
Formation (Colorado), Savrda and Bottjer (1989) interpreted intervals 3.2. Maximum regressive surface
of increased oxygenation to reflect open marine circulation in the
Western Interior Seaway during transgression. Because of differences in The maximum regressive surface occurs at the change from re-
the effect of relative sea level on water column oxygenation depending gression to transgression and marks the maximum basinward shoreline
on the setting, the relative enrichment of redox-sensitive trace metals at position (Helland-Hansen and Martinsen, 1996). Chemostratigraphic
maximum flooding surfaces depends on whether a shift from trans- examples of maximum regressive surfaces in the Haynesville Shale can
gression to regression resulted in increased or decreased water column be found in Ratcliffe et al. (2012c) and Sano et al. (2013) where they
oxygenation. are identified by maxima in Zr/Nb. The terrigenous input profile (sum
of Al, K, Na, Ti) plotted by Ratcliffe et al. (2012c) typically displays
peaks at maximum regressive surfaces, while the V enrichment profile
in Sano et al. (2013) shows that maximum flooding surfaces most often
8
M.T. LaGrange, et al. Earth-Science Reviews 203 (2020) 103137
mark the shift from increasing V enrichment to decreasing V enrich- Turner et al. (2015 and 2016) interpret a third-order transgressive
ment (Fig. 4). The chemostratigraphic data presented in Harris et al. systems tract in the organic-rich Woodford Shale. In this example, the
(2018) shows that the maximum regressive surface identified in the transgressive systems tract is characterized by an overall decline in
Duvernay Formation West Shale Basin by Knapp et al. (2019) is char- terrigenous elements (e.g., Ti, Zr, K, Al) and in redox sensitive elements
acterized by relatively high Al, minimum Mo/Al with values near 0, and (e.g., Mo, V) reflecting decreasing restriction, and often by increasing
low levels of biogenic Si. Si/Al that arises from an increasing proportion of biogenic silica
In marine shelf environments, the maximum regressive surface ty- (Fig. 3). However, in certain intervals, V is concentrated in phosphatic
pically records the shift from coarsening upwards to fining upwards nodules, resulting in V abundance peaks that are not associated with
(Catuneanu, 2002; Embry, 2010), which is reflected in these examples reduced circulation (Turner et al., 2016). Turner et al. (2016) also
by peaks in grain size proxies or detrital elements. Results from Harris identified fourth-order T-R cycles based on trends in Ti and Zr with
et al. (2018) suggest that maximum regressive surfaces are character- trends of decreasing concentration characterizing transgression.
ized by a low abundance of biogenic silica. As previously discussed, Other studies focus on the Upper Jurassic Haynesville Shale of Texas
changes in relative sea level can have variable effects on water column and Louisiana. Ratcliffe et al. (2012c) identified lower-frequency T-R
oxygenation depending on the setting. The shift from increasing to cycles in the Haynesville Formation with higher-frequency T-R cycles
decreasing V enrichment associated with maximum regressive surfaces superimposed on these trends. Higher-rank transgressive packages are
in the Haynesville Shale is likely the product of highest levels of basin characterized by a declining abundance of terrigenous content, which is
isolation at this time. Harris et al. (2018) interpreted that times of the sum of Al2O3, TiO2, Na2O, and K2O. Lower-rank transgressive in-
higher sea level during deposition of the Duvernay Formation were tervals are characterized by declining Zr/Nb and SiO2/Al2O3. Trends in
characterized by lower bottom water oxygenation because of reduced Zr/Nb and SiO2/Al2O3 follow one another, indicating that they are
water column mixing, high productivity, and inflow of oxygen depleted associated with the fraction of silt size sediment and that biogenic silica
bottom waters. This implies that minima in Mo/Al associated with the is proportionately less important in the Haynesville Formation
maximum regressive surface in the Duvernay Formation of the West (Ratcliffe et al., 2012c). This interpretation is supported by the clear
Shale Basin can be explained by highest levels of mixing, lowest pro- positive trend between Zr and Si observed in the Si–Zr cross plot for the
ductivity, and minimum inflow of oxygen depleted bottom waters. Haynesville Formation (Ratcliffe et al., 2012c). Sano et al. (2013)
identified T-R sequences in the same three localities as Ratcliffe et al.,
2012c (Fig. 4). In this case, V enrichment data is also presented, and
4. Observed Geochemical Expression of Systems Tracts
those authors interpreted that anoxia decreases throughout transgres-
sion.
The geochemical expressions of systems tracts observed in fine-
Hammes & Frebourg (2012) also studied the Haynesville Shale. In
grained organic-rich intervals are summarized in Table 4.
this dataset, the transgressive systems tract is expressed by lower Al
concentration than the overlying highstand systems tract. In contrast,
4.1. Transgressive systems tract Si/Al, Ti/Al, and Zr/Al are elevated compared to the overlying high-
stand systems tract. The authors contended that an increased abun-
The transgressive systems tract comprises strata deposited when the dance of Ti and Zr supports the interpretation that bottom waters were
rate of relative sea-level rise surpasses the rate of shoreline sediment euxinic because they are redox sensitive elements. With that said, Ti
supply (Posamentier and Allen, 1999). Examples of the chemostrati- and Zr are generally taken to reflect detrital input (Bhatia and Crook,
graphic expression of the transgressive systems tract include Turner 1986; Plank and Langmuir, 1998; Piper and Calvert, 2009), and have
et al. (2015, 2016), Sano et al. (2013), Ratcliffe et al. (2012c), Hammes not been shown by previous studies to be redox sensitive. The opposite
& Frebourg (2012), Ver Straeten et al. (2011), and Harris et al. (2018).
Table 4
The observed chemostratigraphic expression of systems tracts in fine-grained organic-rich intervals.
Systems tract Observed chemostratigraphic signature Examples
TST Decreasing Al Hammes & Frebourg (2012), Ver Straeten et al. (2011),Turner et al. (2015, 2016), Harris et al.
(2018)
Decreasing sums of terrigenous elements (e.g., Ratcliffe et al. (2012c)
Al + Ti + Na + K)
Variable trends in elements associated with heavy minerals (Ti
and Zr):
Decreases Ver Straeten et al. (2011), Turner et al. (2016)
Increases Hammes & Frebourg (2012), Ver Straeten et al. (2011)
Decreasing grain size proxies Ratcliffe et al. (2012c), Sano et al. (2013)
Elevated proxies for biogenic silica Turner et al. (2015, 2016), Harris et al. (2018)
Variable levels of redox proxies depending on paleohydrography Hammes & Frebourg (2012), Ver Straeten et al. (2011), Sano et al. (2013), Turner et al. (2015,
2016), Harris et al. (2018)
HST Increasing Al Hammes & Frebourg (2012), Turner et al. (2015, 2016), Harris et al. (2018)
Variable trends in elements associated with heavy minerals (Ti
and Zr):
Decreases Hammes & Frebourg (2012)
Increases Turner et al. (2015, 2016)
Variable levels of biogenic silica proxies Harris et al. (2018)
Variable levels of redox proxies depending on paleohydrography Turner et al. (2015, 2016), Harris et al. (2018)
RST Increasing Al Turner et al. (2015, 2016), Ver Straeten et al. (2011)
Variable trends in elements associated with heavy minerals (Ti
and Zr):
Decreases Ver Straeten et al. (2011)
Increases Turner et al. (2015, 2016), Ver Straeten et al. (2011)
Increasing grain size proxies Ratcliffe et al. (2012c), Sano et al. (2013)
Abbreviations: TST–transgressive systems tract; HST–highstand systems tract; RST–regressive systems tracts.
9
M.T. LaGrange, et al. Earth-Science Reviews 203 (2020) 103137
trends in the Al profile compared to the Si/Al, Ti/Al, and Zr/Al suggest sections develop because of a decline in the abundance of terrigenous
decoupling between the sources of Al versus Si, Ti, and Zr, which could sediment input during transgression (Loutit et al., 1988; Galloway,
potentially occur if Si, Ti, and Zr are primarily associated with the 1989). The proportion of biogenic sediment increases in these intervals
wind-blown detrital fraction. Hammes & Frebourg (2012) did propose as dilution from detrital sediment declines (Arthur and Sageman, 2005;
that the detrital fraction in the transgressive strata is likely primarily Ver Straeten et al., 2011). High levels of biogenic silica in transgressive
aeolian. The Mo enrichment profile is also elevated compared to the systems tracts can also be interpreted as the product of increased bio-
overlying highstand systems tract. logical productivity (e.g., Harris et al., 2018). Ver Straeten et al. (2011)
Ver Straeten et al. (2011) identified third-order transgressive sys- interpreted that rising levels of Si/Al and Ti/Al in the transgressive
tems tracts in the Middle Devonian Oatka Creek and Skaneateles for- systems tract of the Skaneateles Formation likely reflects higher aeolian
mations deposited in the Appalachian Basin in western New York or volcanic input.
during the Middle Devonian. These formations are both included in the Trends in redox sensitive trace metal enrichment through the
Hamilton Group and the Oatka Creek Formation is also a member of the transgressive systems tract are variable. Certain studies show de-
Marcellus Subgroup (Arthur and Sageman, 2005). At this location, the creasing paleoredox proxies (e.g., Sano et al., 2013; Ver Straeten et al.,
Oatka Creek Formation dominantly comprises organic-rich shale with a 2011; Turner et al., 2016), whereas others show increases (e.g.,
few limestone interbeds and is interpreted to reflect deposition in pri- Hammes & Frebourg, 2012; Harris et al., 2018), likely because of dif-
marily anoxic conditions (Ver Straeten et al., 2011). In this interval, the ferences between the paleohydrographic conditions of each deposi-
transgressive systems tract is characterized by a small decrease in Al tional setting.
