You are on page 1of 24

A novel integrated AUTHORS

approach for chemofacies Junwen Peng ~ State Key Laboratory of


Petroleum Resources and Prospecting,
China University of Petroleum (Beijing),
characterization of Beijing, China; junwen.peng@cup.edu.cn;
junwen@utexas.edu
organic-rich mudrocks Junwen Peng is a lecturer at China University
of Petroleum (Beijing). He received his B.S.
Junwen Peng and Toti E. Larson degree in resource exploration engineering
(petroleum geology) from China University
of Geosciences (Wuhan) in 2013. In 2016,
he graduated from China University of
ABSTRACT Petroleum (Beijing) with an M.S. degree in
A novel integrated approach for chemofacies characterization of geological resources and geological
engineering (petroleum geology). He
organic-rich mudrocks was developed using principal compo-
obtained his Ph.D. in geology at The
nents analysis of 25 elements from core-based energy-dispersive University of Texas at Austin in 2021. His
x-ray fluorescence (ED-XRF) measurements and k-means cluster- current research interests include shale
ing. Using this approach, three chemofacies were identified that sedimentology, shale petrography, and
capture the spread of the geochemical data sets (5000 ED-XRF sedimentary geochemistry. He is the
measurements from three cores) of the organic-rich Cline Shale corresponding author for this paper.
in the Midland Basin: (1) oxic-suboxic detrital-enriched argilla- Toti E. Larson ~ Bureau of Economic
ceous mudrocks, (2) anoxic siliceous mudrocks, and (3) oxic- Geology (BEG), Jackson School of
suboxic intrabasinal carbonates. Chemofacies defined by the Geosciences, The University of Texas at
covariances of elemental concentrations can be correlated to Austin, Austin, Texas; toti.larson@beg.
depositional environments and primary grain assemblages in a utexas.edu
predictable way, which has significant implications for the evolu- Toti E. Larson is a research associate at the
tion of bulk rock properties, such as total organic carbon (TOC), Bureau of Economic Geology. He received
rock mechanical property, porosity, and permeability. The strati- his Ph.D. from the University of New Mexico
graphically thin anoxic siliceous mudrocks, which are also charac- with a focus on stable isotope and elemental
terized by the highest TOC concentrations, moderate to high geochemistry. His main area of research is
brittleness, and the highest porosity and permeability, are the the integration of geochemistry and
geological characterization to better
“sweet spot” for shale oil production in the Cline Shale. Elevated
understand mudrock systems.
TOC values of the anoxic siliceous mudrocks are most likely
caused by anoxic benthic conditions. This integrated approach for ACKNOWLEDGMENTS
chemofacies characterization of mudrocks is an important step to
fully use all the geochemical data obtained through high- This paper comprises parts of the Ph.D.
project of Junwen Peng at The University of
resolution ED-XRF analyses and provide an economic, efficient,
Texas at Austin and was supported by the
and nondestructive method to study core. State of Texas Advanced Resource Recovery
program at the BEG, the China Scholarship
Council (Grant No. 201606440062), and the
INTRODUCTION Geological Society of America Graduate
Student Research Grant (Grant No.
Deposition of deep-water (i.e., below storm-wave base) organic- 9244823). The Firewheel Energy LLC
rich mudrock has long been considered to be controlled by (Houston, Texas) and Devon Energy
Corporation (Oklahoma City, Oklahoma) are
thanked for donating cores used in this
Copyright ©2022. The American Association of Petroleum Geologists. All rights reserved.
study. The authors thank Patrick Smith at the
Manuscript received December 17, 2020; provisional acceptance April 9, 2021; revised manuscript
received April 28, 2021; final acceptance May 19, 2021.
BEG imaging laboratory for assistance during
DOI:10.1306/05112120210

AAPG Bulletin, v. 106, no. 2 (February 2022), pp. 437–460 437


Downloaded from http://pubs.geoscienceworld.org/aapgbull/article-pdf/106/2/437/5516226/bltn20210.pdf
by China Univ of Geosciences Library user
scanning electron microscope sample pelagic suspension fallout in low-energy environments because of
preparation, Ben Herrmann at Devon Energy the homogenous or structureless depositional textures (e.g.,
Corporation for providing several thin Scholle, 1971; Wynn et al., 2000). However, detailed process-
sections, John Grillo at Premier Oil Field Labs
based studies in recent years suggest that a wide range of transpor-
(Houston, Texas), and Evan Sivil at the BEG
tational and depositional processes, such as low-density turbidity
Mudrock Systems Research Laboratory for
help and guidance during energy-dispersive currents and debris flows, may be responsible for the deposition
x-ray fluorescence analyses and calibration. of deep-water mudrock successions (e.g., Loucks and Ruppel,
Stephen C. Ruppel, Kitty L. Milliken, Robert 2007; Boulesteix et al., 2019; Peng, 2021). Furthermore, the
W. Baumgardner, Scott Hamlin, Robert Reed, addition of new petrographic descriptions (e.g., Milliken et al.,
Nathan Ivicic, and Brandon Williamson at 2012; Alnahwi et al., 2018; Peng et al., 2020a, b), high-resolution
the BEG are thanked for their constructive bulk geochemical characterization (e.g., Rowe et al., 2012;
suggestions and assistance during this study. Alnahwi and Loucks, 2019; Larson et al., 2019; Kabanov et al.,
The publication was authorized by the
2020a, b; Peng et al., 2021), and mechanical characterization
director of the BEG, Jackson School of
Geosciences, The University of Texas at techniques (e.g., Zahm and Enderlin, 2010) in mudrock strata
Austin. indicates that these various properties of mudrocks can vary sig-
nificantly at multiple scales ranging from submillimeter laminae
DATASHARE 145 to outcrops, demonstrating the heterogeneous character of
mudrocks. Understanding such heterogeneity is necessary to
Tables S1 and S2 are available in an
electronic version on the AAPG website
delineate the “sweet spot” that is characterized by high total
(www.aapg.org/datashare) as Datashare organic carbon (TOC) and good reservoir properties (high poros-
145. ity, high permeability, and high content of “brittle minerals”
imparting rigid mechanical properties to the rock), which is
essential for industrial unconventional resources exploration.
Characterization of mudrocks remains hindered mainly by
the required high-resolution petrographic description techniques,
petrophysical detection techniques, or rock mechanical property
analysis, which are extremely time consuming and uneconomic.
Moreover, these traditional laboratory-derived analyses require a
specific size and condition of each mudrock sample, which is
another obstacle to applying those methods at a scale that can
adequately capture the heterogeneity in mudrock successions
with limited samples. Thus, these laboratory-derived methods are
usually impractical to do with the large quantity of samples
required in industrial exploration (e.g., more than thousands).
This study aims to provide a cost-effective and fast way to
characterize mudrocks that can be widely applied in industrial
exploration. Here, we demonstrate a novel integrated protocol
for chemofacies characterization of organic-rich mudrocks based
on energy-dispersive x-ray fluorescence (ED-XRF) elemental
composition measurements, principal components analysis
(PCA), and k-means clustering. Using PCA and k-means cluster-
ing of 5000 ED-XRF measurements from three cores, three
chemofacies were identified from the Upper Pennsylvanian
organic-rich Cline Shale in the Midland Basin, Texas. Each che-
mofacies is characterized by a unique major and trace element
composition and can be correlated to specific depositional envi-
ronments (e.g., sediment provenance, detrital input, redox condi-
tions, and primary productivity) with certain sedimentary struc-
tures and primary grain assemblage composition. In addition,

438 Approach for Mudrock Chemofacies Characterization


Downloaded from http://pubs.geoscienceworld.org/aapgbull/article-pdf/106/2/437/5516226/bltn20210.pdf
by China Univ of Geosciences Library user
each chemofacies characterized by a certain primary sea-level lowstand (Brown et al.,1990; Hentz et al.,
grain assemblage composition will follow predictable 2017; Peng, 2021; Peng et al., 2021). In contrast, car-
sedimentological and diagenetic pathways that have bonate platforms developed on the Eastern shelf and
significant implications for the evolution of bulk rock deep Midland Basin during sea-level highstand
petrophysical and mechanic properties. This study (Brown et al.,1990; Hentz et al., 2017; Peng, 2021;
demonstrates a data analytical approach and protocol Peng et al., 2021).
for mudrock chemofacies characterization. Further-
more, this approach can be applied to a wide range of Problems in Unconventional Resources
geochemical studies that go beyond unconventional Exploration
resource assessment to include any study that
requires large multivariate data sets. For decades, the Upper Pennsylvanian organic-rich
Cline Shale was just considered as a source rock for
hydrocarbon accumulation in the Midland Basin
BACKGROUND (Hamlin and Baumgardner, 2012). Producible shale
oil was recently discovered in the Cline Shale, and
Paleogeography and Depositional good unconventional reservoir properties were
Environment of the Cline Shale observed (Hamlin and Baumgardner, 2012). How-
ever, the high degree of heterogeneity of this marine
During the Late Pennsylvanian, the Midland Basin mudrock succession increases the difficulty of shale
was an epicratonic foreland basin in front of the Oua- oil production (Jacobs, 2013). Specifically, ductile
chita orogenic belt (Figure 1) (Yang and Dorobek, layers with high clay-mineral content are unfavorable
1995). The connection of the Midland Basin to the for hydraulic fracturing. In addition, layers with low
Panthalassa ocean (Figure 1) was restricted by TOC are unfavorable for commercial hydrocarbon in
the serpentine-shaped Great Permian seaway situ accumulation.
(1000 km length), which included several deep
basins and interconnecting shallow-water sills (Figure
1) (Algeo and Heckel, 2008). Thus, the hydrographic DATA AND METHODS
circulation in the Midland Basin was severely
restricted, especially during the sea-level lowstand Three cores discussed in this study are from the
(Peng et al., 2021). Stroman Ranch 5H well, Sterling County, Texas
The rapid subsidence of the Midland Basin dur- (length: 68 m; surface location: 31.9018500 N,
ing the Late Pennsylvanian caused the deposition 101.0751722 W), Adoue 1H well, Mitchell
of the basinal organic-rich Cline Shale (TOC: County, Texas (length: 88 m; surface location:
0.13–9.9 wt. %, mean: 2.64 wt. %; Peng et al., 2021) 32.3093722 N, 100.8947698 W), and Horwood
and mixed siliciclastic-carbonate shelf deposits (Fig- 2151H well, Sterling County, Texas (length: 83 m;
ure 2A) (Brown et al.,1990). The Ouachita orogenic location: 31.9051667 N, 100.9396750 W), respec-
belt to the east of the Eastern shelf is interpreted to tively (Figure 2). All three cores include the entire
have been the dominant siliciclastic provenance for Cline Shale succession. The detailed workflows
the Cline Shale (Brown et al.,1990). Furthermore, for ED-XRF elemental analyses, rock mechanical
previous stratigraphic studies suggest that the high property analyses, and petrographic observation are
frequency and high amplitude glacioeustatic sea-level described below.
variation during the Late Pennsylvanian controlled
the lithofacies cyclicity revealed in the Cline Shale Bulk Geochemical Analyses
(Peng, 2021; Peng et al., 2021). Specifically, silici-
clastic depositional systems transported sediment The ED-XRF measurements for core samples from
westward across the entire Eastern shelf (Figure 2) wells Adoue 1H (1607 measurements; each mea-
from the Ouachita orogenic belt and were responsi- surement contains 25 major and trace elements) and
ble for large amounts of detrital materials deposited Stroman Ranch 5H (1482 measurements; each mea-
in the basin, forming siliciclastic intervals during surement contains 25 major and trace elements) was

