Uploaded by matt.summersgill

Arkoma Zircon Dating Thomas et al 2021

advertisement
Research Paper
GEOSPHERE
GEOSPHERE, v. 17, no. 4
https://doi.org/10.1130/GES02288.1
Detrital-zircon analyses, provenance, and late Paleozoic
sediment dispersal in the context of tectonic evolution of the
Ouachita orogen
William A. Thomas1, George E. Gehrels2, Kurt E. Sundell2, and Mariah C. Romero3
15 figures; 3 tables; 2 supplemental files
Emeritus, University of Kentucky, and Geological Survey of Alabama, P.O. Box 869999, Tuscaloosa, Alabama 35486-9999, USA
Department of Geosciences, University of Arizona, Tucson, Arizona 85721, USA
3
Department of Earth Sciences, Montana State University, Bozeman, Montana 59717-3480, USA
1
2
CORRESPONDENCE: geowat@uky.edu
CITATION: Thomas, W.A., Gehrels, G.E., Sundell, K.E.,
and Romero, M.C., 2021, Detrital-​zircon analyses,
provenance, and late Paleozoic sediment dispersal
in the context of tectonic evolution of the Ouachita
orogen: Geosphere, v. 17, no. 4, p. 1214–​1247, https://​
doi​.org​/10.1130​/GES02288.1.
Science Editor: David E. Fastovsky
Associate Editor: Todd LaMaskin
Received 22 May 2020
Revision received 21 October 2020
Accepted 17 February 2021
Published online 14 May 2020
This paper is published under the terms of the
CC‑BY-NC license.
ABSTRACT
New analyses for U-Pb ages and εHft values,
along with previously published U-Pb ages, from
Mississippian–Permian sandstones in synorogenic
clastic wedges of the Ouachita foreland and nearby
intracratonic basins support new interpretations
of provenance and sediment dispersal along the
southern Midcontinent of North America. Recently
published U-Pb and Hf data from the Marathon
foreland confirm a provenance in the accreted
Coahuila terrane, which has distinctive Amazonia/
Gondwana characteristics. Data from Pennsylvanian–Permian sandstones in the Fort Worth basin,
along the southern arm of the Ouachita thrust belt,
are nearly identical to those from the Marathon
foreland, strongly indicating the same or a similar
provenance. The accreted Sabine terrane, which is
documented by geophysical data, is in close proximity to the Coahuila terrane, suggesting the two
are parts of an originally larger Gondwanan terrane.
The available data suggest that the Sabine terrane
is a Gondwanan terrane that was the provenance of
the detritus in the Fort Worth basin. Detrital-zircon
data from Permian sandstones in the intracratonic
Anadarko basin are very similar to those from the
Fort Worth basin and Marathon foreland, indicating
sediment dispersal from the Coahuila and/or Sabine
terranes within the Ouachita orogen cratonward
from the immediate forelands onto the southern
craton. Similar, previously published data from the
Permian basin suggest widespread distribution
William A. Thomas
https://orcid.org/0000-0003-0550-258X
from the Ouachita orogen. In contrast to the other
basins along the Ouachita-​Marathon foreland, the
Mississippian–Pennsylvanian sandstones in the
Arkoma basin contain a more diverse distribution
of detrital-zircon ages, indicating mixed dispersal
pathways of sediment from multiple provenances.
Some of the Arkoma sandstones have U-Pb age distributions like those of the Fort Worth and Marathon
forelands. In contrast, other sandstones, especially
those with paleocurrent and paleogeographic
indicators of southward progradation of depositional systems onto the northern distal shelf of the
Arkoma basin, have U-Pb age distributions and εHft
values like those of the “Appalachian signature.” The
combined data suggest a mixture of detritus from
the proximal Sabine terrane/Ouachita orogenic
belt with detritus routed through the Appalachian
basin via the southern Illinois basin to the distal
Arkoma basin. The Arkoma basin evidently marks
the southwestern extent of Appalachian-derived
detritus along the Ouachita-Marathon foreland and
the transition southwestward to overfilled basins
that spread detritus onto the southern craton from
the Ouachita-Marathon orogen, including accreted
Gondwanan terranes.
■■ INTRODUCTION
The very thick (~15 km) accumulation of Upper
Mississippian–Middle Pennsylvanian siliciclastic
turbidites in the Ouachita Mountains of Arkansas
and Oklahoma (Figs. 1 and 2) (e.g., Arbenz, 1989a)
has long challenged understanding of sediment
provenance and dispersal. Early interpretations
envisioned a continental borderland, “Llanoria,”
which supplied sediment to the Ouachita geosyncline (e.g., Schuchert, 1923). More recently, Ouachita
orogenesis, along with subsidence and filling of a
synorogenic basin, is understood in the context of
southward subduction of the margin of Laurentian continental crust beneath a volcanic arc and
accreted continental terrane during assembly of
super­continent Pangaea (e.g., summary in Viele and
Thomas, 1989). Interpretations of late Paleozoic sediment provenance and dispersal in the Arkoma and
Fort Worth basins in the foreland of the Ouachita
salient of the late Paleozoic thrust belt (Fig. 1) range
from proximal Ouachita orogenic sources (e.g.,
Brown et al., 1973; Sutherland, in Johnson et al.,
1988; Thompson, in Johnson et al., 1988) to distant
sources in the Appalachian orogen (Graham et al.,
1976), as well as Appalachian sources with other contributions from the Laurentian craton (Xie et al., 2016,
2018a). Other interpretations include a mixture of
proximal orogenic sources in the Ouachita orogen,
including the accreted Sabine terrane, and more distally supplied sediment from the Appalachians (e.g.,
Sharrah, 2006; Alsalem et al., 2018), perhaps passing
through the Illinois basin. Detrital-zircon data from
the Marathon foreland, which is southwest along
the orogen from the Ouachita salient (Fig. 1A), document sediment supply from a temporally diverse
assemblage of rocks in the accreted Coahuila
terrane (Thomas et al., 2019). Paleocurrents in turbidite fans in the Arkoma basin indicate significant
slope-parallel flow with sediment input from both
the cratonic side and the orogen (e.g., Houseknecht,
© 2021 The Authors
GEOSPHERE | Volume 17 | Number 4
Downloaded from http://pubs.geoscienceworld.org/gsa/geosphere/article-pdf/17/4/1214/5362417/1214.pdf
by Idaho State University user
Thomas et al. | Detrital zircons and late Paleozoic sediment dispersal, Ouachita orogen
1214
Research Paper
uth
Appalachian basement (Grenville) massifs
AN
MI
R
AR
KO
Ok
lah
om
P.C.+
a
PA
LA
CH
IA N
INO
IS
AP
35N
A-
O
EE
ANN
SUW
INE
30N
Ou
rift
a
SA
40N
Ozark
B
no
30N
A
UIL
AH
CO
SAMPLE SITES
U-Pb and Hf data reported herein
N
ER
SI
published U-Pb data plotted herein
RA
published U-Pb and Hf data
RE
AD
100W
AN
AD
AR
OK
-5-
KO
GB
2
OK-9-RS2
W
ich
ita
lahoma
35N
FOREST
CITY
KS-5-WH
P
OK-4-W
L
V
Arb
uck
le
ue
ns
te
r
Lla
E
IN
B
A
Wa
co
g
Lu
lin
100W
HS
S
no
30N
35N
O
FO
BS
ST
N
0
95W
Thomas et al. | Detrital zircons and late Paleozoic sediment dispersal, Ouachita orogen
Downloaded from http://pubs.geoscienceworld.org/gsa/geosphere/article-pdf/17/4/1214/5362417/1214.pdf
by Idaho State University user
Aws
ITA
H
C
UA
RT
TX-5-CC
Benton
Ane
Br
o
RT
H
SC
WO
Bend
SC SC
B
BB
B B
AR-6-SV
SP
B
BB
ARKOMA
M
TX-2-PB
W
600 km
80W
91W
AR-1-BV
W
W
AR-2-HC
Ane
Aws
400
85W
95W
Central Ok
Figure 1. Regional map of the structural setting of the Arkoma
and Fort Worth basins in the foreland of the Ouachita orogen,
and the Anadarko and other intracratonic basins (names of
intracratonic basins in all-capital letters; names of intracratonic arches and uplifts in capital and lowercase letters: Cent.
Okla.—Central Oklahoma arch). (A) Map of the Iapetan rift margin (A-O—Alabama-Oklahoma transform fault; Oua—Ouachita
rift) and synrift intracratonic faults of Laurentia (from Thomas,
2014); locations of Gondwanan accreted terranes (from Krogh et
al., 1993; Steiner and Walker, 1996; Stewart et al., 1999; Dickinson and Lawton, 2001; Centeno-García, 2005; Poole et al., 2005;
Hibbard et al., 2007; Hatcher, 2010; Martens et al., 2010; Mueller
et al., 2014; Thomas, 2014); trace of the Appalachian-Ouachita
thrust front, and locations of the Ouachita interior metamorphic belt and Appalachian basement massifs of Grenville-age
rocks (compiled from Thomas et al., 1989; Hatcher, 2010); outlines of late Paleozoic clastic wedges along the Ouachita and
Appalachian orogens (from Thomas, 2006); and location of
the Prairie Creek lamproite pipe (P.C.) (from Clift et al., 2018).
Locations of sample sites discussed in the text; some symbols
represent multiple closely spaced samples. (B) Map enlarged
from Figure 1A with labels for samples. Map shows locations
of late Paleozoic basement uplifts (Wichita, Arbuckle, and
Muenster) along the Southern Oklahoma fault system and
subsurface basement uplifts (Benton, Broken Bow, Waco, and
Luling) beneath the Ouachita thrust belt. Identifications of
the samples reported herein are listed in Table 1; previously
reported samples not listed in Table 1 are: Ane—Atoka Formation northeastward dispersal (Sharrah, 2006); Aws—Atoka
Formation westward and southward dispersal (Sharrah, 2006);
SP—Spiro (Atoka) Sandstone (Sharrah, 2006); B—Bloyd Formation (Xie et al., 2018a); BS—Big Saline Formation (Alsalem
et al., 2018); HS—Hot Springs Sandstone (McGuire, 2017); KS5-WH—Whitehorse Group (Thomas et al., 2020); P—Post Oak
Conglomerate (Thomas et al., 2016); SC—Strawn Group and
Canyon Group (Alsalem et al., 2018); ST—Stanley Shale (McGuire, 2017); V—Vanoss Conglomerate (Thomas et al., 2016);
W—Wedington Sandstone (Xie et al., 2016); some symbols
represent multiple closely spaced samples.
MAYA
95W
100W
ke
nB
ow
105W
200
0
M
published U-Pb data
GEOSPHERE | Volume 17 | Number 4
ITA
OUACH
Lla
PE
Gondwanan accreted terranes
Iapetan rifted margin of Laurentia
sfo
(palinspastic location)
rm
(multiple parallel rift symbols represent
lower-plate extension)
synrift intracratonic basement faults
ern
HON
MARAT
Ouachita-Marathon interior metamorphic belt
tran
SALIN
A
N em
aha
FO
RE
ST
CI
TY
n
So
35N
AD
75W
80W
CA
RO
LIN
IA
Appalachian-Ouachita-Marathon thrust front
Cim
arro
approximate outline of thicker parts of OuachitaMarathon late Paleozoic clastic wedges
AN
Cent. Okla.
approximate present erosional cratonward limit
of Alleghanian synorogenic clastic wedges in
the Appalachian basin (small outliers omitted)
d
outlines of late Paleozoic intracratonic basins
40N
B en
EXPLANATION
outline of limit of Mesozoic–Cenozoic Gulf
Coastal Plain cover
MICHIGAN
90W
95W
100W
ILL
105W
A
200 km
30N
91W
1215
Research Paper
Thurman Ss
Boggy Fm
Savanna Ss
McAlester Shale
Hartshorne Ss
shallow-marine shelf
and deltaic facies
Strawn Gp
Middle
Pennsylvanian
30 km
B
eroded
synorogenic
turbidites
Atokan
synsedimentary
basement faults
(Arkoma only)
ATOKA
shallow-marine shelf and deltaic facies
~313 Ma
~338 Ma
Middle Mississippian to Lower Pennsylvanian
synorogenic JOHNS VALLEY
turbidites
JACKFORK
B1
PENN
ARKOMA
Bloyd Fm
Hale Fm
Marble Falls Fm
MISS
FORT WORTH
A
Ouachita frontal ramps and triangle zones
Ouachita high-amplitude frontal ramps
Pitkin Limestone
Comyn Limestone
Fayetteville Shale
Barnett Shale
Batesville Ss
STANLEY
D-M
~313 Ma
3 km
SOUTH (Arkoma)
EAST (Fort Worth)
NORTH
WEST
passive-margin carbonate shelf
Cambrian to Middle Mississippian
~338 Ma
Precambrian basement rocks
Laurentian continental crust
SIL
C
C1
Cisco Gp
Canyon Gp
~310 Ma
Figure 2. Diagrammatic cross sections of Ouachita stratigraphy in the Arkoma and Fort Worth basins at three
stages of depositional history: (A) passive margin (Cambrian–Middle Mississippian); (B) early synorogenic (Middle
Mississippian–Lower Pennsylvanian); and (C) late synorogenic (Middle Pennsylvanian Atoka Formation). Figure is
generalized from data in Houseknecht (1986); Sutherland,
in Johnson et al. (1988); Arbenz (1989a, 2008); and Viele
and Thomas (1989). Colored dots (symbols as explained
in Fig. 1) show stratigraphic locations of samples for
analyses of detrital zircons. Arrows at top of cross section C show cratonward extent of Ouachita thrust-belt
structures. Box A1 lists stratigraphic units in the Cambrian–
Middle Mississippian passive-margin off-shelf deep-​water
succession; symbol # shows stratigraphic levels of samples for detrital-zircon U-Pb analyses reported in McGuire
(2017); C—Cambrian; SIL—Silurian; D-M—Devonian–
Mississippian; Mtn—Mountain; Ss—Sandstone. Box B1
lists stratigraphic units in the Middle Mississippian–Lower
Pennsylvanian distal shallow-marine shelf succession; colored dots show stratigraphic levels of samples for analyses
of detrital zircons; MISS—Mississippian; PENN—Pennsylvanian. Box C1 lists Middle Pennsylvanian–Lower Permian
units stratigraphically above the Atoka Formation in the
distal shallow-marine shelf succession; colored dots show
stratigraphic levels of samples for analyses of detrital
zircons; Md–Up PENN—Middle–Upper Pennsylvanian;
PR—Permian; Fm—Formation; Gp—Group.
ARKOMA
passive-margin off-shelf,
shale-dominated succession
ORDOVICIAN
Md–Up PENN
PR
FORT WORTH
1986; Sutherland, 1988; Sutherland, in Johnson et al.,
1988; Morris, 1989). Thus, the Arkoma and Fort Worth
basins have become the center of controversy, as
well as the potential key, regarding regional-scale
alternatives to sediment dispersal.
This article reports new U-Pb age data and Hf
isotopic data from detrital zircons and summarizes
previously published data from the Upper Mississippian to Permian detritus in the Arkoma and Fort
Worth foreland basins. The primary objective of
this research is to use the detrital-zircon geochronologic and Hf isotopic data to characterize and
GEOSPHERE | Volume 17 | Number 4
Downloaded from http://pubs.geoscienceworld.org/gsa/geosphere/article-pdf/17/4/1214/5362417/1214.pdf
by Idaho State University user
possibly identify the provenance or provenances
of the clastic sediment, especially to distinguish
between proximal synorogenic sources and more
distant sources in the Laurentian craton or Appalachians. Comparisons with detrital-zircon data from
nearby intracratonic basins (both new data from the
Anadarko basin and previously published data from
other intracratonic basins), as well as possible distant sources of sediment, provide tests of possible
dispersal pathways in the context of stratigraphy
and sedimentology, including both into and out of
the Arkoma and Fort Worth basins.
C
thin transitional or oceanic crust
Alabama-Oklahoma
transform fault (Arkoma)
Ouachita rift (Fort Worth)
Arkansas Novaculite
Missouri Mtn Shale
Blaylock Ss #
Polk Creek Shale #
Bigfork Chert
Womble Shale #
Blakely Ss #
Mazarn Shale
Crystal Mtn Ss #
Collier Shale
A1
■■ EVOLUTION OF THE OUACHITA OROGEN
AND FORELAND: TECTONICS AND
SEDIMENTATION
Ouachita Orogen
The late Paleozoic Appalachian-​
O uachita-​
Marathon orogen rims the eastern and southern
margins of the North American craton (Fig. 1A).
Postorogenic Mesozoic–Cenozoic strata of the Gulf
Coastal Plain cover much of the orogenic belt in
southern North America (Fig. 1A), and geologic
Thomas et al. | Detrital zircons and late Paleozoic sediment dispersal, Ouachita orogen
1216
Research Paper
-30
-40 km
seismic-velocity and gravity crustal models display
a thick mass of deformed sedimentary rocks overlying thin transitional or oceanic crust (Keller et
al., 1989; Mickus and Keller, 1992). Deep drill holes
document depositional ages of Middle Pennsylvanian to Middle Permian (Flawn et al., 1961; Nicholas
and Waddell, 1989).