and Ti/Al, and stable levels of Si/Al in the organic-rich mudstone in-
tervals. This systems tract also shows markedly lower Mo abundance 4.2. Highstand systems tract
than the overlying regressive strata, suggesting a restricted depositional
setting with increased connectivity to the open ocean in the trans- Strata of the highstand systems tract are deposited when the rate of
gressive systems tract. The Appalachian Basin was a restricted epicra- relative sea-level rise decreases such that the rate of sediment supply
tonic basin (Rowe et al., 2008) and, therefore, increased oxygenation equals or exceeds the rate of accommodation generation at the shore-
accompanying the transgressive systems tract is expected. The over- line (Posamentier and Allen, 1999). Examples of the geochemical ex-
lying Skaneateles Formation is composed primarily of interbedded pression of highstand systems tract deposits were presented by Turner
mudstone and organic-rich mudstone and was deposited in a dom- et al. (2015, 2016) where they interpreted a third-order highstand
inantly suboxic environment (Ver Straeten et al., 2011). In this for- systems tract at the top of the Woodford Shale characterized by in-
mation, the transgressive systems tract is expressed chemostrati- creasing Ti, Zr, Al, and K, decreasing Si/Al, and relatively low Mo and V
graphically by overall low Mo, which is expected in a suboxic interval. (Fig. 3). Similarly, Hammes and Frebourg (2011) suggested that the
Here the transgressive systems tract also shows a significant decrease in organic-rich Upper Jurassic Bossier Formation (Texas and Louisiana)
Al while Ti/Al and Si/Al show increasing trends that are opposite to records a second-order highstand systems tract, characterized by in-
trends in Al. Ver Straeten et al. (2011) suggested that increasing Ti/Al creasing Al upwards, and decreasing Si/Al, Ti/Al, and Zr/Al. Highstand
and Si/Al can be attributed to elevated levels of silica from aeolian or systems tracts in the Duvernay Formation of the West Shale Basin show
volcanic sources. Data presented by Harris et al. (2018) shows that increasing Al. In this example, Mo/Al typically remains elevated
transgressive systems tracts in the Devonian Duvernay Formation of the moving from the maximum flooding surface until the mid-highstand
West Shale Basin in Alberta are characterized by decreasing Al2O3, systems tract, where it then begins to decline. The lowermost highstand
rising excess Si, and elevated Mo/Al. systems tract exhibits decreasing biogenic Si, but subsequent highstand
Together, the aforementioned studies suggest that the geochemical systems tract record continuing elevated levels of excess Si following
signature of a transgressive systems tract includes decreasing Al transgressive systems tracts (Harris et al., 2018). It is important to note
(Table 4). Other elements associated with detrital minerals (e.g., Ti, Zr) that the highstand systems tracts in all of the above examples were
may also decrease (e.g., transgressive systems tract in the Oatka Creek identified based on other datasets. At this point, criteria to identify the
Formation presented in Ver Straeten et al., 2011; Turner et al., 2016), highstand systems tract chemostratigraphically are poorly constrained
and composite profiles of terrigenous elements may show declining because no geochemical signature for the basal surface of forced re-
trends (e.g., the terrigenous input profile of Ratcliffe et al., 2012c). gression has been defined. Nonetheless, in some cases it is possible to
Although in some instances, certain detrital elements (e.g., Ti, Zr) show identify the highstand, falling-stage, and lowstand systems tracts in
increasing trends (e.g., transgressive systems tract in the Skaneateles fine-grained unconventional plays using other datasets such as seismic
Formation of Ver Straeten et al. (2011); Hammes & Frebourg, 2012). (e.g., Dominguez and Catuneanu, 2017).
During a transgression, sediment trapping in fluvial and/or coastal The highstand systems tracts described above are all characterized
environments while the shoreline moves landward results in a decrease by increasing Al content, likely reflecting progradation associated with
in the abundance of fluvial detrital sediment that reaches the shelf and relative sea level highstand. Titanium and Zr also increase in some in-
slope (Loutit et al., 1988; Mann and Stein, 1997; Catuneanu, 2006). stances and stable Ti/Al observed in one case likely reflects propor-
This is likely the cause for the observed decreases in the abundance of tional increases in both Ti and Al. However, Zr/Al and Ti/Al decline
Al and of other terrigenous elements where they occur. Turner et al. throughout the highstand systems tracts in other studies, potentially
(2016) observed stronger declines in elements associated with heavy due to falling aeolian or volcanic input. Decreasing Si/Al interpreted as
minerals (e.g., Ti and Zr) compared to elements associated with clay declining biogenic silica moving upwards through highstand systems
minerals (e.g., Al and K) and proposed that this is explained by ter- tracts is likely the product of higher dilution by detrital sediment or
restrial heavy minerals settling out sooner than clay minerals. Increases lower productivity. Conversely, increases in proxies for biogenic silica
in Ti/Al and Zr/Al observed in certain cases may be the product of (e.g., Si/Al, excess Si) through the highstand systems tract can be in-
increased aeolian or volcanic input as suggested by Ver Straeten et al. terpreted as continuing elevated levels of biological productivity (e.g.,
(2011). In marine shelf environments, transgressions are also typically Harris et al., 2018). The signature of redox proxies through the high-
characterized by fining-upwards grain size trends (Catuneanu et al., stand systems tract depends on paleohydrographic conditions. For ex-
2009; Embry, 2010), reflected by declining Zr/Nb and Si/Al in Hay- ample, low levels of Mo and V observed by Turner et al. (2016) during a
nesville Formation (e.g., Ratcliffe et al., 2012c; Sano et al., 2013). highstand systems tract in the Woodford Shale were interpreted as the
These studies all display elevated proxies for excess silica during product of greater oxygenation caused by a higher degree of con-
transgressive systems tracts. In certain cases, this was interpreted as nectivity between the Arkoma Basin and the Palaeotethys at this time.
biogenic silica (e.g., Turner et al., 2016; Harris et al., 2018). Condensed Harris et al. (2018) suggested that the Mo/Al patterns observed in
10
M.T. LaGrange, et al. Earth-Science Reviews 203 (2020) 103137
highstand systems tracts of the Duvernay Formation in the West Shale Bohacs and Schwalbach, 1992; Macquaker and Taylor, 1996; Bohacs,
Basin are attributed to higher input of anoxic nutrient-rich seawater 1998; Schieber, 1998; Williams et al., 2001; Bohacs et al., 2005;
from lower basin restriction leading to higher productivity and more Macquaker et al., 2007). For these reasons, it is not yet possible to
reducing conditions during the upper transgressive systems tracts and define the chemostratigraphic expression of the correlative conformity
lower highstand systems tracts. and basal surface of forced regression. During a transgression, the
transgressive surface of erosion forms from erosion by waves or tides
4.3. Regressive systems tract (Swift, 1975; Catuneanu, 2002), and typically forms in coastal and
upper shoreface settings (Catuneanu, 2002). Similarly, the regressive
The regressive systems tract constitutes all strata deposited when surface of marine erosion is cut by waves during forced regression (Plint
the shoreline moves basinward (Embry, 2002; Embry and Johannessen, and Nummedal, 2000) and this erosion takes place on the inner shelf
1993). The regressive systems tract includes strata of the highstand, (Plint and Nummedal, 2000; Catuneanu, 2002). Until recently, fine-
falling-stage, and lowstand systems tracts (Catuneanu, 2002). Turner grained organic-rich successions were viewed as largely distal, deep-
et al. (2015 and 2016) interpreted fourth-order regressive systems water deposits, which would suggest that the transgressive surface of
tracts superimposed on third-order trends in the Woodford Shale based erosion and regressive surface or marine erosion are not relevant to
on Ti and Zr profiles, which displayed similar signatures to trends in K these units. However, many organic-rich mudstone successions are now
and Al. These regressive intervals are typically characterized by in- being re-interpreted to reflect deposition in more proximal environ-
creasing Ti, Zr, K, and Al (Fig. 3). Lower rank regressive systems tracts ments than originally thought (e.g., Schieber, 1994; Smith et al., 2019).
are interpreted in the Haynesville Shale by Ratcliffe et al. (2012c) and This may lead to increased recognition of these surfaces and allow for
Sano et al. (2013). These regressive trends are characterized by in- future delineation of their chemostratigraphic characteristics.
creasing Zr/Nb and Si/Al in Ratcliffe et al. (2012c) and increasing Zr/ Flooding surfaces and parasequences are not discussed in this work.