PENG AND LARSON 439


Downloaded from http://pubs.geoscienceworld.org/aapgbull/article-pdf/106/2/437/5516226/bltn20210.pdf
by China Univ of Geosciences Library user
Figure 1. Regional paleogeography of the Great Permian seaway (GPS) and Late Pennsylvanian Midcontinent sea (LPMS). The map is
after Peng et al. (2021), which is modified from Blakey (2007) and Algeo et al. (2008). AnB 5 Anadarko Basin; ARM 5 Ancestral Rocky
Mountains, MB 5 Midland Basin.

conducted at 4-cm spacing to obtain high- Horwood 2151H were compiled from Peng et al.
resolution elemental logs. The ED-XRF analyses (2021) (Table S1, supplementary material as AAPG
were conducted using a Bruker Elemental TRACER Datashare 145 at www.aapg.org/datashare). The
IV-SD ED-XRF analyzer at the Bureau of Economic TOC measurements were conducted following the
Geology, The University of Texas at Austin. The modified direct combustion method (Peters and
XRF spectra for major elements were generated Cassa, 1994) using the LECO CS-200 analyzer.
under helium using an Rh tube set at 15 kV and
34.4 mA with 15 s counted time. The XRF spectra Analysis of Rock Mechanical Properties,
for trace elements were analyzed under helium at 40 Porosity, and Permeability
kV and 25 mA (45 s counted time) with an aluminum
(Al)-titanium (Ti)-copper (Cu) filter to attenuate Unconfined compressive strength (UCS) is used as a
characteristic major-element x-rays. All of the raw proxy for rock brittleness in this study (Zahm and
ED-XRF data were then calibrated using proprietary Enderlin, 2010). The microrebound hammer
Bruker calibration software (S1CalProcess; Rowe (MRH)-derived UCS was conducted on the Hor-
et al., 2012). The average unknown standard devia- wood 2151H core for 4-cm spacing (exactly the
tions for different elements obtained from the samples with ED-XRF analyses) (Table S1, supple-
ED-XRF analyses are given in Table 1 (Rowe et al., mentary material as AAPG Datashare 145 at www.
2012). The 4-cm (2-in.) spacing ED-XRF data aapg.org/datashare) using the Equotip Bambino (an
sets of Horwood 2151H well (1649 ED-XRF meas- MRH; Proceq SA, Schwerzenbach, Switzerland) in
urements; each measurement contains 25 major and this study. To avoid the “volume effect” (i.e., smaller
trace elements) are compiled from Peng et al. sample volumes yield significant underestimation of
(2021). the UCS), measurements were only conducted on
A total of 298 bulk TOC measurements were samples with a volume larger than 200 cm3 as sug-
compiled from published literature. Specifically, 115 gested by Brooks et al. (2016). In general, samples
TOC data sets from wells Adoue 1H (43 samples) with a volume less than 200 cm3 are limited and do
and Stroman Ranch 5H (72 samples) were compiled not bias any specific chemofacies. The detailed
from Zheng (2016) (Table S1, supplementary mate- workflow for this measurement follows Zahm and
rial as AAPG Datashare 145 at www.aapg.org/ Enderlin (2010). For each sample (volume >200
datashare). Moreover, 183 TOC data sets from well cm3), 10 measurements were conducted at different

440 Approach for Mudrock Chemofacies Characterization


Downloaded from http://pubs.geoscienceworld.org/aapgbull/article-pdf/106/2/437/5516226/bltn20210.pdf
by China Univ of Geosciences Library user
Figure 2. (A) Map of the Midland Basin showing the distribution of the Cline Shale, regional tectonic unit, and location of studied cores.
The map is after Peng et al. (2020b), which is modified from Wright (2011) and Baumgardner et al. (2016). (B) Upper
Pennsylvanian–lower Permian stratigraphy of the Midland Basin. The map is after Peng et al. (2021), which is modified from Baumgardner
et al. (2016). CBP 5 Central Basin platform; DB 5 Delaware Basin; ES 5 Eastern shelf; HA 5 Horseshoe atoll; MA 5 Matador arch;
MB 5 Midland Basin; NA 5 North America; NM 5 New Mexico; NWS 5 Northwest shelf; OA 5 Ozona arch; SC 5 Sheffield channel;
TEX 5 Texas.

locations along the horizontal bedding. Then, the Petrographic Observation


average value (the highest and lowest value were
removed before calculation) was regarded as the final A total of 265 thin sections (0.9 m spacing on aver-
MRH-derived UCS value for the sample and age) obtained from three cores that captured all
reported as Leeb hardness (HLD; Leeb, 1979). lithofacies were used for conventional transmitted
Fifty-nine Gas Research Institute (GRI) crushed- polarized light microscopy observation. Among
rock porosity and permeability data sets from well them, a total of 32 representative samples were
Horwood 2151H were compiled from Zheng (2016) selected for field-emission scanning electron micro-
(Table S1, supplementary material as AAPG Data- scope (SEM) observation (FEI Nova NanoSEM
share 145 at www.aapg.org/datashare). The GRI 430, Thermo Fisher Scientific, Waltham, MA). The
porosity and permeability were measured via Boyles secondary electron imaging, backscattered electron
Law helium porosimetry on crushed material with imaging, x-ray elemental mapping by energy-
as-received liquids in place (Luffel et al.,1996). dispersive spectroscopy (twin Bruker 30 mm2

PENG AND LARSON 441


Downloaded from http://pubs.geoscienceworld.org/aapgbull/article-pdf/106/2/437/5516226/bltn20210.pdf
by China Univ of Geosciences Library user
Table 1. The Average Unknown Standard Deviations for 2005; Schoepfer et al., 2015). In addition, rare detri-
Elements Obtained from Energy-Dispersive X-Ray Fluorescence tal grains enriched in these two elements (e.g., detri-
Elemental Analyses (Rowe et al., 2012) tal apatite and barite) were identified from the Cline
Shale, and thus, elements Ba and P were used as
Average Unknown Standard
Elements Deviations (Weight Concentration) proxies for primary productivity.
Elements zirconium (Zr), Ti, rubidium (Rb), and
Al, % 0.1 thorium (Th) are enriched in detrital heavy minerals
Si, % 0.2 (e.g., zircon) and generally immobile during the post-
Ca, % 0.01
depositional diagenetic process (Sageman and Lyons,
Mg, % 0.2
2004). Thus, Zr, Ti, Rb, and Th can be good proxies
K, % 0.02
S, % 0.1 for detrital input (Caplan and Bustin, 1999; Doveton
Mn, % 0.002 and Merriam, 2004; Ver Straeten et al., 2011).
Fe, % 0.01 Nine redox-sensitive trace elements (RSTEs;
Ti, % 0.02 cobalt [Co], chromium [Cr], Cu, molybdenum
Zr, ppm 4 [Mo], nickel [Ni], lead [Pb], uranium [U], vanadium
Ba, ppm 147 [V], and zinc [Zn]) and element sulfur (S) exhibit
P, % 0.01
strong enrichment under reducing conditions and
Cu, ppm 14
Ni, ppm 8
considered as redox proxies in this study (Tribovil-
U, ppm 3 lard et al., 2006; Algeo and Liu, 2020). Algeo and
V, ppm 44 Liu (2020) suggest that these nine RSTEs are shown
Zn, ppm 11 to be generally more reliable and superior to other
Cr, ppm 9 redox proxies such as bimetal ratios and the C-S-Fe
Mo, ppm 1 system on the basis of reassessment of redox proxies
Th, ppm 1 from 55 Phanerozoic marine units.
Rb, ppm 3
The other three elements acquired from the
Sr, ppm 1
ED-XRF measurements (strontium, manganese, and
iron) display very mobile behavior in both oxic and
detectors; Bruker Corporation, Billerica, Massachu- anoxic environments (Tribovillard et al., 2006; Wei
setts), and cathodoluminescence (CL) imaging and Algeo, 2020). Thus, these three elements will
(Gatan ChromaCL detector; Gatan, Pleasanton, Cal- not be used as any specific environmental proxies in
ifornia) were integrated during SEM observation. this study.

PCA and k-Means Clustering


Proxies for Chemofacies Characterization
Principal components analysis and k-means clustering
A total of 25 bulk elements can be obtained from the are widely used in geological classification (e.g.,
ED-XRF analyses in this study. These 25 elements Handwerger et al., 2015; Amin et al., 2021). The
can be divided into 5 categories. ED-XRF geochemical data sets were processed using
Six major elements (Al, silicon [Si], calcium the methods outlined in Larson et al. (2019). Briefly,
[Ca], sodium [Na], potassium [K], and magnesium PCA is used to compress the multivariate ED-XRF
[Mg]) compose bulk mineral composition of data sets down to a smaller number of
mudrocks (Baumgardner et al., 2016). Specifically, components (i.e., principal components) that are
we used Al to approximate clay minerals, Si to required to describe the variance and covariance of
approximate quartz, Ca to approximate calcite, Na the major and trace elements. This approach makes
to approximate albite, K to approximate K-feldspar, it much easier to focus on the primary variations of
and Mg to approximate dolomite. the total 25 elements from ED-XRF analysis (Larson
Elements barium (Ba) and phosphorus (P) are et al., 2019). Prior to conducting PCA, the data sets
essential nutrients for marine phytoplankton growth are queried for analytical outliers; samples that have
(Bishop, 1988; Dymond et al., 1992; Schenau et al., concentrations that are greater than four standard