Approximately 100 km south of the abrupt
southern margin of Laurentian continental crust,
the geophysical models show a separate mass of
continental crust identified as the Sabine accreted
terrane (Figs. 1 and 3) (Keller et al., 1989; Mickus
and Keller, 1992), which is inferred to have been
a source of synorogenic clastic sediment to the
Ouachita foreland (e.g., Sharrah, 2006; Alsalem
et al., 2018). No drill holes have penetrated continental basement rocks within the Sabine terrane,
leaving the composition and affinities of the terrane unconfirmed. Xenoliths from the Prairie Creek
lamproite pipe (Fig. 1A), which is near the edge of
the Gulf Coastal Plain and within the subsurface
Gulf Coastal Plain
forearc basin
Sabine terrane
Gondwanan continental crust
oceanic and/or transitional crust
EXPLANATION
SOUTH
continental-margin arc
Triassic graben
successor basin
Ouachita
southern thrust belt
Benton uplift
Laurentian continental crust
Precambrian crystalline rocks
the west-northwest–trending Southern Oklahoma
basement fault system, which intersects and is
overlain directly by the frontal Ouachita thrust
sheets (Fig. 1B). West of the apex of the Ouachita
salient and the Arkoma basin, the Anadarko basin
extends on the craton along the northern side of the
Southern Oklahoma basement fault system (Fig. 1).
The outcrops in the Ouachita Mountains of
Arkansas and Oklahoma, along with deep drill
holes and geophysical data in the subsurface
beneath the Gulf Coastal Plain, provide a profile
of the Ouachita orogen (Fig. 3) (Thomas et al., 1989;
Viele and Thomas, 1989). Thick continental crust of
the Laurentian craton extends southward beneath
the Ouachita sedimentary thrust belt, where gravity
and seismic-velocity models depict an abrupt edge
of the continental crust at the Alabama-Oklahoma
transform fault (Figs. 1–4) (Keller et al., 1989; Mickus
and Keller, 1992; Harry et al., 2003).
South of the Ouachita thrust belt and the underlying abrupt transform margin of Laurentian crust,
AlabamaOklahoma
transform
fault
-20
Ouachita
frontal thrust belt
0
-10
Arkoma basin
NORTH
Maumelle chaotic zone
interpretations rely on drill holes and geophysical
data to supplement observations from the limited
exposures (e.g., Thomas et al., 1989). The orogen
is part of the continent-scale system that resulted
from closure of the oceans that rimmed the Laurentian continent and the late Paleozoic assembly
of supercontinent Pangaea (Thomas, 2019).
The Ouachita salient of the orogen forms a
large-​scale cratonward convex curve with a bend
of ~90° in strike, reflecting inheritance of the trace
of the pre-orogenic Iapetan rifted margin of Laurentia (Fig. 1A) (Thomas, 1977, 2006; Thomas et
al., 2012). The Arkoma foreland basin borders the
apex and northern arm of the Ouachita salient, and
the Fort Worth foreland basin borders the southern
arm (Fig. 1B). Thick (as much as 15 km) Mississippian–Pennsylvanian synorogenic clastic wedges
fill the proximal parts of the foreland basins and
thin dramatically into the distal parts of the basins
(Fig. 2) (Arbenz, 1989a). The two foreland basins are
separated by the Arbuckle basement uplift along
SCALE
0
25
50 km
Middle Pennsylvanian–Permian
successor-basin deposits
Mississippian-Pennsylvanian
synorogenic shallow-marine to deltaic clastic facies
Mississippian–Pennsylvanian
synorogenic turbidites
Cambrian-Middle Mississippian
passive-margin carbonate shelf
Cambrian-Middle Mississippian
passive-margin off-shelf deposits
thrust fault
basement fault
Figure 3. Generalized structural cross section illustrating the components of the Ouachita orogen and the Sabine accreted terrane (compiled from
data and interpretations in Nelson et al., 1982; Lillie et al., 1983; Arbenz, 1989b, 2008; Keller et al., 1989; Thomas et al., 1989; Viele and Thomas, 1989;
Mickus and Keller, 1992). These structural elements constitute the possible proximal orogenic provenance of sediment in the Arkoma basin, the most
proximal part of which is imbricated in the Ouachita thrust belt.
GEOSPHERE | Volume 17 | Number 4
Downloaded from http://pubs.geoscienceworld.org/gsa/geosphere/article-pdf/17/4/1214/5362417/1214.pdf
by Idaho State University user
Thomas et al. | Detrital zircons and late Paleozoic sediment dispersal, Ouachita orogen
1217
Research Paper
GEOSPHERE | Volume 17 | Number 4
Downloaded from http://pubs.geoscienceworld.org/gsa/geosphere/article-pdf/17/4/1214/5362417/1214.pdf
by Idaho State University user
A
100°W
50°N
80°W
90°W
50°N
Superior
>2400 Ma
Trans-Hud
so
2000–180 n
0 Ma
a
0M
10
.1
dc
Mi
kean
Peno 800 Ma
-1
1900
G
13 renv
00 ille
–9
70
M
a
45°N
1
Central Plains
1800–1650 Ma
40°N
e
yolit
-Rh Ma
nite
Gra –1320
0
OK
148
35°N
3
2
S.O
.
539
–53
0
4
eta
n
of rift
La m
(~5 uren argin
30 tia
Ma
)
Ma
renville
-G
o
n
Lla
a
1000 M
1300–
30°N
100°W
Iap
southern part of the Ouachita thrust belt, have U-Pb
zircon ages of 1800–1600 Ma (Clift et al., 2018). The
xenolith ages have been interpreted to represent
the Sabine basement (Clift et al., 2018); however,
the lamproite pipe is near (within ~10 km) the
projected trace of the transform margin of Laurentian crust (e.g., Dunn, 2009) and is ~100 km
north of the northern edge of continental crust
of the Sabine terrane as defined in the geophysical models (Fig. 1A) (Keller et al., 1989; Mickus
and Keller, 1992). The area inferred by Clift et al.
(2018) to be Sabine crust is exactly the same as
the area shown to be thin transitional or oceanic
crust overlain by sedimentary rocks in the seismic-velocity and gravity crustal models (Keller et
al., 1989; Mickus and Keller, 1992). Thus, the data
from the lamproite pipe are most likely applicable
to Laurentian crust along the Alabama-​Oklahoma
transform rather than to the Sabine terrane (Van
Avendonk and Dalziel, 2018). Granite boulders with
ages of 1407–1284 Ma (Bowring, 1984) in Ordovician passive-​margin debris flows in the Ouachita
thrust belt (Stone and Haley, 1977) indicate that the
upper crust along the Laurentian transform margin
is within the Granite-Rhyolite province, consistent
with K-Ar dates (1471–1323 Ma) from the Prairie
Creek xenoliths (Dunn, 2009). The deeper and
older crust represented by the xenoliths may be
equivalent to Yavapai-Mazatzal (or Central Plains)
components of Laurentian crust (Clift et al., 2018).
Following the geophysical models (Keller et al.,
1989; Mickus and Keller, 1992), the Sabine terrane
is isolated along the Laurentian margin, similar to
other accreted terranes (Fig. 1A). The accreted Coahuila terrane within the Marathon salient of the
orogenic belt (Fig. 1A) has basement rocks with
ages of 1232–1214 Ma (Sunsás) and of 580 Ma
(Pan-African/Brasiliano); rocks of Las Delicias volcanic arc have late Paleozoic ages between 331 and
270 Ma (Lopez, 1997; Lopez et al., 2001). Detrital
zircons in Jurassic–Cretaceous proximal sedimentary deposits in local fault-bounded basins within
the Coahuila terrane have distinctive modes of
1485 Ma (Rondonia–San Ignacio), 1254 Ma (Sunsás),
1040 Ma (Sunsás), 562 Ma (Pan-African/Brasiliano),
414 Ma (early–middle Paleozoic), and 282 Ma
(late Paleozoic) (Lawton and Molina-Garza, 2014;
95°W
30°N
0
90°W
500 km
85°W
Ordovician–Silurian
passive-margin deep-water off-shelf
sandstones (n=760)
B
200 400 600 800 1000 1200 1400 1600 1800 2000 2200 2400 2600 2800 3000 3200 3400
Detrital Zircon Age (Ma)
Figure 4. Map of Middle Ordovician sandstones on the Laurentian shelf and detrital-​zircon age
data from pre-orogenic off-shelf sandstones adjacent to the passive margin of Laurentia. (A) Map
of extent and dispersal patterns of Middle Ordovician sandstones in relation to Precambrian
basement provinces (from Van Schmus et al., 1993) and the Iapetan rift margin of Laurentia (from
Thomas, 2014). Red dashed line shows approximate extent of Middle Ordovician sandstone units;
red arrows show inferred dispersal patterns. Red bold ovals show: (1) locations of detrital-zircon
U-Pb age data from Konstantinou et al. (2014); (2 and 3) locations of detrital-zircon U-Pb age
data from Pickell (2012); (4) approximate palinspastic locations to restore Ouachita thrusting of
samples for data shown in Figure 4B. Midc.—Midcontinent rift; OK—Oklahoma; S.O.—Southern
Oklahoma fault system. (B) Composite U-Pb age-probability density plot of data (N = 5, n =
717, from McGuire, 2017; N = 2, n = 43, from Gleason et al., 2002) from stratigraphic units now
displaced in the Ouachita thrust belt from pre-orogenic passive-margin off-shelf sites along the
rift margin (stratigraphic and tectonic setting shown in Fig. 2, box A1).
Thomas et al. | Detrital zircons and late Paleozoic sediment dispersal, Ouachita orogen
1218
Research Paper
Thomas et al., 2019). The Pan-African/Brasiliano
ages indicate that the Coahuila terrane is one of
several Gondwanan terranes, including Sierra
Madre and Maya (Fig. 1A), that may be parts of an
originally larger Oaxaquia terrane (Ortega-Gutierrez
et al., 1995; Lawlor et al., 1999). The Sierra Madre
terrane includes basement rocks with ages of 1235–
978 Ma (Sunsás) and volcanic rocks with ages of
334 Ma (late Paleozoic) (Lawlor et al., 1999; Stewart
et al., 1999; Cameron et al., 2004). The Maya terrane
includes rocks with ages of 554 Ma (Pan-African/
Brasiliano) and of 418–404 Ma (middle Paleozoic)
(Krogh et al., 1993; Steiner and Walker, 1996; Martens et al., 2010). The Sabine terrane (Fig. 1A) may
be either part of Coahuila or Oaxaquia or a separate,
but similar, terrane now inside the Ouachita salient
of the orogenic belt. Palinspastic reconstruction
to restore Triassic rifting and subsequent opening of the Gulf of Mexico places the Maya terrane
(now in Yucatan; Fig. 1A) between and contiguous
to Coahuila on the southwest and Sabine on the
northeast (Pindell and Kennan, 2009). Placing the
Sabine terrane within the assembly of Gondwanan
terranes suggests that the presently undocumented
composition of the Sabine terrane may be generally similar to the well-documented composition
of Coahuila and the compositions of the Maya and
Sierra Madre terranes.
Another possible indicator of the composition
of the Sabine basement is in five successive volcanic tuff units in the Mississippian–Pennsylvanian
Stanley Shale in the Ouachita thrust belt. The distribution of the tuffs indicates dispersal from an
arc complex on the south, consistent with continental-margin arc magmatism associated with
southward subduction beneath a continental terrane (Niem, 1977; Loomis et al., 1994; Shaulis et
al., 2012). The pyroclastic units are pumiceous vitric-crystal tuff and “hammer-ringing hard” siliceous
vitric tuff; the tuffs include a minor component of
shale rip-up clasts, as well as rounded pebbles of
quartzite and basalt (Niem, 1977). U-Pb zircon ages
from the five successive tuff beds are dominantly
328.5 ± 2.7 to 320.7 ± 2.5 Ma, indicating the ages
of tuff crystallization; however, the samples also
include a range of older zircon grains (Shaulis et
al., 2012). In light of the shale rip-up clasts, the
GEOSPHERE | Volume 17 | Number 4
Downloaded from http://pubs.geoscienceworld.org/gsa/geosphere/article-pdf/17/4/1214/5362417/1214.pdf
by Idaho State University user
older ages have been interpreted to be from detrital grains (Shaulis et al., 2012), suggesting that they
represent ages of the provenance of the muddy
turbidites, which was inferred to be Appalachian.
The rounded pebbles of quartzite and basalt, however, imply a more proximal source, probably from
shallow-marine reworking of clasts along the arc
terrane. Alternatively, the lithology and the depositional setting of the ash-flow or ash-fall tuffs (Niem,
1977), particularly the general paucity of lithic clasts
and the very hard vitric-​crystal tuffs, suggest that
the prevolcanic zircons may be xenocrysts from
the magma source at the leading edge of the
Sabine terrane. Even if some of the grains are detrital, the source of the muddy sediment may have
been the proximal Sabine terrane, as suggested
by the rounded pebbles of quartzite and basalt. A
future focused study to evaluate the alternatives
of detrital or xenocrystic zircon grains, as well as
the provenance of the quartzite and basalt pebbles,
is critical to resolving the provenance of this part
of the clastic wedge and the composition of the
Sabine basement.
Drill holes have penetrated rhyolite porphyry
incorporated in the thick mass of Pennsylvanian
sedimentary rocks along the northern leading edge
of the Sabine terrane, further suggesting a continental-margin arc (Nicholas and Waddell, 1989).
Although the Sabine rhyolites are geochemically
distinct from the Stanley tuffs, both have geochemical characteristics of subduction-related magmas
(Loomis et al., 1994).
Pre-Orogenic Tectonic Framework and
Sedimentation
The intersection of the Alabama-Oklahoma
transform fault with the Ouachita rift frames the
Ouachita embayment in the Iapetan rifted margin
of southern Laurentia (Figs. 1 and 4), reflecting the
late stages of continental rifting and breakup of
supercontinent Rodinia (e.g., Thomas, 1977, 2019).
Juvenile synrift igneous rocks along the transform-parallel Southern Oklahoma fault system,
inboard from the rifted margin, indicate the time
of rifting at 539–530 Ma (Wright et al., 1996; Hogan
and Gilbert, 1998; Thomas et al., 2012; Hanson et
al., 2013); coeval early Cambrian synrift clastic and
evaporite deposits in the Argentine Precordillera
terrane, the conjugate to the Ouachita embayment,
confirm the time of rifting (Thomas and Astini, 1996,
1999). Although most of the length of the Laurentian
rift margin (as well as most of the Precordillera) is
within the Grenville province, synrift boulders and
detrital zircons in the Precordillera have ages of the
Granite-​Rhyolite province, indicating that Iapetan
rifting cut across the Grenville Front and excised
part of the Laurentian Granite-Rhyolite province
in the Ouachita embayment (Fig. 4) (Thomas et
al., 2012).
On the passive-margin shelf of southern Laurentia, Cambrian–​
Mississippian deposits are
dominantly shallow-marine carbonates (Fig. 2). The
carbonate succession includes quartzose sand­stone
beds, mostly in the Middle Ordovician, that extend
from the onlap limit on the Canadian Shield (Konstantinou et al., 2014) southward to the shelf margin
around the Ouachita foreland (Fig. 4) (Pickell, 2012;
Thomas et al., 2016). Ages of detrital zircons in the
sandstones represent various provinces of the Laurentian craton, dominantly the Superior province
(Pickell, 2012; Konstantinou et al., 2014).
Cambrian–Mississippian strata in the Ouachita
thrust belt constitute an off-shelf deep-water
mud-​dominated facies that is the counterpart of
the passive-margin shelf carbonates (Fig. 2). The
muddy succession includes carbonate mudstone,
carbonate-​
clast conglomerate, and quartzose
sandstone, all indicating detritus from the shelf
(summaries in Arbenz, 1989a; Viele and Thomas,
1989). The Middle Ordovician Blakely Sandstone
(Fig. 2, box A1) contains boulders of granite and
meta-arkose from basement rocks along scarps
at the rifted continental margin (Stone and Haley,
1977). The granite boulders have U-Pb ages of
1407 ± 13, 1350 ± 30, and 1284 ± 12 Ma; ages of
detrital zircons from the sandstones range from
1350 to 1300 Ma (Bowring, 1984). The range of
ages corresponds to the Granite-Rhyolite and
Grenville provinces of Laurentia (Fig. 4). Samples
from five Ordovician–​Silurian sandstones (Fig. 2,
box A1) have dominant distributions of detrital-zircon ages (Fig. 4) that correspond to the Superior,
Thomas et al. | Detrital zircons and late Paleozoic sediment dispersal, Ouachita orogen
1219
Research Paper
Granite-Rhyolite, and Grenville provinces of the
Laurentian craton (McGuire, 2017). The range of
available ages indicates that the Superior grains
are in grain-flow deposits from the quartzose
sands on the shelf and that the Granite-Rhyolite
and Grenville grains are from the shelf-edge scarps
cut into Laurentian basement rocks. The off-shelf
passive-margin succession includes two prominent
chert units (Fig. 2, box A1)—the Ordovician Bigfork
Chert and the Devonian–Mississippian Arkansas
Novaculite (Arbenz, 1989a)—that indicate low sedimentation rates along the passive margin (Viele
and Thomas, 1989).
Late Paleozoic Orogenesis and
Sedimentation—Arkoma Basin
In the proximal Arkoma basin now imbricated
in the Ouachita thrust belt (Fig. 1B), a thick succession of muddy turbidites (Stanley Shale, 3350 m
thick; Figs. 2B and 5) directly overlies the Arkansas
Novaculite, indicating an abrupt increase in the rate
of deposition of siliciclastic sediment in the off-shelf
deep-water setting and signaling approach to the
trench and initiation of orogeny in Middle Mississippian (Meramecian) time (Viele and Thomas,
1989). The five successive volcanic tuff units with
ages of 328–320 Ma in the Stanley Shale (Fig. 6)
indicate continental-margin arc magmatism on the
leading edge of the Sabine terrane (Niem, 1977;
Shaulis et al., 2012). Very rapid deposition of synorogenic sandy turbidites (Jackfork and Johns
Valley Forma­tions, 2300 m thick; Fig. 2B) continued through the Early Pennsylvanian in a forearc
deep-water off-shelf setting (Morris, 1989).