Nb and V in Sano et al. (2013); Fig. 4). Ratcliffe et al. (2012c) interprets A flooding surface (i.e., parasequence boundary) records an abrupt
that Zr/Nb and Si/Al record grain size variations in the Haynesville water deepening (Van Wagoner et al., 1988), which may or may not
Formation. Ver Straeten et al. (2011) presented geochemical profiles record a change in stratal stacking pattern (Catuneanu, 2019a); as such,
along with a previously established sequence stratigraphic interpreta- it is a surface of allostratigraphy rather than sequence stratigraphy
tion for the Devonian Oatka Creek and Skaneateles formations. They (Catuneanu, 2019a). Certain studies of mudstone intervals still make
use the term ‘early highstand systems tract’ for the highstand systems use of flooding surfaces and parasequences rather than lower rank se-
tract and the term ‘late highstand systems tract’ rather than ‘falling- quence stratigraphic surfaces and sequences (e.g., Bohacs et al., 2014;
stage systems tract’ for the interval deposited during relative sea level Birgenheier et al., 2017; Borcovsky et al., 2017; Turner et al., 2015,
fall, following the Van Wagoner et al. (1988) model. In their inter- 2016), even though the parasequence boundaries no longer comply
pretations, these are grouped together as the highstand systems tract with the original definition of a flooding surface. Therefore, following
rather than separated into the early highstand systems tract and late Catuneanu (2019a), this paper recommends that the use of scale-in-
highstand systems tract. Ver Straeten et al. (2011) suggested that dependent terminology (i.e., the recognition of higher-frequency se-
lowstand systems tracts are not present above these highstand systems quence stratigraphic surfaces and systems tracts) provides a superior
tracts but are instead overlain by transgressive systems tracts. These alternative for stratigraphic mapping and correlation than the deli-
highstand systems tracts are herein considered as regressive systems neation of parasequences (see Catuneanu, 2019a, for a full discussion).
tracts as they may include both normal and forced regressive deposits. The chemostratigraphic expression of lowstand systems tracts and
The third-order regressive systems tract in the Oatka Creek Formation is falling-stage systems tracts have not yet been established, likely for
characterized by increasing Al, decreasing Si/Al, and Ti/Al, and a similar reasons as discussed for the basal surface of forced regression
marked increase in Mo relative to the underlying transgressive systems and correlative conformity. Additionally, although all systems tracts
tract. The regressive systems tract in the overlying Skaneateles For- may be present, higher frequency cycles identified using chemostrati-
mation displays rising Al and Si/Al, with relatively stable Ti/Al. Mo- graphic profiles in these units are commonly limited to the identifica-
lybdenum remains quite low through the entire interval. tion of transgression and regression rather than parsing out normal
These studies suggest that regressive systems tracts are typically from forced regressions (e.g., Ratcliffe et al., 2012c; Sano et al., 2013;
characterized by an increasing abundance of detrital elements and grain Turner et al., 2015, 2016). As a result, only two systems tracts are
size proxies reflecting ongoing progradation. An exception to this is commonly recognized: the transgressive systems tract and regressive
present in the regressive systems tract of the Oatka Creek Formation, systems tract. This is likely the case because in shelfal settings the re-
where Si/Al and Ti/Al decline, which Ver Straeten et al. (2011) inter- gressive systems tract records ongoing progradation, making the basal
preted to have been caused by falling levels of aeolian input. surface of forced regression and correlative conformity cryptic in stu-
dies using log, core, or outcrop data (Catuneanu, 2006). This, therefore,
5. Discussion leads to difficulty in distinguishing the highstand systems tract, low-
stand systems tract, and falling-stage systems tract in these settings. It is
To our knowledge, there are no existing examples of the chemos- also possible that certain systems tracts are not developed in particular
tratigraphic expression of certain surfaces. These include the subaerial depositional settings.
unconformity, correlative conformity, transgressive surface of erosion, One disadvantage of using an undifferentiated regressive systems
regressive surface of marine erosion, and basal surface of forced re- tract is that it amalgamates different genetic types of regressive strata,
gression. The subaerial unconformity is the unconformable portion of which lowers the resolution of the sequence stratigraphic study.
the surface that marks the end of forced regression and is typically Considering normal and forced regressive deposits together may also be
developed in shelf or platform settings (Hunt and Tucker, 1992). This disadvantageous if they have different characteristics related to their
surface is defined by non-marine strata on top (Catuneanu, 2006; unconventional reservoir potential, for example variations in TOC have
Shanmugam, 1988) and is not relevant to the marine mudstone suc- been recorded between forced vs. normal regressive systems tracts in
cessions discussed in this work. As the concept of integrating chemos- unconventional plays (Dominguez et al., 2016; Dominguez and
tratigraphic datasets for sequence stratigraphy is fairly recent, there are Catuneanu, 2017). These are limitations of using chemostratigraphic
relatively few studies that have compared elemental proxies to pre- datasets to identify sequence stratigraphic cycles in fine-grained or-
viously established sequence stratigraphic frameworks or used che- ganic-rich units.
mostratigraphy to help identify surfaces. The recognition of sequence The geochemical characteristics of sequence stratigraphic surfaces
stratigraphic cyclicity within mudstone units is also relatively new (e.g., and systems tracts in organic-rich mudstone intervals are expected to
11
M.T. LaGrange, et al. Earth-Science Reviews 203 (2020) 103137
vary depending on the depositional setting. For example, based on 2017), and for certain elements, poor accuracy relative to accepted
comparison with coarser grained intervals, we expect that there are values or concentrations obtained by other laboratory-based analyses
differences between the chemostratigraphic expression of certain sur- (e.g., Rowe et al., 2012; Hines et al., 2019). The use of helium flow or
faces in more proximal settings influenced by progradation compared to high-performance detectors is now allowing for measurement of lighter
more distal settings beyond the influence of progradation where sedi- elements (e.g., Al and Si), which were traditionally poorly detected by
ment-gravity deposition and pelagic sedimentation are dominant. pXRF, although lower accuracy has been reported for Al compared to
Examples of environments influenced by progradation are the con- elements such as Si, Ti, and in some cases, Zr (Lemiere, 2018). None-
tinental shelf or a ramp setting under the influence of shoreline pro- theless, in a study specific to mudstone intervals, Rowe et al. (2012)
gradation. The slope or abyssal plain of a shelf-slope system or a ramp observed high accuracy for Al and Si. Of the other elements discussed
setting where sedimentation is dominated by sediment-gravity flows herein, Rowe et al. (2012) observed less reliable Cr pXRF measurements
and pelagic fallout are examples of settings that are beyond the influ- when concentrations were below 70 ppm, and unreliable U and Th
ence of shoreline progradation. Herein, environments experiencing results when pXRF data was compared to data collected using wave-
progradation will be referred to as shallow-water settings, whereas length dispersive x-ray fluorescence. Uranium can be difficult to mea-
settings beyond the influence of progradation where sedimentation is sure with XRF because of inter-element interference and low abundance
dominated by sediment-gravity flows and pelagic settling will be re- in many samples (Rowe et al., 2017), and along with Zr, Cr, and Ni, U is
ferred to as deep-water settings. During relative sea-level fall, the deep- one of the elements for which Hines et al. (2019) observed poor cor-
water environment experiences maximum supply of detrital sediment, relations between pXRF data and other laboratory-based measurements
and the detrital sediment supplied to these settings is coarser than the (fusion XRF and ICP-MS).
sediment that accumulates during relative sea-level rise (Posamentier An additional drawback related to the integration of chemostrati-
and Kolla, 2003). Unlike for prograding environments, deposition in graphic datasets for sequence stratigraphic interpretation is the re-
deep-water is less predictable in terms of the locus of accumulation of quirement for fairly continuous measurements from core or outcrop
depositional elements. However, general trends can still be established intervals in order to identify surfaces (Pearce et al., 2010). For example,
at the scale of composite profiles that integrate regional data (e.g., fig. studies that identified sequence stratigraphic surfaces and systems
36 in Catuneanu, 2019b). The main contrasts between the shallow- and tracts (e.g., Ver Straeten et al., 2011; Sano et al., 2013; Turner et al.,
deep-water settings are recorded by the grain-size changes associated 2016) sampled cores and outcrops at intervals ranging from 5 to
with the correlative conformity and the maximum regressive surface. In 120 cm. These high-resolution sampling rates are not possible when
shallow-water environments, the correlative conformity is more diffi- relying on drill cuttings rather than core and outcrop samples because
cult to distinguish because it occurs within a prograding and coarsening drill cuttings are typically taken at coarser intervals (Pearce et al.,
upwards interval (Catuneanu, 2006), whereas the maximum regressive 1999). Given the limitations associated with the use of chemostrati-
surface marks the change from coarsening to fining upwards grain size graphic datasets, chemostratigraphy is more effective when used in
trends (Catuneanu, 2002; Embry, 2010). In coarser grained deep-water conjunction with other datasets (e.g., sedimentological and petrophy-
intervals dominated by sediment-gravity flows, the composite vertical sical) to form robust sequence stratigraphic frameworks in organic-rich
profile coarsens upwards to the correlative conformity during relative mudstone intervals (e.g., Ver Straeten et al., 2011; Hammes and
sea-level fall, and then fines upwards through the lowstand systems Frébourg, 2012).