442 Approach for Mudrock Chemofacies Characterization


Downloaded from http://pubs.geoscienceworld.org/aapgbull/article-pdf/106/2/437/5516226/bltn20210.pdf
by China Univ of Geosciences Library user
deviations from the mean for each element are iden- RESULTS
tified as outliers and excluded from PCA.
The PCA of the 25 major and trace elements General Bulk Geochemical and
derived from ED-XRF is conducted by proprietary PCA Results
software Matlab (Matrix Laboratory version 2018b;
MathWorks, United States), in which each element The results of the PCA of 5000 ED-XRF measure-
was treated as a variable. The principal component ments are summarized in Table S2 (supplementary
algorithm is a singular value decomposition (Jolliffe, material as AAPG Datashare 145 at www.aapg.org/
2002). The robustness of PCA results was further datashare). The first principal component (PC 1)
tested on the basis of two standard procedures, Kaiser- accounted for a much higher fraction of total data-
Meyer-Olkin statistic (Cureton and D’Agostino, base variance (27.8%) than higher-order components
1983) and Bartlett sphericity test (Grossman et al., (Figure 3). The second (PC 2) to fifth principal com-
1991), before reporting. We extracted up to 25 princi- ponents (PC 5) accounted for 14.4%, 11.4%, 8.1%,
pal components (Table S2, supplementary material as and 5.4% of the total variance, respectively (Figure
AAPG Datashare 145 at www.aapg.org/datashare). 3). The first five principal components accounted for
The k-means clustering (Lloyd, 1982) was con- 70% of the total variance (Figure 3). In contrast,
ducted on the basis of the variances of the principal higher-order principal components accounted for
components of each mudrock sample reported from very small amounts of the total variance (each less
PCA. The k-means clustering was also performed on than 5%; Figure 3) and yielded very low factor load-
proprietary software Matlab (Matrix Laboratory ver- ings (Table S2, supplementary material as AAPG
sion 2018b; MathWorks, United States). In each Datashare 145 at www.aapg.org/datashare).
cluster, k-means clustering calculates and minimizes Figure 4 is the crossplot of PC 1 and PC 2 with
the sum of the distances between the centroid and all projections of 25 measured elements. The projec-
samples of the cluster. The squared Euclidean dis- tions of siliciclastic residing elements (i.e., Al, Si, Na,
tance metric is used to determine distances. The and K) are adjacent to the detrital input proxies (Th,
most significant advantage of using k-means cluster- Rb, Ti, and Zr), and all of these elements project
ing is that we can manually assign and adjust the toward the negative direction of the PC 1 axis (Figure
number of final clusters (k) based on geological back- 4). In contrast, the projections of carbonate residing
ground and compare clustering solutions for different elements (i.e., Ca and Mg) exhibit a reverse pattern
values of k to determine an optimal number of clus- and project toward the positive direction of the PC 1
ters for the data sets. axis (Figure 4). The projections of nine RSTEs and S

Figure 3. The variance fractions of each principal component (PC). A total of 25 PCs were extracted. The variance explained by a specific
PC is the ratio between the variance of that PC and the variance of the total database.

PENG AND LARSON 443


Downloaded from http://pubs.geoscienceworld.org/aapgbull/article-pdf/106/2/437/5516226/bltn20210.pdf
by China Univ of Geosciences Library user
Figure 4. Crossplot of the first two principal components (PC 1 and PC 2) of the energy-dispersive x-ray fluorescence–derived 25
elements of the Cline Shale. The PC 1 generally represents the partition of mineral-composition–sensitive elements and detrital input prox-
ies; the positive direction is more Ca and Mg-rich, whereas the negative direction has more detrital input proxies and is Al, Si, K, Na-rich.
The PC 2 generally represents the enrichment of redox-sensitive elements; the positive direction is redox-sensitive elements-rich, whereas
the negative direction is redox-sensitive elements-depleted. The k-means clustering analysis reveals three clusters that represent three dif-
ferent chemofacies: oxic-suboxic detrital-enriched argillaceous mudrocks, anoxic siliceous mudrocks, and oxic-suboxic intrabasinal carbo-
nates. Elements highlighted by five different colors indicate five categories of proxies as defined in the legend, respectively. Two elements
with more adjacent projections in the crossplot suggest that they are more correlatable.

are toward the positive direction of the PC 2 axis The nomenclature of each chemofacies in this
(Figure 4). study is based on the enrichment of RSTEs (i.e.,
In summary, PC 1 generally represents the parti- redox conditions: oxic, suboxic, or anoxic), detrital
tion of mineral-composition-sensitive elements: high input proxies (i.e., sediment provenance: detrital-
PC 1 suggests Ca and Mg rich, whereas low PC 1 sug- enriched or intrabasinal), and mineral-composition-
gests Al, Si, K, Na, Th, Rb, Ti, and Zr rich. The PC 2 sensitive elements (i.e., mineral composition:
generally represents the enrichment of RSTEs: high argillaceous, siliceous, or carbonate). Based on these
PC 2 suggests RSTE rich, whereas low PC 2 suggests criteria, three chemofacies were named oxic-suboxic
RSTE depleted. detrital-enriched argillaceous mudrocks, anoxic
siliceous mudrocks, and oxic-suboxic intrabasinal car-
bonates, respectively (Figure 4). The content of
PCA-Based Chemofacies Classification
representative elements and reservoir properties of
different chemofacies is summarized in Table 2.
Three chemofacies clusters were chosen in this study
because of the following reasons (Figure 4): (1) we
show that three clusters are necessary to capture the Oxic-Suboxic Detrital-Enriched Argillaceous
spread of the ED-XRF data sets of the Cline Shale Mudrocks
after reiterations testing, and (2) geologic considera- The oxic-suboxic detrital-enriched argillaceous mud-
tions (i.e., redox conditions, sediment provenance, rocks are characterized by the highest Al content,
and bulk mineral compositions) provide the additional intermediate to high Si content, and low Ca and Mg
justification for the number of clusters ultimately cho- content (Figure 5; Table 2). The concentration of Al,
sen, besides pure mathematical reasoning. Si, Ca, and Mg in oxic-suboxic detrital-enriched

444 Approach for Mudrock Chemofacies Characterization


Downloaded from http://pubs.geoscienceworld.org/aapgbull/article-pdf/106/2/437/5516226/bltn20210.pdf
by China Univ of Geosciences Library user
Table 2. Variations of Representative Element Concentrations and Bulk Rock Properties of Different Chemofacies in the
Cline Shale

Chemofacies
Elements and Reservoir Oxic-Suboxic Detrital-Enriched Oxic-Suboxic Intrabasinal
Properties Argillaceous Mudrocks Anoxic Siliceous Mudrocks Carbonates

Al, wt. % 3.97–9.54 (6.68; 6.78) 2.64–8.75 (5.87; 5.74) 0–4.29 (1.13; 1.46)
Si, wt. % 15.08–34.39 (24.78; 24.62) 17.98–36.50 (27.17; 26.93) 0–25.75 (7.07; 8.95)
Ca, wt. % 0.02–7.75 (1.97; 2.64) 0.03–3.19 (1.13; 1.47) 2.51–37.59 (22.44; 21.89)
Mg, wt. % 0–3.70 (0.96; 0.72) 0–1.79 (0.88; 0.67) 0.09–10.25 (2.46; 3.24)
Ti, wt. % 0.24–0.48 (0.36; 0.36) 0.20–0.43 (0.32; 0.31) 0–0.25 (0.07; 0.09)
Zr, ppm 83.40–187.44 (132.34; 142.81) 37.64–165 (119; 118) 0–93 (26; 34)
Ni, ppm 0–85 (38.52; 39.63) 34–125 (75.84; 79.97) 0–41.02 (17; 19.35)
Cu, ppm 0–57.65 (26; 25.82) 17–93.3 (52; 55.47) 0–31.91 (0; 6.63)
Mo, ppm 0–14 (6; 5.94) 0–21.62 (10; 11) 0–5.78 (0; 1.31)
TOC, wt. % 0.13–3.62 (1.4; 1.52) 0.67–9.88 (5.07; 5.26) 0.19–1.13 (0.70; 0.83)
Porosity, vol. % 0.52–1.90 (1.22; 1.28) 0.46–3.29 (2.04; 2.08) 0.37–1.69 (0.70; 0.86)
4
Permeability, 10 md 0.38–1.01 (0.76; 0.77) 0.57–2.9 (1.53; 1.53) 0.44–0.78 (0.69; 0.69)
UCS, HLD 322–585 (494; 478) 529–675 (586; 591) 636–866 (776; 759)
The first and second value in parentheses suggest the median value and mean value, respectively.
Abbreviations: HLD = Leeb hardness; TOC = total organic carbon; UCS = unconfined compression strength.

argillaceous mudrocks ranges from 3.9% to 9.54% Oxic-Suboxic Intrabasinal Carbonates


(median: 6.68%), 15.08% to 34.39% (median: The oxic-suboxic intrabasinal carbonates are charac-
24.78%), 0.02% to 7.75% (median: 1.97%), and 0% terized by the highest Ca and Mg content, low Al
to 3.7% (median: 0.96%), respectively (Figure 5; content, and low Si content (Figure 5; Table 2). The
Table 2). Contents of detrital proxies (e.g., Ti and Zr) concentration of Al, Si, Ca, and Mg in oxic-suboxic
are highest in the oxic-suboxic detrital-enriched argil- intrabasinal carbonates ranges from 0% to 4.29%
laceous mudrocks (Figure 6; Table 2). In contrast, (median: 1.13%), 0% to 25.75% (median: 7.07%),
contents of redox proxies (e.g., Ni, Cu, and Mo) are 2.51% to 37.59% (median: 22.44%), and 0.09% to
moderate to low in the oxic-suboxic detrital-enriched 10.25% (median: 2.46%), respectively (Figure 5;
argillaceous mudrocks (Figure 7; Table 2). Table 2). Content of both detrital proxies (e.g., Ti
and Zr) and redox proxies (e.g., Ni, Cu, and Mo) are
Anoxic Siliceous Mudrocks lowest in the oxic-suboxic intrabasinal carbonates
The anoxic siliceous mudrocks are characterized by (Figures 6, 7).
the highest Si content in the data set, intermediate
Al content, and low Ca and Mg content (Figure 5; Petrographic Characteristics
Table 2). The concentration of Al, Si, Ca, and Mg in
anoxic siliceous mudrocks ranges from 2.64% to Oxic-Suboxic Detrital-Enriched Argillaceous
8.75% (median: 5.87%), 17.98%–36.50% (median: Mudrocks
27.17%), 0.03%–3.19% (median: 1.13%), and The oxic-suboxic detrital-enriched argillaceous
0%–1.79% (median: 0.88%), respectively (Figure 5; mudrocks are generally structureless to faint lami-
Table 2). The concentration of detrital proxies (e.g., nated at microfacies (thin-section) level of observa-
Ti and Zr) in the anoxic siliceous mudrocks are inter- tion (Figure 8A, B). In most cases, the primary grain
mediate and lower than that of oxic-suboxic detrital- assemblages in the oxic-suboxic detrital-enriched
enriched argillaceous mudrocks (Figure 6; Table 2). argillaceous mudrocks are predominantly composed
Furthermore, the concentration of redox proxies of detrital clay minerals (Figure 8B, C). Besides detri-
(e.g., Ni, Cu, and Mo) are highest in the anoxic sili- tal clay minerals, silt-sized quartz, detrital carbonate
ceous mudrocks (Figure 7). grains, carbonate allochems (commonly extra-large,

PENG AND LARSON 445


Downloaded from http://pubs.geoscienceworld.org/aapgbull/article-pdf/106/2/437/5516226/bltn20210.pdf
by China Univ of Geosciences Library user
Figure 5. Box-and-whisker plots of energy-dispersive x-ray fluorescence (A) Al, (B) Si, (C) Ca, and (D) Mg content for each chemofacies
of the Cline Shale. The “box” depicts the 25th and 75th percentiles of element concentration, the “whiskers” depict the minimum and max-
imum element concentration, and the central line indicates the median value. The outliers are plotted individually using the red cross sym-
bol. Data sets will be classified as outliers if they are greater than q3 1 w 3 (q3 – q1) or less than q1 – w 3 (q3 – q1), where w is the
maximum whisker length, and q1 and q3 are the 25th and 75th percentiles of the sample data, respectively. n 5 number of samples.