The Middle Mississippian–Lower Pennsylvanian
succession thins dramatically toward the craton
and consists of shallow-marine and deltaic facies
in the distal foreland (Fig. 2B) (Arbenz, 1989a; Viele
and Thomas, 1989). The shallow-marine carbonates and sandstones generally intertongue with
and pinch out southward into a shale succession
that thickens gradually southward toward the
continental margin (Ogren, 1968; Handford, 1986,
1995; Sutherland, 1988; Sutherland, in Johnson
et al., 1988). Distributions of sandstones, as well
GEOSPHERE | Volume 17 | Number 4
Downloaded from http://pubs.geoscienceworld.org/gsa/geosphere/article-pdf/17/4/1214/5362417/1214.pdf
by Idaho State University user
as paleocurrent data, suggest that most of the
sandstones represent southward prograding depositional systems with regional sediment dispersal
across the craton on the north (Sutherland, 1988;
Xie et al., 2016); some shale units suggest localized
partly starved basins, e.g., the Mississippian Fayetteville Shale (Fig. 2, box B1).
In contrast to the pre-Atoka shallow-marine
facies in the distal Arkoma basin, the Middle
Pennsylvanian Atoka Formation is a much thicker
succession (9000 m thick; Fig. 2C) of sandy turbidites,
indicating a rapid increase in subsidence rate and
sediment-accumulation rate in response to advance
of the accretionary prism (Ouachita allochthon) and
tectonic loading of continental crust (Houseknecht,
1986). Paleocurrent indicators in the turbidites show
northeast-directed currents in the western part of
the deep basin and west-​directed currents in the
eastern part (Sharrah, 2006). The deep-water facies
grade upward toward the top of the Atoka Formation
and toward the foreland into shallow-marine to deltaic facies, indicating filling of the late stages of the
peripheral foreland basin (Houseknecht, 1986). The
Atoka Formation thins cratonward to <400 m at the
distal northern margin of the Arkoma foreland basin
(Fig. 2C) (Sutherland, 1988; Sutherland, in Johnson
et al., 1988; Arbenz, 1989a; Morris, 1989).
Post-Atokan Middle Pennsylvanian stratigraphic
units (Fig. 2, box C1) are preserved only in the distal foreland (not in the proximal basin) and record
deposition in the latest stages of subsidence of the
Arkoma basin (Sutherland, in Johnson et al., 1988).
The succession of sandstones and shales includes
deltaic facies prograding southward into the basin,
as well as some sediment from the orogen on the
south. The Middle Pennsylvanian (middle Desmoinesian) Thurman Sandstone (Fig. 2, box C1)
contains chert-pebble conglomerate derived from
the Ordovician and Devonian cherts in the Ouachita
thrust belt (Sutherland, in Johnson et al., 1988).
Late Paleozoic Orogenesis and
Sedimentation—Fort Worth Basin
The Fort Worth basin (Fig. 1B) is a broad homocline that dips gently eastward from the Bend arch,
an intracratonic uplift that separates the orogenic
foreland basin from the intracratonic Permian basin
(Brown et al., 1973; Kier et al., 1979). The Fort Worth
basin dips more steeply eastward beneath the east-​
dipping Ouachita frontal thrust faults (Fig. 1), and
the eastern proximal fill of the basin is imbricated
in the Ouachita thrust belt (Nicholas, in Arbenz et
al., 1989). Along strike, the Fort Worth basin ends
northward at the Muenster uplift and southward
at the intracratonic Llano uplift (Fig. 1B) (Crosby
and Mapel, 1975).
The initial synorogenic deposits along the
southern arm of the Ouachita salient are muddy turbidites of the Mississippian–Lower Pennsylvanian
Stanley Shale (Fig. 2B), now deep in the subsurface
in the Ouachita thrust belt in east Texas (Mapel et
al., 1979). The Stanley Shale directly overlies the
Arkansas Novaculite at the top of the off-shelf passive-margin succession in the Ouachita thrust belt
in east Texas (Fig. 2), as it does in the outcrops in
the Ouachita Mountains (Arbenz, 1989a; Viele and
Thomas, 1989). A very thick succession of sandy
turbidites in the Lower Pennsylvanian Jackfork Formation overlies the Stanley Shale in the Ouachita
thrust belt. Equivalent shelf-facies rocks west of the
Ouachita thrust belt (Fig. 2, box B1) include relatively thin Upper Mississippian shale and overlying
limestones that grade eastward to muddy facies
(Crosby and Mapel, 1975; Mapel et al., 1979).
The thick (>1800 m) Middle Pennsylvanian
Atoka Formation (Fig. 2C), consisting of muddy
and sandy turbidites, records initial rapid subsidence of the Fort Worth foreland basin, now
imbricated in the Ouachita thrust belt. Fan-delta
and deeper water deposits prograded into the
subsiding basin, which bordered the shelf east of
the Bend arch, where a relatively thin Smithwick
Shale was deposited during Atokan time (Kier et
al., 1979). Near the northern end of the Fort Worth
basin, coarse fan-delta deposits prograded across
the subsiding basin and onto the western shelf
from sources in the fault-bounded Muenster uplift
and along the northern part of the Ouachita thrust
belt, near where it impinges on the eastern part of
the basement uplifts associated with the Southern Oklahoma fault system (including the Muenster
uplift; Fig. 1B) (Thompson, 1982).
Thomas et al. | Detrital zircons and late Paleozoic sediment dispersal, Ouachita orogen
1220
Numerical
Age (Ma)
283
299
304
Post Oak Cgl. OK-10-PO1*
Vanoss Cgl. OK-1,2-VN,V2*
*(Thomas et al., 2016)
306
310
Cisco Gp. TX-5-CC
ss above Camp Colorado Ls.
Cisco Gp. TX-2-PB
ss above Bunger Ls.
Canyon Gp.
(S8, Alsalem et al., 2018)
Strawn Gp.
(S4–S7, Alsalem et al., 2018)
Big Saline Fm.
(S3, Alsalem et al., 2018)
316
Savanna Ss. AR-6-SV
Atoka Fm.
(10 samples, Sharrah, 2006)
Bloyd Fm.
(8 samples, Xie et al., 2018a)
5 tuff beds, Stanley Sh.
(Shaulis et al., 2012)
323
Stanley Sh. (McGuire, 2017)
332
Hot Springs Ss. Mbr. Stanley Sh.
(McGuire, 2017)
359
Permian
basin
Delaware Mtn. Gp.
Anadarko
basin
distal
Anadarko
Fort Worth
basin
Rush Springs Ss. Whitehorse Gp.
Bone Spring Fm.
< unconformity >
Gaptank Fm.
Wolfcamp Fm.
Garber Ss.
Wellington Fm.
Post Oak Cgl.
Vanoss Cgl.
< unconformity >
B
Cisco Gp.
Canyon Gp.
306
310
Cisco Gp.
Strawn Gp.
Haymond Fm.
Big Saline Fm.
Dimple Fm.
Arkoma
basin
Shelburn Fm.
Savanna Ss.
Carbondale Fm.
Atoka Fm.
Bloyd Fm.
Tesnus Fm.
Dixon Ss. Mbr.
< unconformity >
Cane Hill Mbr.
323
Illinois basin
332
B
Proctor Ss. Mbr.
Greene Fm.
Washington Fm.
Monongahela Gp.
Conemaugh Gp.
Princess Fm.
Sharp Mountain Mbr.
Tradewater Fm. Cross Mountain Fm.
Norton/Grundy Fm.
Pottsville Fm.
Sharon Ss.
Caseyville Ss. Corbin Ss. Montevallo Raleigh Ss.
Bottom Creek Fm.
Stanley Sh.tuff
Tar Springs Ss.
Stanley Sh.
Hardinsburg Ss.
Wedington Ss. Mbr.
Cypress Ss.
Batesville Ss.
Aux Vases Ss.
Hot Springs Ss. Mbr.
A
Appalachian
basin
Sewanee Ss.
Tumbling Run Mbr.
Lee Fm.
Raccoon Mtn. Fm. Pocahontas Fm.
Bluestone Fm.
Princeton Ss.
Hinton Fm.
Stony Gap Ss. Mbr.
Mauch
Chunk
Gp.
Marathon
foreland
D
316
Wedington Ss. Mbr. Fayetteville Sh.
(6 samples, Xie et al., 2016)
Batesville Ss. AR-1-BV
349
299
C
Cane Hill Mbr. Hale Fm. AR-2-HC
339
273
304
A
Garber Ss. OK-5-GB2
Wellington Fm. OK-4-WL*
E Road Canyon Fm.
283
Arkoma basin
distal shelf
273
successor basin
Desmoinesian–Guadalupian
shallow-marine
siliciclastic and carbonate rocks
Wolfcampian
Series
GuadaStage
Leonardian lupian
Cisuralian
Virgilian
Desmoinesian
Arkoma
proximal foredeep
Rush Springs Ss. OK-9-RS2
Morrowan
Middle
Permian
Lower
Upper
Middle
Pennsylvanian
Lower
Upper
Middle
Mississippian
Lower
Numerical
Age (Ma)
Wolfcampian
Series
GuadaStage
Leonardian lupian
Cisuralian
Virgilian
Desmoinesian
System
Atokan
Fort Worth basin
distal shelf
Anadarko basin
Morrowan
Atokan
KinderMerahookian Osagean mecian Chesterian
Middle
Permian
Lower
Upper
Middle
Pennsylvanian
Lower
Upper
Middle
Mississippian
Lower
Missourian
Missourian
KinderMerahookian Osagean mecian Chesterian
System
Research Paper
339
349
359
Figure 5. Stratigraphic columns of upper Paleozoic units in the Arkoma, Fort Worth, and Anadarko basins and surrounding regions. (A) Stratigraphic columns
showing units sampled for analyses of detrital zircons from the Arkoma, Fort Worth, and Anadarko basins. Cgl.—Conglomerate; Fm.—Formation; Gp.—Group;
Ls.—Limestone; Mbr.—Member; Ss.—Sandstone; Sh.—Shale. (B) Regional stratigraphic columns showing units sampled for analyses of detrital zircons from localities that are relevant to discussion of regional relations of U-Pb age distributions and Hf isotopic data for the Arkoma, Fort Worth, and Anadarko basins. Units
with names in italics are shown for correlation but are not represented by U-Pb age data. Stratigraphic levels of angular unconformities are shown for the Arkoma
basin, Anadarko basin, and Marathon foreland. The Desmoinesian–Guadalupian successor basin deposits overlie Atokan beds at an angular unconformity in the
interior of the Ouachita thrust belt; the Savanna Sandstone is near the top of the preserved succession in the distal foreland of the Arkoma basin. Mtn.—Mountain.
The time scale is International Commission on Stratigraphy (ICS) International Chronostratigraphic Chart 2019 (Cohen et al., 2013, updated), using correlations
from Gradstein et al. (2004). The letters A–E in the Numerical Age column indicate the approximate stratigraphic levels of the maps in Figure 12.
GEOSPHERE | Volume 17 | Number 4
Thomas et al. | Detrital zircons and late Paleozoic sediment dispersal, Ouachita orogen
Downloaded from http://pubs.geoscienceworld.org/gsa/geosphere/article-pdf/17/4/1214/5362417/1214.pdf
by Idaho State University user
1221
Research Paper
Late Paleozoic Tectonics and
Sedimentation—Anadarko Basin
The Anadarko basin is a deep intracratonic basin
along the north side of the Arbuckle and Wichita
basement uplifts, which are bounded by the Southern Oklahoma fault system (Fig. 1B). On the east,
the north-striking Ouachita thrust front cuts across
the west-northwest–striking uplifts, faults, and
basin; the shallow Central Oklahoma arch separates the eastern part of the Anadarko basin from
the Arkoma basin in the Ouachita foreland (Rascoe
and Johnson, in Johnson et al., 1988; Perry, 1989;
Jusczuk, 2002).
The structure of the Southern Oklahoma fault
system (Fig. 1) includes fractures associated with
emplacement of a very large volume of synrift
magma between 539 and 530 Ma (Denison, 1982;
Hogan and Gilbert, 1998; Thomas et al., 2012; Hanson et al., 2013). A transgressive passive-margin
succession of basal sandstone (middle Upper Cambrian) and overlying shallow-marine carbonates
(Upper Cambrian–Middle Ordovician) overlaps the
Cambrian igneous rocks (Denison, in Johnson et al.,
1988). The Middle Ordovician part of the carbonate
succession includes mature quartzose sandstone.
A relatively thin, unconformity-punctuated,
GEOSPHERE | Volume 17 | Number 4
Downloaded from http://pubs.geoscienceworld.org/gsa/geosphere/article-pdf/17/4/1214/5362417/1214.pdf
by Idaho State University user
composite of five tuff units in Stanley
(only grains older than 340 Ma; n=232)
800
Superior
Penokean &
Trans-Hudson
Central Plains
Granite-Rhyolite
1600
1200
2000
2400
2800
Age groups in Arkoma, Fort Worth, and Anadarko basins
400
800
1200
2000
Age provinces in Amazonia (Gondwana)
3036–2387
B
Central
Amazonia
Ventuari–
Tapajós
Rio Negro–
Juruena
1600
Maroni–Itacaiúnas/
Trans-Amazonian/
Eburnian
2224–
1945
1963–1202
Rondonia–
San Ignacio
1276–914
Sunsás–
Oaxaquia
499– 901–500
335
Pan-African/
Brasiliano/
Goiás arc
Late Paleozoic
334–295
400
Grenville
Iapetan synrift
Acadian
Taconic
Alleghanian
Age provinces in Laurentia (Appalachians)
Middle Paleozoic
Early Paleozoic
The Strawn Group (Desmoinesian) includes
fluvial-​deltaic systems that prograded westward,
filling the Fort Worth basin and lapping over the
Bend arch (Kier et al., 1979). The overlying Canyon
and Cisco Groups (Upper Pennsylvanian, Missourian and Virgilian, to Lower Permian; Fig. 2, box
C1; Fig. 5) include numerous cycles of westward
prograding deltaic sandstones and mudstones
and eastwardly transgressive limestones (Gallo­
way and Brown, 1973; Kier et al., 1979). The vertical
succession of sediment-dispersal systems and
thickness distribution of clastic facies suggest
that post-​Atokan sediment was eroded from the
Ouachita orogen and transported westward as the
postorogenic fill of the Fort Worth basin. Chert-clast
conglomerates (Brown et al., 1973) indicate sediment dispersal from erosion of the Ordovician and/
or Devonian cherts in the Ouachita thrust belt.
2400
2800
A
composite of five tuff units in Stanley
(all grains; n=331; Shaulis et al., 2012)
200
400
600
800 1000 1200 1400 1600 1800 2000 2200 2400 2600 2800 3000 3200 3400
Detrital Zircon Age (Ma)
Figure 6. Zircon age data from tuff units in the Stanley Shale and comparison with ages of Laurentian
and Gondwanan provinces, as well as age groups of detrital zircons in the Arkoma, Fort Worth, and
Anadarko basins. (A) U-Pb age-probability density plots of a composite of five separate tuff units in the
Stanley Shale in the Ouachita thrust belt from data published in Shaulis et al. (2012). The lower plot
shows all data with a very strong mode of 328–320 Ma, the crystallization age of the tuffs; the upper plot
shows only data from prevolcanic (older than 340 Ma) grains. (B) Age-correlation diagram comparing
ages of provinces in Laurentia (Appalachians), provinces in Amazonia (Gondwana), and age groups
of detrital zircons in the Arkoma, Fort Worth, and Anadarko basins. Boundaries of the age groups are
shown by dashed black lines, which are defined in Table 3 and used in Figures 7–10.
heterolithic, Upper Ordovician–Mississippian succession represents deposition on a stable shelf
(Amsden, in Johnson et al., 1988).
In contrast to the earlier history, a very thick,
dominantly clastic succession of Pennsylvanian
age reflects the rise of multiple components of
the Arbuckle and Wichita basement uplifts along
large-magnitude basement faults (Johnson et al.,
1988; Denison, 1989; Perry, 1989). Thickness and
facies distributions of a relatively thin succession of
Upper Mississippian to Lower Pennsylvanian (Morrowan) shale, sandstone, and limestone show no
influence of the uplifts along the Southern Oklahoma
fault system; however, beginning abruptly during
the Morrowan and persisting into the Early Permian, deposition of a coarse proximal facies, called
the “Granite Wash,” indicates steep fault-bounded
uplifts along the southwestern edge of the evolving
Anadarko basin (Rascoe and Johnson, in Johnson
et al., 1988). The “Granite Wash” laps onto an erosion surface, which truncates faulted and folded
rocks that range down through the Paleozoic succession to the synrift Cambrian igneous rocks and
the Precambrian basement. The onlapping proximal Upper Pennsylvanian Vanoss Conglomerate
and Lower Permian Post Oak Conglomerate (Fig. 5)
Thomas et al. | Detrital zircons and late Paleozoic sediment dispersal, Ouachita orogen
1222
Research Paper
Supplemental Table S1. Zircon U-Pb geochronologic analyses by Laser-Ablation Multicollector ICP Mass Spectrometry
Isotope ratios
Analysis
U
(ppm)
206Pb
204Pb
U/Th
206Pb*
207Pb*
±
(% 2 )
207Pb*
235U*
±
(% 2 )
Apparent ages (Ma)
206Pb*
238U
±
(% 2 )
error
corr.