tract to the maximum flooding surface (Catuneanu, 2006, 2019a). In
this case, the maximum regressive surface is present within a fining 6. Conclusions
upwards interval and is more difficult to distinguish (Catuneanu,
2019b). These trends relate to the efficiency of transfer of fluvial se- This review of documented elemental stratigraphic changes in
diment across the shelf to the deep-water setting, which is highest marine mudstone successions and their interpretation allows for a
during forced regression (see full discussion in Catuneanu, 2019a, preliminary delineation of the chemostratigraphic expression of certain
2019b). We expect that this would affect the chemostratigraphic sig- sequence stratigraphic surfaces and systems tracts, namely the max-
natures of the correlative conformity and maximum regressive surface imum flooding surface, maximum regressive surface, transgressive
in organic-rich mudstone units depending on the setting, with the systems tract, and regressive systems tract. Chemostratigraphic char-
maximum regressive surface recording maxima in terrigenous proxies acteristics of the highstand systems tract are presented, although this
and grain size proxies in shallow-water settings, but the correlative systems tract has been identified using criteria other than chemos-
conformity recording these maxima in deep-water settings, although tratigraphic in all examples presented in this work, and presently
this has not yet been confirmed. The examples discussed herein seem to cannot be distinguished from the regressive systems tract based on
correspond to the expected expression of the maximum regressive chemostratigraphic proxies alone. Available sequence stratigraphic in-
surface in settings influenced by progradation. Further work is required terpretations of geochemical proxy datasets presented herein demon-
to provide examples of the chemostratigraphic expression of these strate that chemostratigraphic datasets are quite useful in interpreting
surfaces and systems tracts in shallow- versus deep-water settings so transgressive-regressive cycles (regressive systems tracts and trans-
that the chemostratigraphic expression can be confirmed. gressive systems tracts separated by maximum regressive surfaces and
In chemostratigraphic studies of mudstone successions, geochemical maximum flooding surfaces, respectively). Also apparent is the need for
composition data is typically collected using either inductively coupled integration with other datasets to further subdivide the regressive sys-
plasma (ICP) and inductively coupled plasma mass spectrometry (ICP- tems tract and confirm proxy signatures to produce robust sequence
MS) or inductively coupled plasma optical emission spectrometry (ICP- stratigraphic frameworks. To date, the utility of incorporating che-
OES) (e.g., Harris et al., 2013, 2018; Sano et al., 2013; Playter et al., mostratigraphic datasets for the study of sequence stratigraphic cycli-
2018), or by portable energy-dispersive X-ray fluorescence (pXRF) (e.g., city in organic-rich mudstone successions has been demonstrated by a
Hammes and Frébourg, 2012; Sano et al., 2013; Turner et al., 2015, handful of studies. Nonetheless, the field remains quite new with lim-
2016; El Attar and Pranter, 2016; Hines et al., 2019). Advantages of ited published examples of the geochemical expression of surfaces and
pXRF compared to ICP methods include time and cost efficiency as well systems tracts. Furthermore, because of the lack of consensus and on-
as the non-destructive nature of the analysis, potentially allowing for going shift in our understanding of mudstone depositional systems, it is
the collection of higher-resolution datasets (Rowe et al., 2012; Rowe not yet possible to confirm many of the expected differences in the
et al., 2017; Lemiere, 2018; Zhang et al., 2019). There are, however, chemostratigraphic expression of surfaces and systems tracts depending
some drawbacks associated with pXRF, including a more limited suite on the depositional setting. The following are suggestions aimed at
of elements that can be obtained compared to ICP analyses (Rowe et al., advancing the field. First, the publication of studies comparing
12
M.T. LaGrange, et al. Earth-Science Reviews 203 (2020) 103137
geochemical proxies to previously established sequence stratigraphic Schwalbach, J.R., Wegner, M.B., Simo, J.T., 2005. Production, destruction, and di-
frameworks (based on other datasets) for fine-grained organic-rich in- lution—the many paths to source-rock development. In: Harris, N.B. (Ed.), The
Deposition of Organic-Carbon-Rich Sediments: Models, Mechanisms, and
tervals will enable further geochemical characterization of surfaces and Consequences: Society of Economic Paleontologists and Mineralogists Special
systems tracts and allow chemostratigraphy to be better integrated with Publication. 82. pp. 61–101.
other datasets. Incorporating chemostratigraphic datasets while devel- Bohacs, K.M., Lazar, O.R., Demko, T.M., 2014. Parasequence types in shelfal mudstone
strata—quantitative observations of lithofacies and stacking patterns, and conceptual
oping sequence stratigraphic interpretations in organic-rich intervals link to modern depositional regimes. Geology 42 (2), 131–134.
where these frameworks do not yet exist will serve a similar purpose. As Bonjour, J.L., Dabard, M.P., 1991. Ti/Nb ratios of clastic terrigenous sediments used as an
our understanding of depositional systems for these units increases, indicator of provenance. Chem. Geol. 91 (3), 257–267.
Borcovsky, D., Egenhoff, S., Fishman, N., Maletz, J., Boehlke, A., Lowers, H., 2017.
characterizing the differences in chemostratigraphic expression of sur- Sedimentology, facies architecture, and sequence stratigraphy of a Mississippian
faces and systems tracts associated with variations in depositional set- black mudstone succession—the upper member of the Bakken Formation, North
ting will also further the utility of chemostratigraphic data to sequence Dakota, United States. Am. Assoc. Pet. Geol. Bull. 101 (10), 1625–1673.
Brown Jr., L.F., Fisher, W.L., 1977. Seismic stratigraphic interpretation of depositional
stratigraphic studies of mudstone successions.
systems: examples from Brazilian rift and pull apart basins. In: Payton, C.E. (Ed.),
Seismic Stratigraphy––Applications to Hydrocarbon Exploration: American
Declaration of competing interest Association of Petroleum Geologists Memoir. 26. pp. 213–248.
Brumsack, H.J., 2006. The trace metal content of recent organic carbon-rich sediments:
implications for Cretaceous black shale formation. Palaeogeogr. Palaeoclimatol.
None. Palaeoecol. 232, 344–361.
Buckman, J., Mahoney, C., März, C., Wagner, T., Blanco, V., 2017. Identifying biogenic
Acknowledgments silica: mudrock micro-fabric explored through charge contrast imaging. Am. Mineral.
102 (4), 833–844.
Calvert, S.E., Pedersen, T.F., 1993. Geochemistry of recent oxic and anoxic marine se-
The authors would like to thank Kathryn Fiess, Viktor Terlaky, Dave diments: implications for the geological record. Mar. Geol. 113 (1–2), 67–88.
Herbers, Carolyn Furlong, and Sara Biddle for their advice and gui- Canfield, D.E., Thamdrup, B., 2009. Towards a consistent classification scheme for geo-
chemical environments, or, why we wish the term ‘suboxic’would go away.
dance. Maya LaGrange, Brette Harris, and Murray Gingras acknowledge Geobiology 7 (4), 385–392.
the Northwest Territories Geological Survey for generously providing Catuneanu, O., 2002. Sequence stratigraphy of clastic systems: concepts, merits, and
funding of this work. Kurt Konhauser, Octavian Catuneanu, and Murray pitfalls. J. Afr. Earth Sci. 35, 1–43.
Catuneanu, O., 2006. Principles of Sequence Stratigraphy. Elsevier, Amsterdam.
Gingras thank the Natural Sciences and Engineering Research Council Catuneanu, O., 2019a. Model-independent sequence stratigraphy. Earth Sci. Rev. 188,
of Canada (NSERC) for their financial support. We would also like to 312–388.
thank the editor Chris Fielding and reviewers Bruce Hart, Bryan Turner, Catuneanu, O., 2019b. Scale in sequence stratigraphy. Mar. Pet. Geol. 106, 128–159.
Catuneanu, O., Abreu, V., Bhattacharya, J.P., Blum, M.D., Dalrymple, R.W., Eriksson,
and Lauren Birgenheier for providing feedback that greatly improved
P.G., Fielding, C.R., Fisher, W.L., Galloway, W.E., Gibling, M.R., Giles, K.A.,
the quality of the manuscript. Holbrook, J.M., Jordan, R., Kendall, C.G.S., Macurda, B., Martinsen, O.J., Miall, A.D.,
Neal, J.E., Nummedal, D., Pomar, L., Posamentier, H.W., Pratt, B.R., Sarg, J.F.,
References Shanley, K.W., Steel, R.J., Strasser, A., Tucker, M.E., Winker, C., 2009. Towards the
standardization of sequence stratigraphy. Earth Sci. Rev. 92, 1–33.
Catuneanu, O., Galloway, W.E., Kendall, C.G.S., Miall, A.D., Posamentier, H.W., Strasser,
Adachi, M., Yamamoto, K., Sugisaki, R., 1986. Hydrothermal chert and associated silic- A., Tucker, M.E., 2011. Sequence stratigraphy: methodology and nomenclature.
eous rocks from the northern Pacific their geological significance as indication of Newsl. Stratigr. 44, 173–245.
ocean ridge activity. Sediment. Geol. 47 (1-2), 125–148. Chen, H.-F., Yeh, P.-Y., Song, S.-R., Hsu, S.-C., Yang, T.-N., Wang, Y., Chi, Z., Lee, T.-Q.,
Algeo, T.J., Lyons, T.W., 2006. Mo–total organic carbon covariation in modern anoxic Chen, M.-T., Cheng, C.-L., Zou, J., Chang, Y.-P., 2013. The Ti/Al molar ratio as a new
marine environments: Implications for analysis of paleoredox and paleohydrographic proxy for tracing sediment transportation processes and its application in aeolian
conditions. Paleoceanography 21 (1), PA1016. events and sea level change in East Asia. J. Asian Earth Sci. 73, 31–38.
Algeo, T.J., Maynard, J.B., 2004. Trace-element behavior and redox facies in core shales Colodner, D., Sachs, J., Ravizza, G., Turekian, K., Edmond, J., Boyle, E., 1993. The geo-
of Upper Pennsylvanian Kansas-type cyclothems. Chem. Geol. 206 (3–4), 289–318. chemical cycle of rhenium: a reconnaissance. Earth Planet. Sci. Lett. 117 (1-2),
Algeo, T.J., Schwark, L., Hower, J.C., 2004. High-resolution geochemistry and sequence 205–221.
stratigraphy of the Hushpuckney Shale (Swope Formation, eastern Kansas): im- Crusius, J., Calvert, S., Pedersen, T., Sage, D., 1996. Rhenium and molybdenum enrich-
plications for climato-environmental dynamics of the Late Pennsylvanian ments in sediments as indicators of oxic, suboxic and sulfidic conditions of deposi-
Midcontinent Seaway. Chem. Geol. 206 (3–4), 259–288. tion. Earth Planet. Sci. Lett. 145 (1-4), 65–78.