millimeter-scale robust molluscan skeletal frag- observed in the oxic-suboxic detrital-enriched argilla-
ments), mica, heavy minerals, and feldspar can be ceous mudrocks (Figure 8F).
also observed chaotically “floating” in the clay matrix
(Figure 8D, E). In general, rare intrabasinal grain Anoxic Siliceous Mudrocks
assemblages were observed in oxic-suboxic detrital- The anoxic siliceous mudrocks have well-defined
enriched argillaceous mudrocks (Figure 8B–D). planar to wavy lamination (Figure 9A, B). This litho-
Furthermore, bioturbations and burrows can be facies is composed of thin (commonly <10 mm)

446 Approach for Mudrock Chemofacies Characterization


Downloaded from http://pubs.geoscienceworld.org/aapgbull/article-pdf/106/2/437/5516226/bltn20210.pdf
by China Univ of Geosciences Library user
Figure 6. Box-and-whisker plots of energy-dispersive x-ray fluorescence (A) Ti and (B) Zr content for each chemofacies. The “box”
depicts the 25th and 75th percentiles of element concentration, the “whiskers” depict the minimum and maximum element concentration,
and the central line indicates the median value. The outliers are plotted individually using the red cross symbol. Data sets will be classified
as outliers if they are greater than q3 1 w 3 (q3 – q1) or less than q1 – w 3 (q3 – q1), where w is the maximum whisker length, and q1
and q3 are the 25th and 75th percentiles of the sample data, respectively. n 5 number of samples.

normally graded beds with erosive basal division based on SEM-based CL observation and quantita-
comprised of detrital silt quartz and rip-up clasts tive point count (Peng et al., 2020a). In addition, no
(Figure 9C) and homogenized top division composed burrows were observed from any samples of anoxic
of organomineralic aggregates (OMAs; Macquaker siliceous mudrocks in this study.
et al., 2010) (Figure 9D, E). Furthermore, in some
specific samples, a high concentration of biosiliceous Oxic-Suboxic Intrabasinal Carbonates
allochems, such as sponge spicules and radiolarians The oxic-suboxic intrabasinal carbonates are generally
(Figure 9F–H), can be identified. These biosiliceous massive and experienced early dolomitization
allochems are generally showing pale-mauve to dark (Figure 10A, B). Most primary grains were replaced
grayish CL color and low CL intensity (red arrow in by the euhedral rhombohedral dolomite, which
Figure 9H) compared to extrabasinal detrital quartz, commonly obscures the primary grain assemblages
which is characterized by brighter reddish CL color in this chemofacies (Figure 10B). In addition,
(yellow arrows in Figure 9H). In addition, the projec- the matrix of oxic-suboxic intrabasinal carbonates is
tion of Si in Figure 4 is slightly deviated from projec- widely cemented by the euhedral rhomb dolomite
tions of detrital proxies, such as Ti and Zr, reflecting (Figure 10C, D). In some samples without heavy dia-
the presence of authigenic quartz in some samples. genetic overprinting, two major categories of primary
However, the detrital silt-sized quartz is still the grain assemblages can be observed: pelagic allo-
dominant form of quartz in anoxic siliceous chems, such as calcispheres (Figure 10B), conodonts
mudrocks from petrographic observation. These bio- (Figure 10F), peloids, and radiolarians (calcified;
siliceous allochems may only occur in very high Figure 10G, H); and benthic carbonate allochems
abundance in limited samples. This observation is from the shallow-water platform (Figure 10E), such
consistent with previous research that detrital quartz as fusulinids, echinoids, crinoids, and robust mollusk
is the dominant form of quartz in the Cline Shale skeletal fragments.

PENG AND LARSON 447


Downloaded from http://pubs.geoscienceworld.org/aapgbull/article-pdf/106/2/437/5516226/bltn20210.pdf
by China Univ of Geosciences Library user
Figure 7. Box-and-whisker plots of energy-dispersive x-ray fluorescence (A) Ni, (B) Cu, and (C) Mo content for each chemofacies. The
“box” depicts the 25th and 75th percentiles of element concentration, the “whiskers” depict the minimum and maximum element concen-
tration, and the central line indicates the median value. The outliers are plotted individually using the red cross symbol. Data sets will be
classified as outliers if they are greater than q3 1 w 3 (q3 – q1) or less than q1 – w 3 (q3 – q1), where w is the maximum whisker length,
and q1 and q3 are the 25th and 75th percentiles of the sample data, respectively. n 5 number of samples.

Bulk TOC, Petrophysical, and Rock UCS (Figure 11; Table 2). The TOC, UCS, poros-
Mechanical Properties ity, and permeability of oxic-suboxic detrital-
enriched argillaceous mudrocks range from 0.13
Oxic-Suboxic Detrital-Enriched Argillaceous wt. % to 3.62 wt. % (median: 1.4 wt. %), 322
Mudrocks HLD to 585 HLD (median: 494 HLD), 0.52% to
The bulk TOC, petrophysical (porosity and perme- 1.90% (median: 1.22%), and 3.8 · 10 5 md to
ability), and rock mechanical properties (reported as 1.01 · 10 4 md (median: 7.6 · 10 5 md), respec-
UCS) of the Cline Shale exhibit significant variations tively (Figure 11; Table 2).
among different chemofacies (Figure 11; Table 2).
Specifically, the oxic-suboxic detrital-enriched argil- Anoxic Siliceous Mudrocks
laceous mudrocks are characterized by low TOC, The anoxic siliceous mudrocks are characterized
low porosity, low permeability, and the lowest by the highest TOC, the highest porosity, the highest

448 Approach for Mudrock Chemofacies Characterization


Downloaded from http://pubs.geoscienceworld.org/aapgbull/article-pdf/106/2/437/5516226/bltn20210.pdf
by China Univ of Geosciences Library user
Figure 8. Representative photographs of oxic-suboxic detrital-enriched argillaceous mudrocks. (A) Core photograph of oxic-suboxic detri-
tal-enriched argillaceous mudrocks showing homogeneous and structureless texture, Adoue 1H, 2256.6 m. (B) Homogeneous and struc-
tureless texture in the thin section. In general, rare silt- or sand-size detrital siliciclastic particles (e.g., quartz and feldspar) are observed. (C)
Enlarged area of box in (B) suggests the matrix of oxic-suboxic detrital-enriched argillaceous mudrocks are almost exclusively composed of
detrital clay minerals. (D) An energy-dispersive spectroscopy map showing that limited silt-sized grains (less than 10%) are mainly com-
posed of detrital quartz (red) with minor amounts of albite (aqua) and calcite (blue) after Peng et al. (2020a). Silt-size grains (less than
10%) appear to “float” in the clay matrix (green) without a touching framework. Horwood 2151H, 2478.8 m. (E) Outsized robust molluscan
skeletal fragments (arrows) were observed to “float” in the clay matrix. Adoue 1H, 2215.7 m. (F) Millimeter-scale burrows (arrows) can be
observed in oxic-suboxic detrital-enriched argillaceous mudrocks. Horwood 2151H, 2967.1 m. BSE 5 backscattered electron.

permeability, and moderate to high UCS (Figure 11; 10 4 md (median: 1.53· 10 4


md), respectively
Table 2). The TOC, UCS, porosity, and permeability (Figure 11; Table 2).
of anoxic siliceous mudrocks range from 0.67 wt. %
to 9.88 wt. % (median: 5.07 wt. %), 529 HLD to Oxic-Suboxic Intrabasinal Carbonates
675 HLD (median: 586 HLD), 0.46% to 3.29% The oxic-suboxic intrabasinal carbonates are character-
(median: 2.04%), and 5.7 · 10 5 md to 2.9 · ized by the lowest TOC, the lowest porosity, the lowest

PENG AND LARSON 449


Downloaded from http://pubs.geoscienceworld.org/aapgbull/article-pdf/106/2/437/5516226/bltn20210.pdf
by China Univ of Geosciences Library user
Figure 9. Representative photographs of anoxic siliceous mudrocks. (A) Core photograph of anoxic siliceous mudrocks showing wavy to
planar laminations. Adoue 1H, 2422.1 m. (B) Thin normally graded beds (separated by the yellow dashed line and red arrows aside) with
basal silt-sized particle-rich division and top organic-rich homogenized division. The basal silt-sized particle-rich division is commonly char-
acterized by a sharp and erosional base (yellow arrow). Horwood 2151H, 2457.3 m. (C) Basal silt-sized particle-rich divisions are mainly
composed of detrital silt-sized quartz (yellow arrows) and rip-up clasts (red arrows). Adoue 1H, 2242.1 m. (D) Enlarged area of box in (B)
suggests top organic-rich homogenized division is mainly composed of detrital clay matrix and organomineralic aggregates (OMA; arrows).
(E) Backscattered electron (BSE) images of OMA. Horwood 2151H, 2457.3 m. (F) Energy-dispersive spectroscopy map of sponge spicules
(arrow). Horwood 2151H, 2457.5 m. (G) Energy-dispersive spectroscopy map of radiolarian (red arrow) and silt-sized extrabasinal detrital
quartz (yellow arrows). Horwood 2151H, 2460.3 m. (H) Cathodoluminescence (CL) image of the dashed-line area in (G). Radiolarian (red
arrow) shows pale-mauve to dark grayish CL color and low intensity compared to extrabasinal detrital quartz (yellow arrows) characterized
by reddish CL color in the rock matrix. Horwood 2151H, 2460.3 m. (B) Image was created after Peng et al. (2021), and images (C) and (E)
were created after Peng (2021).

450 Approach for Mudrock Chemofacies Characterization


Downloaded from http://pubs.geoscienceworld.org/aapgbull/article-pdf/106/2/437/5516226/bltn20210.pdf
by China Univ of Geosciences Library user
Figure 10. Representative photographs of oxic-suboxic intrabasinal carbonates. (A) Core photograph of oxic-suboxic intrabasinal carbo-
nates showing massive texture. Adoue 1H, 2217.1 m. (B) Enlarged area of box in (A) suggests oxic-suboxic intrabasinal carbonates have
experienced early dolomitization. Primary grain assemblages (possibly calcispheres; arrows) are difficult to accurately identify.
(C) Enlarged area of box in (B) suggests the matrix of oxic-suboxic intrabasinal carbonates are cemented by euhedral rhombus dolomite
crystal, the thin section was stained in red using Alizarin Red S. (D) Energy-dispersive spectroscopy map of oxic-suboxic intrabasinal carbo-
nates showing widely distribution of euhedral rhombus dolomite cementation throughout the rock matrix. Horwood 2151H, 2415.1 m. (E)
Fusulinid (yellow arrow) and echinoids (red arrows) from the shallow-water platform can be observed in some samples without heavy
diagenetic overprinting. Horwood 2151H, 2409.8 m (F) Backscattered electron (BSE) images of conodont (arrow). Stroman Ranch 5H,
2494.0 m. (G) Radiolarians (arrows) can be identified in limited samples without significant dolomitization. Stroman Ranch 5H, 2494.0 m.
(H) BSE images of radiolarian (arrow), created after Peng et al. (2020b). Horwood 2151H, 2457.5 m. BSE.