206Pb*
238U*
±
(Ma 2 )
207Pb*
235U
±
(Ma 2 )
206Pb*
207Pb*
±
(Ma 2 )
Best age
(Ma)
±
(Ma 2 )
Conc
(%)
ARKOMA BASIN
Sample AR-6-SV—Pennsylvanian Savanna Sandstone—35°10’35.9” N, 93°35’21.1” W
-Spot 390
773
14579
1.4
16.9043
1.8
0.3739
-Spot 365
434
14438
1.0
14.5049
1.7
0.5922
-Spot 221
108
9627
3.7
17.8883
1.7
0.5062
-Spot 305
207
14957
1.4
17.7836
1.7
0.5111
-Spot 269
200
11971
2.9
17.6677
1.3
0.5224
-Spot 184
118
10824
1.0
17.8162
1.6
0.5197
-Spot 180
121
11724
2.4
17.7809
2.1
0.5239
-Spot 288
119
20866
1.1
17.8343
1.7
0.5232
-Spot 324
273
14323
1.4
17.9879
1.0
0.5253
-Spot 332
252
11876
1.1
18.0025
1.9
0.5307
-Spot 333
65
11780
1.2
17.3988
2.8
0.5532
-Spot 213
235
24264
0.8
17.9421
1.4
0.5366
-Spot 242
190
8377
0.7
17.8276
1.4
0.5405
-Spot 303
182
28181
1.7
17.5640
1.5
0.5601
-Spot 150
66
8759
1.4
17.4998
3.0
0.5648
-Spot 265
13
1680
0.8
17.6333
6.1
0.5633
-Spot 186
488
36393
1.9
18.2511
1.0
0.5470
-Spot 141
194
17252
1.6
17.7716
1.4
0.5734
-Spot 385
95
10513
0.6
17.8632
2.5
0.5729
-Spot 271
87
47903
1.3
16.8692
1.9
0.6106
-Spot 230
85
11379
1.4
17.7234
2.6
0.5950
-Spot 201
50
5193
0.5
17.4903
3.1
0.6031
-Spot 331
488
25954
1.2
17.1398
0.9
0.6365
-Spot 314
319
34900
143.3
16.4687
1.2
0.7351
-Spot 245
146
21144
1.2
16.7720
1.8
0.7620
-Spot 257
40
7046
1.5
15.9827
2.6
0.8082
-Spot 286
89
11057
1.3
16.5290
1.5
0.8047
-Spot 350
203
36513
1.0
16.2655
1.1
0.8808
-Spot 247
58
8741
2.4
14.3665
1.4
1.5232
-Spot 235
77
10175
1.8
14.1613
1.2
1.5058
-Spot 209
126
15113
1.9
14.0813
1.2
1.5278
-Spot 360
99
10794
2.5
14.0610
1.2
1.4794
-Spot 335
293
50237
0.8
13.9590
0.8
1.5989
-Spot 231
274
35248
3.4
13.9175
0.7
1.6315
-Spot 149
213
17600
8.1
13.8596
0.8
1.6983
-Spot 255
142
20841
1.5
13.8384
0.8
1.6100
-Spot 172
23
4355
1.8
13.8019
2.0
1.7236
-Spot 217
204
28581
2.2
13.7971
0.9
1.6117
-Spot 238
154
42188
2.5
13.7608
1.0
1.6026
-Spot 321
46
12087
0.8
13.7245
1.6
1.6315
-Spot 237
124
18603
6.4
13.6713
1.3
1.7064
-Spot 290
79
10241
2.2
13.6711
1.5
1.7505
-Spot 179
341
30896
2.4
13.6618
0.8
1.7370
-Spot 214
152
26336
3.8
13.6611
1.1
1.7330
-Spot 378
96
12695
0.9
13.6456
1.1
1.7908
-Spot 310
458
37788
5.2
13.6202
0.7
1.7681
-Spot 355
109
22230
2.6
13.6194
1.0
1.8006
-Spot 325
251
44260
2.6
13.6144
0.7
1.7336
-Spot 381
149
26769
1.9
13.6090
1.1
1.7480
-Spot 156
282
345499
2.1
13.6083
0.8
1.7795
-Spot 199
11
4940
0.9
13.6052
3.6
1.8193
-Spot 256
61
54924
1.4
13.5913
1.4
1.6713
-Spot 308
390
62072
4.9
13.5907
0.6
1.8061
-Spot 316
394
673714
5.2
13.5897
0.7
1.8366
-Spot 304
69
9928
0.8
13.5858
1.2
1.7872
-Spot 210
56
16450
3.1
13.5557
1.6
1.7538
-Spot 162
96
9795
2.0
13.5512
1.3
1.7553
-Spot 366
80
9176
1.6
13.5512
2.1
1.7309
-Spot 142
266
32040
3.0
13.5321
0.9
1.8527
-Spot 264
517
56527
13.0
13.5126
0.7
1.7817
-Spot 193
41
39103
314.3
13.4870
1.6
1.7184
-Spot 318
44
21745
1.8
13.4856
1.5
1.7630
-Spot 382
47
11046
1.4
13.4834
1.6
1.6459
-Spot 240
269
53083
119.7
13.4661
0.9
1.7793
-Spot 262
46
8007
1.5
13.4537
1.6
1.7542
-Spot 319
36
5969
1.7
13.4354
1.7
1.8123
-Spot 208
29
13606
0.8
13.4346
1.8
1.7164
-Spot 389
153
13055
2.5
13.4283
0.9
1.9217
-Spot 190
195
27188
2.8
13.4248
0.8
1.6665
-Spot 139
50
13692
2.0
13.4004
1.5
1.7394
-Spot 323
259
20903
1.5
13.3868
0.7
1.7213
-Spot 196
61
22511
3.1
13.3686
1.2
1.7750
-Spot 260
312
30321
6.7
13.3682
0.8
1.8598
-Spot 283
149
44542
3.1
13.3634
0.9
1.8761
-Spot 280
29
7765
1.7
13.3626
2.2
1.7407
-Spot 246
58
9754
1.7
13.3526
1.4
1.7412
-Spot 347
68
17890
2.2
13.3519
1.4
1.7970
-Spot 362
87
11505
1.8
13.3462
1.1
1.7404
-Spot 222
86
25984
0.7
13.3286
1.2
1.8468
-Spot 387
93
12612
1.6
13.2994
1.1
1.7602
-Spot 300
147
24962
2.7
13.2992
0.9
1.8301
-Spot 277
67
8033
0.8
13.2975
1.0
1.8173
-Spot 250
64
10949
2.5
13.2962
1.4
1.7292
-Spot 147
172
185369
2.0
13.2943
0.9
1.8369
-Spot 322
316
47332
3.9
13.2900
0.7
1.8782
-Spot 227
35
14946
1.4
13.2298
1.7
1.7739
-Spot 392
61
12856
1.9
13.2266
1.6
1.8963
-Spot 295
61
15771
1.6
13.2170
1.5
1.8312
-Spot 268
47
17567
0.5
13.2169
1.5
1.7857
-Spot 349
54
12691
2.0
13.2089
1.6
1.8460
-Spot 145
41
40337
1.9
13.1888
1.6
1.7285
-Spot 379
10
4384
1.3
13.1567
3.8
2.0068
-Spot 340
66
41003
0.6
13.1307
1.2
1.8658
-Spot 301
37
13859
1.4
13.1303
1.9
1.8512
-Spot 252
134
137930
2.3
13.1282
0.8
1.9088
-Spot 144
64
13348
1.5
13.1208
1.2
1.8067
-Spot 372
362
56556
0.6
13.1070
0.6
1.9586
-Spot 363
306
26210
0.9
13.0983
0.7
1.8611
-Spot 168
461
49078
14.6
13.0557
0.7
1.7914
-Spot 341
40
10578
0.9
13.0269
2.1
1.8396
-Spot 228
113
14026
2.0
13.0134
0.9
1.8614
-Spot 223
128
27857
1.1
13.0099
1.0
1.8975
-Spot 289
14
5830
3.0
13.0046
3.0
1.8790
-Spot 282
67
12868
1.9
12.9701
1.4
1.9867
-Spot 234
5
1660
1.6
12.9685
4.3
2.0660
-Spot 315
277
60302
2.5
12.9605
0.9
2.0099
-Spot 241
61
9407
1.3
12.9565
1.2
2.0476
-Spot 191
280
76833
2.6
12.9495
0.7
2.0915
-Spot 279
216
63046
2.9
12.9487
0.9
2.0436
-Spot 204
35
12545
0.9
12.9467
1.7
1.8925
-Spot 274
389
24405
1.3
12.9434
1.2
2.0894
-Spot 216
10
3266
1.1
12.9279
3.1
1.9867
-Spot 195
101
18452
2.1
12.9081
1.2
1.8865
-Spot 367
47
13670
0.9
12.8667
1.5
1.9731
3.5
2.3
2.7
2.5
2.0
2.7
2.9
2.4
1.9
2.7
3.4
2.1
2.2
2.4
3.7
6.9
2.2
2.1
3.2
2.7
3.2
4.2
2.1
2.2
2.6
3.5
2.3
2.0
2.3
2.0
2.0
2.1
1.9
1.6
1.6
1.6
3.2
1.8
1.9
2.2
2.1
2.4
1.8
1.8
1.9
2.5
2.0
1.7
2.2
1.8
4.5
2.4
1.7
1.9
2.4
2.4
2.0
2.8
1.7
1.5
2.4
2.4
2.5
1.9
2.3
2.5
2.7
1.8
1.8
2.4
2.0
1.9
1.7
1.6
3.0
2.3
2.3
1.9
2.4
2.1
1.8
2.2
2.4
1.8
1.5
2.6
2.2
2.5
2.2
2.6
2.6
4.8
2.0
2.6
1.9
2.1
2.3
2.5
2.3
2.8
2.4
2.1
3.8
2.4
5.7
1.8
2.2
1.8
2.0
2.6
2.9
3.7
2.3
2.6
0.0458
0.0623
0.0657
0.0659
0.0669
0.0671
0.0676
0.0677
0.0685
0.0693
0.0698
0.0698
0.0699
0.0713
0.0717
0.0720
0.0724
0.0739
0.0742
0.0747
0.0765
0.0765
0.0791
0.0878
0.0927
0.0937
0.0965
0.1039
0.1587
0.1547
0.1560
0.1509
0.1619
0.1647
0.1707
0.1616
0.1725
0.1613
0.1599
0.1624
0.1692
0.1736
0.1721
0.1717
0.1772
0.1747
0.1779
0.1712
0.1725
0.1756
0.1795
0.1647
0.1780
0.1810
0.1761
0.1724
0.1725
0.1701
0.1818
0.1746
0.1681
0.1724
0.1610
0.1738
0.1712
0.1766
0.1672
0.1872
0.1623
0.1690
0.1671
0.1721
0.1803
0.1818
0.1687
0.1686
0.1740
0.1685
0.1785
0.1698
0.1765
0.1753
0.1668
0.1771
0.1810
0.1702
0.1819
0.1755
0.1712
0.1768
0.1653
0.1915
0.1777
0.1763
0.1817
0.1719
0.1862
0.1768
0.1696
0.1738
0.1757
0.1790
0.1772
0.1869
0.1943
0.1889
0.1924
0.1964
0.1919
0.1777
0.1961
0.1863
0.1766
0.1841
3.0
1.6
2.1
1.9
1.6
2.1
1.9
1.7
1.6
1.8
2.0
1.6
1.7
1.9
2.2
3.2
2.0
1.6
2.1
1.9
2.0
2.8
1.9
1.8
1.9
2.3
1.7
1.7
1.9
1.5
1.6
1.7
1.8
1.4
1.3
1.4
2.5
1.5
1.6
1.6
1.7
1.9
1.6
1.5
1.5
2.4
1.7
1.6
1.9
1.6
2.7
2.0
1.6
1.7
2.0
1.9
1.5
1.9
1.4
1.4
1.7
1.9
1.8
1.7
1.6
1.9
2.0
1.6
1.6
1.8
1.9
1.5
1.5
1.4
2.0
1.9
1.9
1.5
2.1
1.7
1.6
1.9
1.9
1.5
1.3
2.0
1.5
2.0
1.6
2.0
2.0
3.0
1.7
1.8
1.7
1.7
2.2
2.4
2.2
1.9
2.3
1.8
2.4
1.9
3.7
1.6
1.8
1.6
1.8
2.0
2.7
2.1
1.9
2.2
0.86
0.70
0.78
0.75
0.79
0.79
0.67
0.71
0.83
0.69
0.59
0.75
0.78
0.78
0.59
0.47
0.89
0.76
0.63
0.71
0.62
0.66
0.91
0.83
0.74
0.67
0.76
0.84
0.80
0.77
0.81
0.82
0.91
0.91
0.86
0.88
0.79
0.88
0.85
0.71
0.80
0.79
0.90
0.82
0.81
0.96
0.86
0.90
0.86
0.89
0.60
0.81
0.93
0.93
0.86
0.76
0.78
0.67
0.86
0.90
0.73
0.80
0.75
0.88
0.71
0.75
0.74
0.88
0.90
0.77
0.93
0.78
0.88
0.83
0.67
0.80
0.81
0.81
0.88
0.84
0.88
0.87
0.80
0.86
0.89
0.76
0.68
0.81
0.73
0.77
0.79
0.61
0.82
0.69
0.90
0.83
0.96
0.96
0.95
0.67
0.92
0.88
0.63
0.80
0.65
0.88
0.83
0.92
0.90
0.76
0.92
0.55
0.84
0.82
288.9
389.6
410.0
411.5
417.7
419.0
421.4
422.1
427.3
431.9
435.0
435.1
435.5
444.2
446.3
448.4
450.6
459.6
461.6
464.4
475.1
475.3
490.9
542.5
571.4
577.3
593.7
637.3
949.6
927.0
934.7
905.8
967.2
982.7
1016.0
965.6
1026.1
963.9
956.5
970.1
1007.7
1031.7
1023.7
1021.5
1051.8
1037.7
1055.3
1018.6
1026.1
1043.1
1064.3
983.1
1056.2
1072.5
1045.6
1025.4
1026.0
1012.7
1077.0
1037.5
1001.6
1025.5
962.1
1032.9
1018.5
1048.4
996.9
1106.0
969.3
1006.9
996.2
1023.7
1068.7
1077.0
1004.9
1004.5
1034.2
1003.7
1058.9
1010.9
1047.9
1041.0
994.2
1051.2
1072.6
1013.3
1077.4
1042.5
1018.6
1049.7
986.4
1129.5
1054.3
1046.7
1076.5
1022.7
1100.7
1049.5
1010.1
1033.0
1043.3
1061.7
1051.8
1104.5
1144.7
1115.6
1134.4
1156.1
1131.8
1054.4
1154.5
1101.2
1048.4
1089.5
8.6
6.1
8.5
7.5
6.5
8.5
7.9
7.0
6.4
7.7
8.6
6.8
7.3
8.1
9.5
13.9
8.7
7.2
9.2
8.6
9.2
12.7
9.1
9.6
10.4
12.8
9.8
10.1
16.6
13.1
14.2
14.4
15.9
13.1
12.7
13.0
24.0
13.9
14.1
14.4
15.6
18.1
15.5
14.2
15.0
23.3
16.8
14.8
18.4
15.0
26.8
17.9
15.5
17.1
19.6
17.6
14.6
17.8
14.1
13.2
16.2
18.4
16.5
16.2
15.2
18.4
18.7
16.3
14.5
17.1
17.6
14.0
15.1
13.4
18.5
17.4
17.7
14.3
20.8
16.2
15.2
18.1
17.6
14.8
13.2
18.4
14.7
19.5
15.3
19.3
18.6
30.7
16.4
16.9
16.8
16.5
22.4
23.3
20.9
18.2
21.8
17.7
23.6
19.1
38.8
16.5
19.1
17.4
18.9
19.4
28.4
20.8
18.5
21.7
322.5
472.2
415.9
419.2
426.7
424.9
427.7
427.3
428.7
432.3
447.1
436.2
438.7
451.6
454.6
453.7
443.0
460.2
459.9
483.9
474.1
479.2
500.1
559.5
575.2
601.5
599.5
641.4
939.8
932.7
941.6
922.0
969.8
982.4
1007.9
974.1
1017.4
974.8
971.2
982.4
1010.9
1027.3
1022.4
1020.9
1042.1
1033.8
1045.7
1021.1
1026.4
1038.0
1052.4
997.7
1047.7
1058.6
1040.8
1028.6
1029.1
1020.1
1064.4
1038.8
1015.4
1031.9
988.0
1037.9
1028.7
1049.9
1014.7
1088.7
995.8
1023.2
1016.5
1036.4
1066.9
1072.7
1023.7
1023.9
1044.4
1023.6
1062.3
1030.9
1056.3
1051.7
1019.5
1058.7
1073.4
1036.0
1079.8
1056.7
1040.2
1062.0
1019.2
1117.8
1069.0
1063.9
1084.2
1047.9
1101.4
1067.4
1042.3
1059.7
1067.5
1080.2
1073.7
1111.0
1137.6
1118.9
1131.5
1146.0
1130.2
1078.5
1145.3
1111.0
1076.3
1106.4
9.7
8.7
9.3
8.6
7.1
9.2
10.1
8.3
6.5
9.4
12.4
7.6
7.9
8.8
13.6
25.3
8.0
7.9
12.0
10.4
12.3
16.0
8.3
9.5
11.4
15.7
10.3
9.5
14.4
12.0
12.4
12.6
12.2
10.0
10.0
10.3
20.6
11.1
11.7
14.1
13.4
15.6
11.7
11.8
12.5
16.3
13.2
11.2
14.5
11.4
29.6
15.4
11.2
12.2
15.4
15.8
12.9
18.1
10.9
10.0
15.3
15.8
15.6
12.5
14.8
16.5
17.5
12.2
11.4
15.4
13.1
12.4
11.5
10.9
19.2
15.0
15.0
12.3
16.0
13.4
11.8
14.1
15.4
11.6
10.0
16.7
14.4
16.4
14.4
17.0
16.7
32.6
13.5
16.9
12.6
13.7
15.4
16.5
15.3
18.7
16.2
13.7
25.4
15.9
38.9
12.4
15.0
12.2
13.9
17.4
20.1
25.1
15.1
17.7
572.8
897.0
448.5
461.5
476.0
457.5
461.8
455.2
436.1
434.3
509.8
441.8
456.1
489.0
497.1
480.3
403.7
463.0
451.6
577.4
469.0
498.3
542.7
629.4
589.9
693.6
621.5
656.1
916.8
946.3
957.9
960.9
975.7
981.8
990.3
993.4
998.7
999.4
1004.7
1010.1
1018.0
1018.0
1019.4
1019.5
1021.8
1025.6
1025.7
1026.4
1027.2
1027.3
1027.8
1029.8
1029.9
1030.1
1030.7
1035.2
1035.8
1035.8
1038.7
1041.6
1045.4
1045.7
1046.0
1048.6
1050.4
1053.2
1053.3
1054.2
1054.8
1058.4
1060.5
1063.2
1063.3
1064.0
1064.1
1065.6
1065.7
1066.6
1069.2
1073.6
1073.7
1073.9
1074.1
1074.4
1075.1
1084.2
1084.7
1086.1
1086.1
1087.3
1090.3
1095.3
1099.2
1099.3
1099.6
1100.7
1102.8
1104.1
1110.6
1115.1
1117.2
1117.7
1118.5
1123.8
1124.0
1125.2
1125.9
1126.9
1127.1
1127.4
1127.9
1130.3
1133.4
1139.7
38.8
34.3
37.9
37.0
27.8
36.4
47.5
37.2
23.2
42.6
60.8
31.3
30.5
33.1
66.0
135.2
22.8
31.2
55.6
41.8
56.7
69.1
18.8
26.6
38.1
54.8
31.7
23.5
28.8
25.6
24.5
24.2
16.3
13.6
16.3
16.1
40.1
17.3
20.1
32.0
25.6
30.3
16.0
21.3
23.0
13.6
21.2
15.0
22.9
16.2
72.9
28.5
12.7
13.6
24.6
31.9
25.3
42.0
17.3
13.7
32.8
29.8
33.3
18.4
32.7
33.3
36.9
17.5
15.9
30.8
14.6
24.1
16.6
18.7
44.5
27.8
27.2
22.6
23.5
22.9
17.4
20.9
28.6
17.9
13.8
33.6
31.7
29.2
30.2
32.8
32.1
76.1
23.4
37.3
16.6
23.4
12.1
13.6
14.2
42.0
18.7
19.7
59.3
28.1
85.9
17.4
24.4
13.8
18.0
33.8
23.0
61.6
24.6
29.5
288.9
389.6
410.0
411.5
417.7
419.0
421.4
422.1
427.3
431.9
435.0
435.1
435.5
444.2
446.3
448.4
450.6
459.6
461.6
464.4
475.1
475.3
490.9
542.5
571.4
577.3
593.7
637.3
916.8
946.3
957.9
960.9
975.7
981.8
990.3
993.4
998.7
999.4
1004.7
1010.1
1018.0
1018.0
1019.4
1019.5
1021.8
1025.6
1025.7
1026.4
1027.2
1027.3
1027.8
1029.8
1029.9
1030.1
1030.7
1035.2
1035.8
1035.8
1038.7
1041.6
1045.4
1045.7
1046.0
1048.6
1050.4
1053.2
1053.3
1054.2
1054.8
1058.4
1060.5
1063.2
1063.3
1064.0
1064.1
1065.6
1065.7
1066.6
1069.2
1073.6
1073.7
1073.9
1074.1
1074.4
1075.1
1084.2
1084.7
1086.1
1086.1
1087.3
1090.3
1095.3
1099.2
1099.3
1099.6
1100.7
1102.8
1104.1
1110.6
1115.1
1117.2