Arsairai, B., Wannakomol, A., Feng, Q., Chonglakmani, C., 2016. Paleoproductivity and Dashtgard, S.E., MacEachern, J.A., 2016. Unburrowed mudstones may record only
paleoredox condition of the Huai Hin Lat Formation in northeastern Thailand. J. slightly lowered oxygen conditions in warm, shallow basins. Geology 44 (5),
Earth Sci. 27 (3), 350–364. 371–374.
Arthur, M.A., Sageman, B.B., 2005. Sea-level control on source-rock development: per- Davis, C., Pratt, L., Sliter, W., Mompart, L., Mural, B., 1999. Factors influencing organic
spectives from the Holocene Black Sea, the mid-Cretaceous Western Interior Basin, carbon and trace metal accumulation in the Upper Cretaceous La Luna Formation of
and the late Devonian Appalachian basin. In: Harris, N.B. (Ed.), The Deposition of the western Maracaibo Basin, Venezuela. In: Barrera, E., Johnson, C.C. (Eds.),
Organic-Carbon-Rich Sediments: Models, Mechanisms, and Consequences: Society of Evolution of the Cretaceous Ocean-Climate System, Boulder, Colorado, Geological
Economic Paleontologists and Mineralogists Special Publication. 82. pp. 35–59. Society of America Special Paper 332.
Aubineau, J., El Abani, A., Bekker, A., Somogyi, A., Bankole, O.M., Macchiarelli, R., Dean, W.E., Arthur, M.A., 1998. Geochemical expressions of cyclicity in Cretaceous pe-
Meunier, A., Riboulleau, A., Reynaud, J.-Y., Konhauser, K.O., 2019. Microbially in- lagic limestone sequences: Niobrara Formation, Western Interior Seaway. In: Dean,
duced potassium enrichment in Paleoproterozoic shales and implications for reverse W.E., Arthur, M.A. (Eds.), Stratigraphy of Paleoenvironments of the Cretaceous
weathering on early Earth. Nat. Commun. 10 (1), 1–9 2670. Western Interior Seaway, USA, SEPM Concepts in Sedimentology and Paleontology
Berry, W.B., Wilde, P., 1978. Progressive ventilation of the oceans; an explanation for the No. 6. SEPM, Tulsa, OK, United States, pp. 227–255.
distribution of the lower Paleozoic black shales. Am. J. Sci. 278 (3), 257–275. DeReuil, A.A., Birgenheier, L.P., 2019. Sediment dispersal and organic carbon preserva-
Bhatia, M.R., Crook, K.A., 1986. Trace element characteristics of graywackes and tectonic tion in a dynamic mudstone‐dominated system, Juana Lopez Member, Mancos Shale.
setting discrimination of sedimentary basins. Contrib. Mineral. Petrol. 92, 181–193. Sedimentology 66 (3), 1002–1041.
Birgenheier, L.P., Horton, B., McCauley, A.D., Johnson, C.L., Kennedy, A., 2017. A de- Dinelli, E., Tateo, F., Summa, V., 2007. Geochemical and mineralogical proxies for grain
positional model for offshore deposits of the lower Blue Gate Member, Mancos Shale, size in mudstones and siltstones from the Pleistocene and Holocene of the Po River
Uinta Basin, Utah, USA. Sedimentology 64 (5), 1402–1438. alluvial plain, Italy. Geol. Soc. Am. Spec. Pap. 420, 25–36.
Bloemsma, M.R., Zabel, M., Stuut, J.B.W., Tjallingii, R., Collins, J.A., Weltje, G.J., 2012. Dominguez, R.F., Catuneanu, O., 2017. Regional stratigraphic framework of the Vaca
Modelling the joint variability of grain size and chemical composition in sediments. Muerta – Quintuco system in the Neuquén Embayment, Argentina. In: Proceedings
Sediment. Geol. 280, 135–148. from: The 20th Geological Congress of Argentina, Tucuman, Argentina.
Bohacs, K.M., 1998. Contrasting expressions of depositional sequences in mudrocks from Dominguez, R.F., Continanzia, M.J., Mykietiuk, K., Ponce, C., Pérez, G., Guerello, R.,
marine to non-marine environs. In: Schieber, J., Zimmerle, W., Sethi, P.S. (Eds.), Lanusse, N.I., Caneva, M., Di Benedetto, M., Catuneanu, O., Cristallini, E., 2016.
Shales and Mudstones: Volume I. E. Schweizerbart’sche Verlagsbuchhandlung Organic-rich stratigraphic units in the Vaca Muerta Formation, and their distribution
(Nägele u. Obermiller), Stuttgart, pp. 33–78. and characterization in the Neuquén Basin (Argentina). In: Proceedings from:
Bohacs, K.M., Schwalbach, J.R., 1992. Sequence stratigraphy of fine-grained rocks with Unconventional Resources Technology Conference, San Antonio, Texas.
special reference to the Monterey Formation. In: Schwalbach, J.R., Bohacs, K.M. Egenhoff, S.O., Fishman, N.S., 2013. Traces in the dark—Sedimentary processes and fa-
(Eds.), Sequence Stratigraphy in Fine-Grained Rocks: Examples from the Monterey cies gradients in the upper shale member of the Upper Devonian–Lower Mississippian
Formation, pp. 7–19 Pacific Section, Society of Economic Paleontologists and Bakken Formation, Williston Basin, North Dakota, USA. J. Sediment. Res. 83 (9),
Mineralogists, Book 70. 803–824.
Bohacs, K.M., Grabowski Jr., G.J., Carroll, A.R., Mankiewicz, P.J., Miskell-Gerhardt, K.J., El Attar, A., Pranter, M.J., 2016. Regional stratigraphy, elemental chemostratigraphy, and
13
M.T. LaGrange, et al. Earth-Science Reviews 203 (2020) 103137
organic richness of the Niobrara Member of the Mancos Shale, Piceance Basin, Knapp, L.J., McMillan, J.M., Harris, N.B., 2017. A depositional model for organic-rich
Colorado. Am. Assoc. Pet. Geol. Bull. 100 (3), 345–377. Duvernay Formation mudstones. Sediment. Geol. 347, 160–182.
Elder, W.P., Gustason, E.R., Sageman, B.B., 1994. Correlation of basinal carbonate cycles Knapp, L.J., Harris, N.B., McMillan, J.M., 2019. A sequence stratigraphic model for the
to nearshore parasequences in the Late Cretaceous Greenhorn seaway, Western organic-rich Upper Devonian Duvernay Formation, Alberta, Canada. Sediment. Geol.
Interior USA. Geol. Soc. Am. Bull. 106 (7), 892–902. 387, 152–181.
Embry, A.F., 2002. Transgressive-regressive (TR) sequence stratigraphy. In: Proceedings Konhauser, K.O., 2007. Introduction to Geomicrobiology. Blackwell, Malden,
from: 22nd Annual Gulf Coast Section SEPM Foundation Bob F. Perkins Research Massachusetts.
Conference, Houston, Texas. Lazar, O.R., Bohacs, K.M., Macquaker, J.H., Schieber, J., Demko, T.M., 2015. Capturing
Embry, A.F., 2009. Practical sequence stratigraphy. Can. Soc. Petrol. Geol. Reserv. key attributes of fine-grained sedimentary rocks in outcrops, cores, and thin sections:
36, 1–6. nomenclature and description guidelines. J. Sediment. Res. 85 (3), 230–246.
Embry, A.F., 2010. Correlation siliciclastic successions with sequence stratigraphy. In: Lemiere, B., 2018. A review of pXRF (field portable X-ray fluorescence) applications for
Ratcliffe, K.T., Zaitlin, B.A. (Eds.), Application of Modern Stratigraphic Techniques: applied geochemistry. J. Geochem. Explor. 188, 350–363.
Theory and Case Histories, pp. 35–53 SEPM (Society for Sedimentary Geology) Lewan, M.D., 1984. Factors controlling the proportionality of vanadium to nickel in crude
Special Publication, no. 94. oils. Geochim. Cosmochim. Acta 48 (11), 2231–2238.
Embry, A.F., Johannessen, E.P., 1993. T–R sequence stratigraphy, facies analysis and Lewan, M.D., Maynard, J.B., 1982. Factors controlling enrichment of vanadium and
reservoir distribution in the uppermost Triassic–Lower Jurassic succession, Western nickel in the bitumen of organic sedimentary rocks. Geochim. Cosmochim. Acta 46
Sverdrup Basin, Arctic Canada. Norw. Petrol. Soc. Spec. Publ. 2, 121–146. (12), 2547–2560.
Erickson, B.E., Helz, G.R., 2000. Molybdenum (VI) speciation in sulfidic waters: stability Li, Z., Schieber, J., 2020. Application of sequence stratigraphic concepts to the Upper
and lability of thiomolybdates. Geochim. Cosmochim. Acta 64 (7), 1149–1158. Cretaceous Tununk Shale Member of the Mancos Shale Formation, south‐central
Frébourg, G., Ruppel, S.C., Rowe, H., 2013. Sedimentology of the Haynesville (Upper Utah: parasequence styles in shelfal mudstone strata. Sedimentology 67 (1), 118–151.
Kimmeridgian) and Bossier (Tithonian) Formations, in the Western Haynesville Liu, Y., Zhang, J., Tang, X., Yang, C., Tang, S., 2016. Weathering characteristics of the
Basin, Texas, U.S.A. In: Hammes, U., Gale, J. (Eds.), Geology of the Haynesville Gas Lower Paleozoic black shale in northwestern Guizhou Province, south China. J. Earth
Shale in East Texas and West Louisiana, U.S.A. pp. 47–67 America Association of Syst. Sci. 125 (5), 1061–1078.