PENG AND LARSON 451


Downloaded from http://pubs.geoscienceworld.org/aapgbull/article-pdf/106/2/437/5516226/bltn20210.pdf
by China Univ of Geosciences Library user
Figure 11. Box-and-whisker plots of (A) total organic carbon (TOC), (B) porosity, (C) permeability, and (D) unconfined compressive
strength (UCS) of each chemofacies. The “box” depicts the 25th and 75th percentiles of element concentration, the “whiskers” depict the
minimum and maximum element concentration, and the central line indicates the median value. The outliers are plotted individually using
the red cross symbol. Data sets will be classified as outliers if they are greater than q3 1 w 3 (q3 – q1) or less than q1 – w 3 (q3 – q1),
where w is the maximum whisker length, and q1 and q3 are the 25th and 75th percentiles of the sample data, respectively. HLD 5 Leeb
hardness; n 5 number of samples.

452 Approach for Mudrock Chemofacies Characterization


Downloaded from http://pubs.geoscienceworld.org/aapgbull/article-pdf/106/2/437/5516226/bltn20210.pdf
by China Univ of Geosciences Library user
permeability, and the highest UCS (Figure 11; Table 2). Anoxic Siliceous Mudrocks
The TOC, UCS, porosity, and permeability of oxic- The anoxic siliceous mudrocks are enriched in RSTEs
suboxic intrabasinal carbonates range from 0.19 wt. % and Si compared to the other two chemofacies
to 1.13 wt. % (median: 0.7 wt. %), 636 HLD to 866 (Figures 5, 7). The highest concentration of Si in
HLD (median: 776 HLD), 0.37% to 1.69% (median: anoxic siliceous mudrocks may attribute to two popu-
0.7%), and 4.4 · 10 5 md to 7.8 · 10 5 md (median: lations of grain assemblages in the basal silt-sized parti-
6.9 · 10 5 md), respectively (Figure 11; Table 2). cle-rich division (Figure 9C): (1) silt-sized detrital
quartz and (2) biosiliceous allochems. The silt-sized
detrital quartz is the dominant form of quartz in the
anoxic siliceous mudrocks (as well as the entire Cline
DISCUSSION Shale) based on petrographic observation. This petro-
graphic evidence is further supported by the moderate
Relationships between Element to high Ti and Zr content in anoxic siliceous mudrocks
Composition, Grain Assemblages, and (Figure 6), suggesting considerable detrital input. This
Depositional Environments discovery suggests that the predominant quartz types
in the Cline Shale are different from some other oil-
Oxic-Suboxic Detrital-Enriched Argillaceous producing anoxic marine shales in the United States,
Mudrocks such as the Mississippian Barnett Shale, Fort Worth
The oxic-suboxic detrital-enriched argillaceous
Basin (Milliken et al., 2012), and the Cretaceous
mudrocks are enriched in Al, K, Si, Na, Th, Zr, Rb,
Mowry Shale, Rocky Mountains (Milliken and Olson,
and Ti and depleted in RSTEs (Table 2). The high 2017), where higher concentrations of biosiliceous
concentration of Al, K, Si, and Na are interpreted to allochems (mainly sponge spicules and radiolarians)
be caused by the high volume of detrital clay miner- were observed. The erosive silt-sized particle-rich
als, as also observed in thin sections (Figure 8). basal division (Figure 9C) was interpreted as distal tur-
Furthermore, the relatively high concentration of Th, bidite during sea-level lowstand by Peng (2021) based
Zr, Rb, and Ti also indicates the detrital siliciclastic on the high concentration of silt-sized detrital quartz,
origin of clayey rock components. The relatively low rip-up clasts, erosional base, planar laminations, and
concentrations of RSTEs indicate an overall suboxic considerable content of detrital proxies. The Ouachita
redox benthic condition, which is consistent with orogenic belt (Figure 1) to the east of the Eastern shelf
burrows encountered in thin sections (Figure 8F) and was an important source of clastic material to the
textures of pyrite framboids (Peng et al., 2021). The basin, and these eroded clastics were redeposited
oxic-suboxic detrital-enriched argillaceous mudrocks increasingly westward, forming distal turbidite in the
are also characterized by moderate to low Ca and Mg basin during sea-level lowstand (Brown et al., 1990;
concentrations. Based on petrographic observation, Peng, 2021). However, between each episodic sub-
Mg and Ca in the oxic-suboxic detrital-enriched aqueous turbidity current, there was sufficient time
argillaceous mudrocks are most likely associated with for the pelagic deposition of the organic-rich homoge-
some specific feldspar, mica, detrital carbonate nized top division as shown in Figure 9D (Peng,
grains, and mollusk skeletal fragments (Figure 8D, 2021). The organic matter in the top division are
E). These robust mollusk skeletal fragments (Figure mainly in the form of OMA and were interpreted to
8E) have been interpreted as debrites derived from be derived from marine algae and associated with pri-
shallow-water shelf (Peng, 2021). mary productivity (Macquaker et al., 2010).
Based on the stratigraphic framework of the The enrichment of RSTEs in sea-level lowstand
Cline Shale established by Peng et al. (2021) and anoxic siliceous mudrocks may be attributed to
Peng (2021), the oxic-suboxic argillaceous mudrocks increased basin restriction and subsequent anoxic ben-
were interpreted as transgressive deposits. These thic conditions (Peng et al., 2021). These seawater sol-
detrital clay minerals are believed to be derived from uble and conservative oxidation states of RSTEs will
siliciclastic transportation systems (e.g., fluvial hyper- transfer into insoluble particle-reactive forms under a
pycnal input) developed on the Eastern shelf during reducing environment and scavenged from the seawa-
this period (Figure 2) (Peng, 2021). ter (Tribovillard et al., 2006; Algeo and Tribovillard,

PENG AND LARSON 453


Downloaded from http://pubs.geoscienceworld.org/aapgbull/article-pdf/106/2/437/5516226/bltn20210.pdf
by China Univ of Geosciences Library user
2009). Furthermore, some specific RSTEs, such as carbonates are interpreted to have been deposited
Mo, Ni, Cu, Zn, and V, are predominantly delivered during sea-level highstand. Siliciclastic transportation
to the sediments in association with organic matter in systems were trapped behind the carbonate platform
the form of organometallic complexes under anoxic on the shelf area (Figure 2). Meager detrital input
conditions (e.g., Helz et al., 1996; Tribovillard et al., and low sedimentation rate favored conditions result-
2004; Algeo and Tribovillard, 2009). Thus, the pres- ing in the widely observed euhedral dolomite cemen-
ence of the high abundance of OMA in the homoge- tation (Figure 10). Suspension fallout and highstand
nized top division of anoxic siliceous mudrocks (Fig- shedding processes (Schlager et al., 1994) were
ure 9D) will further contribute to the enrichment of responsible for the deposition of pelagic allochems
some specific RSTEs. An alternate hypothesis for the (Figure 10F, H) and shelf bioclasts (Figure 10E) in
enrichment of RSTEs in this chemofacies is that the this chemofacies, respectively (Peng, 2021).
RSTEs may be concentrated within rip-up clasts (Fig- It is worth noting that projections of two produc-
ure 9C) that were transported from other parts of the tivity proxies (P and Ba) are not in agreement with
slope or shelf environment. However, these rip-up each other in Figure 4. Furthermore, no certain rela-
clasts (Figure 9C) lack organic matter, which is the tionships between these two elements and other ele-
necessary substance for the scavenge of RSTEs from ments were observed (Figure 4). A likely reason for
the seawater. Furthermore, the contemporary shelf this phenomenon may be caused by remobilization
and slope environment should have been character- of P and Ba after deposition (i.e., during the diagene-
ized by less reducing conditions compared to the basin sis stage). It has been reported that P and Ba in sedi-
floor. Thus, the authors suggest this hypothesis is less ments could be dissolved into pore water in anoxic
likely to have happened. (presence of sulfide) conditions and diffuse upward
from sediment and finally returned to the seawater
Oxic-Suboxic Intrabasinal Carbonates (Brumsack and Gieskes, 1983; Paytan et al., 1993,
The oxic-suboxic intrabasinal carbonates are enriched 2002; Ingall and Jahnke, 1994; Tribovillard et al.,
in Ca and Mg but depleted in RSTEs and detrital 2006). In previous cases reported by Benitez-Nelson
proxies (Figures 5–7). The high content of Ca and (2000), as much as 99% of organic P may be released
Mg in this chemofacies is believed to reflect the high from organic matter decay in anoxic (presence of
abundance of biocalcareous allochems and euhedral sulfide) conditions and then back into the water col-
dolomite cement (Figure 10). Extremely low Zr and umn. Thus, although high primary marine productiv-
Ti content and widespread dolomite cementation ity occurred, Ba and P were not necessarily enriched
suggest limited detrital input during the deposition of in sediments where benthic conditions are anoxic
oxic-suboxic intrabasinal carbonates, which has been (Tribovillard et al., 2006). In addition, some authi-
interpreted as rising sea level and retrogradation of sil- genic minerals (e.g., apatite and barite) that filled
iciclastic systems on the Eastern shelf in previous fractures of the Cline Shale and formed some specific
studies (Brown et al.,1990; Peng, 2021; Peng et al., nodules (e.g., barite and phosphate nodule) linked to
2021). Based on petrographic observation, the pri- biological reactions (Lash, 2015) may also likely have
mary grain assemblages in oxic-suboxic intrabasinal caused some unusual Ba and P enrichment pattern.
carbonates, such as radiolarians, calcispheres, peloids, In addition, the Ba enrichment pattern in the core
and conodonts (Figure 10F–H), mainly originated samples may also likely be affected by drilling fluid.
from the water column (Peng, 2021). Moreover,
minor shallow-water carbonate bioclasts (e.g., Figure Evolution of Bulk Petrophysical and Rock
10E) derived from the carbonate platform developed Mechanical Properties
on the Eastern shelf during this period were deposited
through gravity flows (Schlager et al., 1994; Peng, Oxic-Suboxic Detrital-Enriched Argillaceous
2021). A low concentration of RSTEs suggests overall Mudrocks
oxic-suboxic bottom water conditions. This clay matrix-rich chemofacies (Figure 8) can be
Based on the established stratigraphic framework easily mechanically compacted during the diagenetic
by Peng (2021), the oxic-suboxic intrabasinal process, and thus, rare intergranular or intragranular