1117.7
1118.5
1123.8
1124.0
1125.2
1125.9
1126.9
1127.1
1127.4
1127.9
1130.3
1133.4
1139.7
8.6
6.1
8.5
7.5
6.5
8.5
7.9
7.0
6.4
7.7
8.6
6.8
7.3
8.1
9.5
13.9
8.7
7.2
9.2
8.6
9.2
12.7
9.1
9.6
10.4
12.8
9.8
10.1
28.8
25.6
24.5
24.2
16.3
13.6
16.3
16.1
40.1
17.3
20.1
32.0
25.6
30.3
16.0
21.3
23.0
13.6
21.2
15.0
22.9
16.2
72.9
28.5
12.7
13.6
24.6
31.9
25.3
42.0
17.3
13.7
32.8
29.8
33.3
18.4
32.7
33.3
36.9
17.5
15.9
30.8
14.6
24.1
16.6
18.7
44.5
27.8
27.2
22.6
23.5
22.9
17.4
20.9
28.6
17.9
13.8
33.6
31.7
29.2
30.2
32.8
32.1
76.1
23.4
37.3
16.6
23.4
12.1
13.6
14.2
42.0
18.7
19.7
59.3
28.1
85.9
17.4
24.4
13.8
18.0
33.8
23.0
61.6
24.6
29.5
NA
NA
91.4
89.2
87.7
91.6
91.2
92.7
98.0
99.4
85.3
98.5
95.5
90.8
89.8
93.4
111.6
99.3
102.2
80.4
101.3
95.4
90.4
86.2
96.9
83.2
95.5
97.1
103.6
98.0
97.6
94.3
99.1
100.1
102.6
97.2
102.7
96.4
95.2
96.0
99.0
101.3
100.4
100.2
102.9
101.2
102.9
99.2
99.9
101.5
103.6
95.5
102.5
104.1
101.5
99.1
99.0
97.8
103.7
99.6
95.8
98.1
92.0
98.5
97.0
99.5
94.6
104.9
91.9
95.1
93.9
96.3
100.5
101.2
94.4
94.3
97.0
94.1
99.0
94.2
97.6
96.9
92.6
97.8
99.8
93.5
99.3
96.0
93.8
96.5
90.5
103.1
95.9
95.2
97.9
92.9
99.8
95.0
90.9
92.6
93.4
95.0
94.0
98.3
101.8
99.1
100.8
102.6
100.4
93.5
102.4
97.4
92.5
95.6
Supplemental Material. Table S1. Zircon U-Pb geochronologic analyses. Please visit https://doi​.org​
/10.1130​/GEOS​.S​.14049965 to access the supplemental material, and contact editing@geosociety.org with
any questions.
1
2
Supplemental Material. Table S2. Hf isotopic data.
Please visit https://doi.org/10.1130/GEOS.S.14049980
to access the supplemental material, and contact
editing@geosociety.org with any questions.
have detrital-zircon age distributions from recycling
of earlier Paleozoic sedimentary rocks and from
local primary sources in synrift Cambrian igneous
rocks, respectively (Thomas et al., 2016). The proximal Granite Wash is limited to a relatively narrow
area adjacent to the Arbuckle-Wichita uplifts and
interfingers basinward into a succession of shale,
sandstone, and limestone. A maximum thickness of
>4500 m of Atokan through Virgilian beds indicates
rapid subsidence of the Anadarko basin (Rascoe
and Johnson, in Johnson et al., 1988). Subsidence
slowed in Permian time, and the supply of coarse
clastic sediment from the uplifts diminished. By
Middle Permian time, the uplifts were largely covered by a more blanket-like succession of red beds
and evaporites (Johnson et al., 1988).
The Anadarko basin is highly asymmetric, with
a long, gentle, southwest-dipping limb that bends
abruptly upward along the Southern Oklahoma
fault system (Fig. 1B), defining a deep trough in
the southern part of the basin (Rascoe and Johnson, in Johnson et al., 1988). The basin is doubly
plunging, up plunge onto the Cimarron arch on
the northwest and up plunge onto the southern
part of the Nemaha uplift or Central Oklahoma
arch on the east (Fig. 1A). The Central Oklahoma
arch separates the eastern up-plunge end of the
Anadarko basin from the Arkoma foreland basin,
which had a distinctly different subsidence history,
both mechanically and temporally.
Late Paleozoic Ouachita Orogen
In the Ouachita Mountains, along the northern arm of the Ouachita thrust belt, facing the
Arkoma basin, the Cambrian–Mississippian offshelf passive-​margin succession and overlying
Mississippian–Pennsylvanian synorogenic muddy
to sandy turbidites are imbricated in the Ouachita
allochthon and thrust over the Cambrian–Mississippian passive-margin carbonate-shelf facies, which
remained in the Ouachita footwall (Fig. 3) (Arbenz,
1989a, 2008; Viele and Thomas, 1989). The Ouachita
detachment generally ramps upward toward the
foreland from the lower part of the Cambrian–Mississippian off-shelf passive-margin succession into
GEOSPHERE | Volume 17 | Number 4
Downloaded from http://pubs.geoscienceworld.org/gsa/geosphere/article-pdf/17/4/1214/5362417/1214.pdf
by Idaho State University user
the Middle Pennsylvanian (Atokan) part of the synorogenic succession (Fig. 3). The Maumelle chaotic
zone (along the trailing side of the frontal thrust belt;
Fig. 3) of broken Mississippian–Pennsylvanian sandstone and shale is interpreted to be the leading edge
of the accretionary prism, which was thrust onto the
pre-orogenic shelf (Viele and Thomas, 1989). During
the later stages of contraction, basement-rooted
thrusts inserted the deep subsurface Benton and
Broken Bow uplifts (Figs. 1B and 3) beneath the
sedimentary thrust belt, raising internally folded,
disharmonic thrust sheets of Cambrian–Mississippian off-shelf passive-margin mudstone, chert,
and sandstone in large-scale frontal thrust ramps
(Viele and Thomas, 1989; Arbenz, 2008). Rare tectonically bounded slivers of ultramafic rocks within the
deformed sedimentary rocks along the uplifts are
interpreted to be fragments of the oceanic crust of
the Laurentian plate that were picked off from the
down-going slab (Morris and Stone, 1986; Nielsen
et al., 1989; Viele and Thomas, 1989).
South of the intersection of the Ouachita thrust
front with basement structures along the Southern
Oklahoma fault system, postorogenic Mesozoic–​
Cenozoic strata of the Gulf Coastal Plain cover
the Ouachita orogen and the proximal part of the
Fort Worth basin in the footwall of the orogen
(Fig. 1) (e.g., Thomas et al., 1989). Drill data and
seismic-reflection profiles document Upper Cambrian to Lower Mississippian off-shelf deep-water
passive-margin strata and Middle Mississippian to
Middle Pennsylvanian synorogenic clastic facies in
the Ouachita frontal trust belt (Fig. 2), which have
been thrust over the shelf edge onto the Cambrian–
Mississippian passive-margin shelf-carbonate
facies (e.g., King, in Flawn et al., 1961; Nicholas
and Rozendal, 1975; Nicholas, in Arbenz et al., 1989;
Nicholas and Waddell, 1989; Thomas et al., 1989;
Viele and Thomas, 1989; Culotta et al., 1992). The
frontal structures and stratigraphy in the subsurface east of the Fort Worth basin (southern arm of
Ouachita salient; Fig. 1) are like those in the outcrops in the Ouachita Mountains in the eastern arm
of the Ouachita salient, except that the detachment
does not ramp up stratigraphically at the front of
the southern arm. The Waco and Luling basement
uplifts (Fig. 1B) underlie tectonically thickened and
deformed clastic rocks along the trailing edge of
the sedimentary thrust belt (Rozendal and Erskine,
1971; Nicholas and Waddell, 1989; Culotta et al.,
1992). Trailing the Ouachita sedimentary thrust belt,
the “Ouachita interior metamorphic belt” (Fig. 1)
consists of low-grade metasedimentary rocks and
extends along strike southwestward around the
Texas recess to the Marathon salient (Nicholas
and Waddell, 1989; Thomas et al., 1989; Viele and
Thomas, 1989). The trailing part of the Ouachita
orogen east of the Fort Worth basin may have been
impacted by accretion of the Sabine terrane (Fig. 1),
the extent of which is better defined south of the
Ouachita orogen in Arkansas and Louisiana (Keller
et al., 1989; Mickus and Keller, 1992); however, no
definitive data are available for the location of the
leading edge of the Sabine terrane in east Texas.
The timing of the end of Ouachita thrusting in
the accretionary prism is well constrained in the
subsurface, where an angular unconformity (Fig. 3)
separates deformed “Ouachita facies” rocks as
young as early Middle Pennsylvanian (Atokan)
from overlying undeformed shallow-marine successor-basin strata of late Middle Pennsylvanian
(Desmoinesian) to Middle Permian (Guadalupian)
age (Nicholas and Waddell, 1989). In the foreland,
rocks as young as middle Desmoinesian (Thurman
Sandstone; Fig. 2, box C1) are involved in folds that
parallel the Ouachita thrust front (Denison, 1989).
These observations bracket the age of the end of
Ouachita deformation at approximately 310 Ma
(end Atokan) in the interior structures and the end
of contraction on the frontal faults at approximately
309 Ma (middle Desmoinesian) in the foreland.
■■ SAMPLES FOR DETRITAL-ZIRCON
ANALYSES
Detrital zircons from three samples of Upper
Mississippian to Middle Pennsylvanian sandstones in the distal northern fringe of the Arkoma
basin were analyzed for U-Pb ages and Hf isotopic
compositions (Figs. 1B, 2, and 5; Tables 1 and 2;
Supplemental Tables S11 and S22). In addition, previously published U-Pb analyses from two other
sandstones in the distal Arkoma basin (Xie et al.,
Thomas et al. | Detrital zircons and late Paleozoic sediment dispersal, Ouachita orogen
1223
Research Paper
TABLE 1. LIST OF SAMPLES WITH DEPOSITIONAL AGES AND ENVIRONMENTS
Sample
Stratigraphic unit
Approximate
depositional age*
(Ma)
Depositional
environment
Reference
Savanna Sandstone
Cane Hill Member of the Hale Formation
Batesville Sandstone
309
322
330
Fluvial-deltaic
Fluvial-deltaic
Shallow marine
Sutherland (1988)
Sutherland (1988)
Sutherland (1988)
Cisco Group, above Camp Colorado
Limestone
Cisco Group, above Bunger Limestone
298
Fluvial-deltaic
Brown et al. (1973)
302
Fluvial-deltaic
Brown et al. (1973)
Rush Springs Sandstone
Garber Sandstone
Wellington Formation
265
281
282
Eolian
Fluvial-deltaic
Fluvial-deltaic
Johnson et al. (1988)
Johnson et al. (1988)
Johnson et al. (1988)
Arkoma basin
AR-6-SV
AR-2-HC
AR-1-BV
Fort Worth basin
TX-5-CC
TX-2-PB
Anadarko basin
OK-9-RS2
OK-5-GB2
OK-4-WL
*Approximate depositional age extrapolated by correlation to International Chronostratigraphic Chart, version 2019/05 (Cohen
et al., 2013, updated).
TABLE 2. SUMMARY OF εHft VALUES FOR U-Pb AGE GROUPS
Sampled unit
εHft values for U-Pb zircon age groups
Sample number
334–295 Ma
499–335 Ma
898–500 Ma
1276–914 Ma
AR-6-SV
AR-2-HC
AR-1-BV
+8.3 to –15.6
+10.0 to –19.2
+6.1 to –30.5
–2.0 to –10.0
+6.4 to –10.1
+7.2 to –4.4
+7.5 to –0.7
+11.2 to –5.0
+7.3 to –3.9
TX-5-CC
TX-2-PB
+4.2 to 0
–1.7 to –7.1
+9.0 to –14.4
+9.1 to –31.7
+13.7 to –15.3
+7.9 to –16.4
+1.1
+11.4 to –15.4
+6.2 to –8.1
+6.5 to –13.4
+6.6 to –10.8
+7.1 to –10.6
+6.3 to –5.7
+4.2 to –8.5
+6.5 to –32.2
+13.7 to –4.0
+7.3 to –2.0
+9.6 to –3.5
+12.1 to –6.4
+4.8 to –5.8
+11.4 to –30.5
+9.1 to –32.2
+13.7 to –16.4
Arkoma basin
Savanna
Cane Hill
Batesville
Fort Worth basin
Camp Colorado
Bunger
Anadarko basin
Whitehorse
Rush Springs
Garber
Wellington
Total range
GEOSPHERE | Volume 17 | Number 4
Downloaded from http://pubs.geoscienceworld.org/gsa/geosphere/article-pdf/17/4/1214/5362417/1214.pdf
by Idaho State University user
KS-5-WH
OK-9-RS2
OK-5-GB2
OK-4-WL
+4.8 to –5.8
+0.3 to –1.8
Thomas et al. | Detrital zircons and late Paleozoic sediment dispersal, Ouachita orogen
1224
Research Paper
2016, 2018a) are discussed here for comparison
with the new data (Figs. 1B, 2, and 5). Previously
published U-Pb analyses from the Mississippian–
Pennsylvanian Stanley Shale (McGuire, 2017)
and from the Pennsylvanian Atoka Formation
(Sharrah, 2006) characterize the more proximal
basin in the Ouachita thrust belt (Figs. 1B, 2, and
5). Published U-Pb data from tuff beds (Shaulis
et al., 2012) in the Mississippian–Pennsylvanian
Stanley Shale in the western part of the proximal
Arkoma basin in the Ouachita thrust belt provide
a summary of ages of volcanism, as well as possible xenocrysts from basement rocks of the Sabine
terrane (Figs. 5 and 6).
New analyses for U-Pb ages and Hf isotopic
compositions of two samples from Upper Pennsylvanian and Lower Permian sandstones characterize
the upper part of the late synorogenic to early
postorogenic fill of the Fort Worth basin (Figs. 1B,
2, and 5; Tables 1 and 2; Supplemental Tables S1
and S2). Previously published U-Pb data from Middle to Upper Pennsylvanian sandstones in the Fort
Worth basin (Alsalem et al., 2018) are discussed
here for comparison (Figs. 1B and 5). The early,
proximal deposits in the Fort Worth basin are now
deep in the subsurface beneath the thrust belt, and
the available samples are from the stratigraphically
higher, deltaic to shallow-marine deposits on the
distal shelf (Fig. 2C).