Petroleum Geologists Memoir 105. Loutit, T.S., Hardenbol, J., Vail, P.R., Baum, G.R., 1988. Condensed sections: the key to
Froelich, P., Klinkhammer, G.P., Bender, M.L., Luedtke, N.A., Heath, G.R., Cullen, D., age determination and correlation of continental margin sequences. In: Wilgus, C.K.,
Dauphin, P., Hammond, D., Hartman, B., Maynard, V., 1979. Early oxidation of or- Hastings, B.S., Kendall, C.G.S., Posamentier, H.W., Ross, C.A., Van Wagoner, J.C.
ganic matter in pelagic sediments of the eastern equatorial Atlantic: suboxic diag- (Eds.), Sea Level Changes – An Integrated Approach: Society of Economic
enesis. Geochim. Cosmochim. Acta 43 (7), 1075–1090. Paleontologists and Geologists Special Publication. 42. pp. 183–213.
Fru, E.C., Rodríguez, N.P., Partin, C.A., Lalonde, S.V., Andersson, P., Weiss, D.J., El MacEachern, J.A., Bann, K.L., Pemberton, S.G., Gingras, M.K., 2009a. The ichnofacies
Albani, A., Rodushkin, I., Konhauser, K.O., 2016. Cu isotopes in marine black shales paradigm: high-resolution paleoenvironmental interpretation of the rock record.
record the Great Oxidation Event. Proc. Natl. Acad. Sci. 113 (18), 4941–4946. SEPM Soc. Sediment. Geol. App. Ichnol. SC52, 27–64.
Galloway, W.E., 1989. Genetic stratigraphic sequences in basin analysis I: architecture MacEachern, J.A., Pemberton, S.G., Bann, K.L., Gingras, M.K., 2009b. Departures from
and genesis of flooding-surface bounded depositional units. Am. Assoc. Pet. Geol. the archetypal ichnofacies: effective recognition of physico-chemical stresses in the
Bull. 73 (2), 125–142. rock record. Soc. Sediment. Geol. Appl. Ichnol. SC52, 65–93.
Gambacorta, G., Trincianti, E., Torricelli, S., 2016. Anoxia controlled by relative sea-level Macquaker, J.H., Taylor, K.G., 1996. A sequence stratigraphic interpretation of a mud-
changes: an example from the Mississippian Barnett Shale Formation. Palaeogeogr. stone-dominated succession: the Lower Jurassic Cleveland ironstone Formation, U.K.
Palaeoclimatol. Palaeoecol. 459, 306–320. J. Geol. Soc. Lond. 153, 759–770.
Grosjean, E., Adam, P., Connan, J., Albrecht, P., 2004. Effects of weathering on nickel and Macquaker, J.H., Taylor, K.G., Gawthorpe, R.L., 2007. High-resolution facies analyses of
vanadyl porphyrins of a Lower Toarcian shale of the Paris basin. Geochim. mudstones: implications for paleoenvironmental and sequence stratigraphic inter-
Cosmochim. Acta 68 (4), 789–804. pretations of offshore ancient mud-dominated successions. J. Sediment. Res. 77 (4),
Gutierrez Parades, H.C., Catuneanu, O., Romano, U.H., 2017. Sequence stratigraphy of 324–339.
the Miocene section, southern Gulf of Mexico. Mar. Pet. Geol. 86, 711–732. Mann, U., Stein, R., 1997. Organic facies variations, source rock potential, and sea level
Hammes, U., Frébourg, G., 2012. Haynesville and Bossier mudrocks: a facies and se- changes in Cretaceous black shales of the Quebrada Ocal, Upper Magdalena Valley,
quence stratigraphic investigation, East Texas and Louisiana, USA. Mar. Pet. Geol. 31 Colombia. Am. Assoc. Pet. Geol. Bull. 81 (4), 556–576.
(1), 8–26. Marynowski, L., Pisarzowska, A., Derkowski, A., Rakociński, M., Szaniawski, R., Środoń,
Hammes, U., Hamlin, H.S., Ewing, T.E., 2011. Geologic analysis of the Upper Jurassic J., Cohen, A.S., 2017. Influence of palaeoweathering on trace metal concentrations
Haynesville Shale in east Texas and west Louisiana. Am. Assoc. Pet. Geol. Bull. 95 and environmental proxies in black shales. Palaeogeogr. Palaeoclimatol. Palaeoecol.
(10), 1643–1666. 472, 177–191.
Harazim, D., McIlroy, D., Edwards, N.P., Wogelius, R.A., Manning, P.L., Poduska, K.M., McLennan, S.M., Hemming, S., McDaniel, D.K., Hanson, G.N., 1993. Geochemical ap-
Layne, G.D., Sokaras, D., Alonso-Mori, R., Bergmann, U., 2015. Bioturbating animals proaches to sedimentation, provenance, and tectonics. Geol. Soc. Am. Spec. Pap. 284,
control the mobility of redox-sensitive trace elements in organic-rich mudstone. 21–40.
Geology 43 (11), 1007–1010. McManus, J., Berelson, W.M., Klinkhammer, G.P., Hammond, D.E., Holm, C., 2005.
Harris, N.B., Mnich, C.A., Selby, D., Korn, D., 2013. Minor and trace element and Re–Os Authigenic uranium: relationship to oxygen penetration depth and organic carbon
chemistry of the Upper Devonian Woodford Shale, Permian Basin, west Texas: in- rain. Geochim. Cosmochim. Acta 69 (1), 95–108.
sights into metal abundance and basin processes. Chem. Geol. 356, 76–93. Mongelli, G., Critelli, S., Perri, F., Sonnino, M., Perrone, V., 2006. Sedimentary recycling,
Harris, N.B., McMillan, J.M., Knapp, L.J., Mastalerz, M., 2018. Organic matter accumu- provenance and paleoweathering from chemistry and mineralogy of Mesozoic con-
lation in the Upper Devonian Duvernay Formation, Western Canada Sedimentary tinental redbed mudrocks, Peloritani Mountains, southern Italy. Geochem. J. 40 (20),
Basin, from sequence stratigraphic analysis and geochemical proxies. Sediment. Geol. 197–209.
376, 185–203. Morford, J.L., Russell, A.D., Emerson, S., 2001. Trace metal evidence for changes in the
Hart, B.S., 2015. The Greenhorn Cyclothem of the Cretaceous Western Interior Seaway: redox environment associated with the transition from terrigenous clay to diato-
lithology trends, stacking patterns, log signatures, and, and application to the Eagle maceous sediment, Saanich Inlet, BC. Mar. Geol. 174 (1–4), 355–369.
Ford of West Texas. Gulf Coast Assoc. Geol. Soc. Trans. 65, 155–174. Morrison, J.M., Codispoti, L.A., Smith, S.L., Wishner, K., Flagg, C., Gardner, W.D., Gaurin,
Helland-Hansen, W., Martinsen, O.J., 1996. Shoreline trajectories and sequences: de- S., Naqvi, S.W.A., Manghnani, V., Prosperie, L., Gundersen, J.S., 1999. The oxygen
scription of variable depositional-dip scenarios. J. Sediment. Res. 66 (4), 670–688. minimum zone in the Arabian Sea during 1995. Deep-Sea Res. II Top. Stud. Oceanogr.
Helly, J.J., Levin, L.A., 2004. Global distribution of naturally occurring marine hypoxia 46 (8–9), 1903–1931.
on continental margins. Deep-Sea Res. I Oceanogr. Res. Pap. 51 (9), 1159–1168. Myers, K.J., Wignall, P.B., 1987. Understanding Jurassic organic-rich mudrocks—new
Helz, G.R., Miller, C.V., Charnock, J.M., Mosselmans, J.F.W., Pattrick, R.A.D., Garner, concepts using gamma-ray spectrometry and palaeoecology: examples from the
C.D., Vaughan, D.J., 1996. Mechanism of molybdenum removal from the sea and its Kimmeridge Clay of Dorset and the Jet Rock of Yorkshire. In: Legget, J.K., Zuffa, G.G.
concentration in black shales: EXAFS evidence. Geochim. Cosmochim. Acta 60 (19), (Eds.), Marine Clastic Sedimentology. Springer, Dordrecht, pp. 172–189.
3631–3642. Nadeau, P.H., Wilson, M.J., McHardy, W.J., Tait, J.M., 1985. The conversion of smectite
Hines, B.R., Gazley, M.F., Collins, K.S., Bland, K.J., Crampton, J.S., Ventura, G.T., 2019. to illite during diagenesis: evidence from some illitic clays from bentonites and
Chemostratigraphic resolution of widespread reducing conditions in the southwest sandstones. Mineral. Mag. 49 (352), 393–400.
Pacific Ocean during the Late Paleocene. Chem. Geol. 504, 236–252. Nesbitt, H.W., Young, G.M., 1982. Early Proterozoic climates and plate motions inferred
Huerta-Diaz, M.A., Morse, J.W., 1992. Pyritization of trace metals in anoxic marine se- from major element chemistry of lutites. Nature 299 (5885), 715–717.
diments. Geochim. Cosmochim. Acta 56 (7), 2681–2702. Nyhuis, C.J., Riley, D., Kalasinska, A., 2016. Thin section petrography and chemostrati-
Hunt, D., Tucker, M.E., 1992. Stranded parasequences and the forced regressive wedge graphy: integrated evaluation of an upper Mississippian mudstone dominated suc-
systems tract: deposition during base-level fall. Sediment. Geol. 81 (1–2), 1–9. cession from the southern Netherlands. Neth. J. Geosci. 95 (1), 3–22.