454 Approach for Mudrock Chemofacies Characterization


Downloaded from http://pubs.geoscienceworld.org/aapgbull/article-pdf/106/2/437/5516226/bltn20210.pdf
by China Univ of Geosciences Library user
pores can be preserved (Milliken et al., 2012; Emmings et al., 2020) and retaining the pore space.
Milliken, 2014). Reed and Roush (2016) and Peng Finally, the Cline Shale is dominated by organic
et al. (2020b) indicate that very limited intragranular matter-hosted pores (Reed and Roush, 2016; Peng
and intergranular pores can be observed from these et al., 2020a), and thus, the higher concentration of
clay matrix-rich samples of the Cline Shale because organic matter in anoxic siliceous mudrocks will
of intensified mechanical compaction. In agreement increase the bulk porosity and permeability.
with previous studies, relatively low bulk crushed-
rock porosity (median: 1.22%) and permeability Oxic-Suboxic Intrabasinal Carbonates
(median: 7.6 · 10 5 md) were detected in the oxic- Widely distributed carbonate cement (Figure 10) in
suboxic detrital-enriched argillaceous mudrocks in oxic-suboxic intrabasinal carbonates acts as a binder
this study (Figure 11B, C). of other rigid grains and significantly increases rock
Furthermore, rigid grains (generally <10 vol. % of brittleness (medium UCS: 776 HLD; Figure 11D).
quartz, feldspar, and carbonate grains based on petro- However, this widely distributed cement also causes
graphic estimation) in the oxic-suboxic detrital- a dramatic decrease of bulk porosity (median: 0.7%)
enriched argillaceous mudrocks are “floating” in the and permeability (median: 6.9 · 10 5 md) (Figures
ductile clay matrix (Figure 8D). These rigid grains do 11B, C).
not form a touching framework (Figure 8D). As a
result, compaction stress can be accommodated and Mechanisms for Organic Matter
absorbed by particle rearrangement within the duc- Enrichment
tile clay matrix (Milliken et al., 2018). This inference
is consistent with the lowest bulk UCS (median: 494 In general, the highest bulk TOC was detected in
HLD) of oxic-suboxic detrital-enriched argillaceous anoxic siliceous mudrocks (Figure 11A), suggesting
mudrocks as reported in this study (Figure 11D). an overall redox control on bulk TOC enrichment.
The falling sea level (>100 m glacioeustatic falls dur-
Anoxic Siliceous Mudrocks ing the Late Pennsylvanian; Rygel et al., 2008) during
The anoxic siliceous mudrocks are characterized the deposition of anoxic siliceous mudrocks would
by volumetrically significant rigid grains (e.g., silt have resulted in the severe isolation of the Midland
quartz, feldspar, calcite, and biosilicoeus allochems; Basin from open marine waters (Panthalassic ocean
Figure 9). These rigid grains are likely to form a in Figure 1) (Nance and Rowe, 2015; Peng et al.,
touching framework, which could increase rock brit- 2021). This increased basin restriction can result in
tleness by creating force chains that pass through stratified water-column and consistent anoxic bot-
rigid grain contacts (Daniels and Hayman, 2008; tom water conditions, which were favorable for
Emmings et al., 2020; Peng et al., 2020a). This infer- organic matter preservation in deep basins along the
ence is consistent with the relatively high bulk UCS Great Permian seaway (Nance and Rowe, 2015;
(median: 586 HLD) detected in the anoxic siliceous Peng et al., 2021). A similar organic matter enrich-
mudrocks (Figure 11D). ment pattern (i.e., high TOC successions deposited
The highest bulk crushed-rock porosity (median: during sea-level lowstand with increased basin
2.04%) and permeability (median: 1.53 · 10 4 md) restriction) was also reported in the lower Permian
of anoxic siliceous mudrocks (Figures 11B, C) are mudrock successions in the Delaware Basin (west of
likely caused by three factors. First of all, the touch- the Midland Basin; Figure 1) by Nance and Rowe
ing framework of rigid grains will contribute to the (2015).
preservation of intergranular pores by creating pack- Although anoxic siliceous mudrocks contain the
ing flaws below those contacts during mechanical highest bulk TOC (Figure 11A), organic matter is
compactions (Schneider et al., 2011). Second, the not homogeneously distributed throughout the rock
early dissolution of chemically unstable biosiliceous matrix. Based on SEM observation, organic matter in
allochems and precipitation of matrix dispersed the anoxic siliceous mudrocks are almost exclusively
authigenic microquartz in the anoxic siliceous in the form of OMA (Macquaker et al., 2010) and
mudrocks as identified by Peng et al. (2020a) will only enriched in the homogenized top division of
suppress pore collapse (i.e., “buttress effect”; anoxic siliceous mudrocks (Figure 9D). The OMA

PENG AND LARSON 455


Downloaded from http://pubs.geoscienceworld.org/aapgbull/article-pdf/106/2/437/5516226/bltn20210.pdf
by China Univ of Geosciences Library user
was interpreted to have originated from primary 10 5 md), caused by wide dolomitic cementation
algae blooms delivered to the seafloor as suspension throughout the rock matrix, are unfavorable for oil
fallout (“marine snow” of Macquaker et al. [2010]). enrichment in this chemofacies. The Rock-Eval data
The interlaminated fabric of erosive basal silt-sized sets of the Cline Shale from Zheng (2016) suggest
particle-rich division (distal turbidite; Peng, 2021) that oxic-suboxic intrabasinal carbonates are charac-
and organic-rich homogenized top division (pelagic terized by the lowest amount of hydrocarbon that
deposits; Peng, 2021) indicates a highly variable and can be volatilized from the rock (S1)/TOC compared
dynamic depositional process and environment for to the other two chemofacies, indicating limited free
the organic-rich anoxic siliceous mudrocks. Similar hydrocarbon in this chemofacies. Thus, the oxic-
millimeter-scale interlaminated fabric (i.e., interlami- suboxic intrabasinal carbonates are also unlikely to be
nated turbidite and organic-rich pelagic deposits) was the sweet spot for shale oil production.
also identified in many other marine shale systems, The anoxic siliceous mudrocks are characterized
such as Whitby Mudstone Formation, Yorkshire by the highest bulk TOC (median: 5 wt. %), mod-
Coast, England, Kimmeridge Clay Formation, Kim- erate to high UCS (median: 586 HLD), the highest
meridge Bay, England, and Hue Shale, North Slope porosity (median: 2.04%), and the highest perme-
of Alaska (Macquaker et al., 2010). These discoveries ability (median: 1.5 · 10 4 md; Figure 11), indicating
challenge the conventional model that organic-rich relatively good hydrocarbon generation potentials,
mudrocks are deposited by background “rainout” in a hydrocarbon enrichment capacity, and reservoir frac-
consistent quiescent environment with rare detrital turing conditions. The anoxic siliceous mudrocks also
input. exhibit the highest S1/TOC among the other two
chemofacies (Zheng, 2016), suggesting a relatively
“Sweet Spot” for Shale Oil Production in the high concentration of free hydrocarbon in this che-
Cline Shale mofacies. Thus, we suggest the anoxic siliceous
The oxic-suboxic detrital-enriched argillaceous mudrocks of the Cline Shale are the most favorable
mudrocks are characterized by moderate to low bulk sweet spot for shale oil production.
TOC (median <2 wt. %). In addition, the low poros-
ity (median: 1.22%) and permeability (median: 7.5 · Implications for Industrial
10 5 md) of the oxic-suboxic detrital-enriched argil- Unconventional Resources Exploration
laceous mudrocks (Figure 11) are unfavorable for oil
enrichment. The low UCS (median: 494 HLD; The ED-XRF elemental measurement is not destruc-
Figure 11) of oxic-suboxic detrital-enriched argilla- tive for core samples and provides an efficient and
ceous mudrocks suggests a ductile mechanical prop- cost-saving way to obtain large multivariate data sets
erty, which is unfavorable for hydraulic fracturing as needed in mudrocks research. These elemental
and oil production. Thus, the detrital-enriched argil- compositions contain information of mineral compo-
laceous mudrocks are unlikely to be the sweet spot sition (Al, Si, K, Na, Mg, Ca), detrital input (Rb, Th,
for industrial shale oil production, consistent with Ti, and Zr), redox condition (Co, Cr, Cu, Mo, Ni,
the results of previous industrial exploration drilling Pb, U, V, and Zn), and paleoproductivity (Ba and P),
(Jacobs, 2013) and the well-known low value of duc- which can be linked to a certain deposition environ-
tile mudrocks as shale hydrocarbon reservoirs. ment and composition of primary grain assemblages.
The oxic-suboxic intrabasinal carbonates are Moreover, the primary grain assemblages in the
characterized by the highest UCS (median: 777 mudrocks further determined potential diagenetic
HLD; Figure 11), suggesting a rigid mechanical prop- pathways and evolution of bulk petrophysical and
erty and favorable for oil production. However, the mechanic properties (Milliken et al., 2012; Milliken
oxic-suboxic intrabasinal carbonates are also charac- 2014; Peng et al., 2020a, b). Thus, rocks character-
terized by the lowest TOC (median: 0.5 wt. %; ized by significantly different bulk elemental
Figure 11), suggesting limited hydrocarbon genera- compositions (i.e., different chemofacies) are always
tion potential and, thus, unfavorable for in situ oil correlated to different bulk mineral compositions,
enrichment. Furthermore, the extremely low poros- TOC, UCS, porosity, and permeability as shown in
ity (median: 0.7%) and permeability (median: 6.9 · this study.

456 Approach for Mudrock Chemofacies Characterization


Downloaded from http://pubs.geoscienceworld.org/aapgbull/article-pdf/106/2/437/5516226/bltn20210.pdf
by China Univ of Geosciences Library user
Figure 12. Stratigraphical proportions of each chemofacies of the Cline Shale. Note that the anoxic siliceous mudrocks, which are most
favorable for shale oil production, only account for 21% stratigraphic volume of the Cline Shale.