Analyses for U-Pb ages and Hf isotopic compositions of three samples (one previously published;
Thomas et al., 2016) from the southern part of the
intracratonic Anadarko basin (Figs. 1B and 5;
Tables 1 and 2; Supplemental Tables S1 and S2) represent the sedimentary cover over the Wichita and
Arbuckle uplifts, in contrast to the conglomerates
that rest unconformably on the Precambrian–Cambrian crystalline rocks in the uplifts (Thomas et al.,
2016). Published U-Pb and Hf data from a Permian sandstone (Thomas et al., 2020) on the divide
between the Anadarko basin and the intracratonic
Forest City basin on the north (Figs. 1 and 5B) provide a test of regional dispersal systems. Published
U-Pb data from the Permian basin (Soreghan and
Soreghan, 2013; Xie et al., 2018b; Liu and Stockli,
2019) provide comparison with a more distant intracratonic basin (Figs. 1A and 5B).
GEOSPHERE | Volume 17 | Number 4
Downloaded from http://pubs.geoscienceworld.org/gsa/geosphere/article-pdf/17/4/1214/5362417/1214.pdf
by Idaho State University user
■■ ANALYTICAL METHODS
Sample Collection and Processing
Each detrital-zircon sample, consisting of
approximately 12 kg of medium- to coarse-grained
sandstone, was collected from a restricted stratigraphic interval of a few successive beds and then
was processed utilizing methods outlined by Gehrels (2000), Gehrels et al. (2008), Gehrels and Pecha
(2014), and Thomas et al. (2017). Zircon grains were
extracted using traditional methods of jaw crushing and pulverizing, followed by density separation
using a Wilfley table. The resulting heavy-mineral fraction was further purified using a Frantz
LB-1 magnetic barrier separator and heavy liquids.
A representative split of the zircon yield was incorporated into a 2.54 mm epoxy mount along with
multiple fragments of the U-Pb primary standards
Sri Lanka SL-F, FC-1, and R33, and Hf standards R33,
Mud Tank, FC-1, Plesovice, Temora, and 91500. The
mounts were sanded down ~20 μm, polished to 1
μm, and imaged by backscattered electron (BSE)
and cathodoluminescence (CL) imaging techniques using a Hitachi 3400N scanning electron
microscope (SEM) and a Gatan Chroma CL2 detector system at the Arizona LaserChron SEM Facility
(www.laserchron.org). Prior to isotopic analysis,
mounts were cleaned in an ultrasonic bath of 1%
HNO3 and 1% HCl in order to remove contaminants
and surficial common Pb.
U-Pb Geochronologic Analysis
Uranium-lead (U-Pb) geochronologic analyses
of individual zircon crystals were conducted by
laser ablation–inductively coupled plasma–mass
spectrometry (LA-ICP-MS) at the Arizona LaserChron Center (www.laserchron.org). The isotopic
analyses involved ablation of zircon using a Photon Machines Analyte G2 excimer laser coupled to
either a Thermo Element2 single-collector ICP-MS
or a Nu Instruments multicollector ICP-MS. Ultrapure helium carried the ablated material from the
HelEx cell into the plasma source of each ICP-MS
instrument.
Analyses conducted with the Nu ICP-MS utilized Faraday collectors for measurement of 238U
and 232Th, either Faraday collectors or ion counters
for 208Pb, 207Pb, and 206Pb, and ion counters for 204(Pb,
Hg) and 202Hg (see Supplemental Table S1 for specific methods used for each sample), depending on
grain size. For larger grains, a 30-μm-diameter spot
was used, and masses 206, 207, 208, 232, and 238
were measured with Faraday detectors, whereas
the smaller 202 and 204 ion beams were measured
with ion counters. The acquisition routine included
a 15 s integration on peaks with the laser off (for
backgrounds), fifteen 1 s integrations with the laser
firing, and a 30 s delay to ensure that the previous
sample was completely purged from the system.
Smaller grains were analyzed with all Pb isotopes
in ion counters, using a 20 μm beam diameter, and
consisted of a 12 s integration on peaks with the
laser off (for backgrounds), twelve 1 s integrations
with the laser firing, and a 30 s delay to purge the
previous sample.
Analyses conducted with the Element2 ICP-MS
utilized a single scanning electron multiplier
that sequenced rapidly through U, Th, Pb, and
Hg isotopes. Ion intensities were measured in
pulse-counting mode for signals less than 50,000
cps, in both pulse-counting and analog mode for
signals between 50,000 and 5,000,000 cps, and in
analog mode above 5,000,000 cps. The calibration
between pulse-counting and analog signals was
determined line-by-line for signals between 50,000
and 5,000,000 cps and was applied to signals
>5,000,000 cps. Four intensities were determined
and averaged for each isotope, with dwell times of
0.0052 s for 202, 0.0075 s for 204, 0.0202 s for 206,
0.0284 s for 207, 0.0026 s for 208, 0.0026 s for 232,
and 0.0104 s for 238. With the laser set at an energy
density of ~5 J/cm2, a repetition rate of 8 Hz, and
an ablation time of 10 s, ablation pits were ~12 μm
in depth. Sensitivity with these settings was ~5000
cps/ppm. Each analysis consisted of 5 s on peaks
with the laser off (for backgrounds), 10 s with the
laser firing (for peak intensities), and a 20 s delay to
purge the previous sample and save files.
Analyses were conducted with one U-Th-Pb
measurement per grain (numbers of grains per
sample varied and are reported in Supplemental
Thomas et al. | Detrital zircons and late Paleozoic sediment dispersal, Ouachita orogen
1225
Research Paper
Table S1). Grains were selected in random fashion;
crystals were rejected only if they contained cracks
or inclusions or were too small to be analyzed. The
use of high-resolution BSE and CL images provided
assistance in grain selection and spot placement.
Data reduction was accomplished using
AgeCalc (a Microsoft Excel macro), which is the
standard Arizona LaserChron Center reduction
protocol (Gehrels et al., 2008; Gehrels and Pecha,
2014). Data were filtered for discordance, 206Pb/238U
precision, and 206Pb/207Pb precision as indicated in
the notes in Supplemental Table S1. Data are presented on normalized age-probability diagrams,
which sum all relevant analyses and uncertainties
and divide each curve by the number of analyses
such that all curves contain the same area. Age
groups are characterized by the ages of modes in
age probability and by the range of constituent
ages (Fig. 6; Table 3).
Age-probability diagrams were quantitatively
compared using multidimensional scaling (MDS;
Vermeesch, 2013). MDS facilitates the comparison
of detrital age distributions by converting sample
dissimilarity into a Cartesian plot where distance
is a proxy for dissimilarity (i.e., more dissimilar
age-probability diagrams plot farther apart, and
similar age-probability diagrams cluster together).
Mismatch (Amidon et al., 2005) was used to calculate dissimilarity, and the program DZmds
(Saylor et al., 2018) was used to produce MDS plots.
Transformation to Cartesian distance was calculated
using metric squared stress.
Hf Isotopic Analysis
Hafnium (Hf) isotopic analyses were conducted
utilizing the Nu multicollector LA-ICP-MS system
at the Arizona LaserChron Center following methods reported by Cecil et al. (2011) and Gehrels and
Pecha (2014). An average of 56 Hf analyses was
conducted for each sample. Grains were selected
to represent each of the main age groups and to
avoid crystals with discordant or imprecise ages. CL
images were utilized to ensure that all Hf analyses
TABLE 3. AGE GROUPS OF DETRITAL ZIRCONS FROM MISSISSIPPIAN–PERMIAN SANDSTONES IN THE ARKOMA, FORT WORTH, AND ANADARKO
BASINS; CHARACTERISTICS OF POSSIBLE APPALACHIAN (LAURENTIAN) OR OUACHITA-MARATHON (GONDWANAN) PROVENANCE*
Age group
(Ma)
Appalachian provenance
Ouachita-Marathon provenance
334–295
Primary: synorogenic Alleghanian plutons.
Dispersal: very rare (7 grains of 3564) in synorogenic Alleghanian clastic wedges.
Primary: synorogenic magmatic arc rocks in Coahuila and Sabine Gondwanan accreted
terranes.
Dispersal: in Marathon clastic wedge.
499–335
Primary: synorogenic Taconic and Acadian plutons.
Recycled: from Acadian synorogenic clastic wedge.
Dispersal: abundant in Alleghanian clastic wedges.
Primary and/or recycled: early and middle Paleozoic synorogenic plutons and detrital
zircons in Coahuila, Maya, and Sierra Madre accreted terranes.
Dispersal: abundant in Marathon clastic wedge.
901–500
Primary: Pan-African/Brasiliano basement rocks in Gondwanan accreted terranes.
Dispersal: rare in Alleghanian clastic wedges.
Primary: Pan-African/Brasiliano basement rocks in Gondwanan accreted terranes.
Dispersal: generally abundant in Marathon clastic wedge.
1276–914
Primary: Grenville rocks in external and internal Appalachian basement massifs.
Recycled: post-Grenville synrift, passive-margin, and synorogenic clastic rocks.
Dispersal: abundant in Alleghanian clastic wedges.
Primary: Sunsás basement rocks in Gondwanan accreted terranes.
Dispersal: abundant in Marathon clastic wedge.
1963–1202
Primary: scattered enclaves of older rocks within Grenville-age rocks in basement massifs.
Recycled: synrift and passive-margin deposits, originally supplied to the Laurentian rifted
margin and passive-margin shelf from the Penokean–Trans-Hudson, Central Plains–
Yavapai–Mazatzal, and Granite-Rhyolite provinces of the Laurentian craton.
Dispersal: common in Alleghanian clastic wedges.
Primary: Ventuari-Tapajós, Rio Negro–Juruena, and Rondonia–San Ignacio provinces of
Gondwanan Coahuila accreted terrane.
Recycled: passive-margin deposits along the Laurentian rifted margin, especially slope
deposits from Granite-Rhyolite rocks at the Alabama-Oklahoma transform margin.
Dispersal: common in Marathon clastic wedge.
2224–1945
Primary: Trans-Amazonian/Eburnian basement rocks in Gondwanan accreted terranes.
Dispersal: rare in Alleghanian clastic wedges.
Primary: Maroni-Itacaiúnas and Trans-Amazonian/Eburnian provinces of Gondwanan
Coahuila accreted terrane.
Dispersal: minor in Marathon clastic wedge.
3036–2387
Recycled: synrift and passive-margin detritus from the Superior province of the Canadian
Shield.
Dispersal: minor to rare in Alleghanian clastic wedges.
Primary and/or recycled: Central Amazonian province of Gondwanan accreted terranes.
Dispersal: minor in Marathon clastic wedge.
*Compiled from summary and references cited in Thomas et al. (2017, 2019).
GEOSPHERE | Volume 17 | Number 4
Downloaded from http://pubs.geoscienceworld.org/gsa/geosphere/article-pdf/17/4/1214/5362417/1214.pdf
by Idaho State University user
Thomas et al. | Detrital zircons and late Paleozoic sediment dispersal, Ouachita orogen
1226
Research Paper
were within the same growth domain as the U-Pb
pit, although in most analyses, Hf laser pits were
located directly on top of the U-Pb analysis pits.
Complete Hf isotopic data and Hf evolution plots
of individual samples are presented in Supplemental Table S2.
Hafnium data are presented using Hf evolution diagrams, where initial 176Hf/177Hf ratios are
expressed in εHft notation, which represents the
Hf isotopic composition at the time of zircon crystallization relative to the chondritic uniform reservoir
(CHUR) (Bouvier et al., 2008). Internal precision for
176
Hf/177Hf and εHft is reported for each analysis on
Hf evolution plots in Supplemental Table S2 and
as the average for all analyses (2.2 epsilon units
at 2σ) on εHft evolution diagrams. On the basis of
the in-run analysis of zircon standards, the external
precision was 2–2.5 epsilon units (2σ). Hf isotopic
evolution of typical continental crust is shown with
arrows on εHft evolution diagrams, which are based
on a 176Lu/177Hf ratio of 0.0115 (Vervoort and Patchett,
1996; Vervoort et al., 1999).
Our Hf isotope data were interpreted within the
standard framework of juvenile (positive) values
indicating magma consisting mainly of material
extracted from the mantle during or immediately
prior to magmatism versus more evolved (negative) values recording incorporation of significantly
older crust. Vertical arrays on εHft diagrams are
interpreted to represent magmas that contain both
materials derived from the mantle during (or immediately prior to) magmatism and significantly older
crustal materials.
Hf isotope data are plotted as bivariate kernel
density estimates (KDEs) using HafniumPlotter
(Sundell et al., 2019; github.com/kurtsundell/HafniumPlotter). Bivariate KDEs facilitate comparison
of two-dimensional (2-D, U-Pb–Hf) data by converting sample density into a color-coded intensity plot,
where higher intensity corresponds to higher data
density. Density estimation uses a diagonal kernel
bandwidth matrix and generates bivariate KDEs
by (1) applying a discrete cosine transform to the
2-D data, (2) multiplying the resulting 2-D matrix
of points by a Gaussian function generated on the
basis of set kernel bandwidths (25 m.y. for U-Pb
age and 2 epsilon units for εHft), (3) performing
GEOSPHERE | Volume 17 | Number 4
Downloaded from http://pubs.geoscienceworld.org/gsa/geosphere/article-pdf/17/4/1214/5362417/1214.pdf
by Idaho State University user
an inverse discrete cosine transform of the
Gaussian-scaled matrix, and (4) normalizing the
three-dimensional (3-D) bivariate KDE volume sum
to 1. The bivariate KDE 3-D surface is color-coded to
intensity based on z-axis height and viewed parallel
to the z axis to visualize density in 2-D. Results are
clipped at the 95% confidence level from relative
peak data density.
ages of 487–380 Ma and subequal modes of 458
and 433 Ma. A group with ages between 1935 and
1295 Ma has a strong mode of 1618 Ma and secondary modes of 1821, 1505, and 1339 Ma (Fig. 7).
Five grains have ages between 682 and 532 Ma,
and one grain has an age of 2041 Ma. A secondary
age group between 2957 and 2550 Ma has a lesser
mode of 2820 Ma (Fig. 7).
■■ DETRTIAL-ZIRCON DATA FROM
ARKOMA BASIN
Savanna Sandstone (Sample AR-6-SV)
Results of U-Pb Analyses
Batesville Sandstone (Sample AR-1-BV)
The stratigraphically lowest sample was collected from the Batesville Sandstone, which is
one of two quartzose sandstones within the lower
Upper Mississippian succession of shallow-marine
sandstones, limestones, and shales in the northern
fringe of the Arkoma basin (Figs. 1B, 2, and 5A).
The most dominant group of detrital-zircon ages
is between 488 and 339 Ma with a dominant mode
of 434 Ma (Fig. 7). Another prominent group with
ages of 1210–942 Ma has subequal modes of 1161
and 1042 Ma. Secondary modes of 1898, 1788, 1650,
1479, and 1397 Ma are broadly spaced through a
distribution from 1907 to 1210 Ma (Fig. 7). A few
grains have ages of 2176–1996 and of 762–502 Ma.
A subordinate age group of 2909–2634 Ma has a
mode of 2723 Ma (Fig. 7).
Cane Hill Member of the Hale Formation
(Sample AR-2-HC)
A deltaic to shallow-marine sandstone at the
base of the Cane Hill Member of the Lower Pennsylvanian Hale Formation (Sutherland, 1988) overlies
truncated Mississippian strata at a regional unconformity in the distal northern part of the Arkoma
basin (Figs. 1B, 2, and 5A). The dominant age group
of detrital zircons between 1265 and 935 Ma has
a mode of 1045 Ma with a secondary shoulder of
1131 Ma (Fig. 7). Another prominent group has
The upper Middle Pennsylvanian (lower
Desmoinesian) Savanna Sandstone is within a
succession of southward prograding fluvial-deltaic
systems (Sutherland, 1988) in the distal northern
part of the Arkoma basin (Figs. 1B, 2, and 5A); in
contrast, the stratigraphically higher middle Desmoinesian Thurman Sandstone (Fig. 2, box C1)
contains chert clasts transported northwestward
from the Ouachita thrust belt (Sutherland, 1988).
A sample from the stratigraphically highest sandstone preserved in the Savanna Sandstone has a
dominant detrital-zircon age group between 1208
and 946 Ma with a primary mode of 1055 Ma and
a secondary mode of 1183 Ma (Fig. 7). Another
important group with ages of 491–390 Ma has a
mode of 420 Ma. The youngest grain in the sample
is 390 Ma; however, one grain (289 Ma) was omitted from the plot because it is younger than the
depositional age. A secondary age group between
1847 and 1606 Ma has modes of 1798, 1742, and
1629 Ma (Fig. 7). A lesser group of ages between
1549 and 1224 Ma has minor modes of 1455, 1334,
and 1229 Ma (Fig. 7). A minor age group between
637 and 543 Ma has a mode of 589 Ma, and three
grains are scattered between 2098 and 1950 Ma.
Ages scattered between 2916 and 2503 Ma have
minor modes of 2723 and 2569 Ma (Fig. 7).
Results of Hf Isotopic Analyses
Thomas et al. | Detrital zircons and late Paleozoic sediment dispersal, Ouachita orogen
Hf isotopic data are similar through the set of
three samples (AR-1-BV, AR-2-HC, and AR-6-SV)
from the Arkoma basin (Fig. 7; Table 2). Zircon
grains in the age range 1265–935 Ma have εHft
1227
Research Paper
Summary of Previously Published U-Pb
Age Data
Hot Springs Sandstone Member of the Stanley
Shale (McGuire, 2017)
20
15
B
2σ uncertainty
10
Epsilon Hf
values ranging from +11.2 to −5.0, but most values
are concentrated between +6.2 and −1.0 (Fig. 7).