Jaffe, L.A., Peucker-Ehrenbrink, B., Petsch, S.T., 2002. Mobility of rhenium, platinum Over, D.J., 1990. Trace metals in burrow walls and sediments, Georgia Bight, USA.
group elements and organic carbon during black shale weathering. Earth Planet. Sci. Ichnos. Int. J. Plant Anim. 1 (1), 31–41.
Lett. 198 (3–4), 339–353. Partin, C.A., Bekker, A., Planavsky, N.J., Scott, C.T., Gill, B.C., Li, C., Podkovyrov, V.,
Kendall, B., Reinhard, C.T., Lyons, T.W., Kaufman, A.J., Poulton, S.W., Anbar, A.D., 2010. Maslov, V., Konhauser, K.O., Lalonde, S.V., Love, G.D., Love, G.D., Poulton, S.W.,
Pervasive oxygenation along late Archaean ocean margins. Nat. Geosci. 3 (9), 647. Lyons, T.W., 2013. Large-scale fluctuations in Precambrian atmospheric and oceanic
Kimura, H., Watanabe, Y., 2001. Oceanic anoxia at the Precambrian-Cambrian boundary. oxygen levels from the record of U in shales. Earth Planet. Sci. Lett. 369, 284–293.
Geology 29 (11), 995–998. Patchett, P.J., White, W.M., Feldmann, H., Kielinczuk, S., Hofmann, A.W., 1984.
Klinkhammer, G.P., Palmer, M.R., 1991. Uranium in the oceans: where it goes and why. Hafnium/rare earth element fractionation in the sedimentary system and crustal
Geochim. Cosmochim. Acta 55 (7), 1799–1806. recycling into the Earth’s mantle. Earth Planet. Sci. Lett. 69 (2), 365–378.
14
M.T. LaGrange, et al. Earth-Science Reviews 203 (2020) 103137
Pearce, T.J., Besly, B.M., Wray, D.S., Wright, D.K., 1999. Chemostratigraphy: a method to Zhang, J., Liang, Q., Sivil, E., 2017. Chemostratigraphic insights into fluvio-lacustrine
improve interwell correlation in barren sequences—a case study using onshore deposition, Yanchang Formation, Upper Triassic, Ordos Basin, China. Interpretation 5
Duckmantian/Stephanian sequences (West Midlands, UK). Sediment. Geol. 124 (2), SF149–SF165.
(1–4), 197–220. Sageman, B.B., Lyons, T.W., 2004. Geochemistry of fine-grained sediments and sedi-
Pearce, T.J., Wray, D., Ratcliffe, K., Wright, D.K., Moscariello, A., 2005. mentary rocks. In: Mackenzie, F.T. (Ed.), Sediments, Diagenesis, and Sedimentary
Chemostratigraphy of the Upper Carboniferous Schooner Formation, southern North Rocks: Treatise on Geochemistry, Second Edition, Volume 7. Elsevier Pergamon,
Sea. In: Collinson, J.D., Evans, D.J., Holliday, D.W. (Eds.), Carboniferous Oxford, United Kingdom, pp. 116–148.
Hydrocarbon Resources: The Southern North Sea and Surrounding Onshore Areas: Sano, J.L., Ratcliffe, K.T., Spain, D.R., 2013. Chemostratigraphy of the Haynesville Shale.
Occasional Publication Series of the Yorkshire Geological Society. 7. pp. 147–164. In: Hammes, U., Gales, J. (Eds.), Geology of the Hanyesville Gas Shale in East Texas
Pearce, T.J., Martin, J.H., Cooper, D., Wray, D.S., 2010. Chemostratigraphy of upper and West Louisiana, U.S.A. pp. 137–154 American Association of Petroleum
carboniferous (Pennsylvanian) sequences from the Southern North Sea (United Geologists Memoir 105.
Kingdom). In: Application of Modern Stratigraphic Techniques: Theory and Case Savrda, C.E., Bottjer, D.J., 1989. Trace-fossil model for reconstructing oxygenation his-
Histories: SEPM (Society for Sedimentary Geology) Special Publication, 94, pp. tories of ancient marine bottom waters: application to Upper Cretaceous Niobrara
109–127. Formation, Colorado. Palaeogeogr. Palaeoclimatol. Palaeoecol. 74 (1–2), 49–74.
Peucker-Ehrenbrink, B., Hannigan, R.E., 2000. Effects of black shale weathering on the Schieber, J., 1994. Evidence for high-energy events and shallow-water deposition in the
mobility of rhenium and platinum group elements. Geology 28 (5), 475–478. Chattanooga Shale, Devonian, central Tennessee, USA. Sediment. Geol. 93 (3-4),
Piper, D.Z., Calvert, S.E., 2009. A marine biogeochemical perspective on black shale 193–208.
deposition. Earth Sci. Rev. 95 (1–2), 63–96. Schieber, J., 1996. Early diagenetic silica deposition in algal cysts and spores; a source of
Plank, T., Langmuir, C.H., 1998. The chemical composition of subducting sediment and sand in black shales? J. Sediment. Res. 66 (1), 175–183.
its consequences for the crust and mantle. Chem. Geol. 145 (3), 325–394. Schieber, J., 1998. Developing a sequence stratigraphic framework for the Late Devonian
Playter, T., Corlett, H., Konhauser, K., Robbins, L., Rohais, S., Crombez, V., Maccormack, Chattanooga Shale of the southeastern USA: relevance for the Bakken Shale. In:
K., Rokosh, D., Prenoslo, D., Furlong, C.M., Pawlowicz, J., Gingras, M., Lalonde, S., Christopher, J.E., Gilboy, C.F., Paterson, D.F., Bends, S.L. (Eds.), Eighth International
Lyster, S., Zonneveld, J.-P., 2018. Clinoform identification and correlation in fine‐- Williston Basin Symposium, pp. 58–68 Saskatchewan Geological Society Special
grained sediments: a case study using the Triassic Montney Formation. Publication No. 13.
Sedimentology 65 (1), 263–302. Schieber, J., Krinsley, D., Riciputi, L., 2000. Diagenetic origin of quartz silt in mudstones
Plint, A.G., Nummedal, D., 2000. The falling stage systems tract: recognition and im- and implications for silica cycling. Nature 406 (6799), 981.
portance in sequence stratigraphic analysis. Geol. Soc. Lond., Spec. Publ. 172 (1), Schönfeld, J., Kuhnt, W., Erdem, Z., Flögel, S., Glock, N., Aquit, M., Frank, M., Holbourn,
1–17. A., 2015. Records of past mid-depth ventilation: Cretaceous ocean anoxic event 2 vs.
Posamentier, H.W., Allen, G.P., 1999. Siliciclastic sequence stratigraphy: concepts and Recent oxygen minimum zones. Biogeosciences 12 (4), 1169–1189.
applications. SEPM Soc. Sediment. Geol. Concepts Sedimentol. Paleontol. 7, 210. Scott, C., Lyons, T.W., 2012. Contrasting molybdenum cycling and isotopic properties in
Posamentier, H.W., Kolla, V., 2003. Seismic geomorphology and stratigraphy of deposi- euxinic versus non-euxinic sediments and sedimentary rocks: refining the paleo-
tional elements in deep-water settings. J. Sediment. Res. 73 (3), 367–388. proxies. Chem. Geol. 324–325, 19–27.
Pyle, L.J., Gal, L.P., 2016. Reference Section for the Horn River Group and Definition of Shanmugam, G., 1988. Origin, recognition, and importance of erosional unconformities
the Bell Creek Member, Hare Indian Formation in central Northwest Territories. Bull. in sedimentary basins. In: New Perspectives in Basin Analysis. Springer, New York,
Can. Petrol. Geol. 64 (1), 67–98. NY, pp. 83–108.
Ratcliffe, K.T., Wright, A.M., Hallsworth, C., Morton, A., Zaitlin, B.A., Potocki, D., Wray, Shen, J., Zhou, L., Feng, Q., Zhang, M., Lei, Y., Zhang, N., Yu, J., Gu, S., 2014. Paleo-
D., 2004. An example of alternative correlation techniques in a low-accommodation productivity evolution across the Permian-Triassic boundary and quantitative cal-
setting, nonmarine hydrocarbon system: the (Lower Cretaceous) Mannville Basal culation of primary productivity of black rock series from the Dalong Formation,
Quartz succession of southern Alberta. Am. Assoc. Pet. Geol. Bull. 88 (10), South China. Sci. China Earth Sci. 57 (7), 1583–1594.
1419–1432. Slatt, R.M., Rodriguez, N.D., 2012. Comparative sequence stratigraphy and organic
Ratcliffe, K.T., Martin, J., Pearce, T.J., Hughes, A.D., Lawton, D.E., Wray, D.S., Bessa, F., geochemistry of gas shales: Commonality or coincidence? J. Nat. Gas Sci. Eng. 8,
2006. A regional chemostratigraphically-defined correlation framework for the late 68–84.
Triassic TAG-I Formation in Blocks 402 and 405a, Algeria. Pet. Geosci. 12 (1), 3–12. Smith, L.B., Schieber, J., Wilson, R.D., 2019. Shallow-water onlap model for the de-
Ratcliffe, K., Wright, M., Montgomery, P., Palfrey, A., Vonk, A., Vermeulen, J., Barrett, position of Devonian black shales in New York, USA. Geology 47 (3), 279–283.
M., 2010. Application of chemostratigraphy to the Mungaroo Formation, the Gorgon Swift, D.J., 1975. Barrier-island genesis: evidence from the central Atlantic shelf, eastern
field, offshore northwest Australia. APPEA J. 50 (1), 371–388. USA. Sediment. Geol. 14 (1), 1–43.