The PCA and k-means clustering are important mudrocks, anoxic siliceous mudrocks, and oxic-
and efficient enough to classify ED-XRF measure- suboxic intrabasinal carbonates, were identified from
ments typical in a big database (e.g., 5000 ED-XRF the Cline Shale. The stratigraphically minor anoxic
measurements; each measurement contains 25 major siliceous mudrocks that are characterized by high
and trace elements in this study). This approach effi- TOC, UCS, porosity, and permeability are inter-
ciently delineates the anoxic siliceous mudrocks as preted as the most favorable target for shale oil pro-
the sweet spots for shale oil production, which only duction. Anoxic benthic conditions are likely the
account for minor stratigraphic volume (21%) of the dominant factors controlling bulk TOC enrichment.
entire Cline Shale (80 m thick in the study area; Interlaminated fabrics in the anoxic siliceous
Figure 12). This rapid, economic, and effective inte- mudrocks suggest a highly variable depositional envi-
grated approach for chemofacies characterization ronment, which challenged the traditional view that
proposed here informs a new idea to discriminate organic-rich mudrocks were deposited in a consis-
favorable sweet spots for the Cline Shale as well as tently deep and quiet environment.
other shale systems in industrial unconventional
exploration. REFERENCES CITED

Algeo, T. J., and J. Liu, 2020, A re-assessment of elemental


CONCLUSIONS proxies for paleoredox analysis: Chemical Geology,
v. 540, p. 119549, doi:10.1016/j.chemgeo.2020.119549.
The ED-XRF-derived elemental composition con- Algeo, T. J., and P. H. Heckel, 2008, The Late Pennsylvanian
tains information of mineral composition (Al, Si, K, Midcontinent Sea of North America: A review: Palaeo-
geography, Palaeoclimatology, Palaeoecology, v. 268, no.
Na, Mg, Ca), detrital input (Rb, Th, Ti, and Zr), 3–4, p. 205–221, doi:10.1016/j.palaeo.2008.03.049.
redox condition (Co, Cr, Cu, Mo, Ni, Pb, U, V, and Algeo, T. J., P. H. Heckel, J. B. Maynard, R. Blakey, and H.
Zn), and paleo-productivity (Ba and P). The PCA Rowe, 2008, Modern and ancient epicratonic seas and
and k-means clustering analyses of 25 elements from the superestuarine circulation model of marine anoxia,
in B. R. Pratt and C. Holmden, eds., Dynamics of
core-based ED-XRF measurements help minimize
epeiric seas: Sedimentological, paleontological and geo-
the number of components necessary to describe all chemical perspectives: Geological Association of Can-
these element variations in a big database and effi- ada Special Publication 48, p. 7–38.
ciently discriminate different chemofacies. This novel Algeo, T. J., and N. Tribovillard, 2009, Environmental analy-
integrated protocol for chemofacies characterization sis of paleoceanographic systems based on molybdenum-
uranium covariation: Chemical Geology, v. 268, no. 3–4,
as demonstrated by the Cline Shale in this study is p. 211–225, doi:10.1016/j.chemgeo.2009.09.001.
cost-effective, efficient, and nondestructive to core Alnahwi, A., and R. G. Loucks, 2019, Mineralogical composi-
samples, which exhibits huge potential for wide tion and total organic carbon quantification using x-ray
applications in industrial unconventional resources fluorescence data from the Upper Cretaceous Eagle Ford
Group in southern Texas: AAPG Bulletin, v. 103, no. 12,
exploration. p. 2891–2907, doi:10.1306/04151918090.
Using this approach, three chemofacies, includ- Alnahwi, A., R. G. Loucks, S. C. Ruppel, R. W. Scott, and N.
ing oxic-suboxic detrital-enriched argillaceous Tribovillard, 2018, Dip-related changes in stratigraphic

PENG AND LARSON 457


Downloaded from http://pubs.geoscienceworld.org/aapgbull/article-pdf/106/2/437/5516226/bltn20210.pdf
by China Univ of Geosciences Library user
architecture and associated sedimentological and geo- Doveton, J. H., and D. F. Merriam, 2004, Borehole petrophysi-
chemical variability in the Upper Cretaceous Eagle cal chemostratigraphy of Pennsylvanian black shales in the
Ford Group in south Texas: AAPG Bulletin, v. 102, no. Kansas subsurface: Chemical Geology, v. 206, no. 3–4,
12, p. 2537–2568, doi:10.1306/05111817310. p. 249–258, doi:10.1016/j.chemgeo.2003.12.027.
Amin, S., M. Wehner, Z. Heidari, and M. M. Tice, 2021, Dymond, J., E. Suess, and M. Lyle, 1992, Barium in deep-sea
Rock classification in the Eagle Ford formation through sediment: A geochemical proxy for paleoproductivity:
integration of petrophysical, geological, geochemical, and Paleoceanography, v. 7, no. 2, p. 163–181, doi:10.1029/
geomechanical characterization: AAPG Bulletin, v. 105, 92PA00181.
no. 7, p. 1357–1381, doi:10.1306/12222016520. Emmings, J. F., P. J. Dowey, K. G. Taylor, S. J. Davies,
Baumgardner, R. W. Jr., H. S. Hamlin, and H. D. Rowe, C. H. Vane, V. Moss-Hayes, and J. Rushton, 2020,
2016, Lithofacies of the Wolfcamp and Lower Leonard Origin and implications of early diagenetic quartz in
Intervals, Southern Midland Basin, Texas: Austin, Texas, the Mississippian Bowland Shale Formation, Craven
The University of Texas at Austin, Bureau of Economic Basin, UK: Marine and Petroleum Geology, v. 120,
Geology Report of Investigations 281, 66 p. 104567, 15 p., doi:10.1016/j.marpetgeo.2020.104567.
Benitez-Nelson, C. R., 2000, The biogeochemical cycling of Grossman, G. D., D. M. Nickerson, and M. C. Freeman,
phosphorus in marine systems: Earth-Science Reviews, 1991, Principal component analyses of assemblage struc-
v. 51, no. 1–4, p. 109–135, doi:10.1016/S0012-8252 ture data: Utility of tests based on eigenvalues: Ecology,
(00)00018-0. v. 72, no. 1, p. 341–347, doi:10.2307/1938927.
Bishop, J. K. B., 1988, The barite-opal-organic carbon associa- Hamlin, H. S. Jr., and R. W. Baumgardner, 2012, Wolfberry
tion in oceanic particulate matter: Nature, v. 332, no. (Wolfcampian–Leonardian) deep-water depositional sys-
6162, p. 341–343, doi:10.1038/332341a0. tems in the Midland Basin: Stratigraphy, lithofacies, res-
Blakey, R. C., 2007, Carboniferous-permianpaleogeography ervoirs, and source rocks: Austin, The University of
of the assembly of Pangaea, in T. E. Wong, ed., Proceed- Texas at Austin, Bureau of Economic Geology Report of
ings of the XVth International Congress on Carbonifer- Investigations 277, 61 p.
ous and Permian Stratigraphy: Amsterdam, Royal Dutch Handwerger, D. A., R. Casta~ neda-Aguilar, G. V. Dahl, H.
Academy of Arts and Sciences, p. 443–456. Borgos, J. Zacharski, D. Krawiec, A. Buniak, W. Prugar,
Boulesteix, K., M. Poyatos-More, S. S. Flint, K. G. Taylor, and R. Suarez-Rivera, 2015, Multivariate classification
D. M. Hodgson, and S. T. Hasiotis, 2019, Transport and for the integration of core, log and seismic data on inter-
deposition of mud in deep-water environments: Pro- secting pre-stack inverted 2D seismic lines, Lublin Basin,
cesses and stratigraphic implications: Sedimentology, Poland: Society of Petroleum Engineers/AAPG/Society
v. 66, no. 7, p. 2894–2925, doi:10.1111/sed.12614. of Exploration Geophysicists Unconventional Resources
Brooks, D., X. Janson, and C. Zahm, 2016, The effect of sam- Technology Conference, San Antonio, Texas, July
ple volume on micro-rebound hammer UCS measure- 20–22, 2015, doi: 10.15530/URTEC-2015-2179307.
ments in Gulf Coast cretaceous carbonate cores: Helz, G. R., C. V. Miller, J. M. Charnock, J. F. W. Mosselmans,
GCAGS Journal, v. 5, p. 189–202. R. A. D. Pattrick, C. D. Garner, and D. J. Vaughan, 1996,
Brown, L. F. Jr., R. F. Solis-Iriarte, and D. A. Johns, 1990, Mechanism of molybdenum removal from the sea and its
Regional depositional systems tracts, paleogeography, concentration in black shales: EXAFS evidence: Geochi-
and sequence stratigraphy, Upper Pennsylvanian and mica et Cosmochimica Acta, v. 60, no. 19, p. 3631–3642,
Lower Permian strata, north- and west-central Texas: doi:10.1016/0016-7037(96)00195-0.
Austin, Texas, The University of Texas at Austin, Hentz, T. F., W. A. Ambrose, and H. S. Hamlin, 2017, Upper
Bureau of Economic Geology Report of Investigations Pennsylvanian and Lower Permian shelf-to-basin facies
197, 116 p. architecture and trends, Eastern Shelf of the Southern
Brumsack, H. J., and J. M. Gieskes, 1983, Interstitial water Midland Basin, West Texas: Austin, The University of
trace-metal chemistry of laminated sediments from the Texas at Austin, Bureau of Economic Geology Report of
Gulf of California, Mexico: Marine Chemistry, v. 14, Investigations 282, 68 p.
no. 1, p. 89–106, doi:10.1016/0304-4203(83)90072-5. Ingall, E., and R. Jahnke, 1994, Evidence for enhanced phos-
Caplan, M. L., and R. M. Bustin, 1999, Palaeoceanographic phorus regeneration from marine sediments overlain by
controls on geochemical characteristics of organic-rich oxygen depleted waters: Geochimica et Cosmochimica
Exshaw mudrocks: Role of enhanced primary produc- Acta, v. 58, no. 11, p. 2571–2575, doi:10.1016/0016-
tion: Organic Geochemistry, v. 30, no. 2–3, p. 161–188, 7037(94)90033-7.
doi:10.1016/S0146-6380(98)00202-2. Jacobs, T., 2013, Cracking the Cline: A new shale play
Cureton, E. E., and R. B. D’Agostino, 1983, Factor analysis: develops in the Permian Basin: Journal of Petroleum
An applied approach: Hillside, New Jersey, Lawrence Technology, v. 65, no. 11, p. 70–77, doi:10.2118/1113-
Erlbaum Associates, 480 p. 0070-JPT.
Daniels, K. E., and N. W. Hayman, 2008, Force chains in seis- Jolliffe, I. T., 2002, Principal component analysis, 2nd ed.:
mogenic faults visualized with photoelastic granular Berlin, Springer, 488 p.
shear experiments: Journal of Geophysical Research, Kabanov, P., R. Vandenberg, P. Pelchat, M. Cameron, and
v. 113, no. B11, B11411, 13 p., doi:10.1029/ K. Dewing, 2020a, Lithostratigraphy of Devonian
2008JB005781. basinal mudrocks in frontier areas of northwestern