Zircon grains in the age range of 762–502 Ma have
εHft values between +7.2 and −10.1. Although zircon
grains with ages of 488–390 Ma have εHft values
spread between +10.0 and −30.5, most are tightly
clustered between +6.0 and −11.2 (Fig. 7). The range
of εHft values indicates mixing of magma sources,
including evolved crust and juvenile magmas.
DM
5
0
CHUR
-5
-10
-15
-20
A
Savanna AR-6-SV (n=261)
Atoka NE (n=586)
The Mississippian Hot Springs Sandstone Member in the lowermost part of the Stanley Shale in
the more proximal part of the Arkoma basin and
now in the Ouachita thrust belt (Figs. 1B, 2, and 5A)
has been interpreted to be a marginal turbidite fan
from the northern (cratonic) margin of the Ouachita
foredeep (Morris, 1989). Previously reported results
of U-Pb analyses (McGuire, 2017) of a sample from
the Hot Springs Sandstone Member include a dominant age group between 1195 and 994 Ma with a
dominant mode of 1050 Ma (Fig. 7). A prominent
secondary group with ages of 484–373 Ma has a
mode of 420 Ma. A lesser group with ages between
1809 and 1231 Ma has a minor mode of 1457 Ma
(Fig. 7). Two grains have ages of 767 and 551 Ma,
and two grains have ages of 2287 and 2263 Ma.
Three older grains have ages of 2739–2704 Ma.
Atoka WS (n=583)
Spiro (n=299)
Bloyd (n=850)
Cane Hill AR-2-HC (n=306)
Stanley tuffs >340 Ma (n=232)
Stanley (n=92)
Wedington (n=559)
Batesville AR-1-BV (n=308)
Hot Springs (n=99)
200 400 600 800 1000 1200 1400 1600 1800 2000 2200 2400 2600 2800 3000 3200 3400
Wedington Sandstone Member of the
Fayetteville Shale (Xie et al., 2016)
Previously published results (Xie et al., 2016)
of U-Pb analyses of six samples from southward
prograding deltaic deposits of the Upper Mississippian Wedington Sandstone in the distal northern
part of the Arkoma basin (Figs. 1B, 2, and 5A) are
similar through the sample set and are summarized here in a composite plot (Fig. 7). A strongly
dominant group of ages between 1209 and 937 Ma
GEOSPHERE | Volume 17 | Number 4
Downloaded from http://pubs.geoscienceworld.org/gsa/geosphere/article-pdf/17/4/1214/5362417/1214.pdf
by Idaho State University user
Detrital Zircon Age (Ma)
Figure 7. U-Pb age-probability density plots (A) and Hf-evolution diagram (B) showing results from analyses of Mississippian–Pennsylvanian sandstones in the Arkoma basin. (A) U-Pb age-probability density
plots for three analyzed samples (data in Supplemental Table S1 [see text footnote 1]) and from published data as follows: Wedington (Xie et al., 2016); Hot Springs and Stanley (McGuire, 2017); Stanley
tuffs (Shaulis et al., 2012); Bloyd (Xie et al., 2018a); Spiro, Atoka WS—westward and southward dispersal,
and Atoka NE—northeastward dispersal (Sharrah, 2006). Vertical, dashed, black lines mark boundaries of
detrital-zircon age groups (Fig. 6; Table 3). (B) εHft data for three samples (data in Supplemental Table S2
[see text footnote 2]). Data points are color coded as in panel A. The average uncertainty of Hf isotopic
analyses is 2.6 epsilon units at 2σ. The Hf-evolution diagram shows the Hf isotopic composition at the
time of zircon crystallization, in epsilon units, relative to the chondritic uniform reservoir (CHUR) (Bouvier
et al., 2008) and to the depleted mantle (DM) (Vervoort et al., 1999).
Thomas et al. | Detrital zircons and late Paleozoic sediment dispersal, Ouachita orogen
1228
Research Paper
has a primary mode 1051 Ma with a secondary
shoulder of 1120 Ma (Fig. 7). The most prominent
secondary group with ages of 496–388 Ma has a
mode of 471 Ma. A spread of relatively abundant
grains between 1883 and 1215 Ma has secondary
modes of 1873, 1802, 1637, and 1480 Ma (Fig. 7).
Minor numbers of grains are scattered between
2186 and 1945 Ma and between 687 and 501 Ma.
Another minor group of ages extends from 3036
to 2466 Ma.
Sandstone in the Stanley Shale (McGuire, 2017)
A sample of a sandstone in the upper part of
the succession of mud-dominated turbidites in the
Mississippian–Pennsylvanian Stanley Shale in the
more proximal part of the Arkoma basin, which is
now in the Ouachita southern thrust belt (Figs. 1B,
2, and 5A), was analyzed for U-Pb ages of detrital
zircons (McGuire, 2017), and the results are shown
in Figure 7. The most dominant group with ages of
874–501 Ma has a dominant mode of 547 Ma and
secondary modes of 856 and 622 Ma; this dominant
age group contrasts notably with other samples
from the Arkoma basin (Fig. 7). Another important group with ages of 1197–1018 Ma has a mode
of 1054 Ma. Twelve grains are scattered between
2185 and 1953 Ma and define minor modes of 2053
and 1969 Ma. Another secondary group has ages
of 1885–1637 Ma (Fig. 7). The sample has only one
grain (1322 Ma) with an age between 1637 and
1197 Ma, an age interval that is well represented in
the other samples from the Arkoma basin. A secondary age group between 486 and 336 Ma has
a mode of 482 Ma. A few older grains have ages
scattered between 3228 and 2358 Ma.
Tuff Beds in the Stanley Shale (Shaulis et
al., 2012)
Published U-Pb data from five tuff beds (Shaulis
et al., 2012) in the Mississippian–Pennsylvanian
Stanley Shale in the western part of the proximal
Arkoma basin in the Ouachita thrust belt (Fig. 6) are
dominated by grains with ages of the volcanism
GEOSPHERE | Volume 17 | Number 4
Downloaded from http://pubs.geoscienceworld.org/gsa/geosphere/article-pdf/17/4/1214/5362417/1214.pdf
by Idaho State University user
(328.5 ± 2.7 to 320.7 ± 2.5 Ma). In addition, the samples include a range of older grains (Figs. 6 and 7)
that have been interpreted as detrital (Shaulis et al.,
2012), but the depositional setting of the ash-flow
or ash-fall tuffs (Niem, 1977) within a succession
of deep-water muddy turbidites suggests that
the zircons may be xenocrysts from the magma
source in the Sabine terrane. A plot of the ages
of prevolcanic zircons (Figs. 6 and 7) is included
here to evaluate the possible xenocrystic grains by
comparison with detrital-​zircon age distributions
and to illustrate the composition of the possible
Sabine magma chamber as a component of the
provenance for synorogenic clastic sediment. The
dominant zircon age group of 1276–928 Ma has
modes of 1177 and 1068 Ma, and a strong secondary age group extends between 1577 and 1283 Ma
with modes of 1440 and 1353 Ma (Figs. 6 and 7). A
secondary age group of 787–506 Ma has a mode
of 613 Ma, and three grains are scattered between
2129 and 1997 Ma. Another secondary group with
ages of 483–409 Ma has a mode of 451 Ma, but
the age-probability plot of the young boundary of
that distribution is obscured by the strongly dominant mode of the volcanic crystallization ages
(Fig. 6). A few older grains have ages between
2889 and 2503 Ma.
Middle Sandstone Unit in the Bloyd Formation
(Xie et al., 2018a)
Previously published results (Xie et al., 2018a)
of U-Pb analyses of eight samples from the middle
sandstone unit in the Lower Pennsylvanian (upper
Morrowan) Bloyd Formation in the distal northern
part of the Arkoma basin (Figs. 1B, 2, and 5A) are
summarized here in a composite plot (Fig. 7). The
dominant detrital-zircon age group of 1200–972 Ma
is bimodal with modes of 1165 and 1052 Ma (Fig. 7).
A group of ages spread between 1923 and 1202 Ma
has a secondary mode of 1635 Ma (Fig. 7). Another
secondary age group of 495–374 Ma has a mode of
434 Ma. Seven grains have ages between 618 and
528 Ma, and six grains have ages between 2190 and
1979 Ma. A secondary age group between 3146 and
2636 Ma has a mode of 2713 Ma (Fig. 7).
Spiro Sandstone in Lower Atoka Formation
(Sharrah, 2006)
The Spiro and related sandstones in the lower
Atoka Formation (Figs. 1B, 2, and 5A) include southward trending channel sands that grade southward
into deltaic facies and marine-reworked facies
along the northern (cratonic) side of the Ouachita
foredeep (Sutherland, in Johnson et al., 1988). A
composite plot (Fig. 7) of two samples (N = 2, n =
299) of Spiro sandstone (Sharrah, 2006) shows a
dominant age group between 1203 and 933 Ma with
a prominent mode of 1026 Ma and a minor shoulder
mode of 1082 Ma (Fig. 7). A secondary group with
ages of 492–353 Ma has a mode of 418 Ma. The
youngest grain in the two samples is 353 Ma; however, one grain with an age of 295 Ma was omitted
from the age-probability plot because it is younger
than the depositional age of the sandstone. A scatter of ages between 1894 and 1221 Ma has a strong
secondary mode of 1609 Ma and a minor mode of
1308 Ma. Three grains have ages of 656–541 Ma,
and three ages are scattered between 2060 and
1956 Ma (Fig. 7). Older grains between 2978 and
2342 Ma have a minor mode of 2699 Ma.
Sandstones with Westward and Southward
Dispersal in the Atoka Formation (Sharrah, 2006)
Sandstones in a thick succession of turbidites
in the Atoka Formation (Figs. 1B, 2, and 5A) with
paleocurrent indicators of westward and southward
sediment dispersal are now mostly in the more
easterly part of the frontal Ouachita thrust belt in
Arkansas but extend farther west (Sharrah, 2006).
A composite plot (Fig. 7) of data from four samples
(N = 4, n = 583) has a dominant group of detrital-zircon ages of 1205–934 Ma and two prominent
modes of 1062 and 1031 Ma (Fig. 7). A secondary
age group of 481–336 Ma has modes of 441 and
412 Ma. The single youngest grain in the four samples is 313 Ma; however, three grains with ages
(309–248 Ma) that are younger than the depositional age of the sandstone were omitted from the
age-probability plot. A scatter of ages between 1935
and 1214 Ma has a secondary mode of 1607 Ma.
Thomas et al. | Detrital zircons and late Paleozoic sediment dispersal, Ouachita orogen
1229
Research Paper
Nine grains have ages scattered between 799 and
534 Ma, and 10 ages are scattered between 2197
and 1968 Ma (Fig. 7). Older grains between 2999
and 2546 Ma have a minor mode of 2711 Ma.
Sandstones with Northeastward Dispersal in
the Atoka Formation (Sharrah, 2006)
Sandstones in a thick succession of turbidites
in the Atoka Formation (Figs. 1B, 2, and 5A) with
paleocurrent indicators of northeastward sediment
dispersal are now in the more westerly part of the
frontal Ouachita thrust belt in Oklahoma (Sharrah,
2006). A composite plot of data from four samples (N = 4, n = 586) has a dominant age group of
1195–937 Ma with a prominent mode of 1040 Ma
(Fig. 7). A secondary group with ages of 492–354 Ma
has modes of 408 and 384 Ma. A minor but distinct
younger group has two grains with ages of 331–
327 Ma, but two other grains with ages of 295 and
293 Ma were omitted from the age-probability plot
because they are younger than the depositional
age. A scatter of ages between 1947 and 1210 Ma
has secondary modes of 1613 and 1455 Ma. An
important group of 34 grains has ages between
799 and 505 Ma with a minor mode of 632 Ma, and
15 ages are scattered between 2218 and 1962 Ma
(Fig. 7). Older grains between 2932 and 2523 Ma
have a minor mode of 2735 Ma.
■■ DETRITAL-ZIRCON DATA FROM FORT
WORTH BASIN
Results of U-Pb Analyses
Sandstone above the Bunger Limestone, Cisco
Group (Sample TX-2-PB)
The Pennsylvanian–Permian Cisco Group
includes numerous cyclic westward prograding
fluvial-deltaic systems and transgressive shallow-marine limestones in the distal shelf of the
Fort Worth basin (Brown et al., 1973). A sample
was collected from a road-cut exposure of the
sandstone above the Bunger Limestone in the
GEOSPHERE | Volume 17 | Number 4
Downloaded from http://pubs.geoscienceworld.org/gsa/geosphere/article-pdf/17/4/1214/5362417/1214.pdf
by Idaho State University user
Upper Pennsylvanian (Virgilian) Cisco Group
(Figs. 1B, 2, and 5A). The sandstone includes beds
of conglomerate with subrounded clasts of white
to light-gray chert as much as 4 cm in diameter
in channel-fill deposits (Brown et al., 1973). The
sample has two dominant age groups of detrital
zircons: one between 898 and 501 Ma with modes
of 644 and 532 Ma, and the other between 1255
and 919 Ma with two modes of 1051 and 1009 Ma
and a secondary shoulder mode of 1161 Ma (Fig. 8).
Another prominent group has ages of 491–339 Ma
and modes of 421 and 344 Ma (Fig. 8); three grains
with ages of 334–325 Ma document a distinctively
young age range. A group of detrital-zircon ages
extends from 1963 to 1280 Ma and has a prominent
mode of 1466 Ma along with several minor modes
(Fig. 8). A group of ages between 2121 and 1981 Ma
has modes of 2073 and 1984 Ma. An older minor
group with ages of 2856–2519 Ma has a minor
mode of 2712 Ma.
Results of Hf Isotopic Analyses
Hf isotopic data from the Fort Worth basin are
generally similar in the two samples TX-5-CC and
TX-2-PB (Fig. 8; Table 2). The older Precambrian
grains generally conform to crustal evolution trends
with some exceptions toward more positive εHft values (Fig. 8). Zircon grains with ages of 1252–969 Ma
have εHft values between +13.7 and −16.4 (Fig. 8).
Zircon grains in the age range of 871–524 Ma have
εHft values between +9.1 and −31.7, although grains
with ages older than 600 Ma are mostly between
+9.0 and −4.9, whereas those with ages younger
than 600 Ma are mostly between +4.9 and −5.6, indicating heterogeneous sources. Zircon grains in the
age range of 474–348 Ma have εHft values tightly
clustered between +4.2 and −7.1 (Fig. 8).
Summary of Previously Published U-Pb
Age Data
Sandstone above Camp Colorado Limestone,
Cisco Group (Sample TX-5-CC)
Sample from Middle Pennsylvanian Big Saline
Formation (Alsalem et al., 2018)
A sample of the sandstone stratigraphically
above the Camp Colorado Limestone (Figs. 1B, 2,
and 5A) is from part of the cyclic succession of
westward prograding fluvial-deltaic sandstones
and transgressive limestones in the Lower Permian part of the Cisco Group (Brown et al., 1973).
The most prominent group of detrital-zircon ages is
between 861 and 511 Ma and has modes of 731, 657,
588, and 531 Ma (Fig. 8). The second most prominent group has ages of 1256–935 Ma and modes
of 1213 and 1034 Ma. Notably, the mode of 588 Ma
exceeds the Mesoproterozoic-age modes in this
sample (Fig. 8). Another prominent group has ages
of 498–362 Ma and a mode of 419 Ma. The youngest
grain in the sample has an age of 327 Ma; however,
one grain with an age of 274 Ma was omitted from
the age-probability plot because it is younger than
the depositional age of the sandstone. A group with
detrital-zircon ages of 1959–1279 Ma has modes of
1668 and 1503 Ma. A secondary group with ages
of 2224–1997 Ma has a mode of 2092 Ma. A minor
age group of 2859–2395 Ma has a mode of 2750 Ma.
A sample of a sandstone in the Big Saline Forma­
tion (Figs. 1B, 2, and 5A) of the Bend Group is from
the lower part of the Middle Pennsylvanian (Atokan) succession in the southwestern part of the
Fort Worth basin (Alsalem et al., 2018). A group of
ages between 1214 and 914 Ma has a dominant
mode of 1029 Ma (Fig. 8). The sample includes two
prominent secondary age groups: 795–514 Ma with
modes of 616 and 538 Ma, and 498–354 Ma with a
mode of 421 Ma (Fig. 8). Ages are scattered between
1885 and 1228 Ma with modes of 1728 and 1668 Ma.
A minor age group of 2188–1952 Ma has a mode of
1976 Ma. A group of older grains ranges from 2801
to 2440 Ma and has a mode of 2687 Ma (Fig. 8).
Samples from Upper Middle to Lower
Upper Pennsylvanian Formations
(Alsalem et al., 2018)
Thomas et al. | Detrital zircons and late Paleozoic sediment dispersal, Ouachita orogen
Analytical data from five samples from multiple
formations in the upper Middle (Desmoinesian) to
1230
Research Paper
20
15
B
2σ uncertainty
10
Epsilon Hf
lower Upper Pennsylvanian (Missourian) succession are from closely spaced sites in a relatively
small area in the northwestern part of the Fort
Worth basin (Figs. 1B, 2, and 5A) and have very
similar age distributions (Alsalem et al., 2018); the
data from all five samples are shown in a composite plot (Fig. 8). The most prominent age group of
1237–920 Ma has a dominant mode of 1047 Ma
(Fig. 8). Another age group, nearly as strong, of
828–500 Ma has modes of 616 and 548 Ma; a
minor group with ages of 2188–1981 Ma has a
mode of 2066 Ma. A secondary group has ages of
499–335 Ma; a younger group of three grains has
ages of 329–318 Ma. Ages are scattered between
1913 and 1258 Ma with modes of 1745, 1637, 1509,
and 1369 Ma. An older group between 2893 and
2460 Ma has a mode of 2706 Ma (Fig. 8).
DM
5
0
CHUR
-5
-10
-15
-20
-25
-30
■■ DETRITAL-ZIRCON DATA FROM
ANADARKO BASIN
-35
A
Results of U-Pb Analyses
Camp Colorado TX-5-CC (n=228)
Wellington Formation (Sample OK-4-WL)
The Lower Permian Wellington Formation
(Figs. 1B and 5A) is the lowermost part of a
widespread succession of alluvial to deltaic red
sandstones and mudstones covering the proximal
conglomerates and angular unconformity over the
Arbuckle and Wichita uplifts at the southern margin of the Anadarko basin (Johnson et al., 1988).