Ratcliffe, K.T., Wright, A.M., Schmidt, K., 2012a. Application of inorganic whole-rock Taylor, K.G., Macquaker, J.H.S., 2014. Diagenetic alterations in a silt-and clay-rich
geochemistry to shale resource plays: an example from the Eagle Ford Shale mudstone succession: an example from the Upper Cretaceous Mancos Shale of Utah,
Formation, Texas. Sediment. Rec. 10 (2), 4–9. USA. Clay Miner. 49 (2), 213–227.
Ratcliffe, K.T., Woods, J., Rice, C., 2012b. Determining well-bore pathways during mul- Taylor, S.R., McLennan, S.M., 1985. The Continental Crust: Its Composition and
tilateral drilling campaigns in shale resource plays: an example using chemostrati- Evolution. Blackwell, Malden, Massachusetts.
graphy from the Horn River Formation, British Columbia, Canada. In: Proceedings Thomson, J., Higgs, N.C., Croudace, I.W., Colley, S., Hydes, D.J., 1993. Redox zonation of
from: Eastern Australasian Basin Symposium IV, Brisbane, Australia. elements at an oxic/post-oxic boundary in deep-sea sediments. Geochim. Cosmochim.
Ratcliffe, K.T., Wright, A.M., Spain, D.R., 2012 April/Mayc. Unconventional methods for Acta 57 (3), 579–595.
unconventional plays: using elemental data to understand shale resource plays. Thomson, J., Higgs, N.C., Wilson, T.R.S., Croudace, I.W., De Lange, G.J., Van Santvoort,
Petrol. Explor. Soc. Aust. News Resour. 117, 50–54. P.J.M., 1995. Redistribution and geochemical behaviour of redox-sensitive elements
Reichart, G.J., Lourens, L.J., Zachariasse, W.J., 1998. Temporal variability in the northern around S1, the most recent eastern Mediterranean sapropel. Geochim. Cosmochim.
Arabian Sea Oxygen Minimum Zone (OMZ) during the last 225,000 years. Acta 59 (17), 3487–3501.
Paleoceanography 13 (6), 607–621. Thomson, J., Jarvis, I., Green, D.R., Green, D., 1998. Oxidation fronts in Madeira Abyssal
Reinhard, C.T., Planavsky, N.J., Robbins, L.J., Partin, C.A., Gill, B.C., Lalonde, S.V., Plain turbidites: persistence of early diagenetic trace-element enrichments during
Bekker, A., Konhauser, K.O., Lyons, T.W., 2013. Proterozoic ocean redox and bio- burial, site 9501. In: Weaver, P.P.E., Schmincke, H.-U., Duffield, W. (Eds.),
geochemical stasis. Proc. Natl. Acad. Sci. 110 (14), 5357–5362. Proceedings of the Ocean Drilling Program, Scientific Results. Vol. 157. pp. 559–571.
Revsbech, N.P., Larsen, L.H., Gundersen, J., Dalsgaard, T., Ulloa, O., Thamdrup, B., 2009. Tribovillard, N., Algeo, T.J., Lyons, T., Riboulleau, A., 2006. Trace metals as paleoredox
Determination of ultra-low oxygen concentrations in oxygen minimum zones by the and paleoproductivity proxies: an update. Chem. Geol. 232 (1–2), 12–32.
STOX sensor. Limnol. Oceanogr. Methods 7, 371–381. Turner, B.W., Slatt, R.M., 2016. Assessing bottom water anoxia within the Late Devonian
Robbins, L.J., Lalonde, S.V., Planavsky, N.J., Partin, C.A., Reinhard, C.T., Kendall, B., Woodford Shale in the Arkoma Basin, southern Oklahoma. Mar. Pet. Geol. 78,
Scott, C., Hardisty, D.S., Gill, B.C., Alessi, D.S., Dupont, C.L., Saito, M.A., Crowe, S.A., 536–546.
Poulton, S.W., Bekker, A., Lyons, T.W., Konhauser, K.O., 2016. Trace elements at the Turner, B.W., Molinares-Blanco, C.E., Slatt, R.M., 2015. Chemostratigraphic, palynos-
intersection of marine biological and geochemical evolution. Earth Sci. Rev. 163, tratigraphic, and sequence stratigraphic analysis of the Woodford Shale, Wyche Farm
323–348. Quarry, Pontotoc County, Oklahoma. Interpretation 3 (1), SH1–SH9.
Rosenthal, Y., Lam, P., Boyle, E.A., Thomson, J., 1995. Authigenic cadmium enrichments Turner, B.W., Tréanton, J.A., Slatt, R.M., 2016. The use of chemostratigraphy to refine
in suboxic sediments: precipitation and postdepositional mobility. Earth Planet. Sci. ambiguous sequence stratigraphic correlations in marine mudrocks. An example from
Lett. 132 (1–4), 99–111. the Woodford Shale, Oklahoma, USA. J. Geol. Soc. 173, 854–868.
Ross, D.J., Bustin, R.M., 2009. Investigating the use of sedimentary geochemical proxies Tyson, R.V., Pearson, T.H., 1991. Modern and ancient continental shelf anoxia: an
for paleoenvironment interpretation of thermally mature organic-rich strata: ex- overview. Geol. Soc. Lond., Spec. Publ. 58 (1), 1–24.
amples from the Devonian–Mississippian shales, Western Canadian Sedimentary Van der Weijden, C.H., 2002. Pitfalls of normalization of marine geochemical data using a
Basin. Chem. Geol. 260 (1–2), 1–19. common divisor. Mar. Geol. 184 (3–4), 167–187.
Rowe, H., Hughes, N., Robinson, K., 2012. The quantification and application of handheld Van Wagoner, J.C., Posamentier, H.W., Mitchum, R.M.J., Vail, P.R., Sarg, J.F., Loutit,
energy-dispersive x-ray fluorescence (ED-XRF) in mudrock chemostratigraphy and T.S., Hardenbol, J., 1988. An overview of the fundamentals of sequence stratigraphy
geochemistry. Chem. Geol. 324–325, 122–131. and key definitions. In: Wilgus, C.K., Hastings, B.S., Kendall, C.G.S., Posamentier,
Rowe, H.D., Loucks, R.G., Ruppel, S.C., Rimmer, S.M., 2008. Mississippian Barnett H.W., Ross, C.A., Van Wagoner, J.C. (Eds.), Sea Level Changes–An Integrated
Formation, Fort Worth Basin, Texas: bulk geochemical inferences and Mo–TOC Approach, pp. 39–45 SEPM (Society for Sedimentary Geology) Special
constraints on the severity of hydrographic restriction. Chem. Geol. 257 (1–2), 16–25. Publication 42.
Rowe, H., Ruppel, S., Rimmer, S., Loucks, R., 2009. Core-based chemostratigraphy of the Ver Straeten, C.A., Brett, C.E., Sageman, B.B., 2011. Mudrock sequence stratigraphy: a
Barnett Shale, Permian Basin, Texas. Gulf Coast Assoc. Geol. Soc. Trans. 59, 675–686. multi-proxy (sedimentological, paleobiological and geochemical) approach,
Rowe, H., Wang, X., Fan, B., Zhang, T., Ruppel, S.C., Milliken, K.L., Loucks, R., Shen, Y., Devonian Appalachian Basin. Palaeogeogr. Palaeoclimatol. Palaeoecol. 304 (1–2),
15
M.T. LaGrange, et al. Earth-Science Reviews 203 (2020) 103137
54–73. Middle Devonian Geneseo Formation of New York, USA. J. Sediment. Res. 85 (11),
Wanty, R.B., Goldhaber, M.B., 1992. Thermodynamics and kinetics of reactions involving 1393–1415.
vanadium in natural systems: accumulation of vanadium in sedimentary rocks. Wyrtki, K., 1962. The oxygen minima in relation to ocean circulation. Deep-Sea Res.
Geochim. Cosmochim. Acta 56 (4), 1471–1483. Oceanogr. Abstr. 9 (1-2), 11–23.
Wedepohl, K.H., 1971. Environmental influences on the chemical composition of shales Yamashita, Y., Takahashi, Y., Haba, H., Enomoto, S., Shimizu, H., 2007. Comparison of
and clays. Phys. Chem. Earth 8, 307–333. reductive accumulation of Re and Os in seawater–sediment systems. Geochim.
Wignall, P.B., Twitchett, R.J., 1996. Oceanic anoxia and the end Permian mass extinction. Cosmochim. Acta 71 (14), 3458–3475.
Science 272 (5265), 1155–1158. Zhang, J., Zeng, Y., Slatt, R., 2019. XRF (X-ray fluorescence) applied to characterization
Wilde, P., Quinby-Hunt, M.S., Erdtmann, B.D., 1996. The whole-rock cerium anomaly: a of unconventional Woodford Shale (Devonian, USA) lateral well heterogeneity. Fuel
potential indicator of eustatic sea-level changes in shales of the anoxic facies. 254, 115565.
Sediment. Geol. 101 (1–2), 43–53. Zhou, C., Jiang, S.Y., 2009. Palaeoceanographic redox environments for the lower
Williams, C.J., Hesselbo, S.P., Jenkyns, H.C., Morgans-Bell, H.S., 2001. Quartz silt in Cambrian Hetang Formation in South China: evidence from pyrite framboids, redox
mudrocks as a key to sequence stratigraphy (Kimmeridge Clay Formation, Late sensitive trace elements, and sponge biota occurrence. Palaeogeogr. Palaeoclimatol.
Jurassic, Wessex Basin, UK). Terra Nova 13, 449–455. Palaeoecol. 271 (3–4), 279–286.
Wilson, R.D., Schieber, J., 2015. Sedimentary facies and depositional environment of the
16