458 Approach for Mudrock Chemofacies Characterization


Downloaded from http://pubs.geoscienceworld.org/aapgbull/article-pdf/106/2/437/5516226/bltn20210.pdf
by China Univ of Geosciences Library user
Canada augmented with ED-XRF technique: Arktos, v. mudstones, Mowry Shale (Cretaceous), Rocky Moun-
6, p. 39–52, doi:10.1007/s41063-020-00074-z. tains, U.S.A.: Journal of Sedimentary Research, v. 87, no.
Kabanov, P., R. Vandenberg, P. Pelchat, M. Cameron, and K. 4, p. 366–387, doi:10.2110/jsr.2017.24.
Dewing, 2020b, Correction to: Lithostratigraphy of Nance, H. S., and H. Rowe, 2015, Eustatic controls on stratig-
Devonian basinal mudrocks in frontier areas of north- raphy, chemostratigraphy, and water mass evolution pre-
western Canada augmented with ED-XRF technique: served in a Lower Permian mudrock succession,
Arktos, v. 6, p. 53–54, doi:10.1007%2Fs41063-020- Delaware Basin, west Texas, USA: Interpretation, v. 3,
00076-x. no. 1, p. SH11–SH25, doi:10.1190/INT-2014-0207.1.
Larson, T., E. Sivil, K. Hattori, R. Loucks, and S. Ruppel, Paytan, A., M. Kastner, E. E. Martin, J. D. Macdougall, and T.
2019, Estimating lithologic facies in argillaceous and Herbert, 1993, Marine barite as a monitor of seawater
carbonate-rich mudrocks using x-ray fluorescence meas- strontium isotope composition: Nature, v. 366, no. 6454,
urements and multivariate statistics: AAPG Annual p. 445–449, doi:10.1038/366445a0.
Convention and Exhibition, San Antonio, Texas, May Paytan, A., S. Mearon, K. Cobb, and M. Kastner, 2002, Origin
19–22, 2019, accessed January 10, 2022, https://www. of marine barite deposits: Sr and S isotope characteriza-
searchanddiscovery.com/abstracts/html/2019/ace2019/ tion: Geology, v. 30, no. 8, p. 747–750, doi:10.1130/
abstracts/2447.html. 0091-7613(2002)030<0747:OOMBDS>2.0.CO;2.
Lash, G. G., 2015, Authigenic barite nodules and carbonate Peng, J., 2021, Sedimentology of the Upper Pennsylvanian
concretions in the Upper Devonian shale succession of organic-rich Cline Shale, Midland Basin: From gravity
western New York–a record of variable methane flux flows to pelagic suspension fallout: Sedimentology, v. 68,
during burial: Marine and Petroleum Geology, v. 59, no. 2, p. 805–833, doi:10.1111/sed.12811.
p. 305–319, doi:10.1016/j.marpetgeo.2014.09.009. Peng, J., Q. Fu, T. Larson, and X. Janson, 2021, Trace-ele-
Leeb, D., 1979, Dynamic hardness testing of metallic materi- mental and petrographic constraints on the severity of
als (abs.): NDT International, v. 12, no. 6, (December hydrographic restriction in the silled Midland Basin dur-
1979), p 274–278, doi:10.1016/0308-9126(79)90087-7. ing the Late Paleozoic Ice Age: GSA Bulletin, v. 133, no.
Lloyd, S., 1982, Least squares quantization in PCM: IEEE
1–2, p. 57–73, doi:10.1130/B35336.1.
Transactions on Information Theory, v. 28, no. 2, Peng, J., K. L. Milliken, and Q. Fu, 2020a, Quartz types in the
p. 129–137, doi:10.1109/TIT.1982.1056489.
Upper Pennsylvanian organic-rich Cline Shale (Wolf-
Loucks, R. G., and S. C. Ruppel, 2007, Mississippian Barnett
camp D), Midland Basin, Texas: Implications for silica
Shale: Lithofacies and depositional setting of a deep-
diagenesis, porosity evolution, and rock mechanical prop-
water shale-gas succession in the Fort Worth Basin,
erties: Sedimentology, v. 67, no. 4, p. 2040–2064, doi:
Texas: AAPG Bulletin, v. 91, no. 4, p. 579–601, doi:10.
10.1111/sed.12694.
1306/11020606059.
Peng, J., K. L. Milliken, Q. Fu, X. Janson, and H. S. Hamlin,
Luffel, D. L., F. K. Guidry, and J. B. Curtis, 1996, Develop-
2020b, Grain assemblages and diagenesis in organic-rich
ment of laboratory and petrophysical techniques for eval-
mudrocks, Upper Pennsylvanian Cline Shale (Wolfcamp
uating shale reservoirs: Final report (October
1986–September 1993) GRI-95/0496: Chicago, Illinois, D), Midland Basin, Texas: AAPG Bulletin, v. 104, no. 7,
Gas Research Institute, 301 p. p. 1593–1624, doi:10.1306/03022018240.
Macquaker, J. H. S., M. A. Keller, and S. J. Davies, 2010, Peters, K. E., and M. R. Cassa, 1994, Applied source rock geo-
Algal blooms and “marine snow”: Mechanisms that chemistry, in L. B. Magoon and W. G. Dow, eds., The
enhance preservation of organic carbon in ancient fine- petroleum system from source to trap: AAPG Memoir
grained sediments: Journal of Sedimentary Research, 60, p. 93–120.
v. 80, no. 11, p. 934–942, doi:10.2110/jsr.2010.085. Reed, R. M., and R. S. Roush, 2016, Pore Systems of the Cline
Milliken, K. L., 2014, A compositional classification for grain Shale, Midland Basin, West Texas: Society of Petroleum
assemblages in fine-grained sediments and sedimentary Engineers/AAPG/Society of Exploration Geophysicists
rocks: Journal of Sedimentary Research, v. 84, no. 12, Unconventional Resources Technology Conference, San
p. 1185–1199, doi:10.2110/jsr.2014.92. Antonio, Texas, August 1–3, 2016, doi:10.15530/urtec-
Milliken, K. L., W. L. Esch, R. M. Reed, and T. Zhang, 2012, 2016-2423781.
Grain assemblages and strong diagenetic overprinting in Rowe, H., N. Hughes, and K. Robinson, 2012, The quantifica-
siliceous mudrocks, Barnett Shale (Mississippian), Fort tion and application of handheld energy dispersive x-ray
Worth Basin, Texas, USA: AAPG Bulletin, v. 96, no. 8, fluorescence (ED-XRF) in mudrock chemostratigraphy
p. 1553–1578, doi:10.1306/12011111129. and geochemistry: Chemical Geology, v. 324–325,
Milliken, K. L., D. K. McCarty, and A. Derkowski, 2018, p. 122–131, doi:10.1016/j.chemgeo.2011.12.023.
Grain assemblages and diagenesis in the tarl-dominated Rygel, M. C., C. R. Fielding, T. D. Frank, and L. P. Birgenhe-
Lower Silurian mudrock succession of the western mar- ier, 2008, The magnitude of late Paleozoic glacioeustatic
gin of the east European craton in Poland and Lithuania: fluctuations: A synthesis: Journal of Sedimentary Re-
Sedimentary Geology, v. 374, p. 115–133, doi:10.1016/ search, v. 78, no. 8, p. 500–511, doi:10.2110/jsr.2008.
j.sedgeo.2018.07.011. 058.
Milliken, K. L., and T. Olson, 2017, Silica diagenesis, porosity Sageman, B. B., and T. W. Lyons, 2004, Geochemistry
evolution, and mechanical behavior in siliceous of fine-grained sediments and sedimentary rocks, in

PENG AND LARSON 459


Downloaded from http://pubs.geoscienceworld.org/aapgbull/article-pdf/106/2/437/5516226/bltn20210.pdf
by China Univ of Geosciences Library user
F. Mackenzie, ed., Treatise on geochemistry, vol. 7: Ver Straeten, C. A., C. E. Brett, and B. B. Sageman, 2011,
New York, Elsevier, p. 115–158. Mudrock sequence stratigraphy: A multi-proxy (sedi-
Schenau, S. J., G. J. Reichart, and G. J. De Lange, 2005, Phos- mentological, paleobiological and geochemical)
phorus burial as a function of paleoproductivity and approach, Devonian Appalachian Basin: Palaeogeogra-
redox conditions in Arabian Sea sediments: Geochimica phy, Palaeoclimatology Palaeoecology, v. 304, no. 1–2,
et Cosmochimica Acta, v. 69, no. 4, p. 919–931, doi:10. p. 54–73, doi:10.1016/j.palaeo.2010.10.010.
1016/j.gca.2004.05.044. Wei, W., and T. J. Algeo, 2020, Elemental proxies for paleosa-
Schlager, W., J. J. G. Reijmer, and A. Droxler, 1994, High-
linity analysis of ancient shales and mudrocks: Geochi-
stand shedding of carbonate platforms: Journal of Sedi-
mica et Cosmochimica Acta, v. 287, p. 341–366, doi:10.
mentary Research, v. 64, no. 3B, p. 270–281, doi:10.
1016/j.gca.2019.06.034.
1306/D4267FAA-2B26-11D7-8648000102C1865D.
Wright, W. R., 2011, Pennsylvanian paleodepositional evolu-
Schneider, J., P. B. Flemings, R. J. Day-Stirrat, and J. T.
Germaine, 2011, Insights into pore-scale controls on tion of the greater Permian Basin, Texas and New Mex-
mudstone permeability through resedimentation experi- ico: Depositional systems and hydrocarbon reservoir
ments: Geology, v. 39, no. 11, p. 1011–1014, doi:10. analysis: AAPG Bulletin, v. 95, no. 9, p. 1525–1555, doi:
1130/G32475.1. 10.1306/01031110127.
Schoepfer, S. D., J. Shen, H. Y. Weil, R. V. Tyson, E. Ingall, Wynn, R. B., D. G. Masson, D. A. V. Stow, and P. P. E. Weaver,
and T. J. Algeo, 2015, Total organic carbon, organic phos- 2000, The Northwest African slope apron: A modern ana-
phorus, and biogenic barium fluxes as proxies for paleo- logue for deep water systems with complex seafloor topog-
marine productivity: Earth-Science Reviews, v. 149, raphy: Marine and Petroleum Geology, v. 17, no. 2,
p. 23–52, doi:10.1016/j.earscirev.2014.08.017. p. 253–265, doi:10.1016/S0264-8172(99)00014-8.
Scholle, P. A., 1971, Sedimentology of fine-grained Yang, K. M., and S. L. Dorobek, 1995, The Permian Basin
deepwater carbonate turbidites, Monte Antola Flysch of west Texas and New Mexico: Flexural modeling
(upper cretaceous), Northern Apennines, Italy: GSA and evidence for lithospheric heterogeneity across
Bulletin, v. 82, no. 3, p. 629–658, doi:10.1130/0016-
the Marathon Foreland, in S. L. Dorobek and
7606(1971)82[629:SOFDCT]2.0.CO;2.
G. M. Ross, eds., Stratigraphic evolution of foreland
Tribovillard, N., T. J. Algeo, T. Lyons, and A. Riboulleau,
basins: Tulsa, Oklahoma, SEPM Special Publication 52,
2006, Trace metals as paleoredox and paleoproductivity
proxies: An update: Chemical Geology, v. 232, no. 1–2, p. 149–174.
p. 12–32, doi:10.1016/j.chemgeo.2006.02.012. Zahm, C. K., and M. Enderlin, 2010, Characterization of rock
Tribovillard, N., A. Riboulleau, T. Lyons, and F. Baudin, strength in Cretaceous strata along the Stuart City Trend,
2004, Enhanced trapping of molybdenum by Texas: GCAGS Journal, v. 60, p. 693–702.
sulfurized marine organic matter of marine origin in Zheng, H., 2016, Sedimentology and Reservoir Characteriza-
Mesozoic limestones and shales: Chemical Geology, tion of the Upper Pennsylvanian Cline Shale, Midland
v. 213, p. 385–401, doi:10.1016/j.chemgeo.2004.08. Basin, Texas, Master’s thesis, Austin, Texas, The Univer-
011. sity of Texas at Austin, 141 p.

460 Approach for Mudrock Chemofacies Characterization


Downloaded from http://pubs.geoscienceworld.org/aapgbull/article-pdf/106/2/437/5516226/bltn20210.pdf
by China Univ of Geosciences Library user

You might also like