Detrital-zircon grains of the Wellington Formation
(previously reported; Thomas et al., 2016) have
three dominant age groups: 1210–924 Ma with
modes of 1145 and 1032 Ma, 857–504 Ma with a
mode of 558 Ma, and 499–339 Ma with a mode
of 435 Ma (Fig. 9). Two grains have distinctively
young ages of 334–329 Ma. A secondary group with
ages of 2156–1960 Ma has a secondary mode of
1976 Ma (Fig. 9). Age groups of 1922–1602 Ma and
1532–1223 Ma have multiple secondary modes of
1852, 1744, and 1648 Ma and of 1453, 1326, and
1253 Ma, respectively (Fig. 9). An older secondary
group has ages of 2905–2305 Ma and a secondary
mode of 2721 Ma.
GEOSPHERE | Volume 17 | Number 4
Downloaded from http://pubs.geoscienceworld.org/gsa/geosphere/article-pdf/17/4/1214/5362417/1214.pdf
by Idaho State University user
Bunger TX-2-PB (n=318)
Strawn & Canyon S4–S8 (n=510)
Big Saline S3 (n=283)
200 400 600 800 1000 1200 1400 1600 1800 2000 2200 2400 2600 2800 3000 3200 3400
Detrital Zircon Age (Ma)
Figure 8. U-Pb age-probability density plots (A) and Hf-evolution diagram (B) showing results from analyses of Pennsylvanian–Permian sandstones in the Fort Worth basin. (A) U-Pb age-probability density
plots for two analyzed samples (data in Supplemental Table S1 [see text footnote 1]) and from published
data for Big Saline and composite Strawn–Canyon (Alsalem et al., 2018). Vertical, dashed, black lines
mark boundaries of detrital-zircon age groups (Fig. 6; Table 3). (B) εHft data for two samples (data in
Supplemental Table S2 [see text footnote 2]). Data points are color coded as in panel A. Specifications
for the Hf-evolution diagram are described in the caption for Figure 7. CHUR—chondritic uniform reservoir; DM—depleted mantle.
Thomas et al. | Detrital zircons and late Paleozoic sediment dispersal, Ouachita orogen
1231
Research Paper
Garber Sandstone (Sample OK-5-GB2)
20
15
B
2σ uncertainty
10
Epsilon Hf
A sample from a sandstone in the Lower Permian Garber Sandstone (Figs. 1B and 5A) is part of
a succession of fluvial to deltaic red beds (Johnson et al., 1988). The sample has a dominant group
with ages of 1228–932 Ma and a prominent mode of
1035 Ma and a lesser mode of 1141 Ma (Fig. 9). The
sample includes three strong secondary age groups:
1794–1621 Ma with modes of 1745 and 1645 Ma;
1520–1240 Ma with modes of 1490, 1384, and
1347 Ma; and 499–354 Ma with modes of 454 and
415 Ma (Fig. 9). A lesser group has ages between
820 and 510 Ma and multiple minor modes; four
grains are scattered between 2137 and 1976 Ma.
A minor age group of 2992–2401 Ma has a minor
mode of 2715 Ma (Fig. 9).
DM
5
0
CHUR
-5
-10
-15
-20
-25
-30
Rush Springs Sandstone (Sample OK-9-RS2)
A sample of the Middle Permian Rush Springs
Sandstone (Figs. 1B and 5A) was collected from
a southwest-directed eolian sandstone within an
erg system (Poland and Simms, 2012). Two age
groups of detrital zircons dominate the sample age
distribution: ages of 1200–940 Ma with modes of
1138, 1070, and 1023 Ma; and ages of 496–337 Ma
with modes of 420 and 343 Ma (Fig. 9). A distinctive
younger group has ages of 327–295 Ma. The next
most prominent group with ages of 897–502 Ma
(mostly between 649 and 502 Ma) has modes of
630 and 590 Ma, and a secondary age group of
2182–1981 Ma has a minor mode of 2079 Ma (Fig. 9).
Other strong secondary groups have ages of 1913–
1626 Ma with modes of 1737 and 1637 Ma, and ages
of 1524–1221 Ma with modes of 1490, 1377, and
1339 Ma. Older zircons with ages between 2925 and
2438 Ma have a minor mode of 2645 Ma (Fig. 9).
Results of Hf Isotopic Analyses
Hf isotopic data for older Precambrian grains in
samples OK-4-WL, OK-5-GB2, and OK-9-RS2 generally conform to crustal evolution trends (Fig. 9).
Zircon grains with ages of 1162–932 Ma have εHft
GEOSPHERE | Volume 17 | Number 4
Downloaded from http://pubs.geoscienceworld.org/gsa/geosphere/article-pdf/17/4/1214/5362417/1214.pdf
by Idaho State University user
-35
A
Rush Springs OK-9-RS2 (n=327)
Garber OK-5-GB2 (n=315)
Wellington OK-4-WL (n=329)
200 400 600 800 1000 1200 1400 1600 1800 2000 2200 2400 2600 2800 3000 3200 3400
Detrital Zircon Age (Ma)
Figure 9. U-Pb age-probability density plots (A) and Hf-evolution diagram (B) showing results from
analyses of Permian sandstones in the Anadarko basin. (A) U-Pb age-probability density plots for two
analyzed samples (data in Supplemental Table S1 [see text footnote 1]) and from published data for
Wellington (Thomas et al., 2016). Vertical, dashed, black lines mark boundaries of detrital-zircon age
groups (Fig. 6; Table 3). (B) εHft data for two samples (data in Supplemental Table S2 [see text footnote 2])
and for published data for Wellington (Thomas et al., 2016). Data points are color coded as in panel A.
Specifications for the Hf-evolution diagram are described in the caption for Figure 7. CHUR—chondritic
uniform reservoir; DM—depleted mantle.
Thomas et al. | Detrital zircons and late Paleozoic sediment dispersal, Ouachita orogen
1232
Research Paper
values between +12.1 and −6.4 (Fig. 9; Table 2). For
zircon grains in the age range 857–504 Ma, εHft
values range from +6.5 to −32.2 and cluster around
+3.2 to −0.9, consistent with involvement of older
crust in the source of the magmas. Zircon grains
in the age range of 464–337 Ma have εHft values
between +6.6 and −13.4 (Fig. 9), indicating variable
sources. The few young grains in the age range of
327–308 Ma have εHft values of +1.1 to −1.8.
A - Appalachian basin (n=3564)
B - Illinois basin south (n=2159)
C - Permian basin (n=3357)
D - Coahuila terrane (n=1265)
E - Marathon foreland (n=545)
■■ DISCUSSION
F - passive-margin off-shelf (n=760)
Age Groups of Detrital Zircons and
Implications for Sediment Provenance
The U-Pb age distributions of detrital zircons
from sandstones in the Arkoma, Fort Worth, and
Anadarko basins define several general groups
(Figs. 6 and 10; Table 3). The distributions within
each age group (Figs. 6 and 10; Table 3), along with
corresponding ranges of εHft values (Fig. 11; Table 2),
can be used to characterize the provenance(s). The
primary alternatives for provenance are either
the Appalachian orogen, generally with dispersal
through the southern Illinois basin, or the Ouachita
orogen, including accreted Gondwanan terranes
(Table 3). Lesser alternatives for provenance include
synrift igneous rocks in the Arbuckle-Wichita uplifts,
other intracratonic uplifts, or the Canadian Shield.
Variability between samples suggests variations in
provenance.
334–295 Ma
This age group (Table 3) is sparsely represented
or lacking in most of the samples but is a distinctive minor component in some (e.g., samples
OK-9-RS2 and OK-4-WL; Fig. 9). The age range
corresponds to the age of the Alleghanian orogeny, including the ages of plutons (330–270 Ma)
in the Appalachians and tonsteins in coal beds
(316–311 Ma) in the Appalachian basin (summary
in Thomas et al., 2017, and references therein).
Grains of this age range, however, are nearly
lacking among detrital zircons (7 grains of 3564)
GEOSPHERE | Volume 17 | Number 4
Downloaded from http://pubs.geoscienceworld.org/gsa/geosphere/article-pdf/17/4/1214/5362417/1214.pdf
by Idaho State University user
G - Whitehorse KS-5-WH (n=224)
H - Anadarko basin (n=1065)
I - Fort Worth basin (n=1339)
J - Arkoma basin (n=4170)
200 400 600 800 1000 1200 1400 1600 1800 2000 2200 2400 2600 2800 3000 3200 3400
Detrital Zircon Age (Ma)
Figure 10. U-Pb age-probability density plots of composites of data from selected basins, including
those along potential dispersal pathways, as well as some potential provenances. Data sources are as
follows: (A) Appalachian basin (composite plot from Thomas et al., 2017, which includes original analytical data from seven samples plus previously published data from Gray and Zeitler, 1997; Eriksson
et al., 2004; Thomas et al., 2004; Becker et al., 2005, 2006; Dodson, 2008; Park et al., 2010; Grimm et al.,
2013); (B) southern Illinois basin (composite of original analytical data from Thomas et al., 2020); (C) Permian basin (composite of published data from Soreghan and Soreghan, 2013; Xie et al., 2018b; Liu and Stockli,
2019); (D) Coahuila terrane (composite plot from Thomas et al., 2019, which includes original analytical
data from 10 samples plus previously published data from Lawton and Molina-Garza, 2014); (E) Marathon foreland (composite plot from Thomas et al., 2019, which includes original analytical data from two
samples plus previously published data from Gleason et al., 2007); (F) Ordovician–​Silurian of Ouachita
passive-​margin off-shelf deep-water sandstones (composite of data from McGuire, 2017; Gleason et al.,
2002); (G) Whitehorse Group KS-5-WH, distal Anadarko basin (from Thomas et al., 2020); (H) Anadarko
basin (composite from Fig. 9); (I) Fort Worth basin (composite from Fig. 8); (J) Arkoma basin (composite from Fig. 7).
from synorogenic Alleghanian sandstones of the
Appalachian basin (Fig. 10A) (Thomas et al., 2004,
2017), indicating that Appalachian dispersal systems are not the likely source of these grains. This
age group is well represented in magmatic arc
rocks of the accreted Coahuila (Gondwanan) terrane within the Marathon orogen (Lopez, 1997;
Lopez et al., 2001) and brackets the ages of tuff
beds in the Ouachita synorogenic clastic wedge
(Fig. 6). These ages are represented also in detrital
zircons from sandstones in the Marathon foreland (Fig. 10E). The small numbers of grains of
this age group in multiple samples favor a provenance in the Coahuila terrane or another accreted
Gondwanan terrane in the interior of the Ouachita-​
Marathon orogen.
Thomas et al. | Detrital zircons and late Paleozoic sediment dispersal, Ouachita orogen
1233
Research Paper
Density contours
A) Arkoma basin
20
Wichita Granite
15
68% (1σ)
95% (2σ)
Appalachian
basin
10
5
20
5
-5
DM
CHUR
0
εHft
εHft
Appalachian
basin
10
-5
-10
-10
-15
Arkoma basin
Amazonia
Marathon foreland
-20
-25
Maximum Intensity
Wichita Granite
15
DM
CHUR
0
Minimum Intensity
400
800
1200
1600
2000
Detrital Zircon Age (Ma)
l
sta
cru tion
u
l
o
ev
2400
Arkoma basin
Amazonia
Marathon foreland
-15
-20
2800
-25
400
800
1200
1600
2000
Detrital Zircon Age (Ma)
l
sta
cru tion
u
l
o
ev
2400
2800
B) Fort Worth basin
20
20
Wichita Granite
15
Appalachian
basin
10
5
5
εHft
εHft
DM
CHUR
0
-5
-5
-10
Fort Worth
basin
Amazonia
Marathon foreland
-15
-20
-25
Appalachian
basin
10
DM
CHUR
0
Wichita Granite
15
400
800
1200
1600
2000
Detrital Zircon Age (Ma)
-10
l
sta
cru tion
lu
o
ev
2400
Fort Worth
basin
Amazonia
Marathon foreland
-15
-20
2800
-25
400
800
1200
1600
2000
Detrital Zircon Age (Ma)
l
sta
cru tion
lu
o
ev
2400
2800
C) Anadarko basin
20
20
Wichita Granite
15
Appalachian
basin
10
5
5
DM
CHUR
0
εHft
εHft
-5
-5
-10
Anadarko
basin
-15
Amazonia
Marathon foreland
-20
-25
Appalachian
basin
10
DM
CHUR
0
Wichita Granite
15
Figure 11. Comparisons of εHft data for
the (A) Arkoma (Fig. 7), (B) Fort Worth
(Fig. 8), and (C) Anadarko (Fig. 9) basins with respect to the Appalachian
basin, Marathon foreland, and Amazonia (from Thomas et al., 2017, 2019).
Age-εHft data are plotted as bivariate
kernel density estimates (KDEs), using
HafniumPlotter version 1.5 (github.com​
/kurtsundell/​ HafniumPlotter; Sundell et
al., 2019). This construction of bivariate
KDEs follows a similar formulation as
standard one-dimensional KDEs (Silverman, 1986) and incorporates set kernel
bandwidths of 25 m.y. along the x axis
(U-Pb age) and 2 epsilon units along the
y axis (εHft value). The bivariate KDEs
were constructed on a two-dimensional
x-y grid of 512 × 512 cells, in which each
cell had a corresponding density value
on the z axis. When viewed parallel to
the z axis, bivariate KDEs produce a
density map that is contoured to specified intervals. Contours are plotted at
68% (solid lines) and 95% (dashed lines)
(1σ and 2σ, respectively) of peak density;
density is the height of the z axis and
is shown by a color gradient, in which
hotter colors represent larger (higher)
z-axis values. See text for discussion of
interpretations. CHUR—chondritic uniform reservoir; DM—depleted mantle.
400
800
1200
1600
2000
Detrital Zircon Age (Ma)
GEOSPHERE | Volume 17 | Number 4
Downloaded from http://pubs.geoscienceworld.org/gsa/geosphere/article-pdf/17/4/1214/5362417/1214.pdf
by Idaho State University user
-10
l
sta
cru tion
u
ol
ev
2400
Anadarko
basin
-15
Amazonia
Marathon foreland
-20
2800
-25
400
800
1200
1600
2000
Detrital Zircon Age (Ma)
l
sta
cru tion
olu
v
e
2400
Thomas et al. | Detrital zircons and late Paleozoic sediment dispersal, Ouachita orogen
2800
1234
Research Paper
499–335 Ma
This age group (Table 3) is represented in all
of the samples and is prominent in many. These
ages, corresponding to the ages of the Taconic and
Acadian orogenies, are generally abundant in sandstones of the Appalachian basin (Fig. 10A) (Thomas
et al., 2017, and references therein). Likewise, these
ages are common in Gondwanan accreted terranes
adjacent to the southern margin of Laurentia, and
the ages are generally abundant among detrital zircons in the Marathon foreland (Fig. 10E) (Thomas
et al., 2019). This age group offers no clear distinction between an Appalachian or Gondwanan
provenance.
901–500 Ma
This age group (Table 3) is one of the most
distinctive, in that it is a dominant group in many
samples but is only sparsely represented in others. The age range corresponds to the Pan-African/
Brasiliano components of accreted terranes of
Gondwanan origin, along both the Appalachians
and the Ouachita-Marathon orogen (Thomas et al.,
2017, 2019). Alleghanian synorogenic sandstones
in the Appalachian basin have persistent but only
minor numbers of detrital-zircon grains of this age
group (Fig. 10A), indicating limited dispersal of
grains of this age through the Appalachian basin
(Thomas et al., 2017). The Coahuila and other
Gondwanan accreted terranes along the Ouachita-​
Marathon orogen include Pan-African/Brasiliano
rocks, and this age group is represented in dominant modes of detrital-zircon ages in the Marathon
foreland (Fig. 10E) (Thomas et al., 2019). The possibility that Iapetan synrift volcanic-magmatic
rocks (539–530 Ma; e.g., Wichita Granite Group)
along the Southern Oklahoma fault system might
have contributed to this age group (e.g., Sharrah,
2006) is eliminated by the contrast between the
εHft values of detritus from the Wichita Granite
Group (+10.1 to +4.7; Thomas et al., 2016) and
those of the detrital zircons in the Ouachita foreland (+8.3 to −30.5, mostly +4.9 to −5.6 in grains
<600 Ma; Figs. 7–9 and 11). Abundant grains in
GEOSPHERE | Volume 17 | Number 4
Downloaded from http://pubs.geoscienceworld.org/gsa/geosphere/article-pdf/17/4/1214/5362417/1214.pdf
by Idaho State University user
this age group along the Ouachita foreland favor
a provenance in the accreted Gondwanan terranes,
including Sabine, along the Ouachita-Marathon orogen.
1276–914 Ma
This age group (Table 3) is represented in all
of the samples and is the dominant age group in
many. This is the age range of the Grenville orogen in the Appalachians, and Grenville-age detrital
zircons are abundant in late Paleozoic sandstones
in the Appalachian basin (Fig. 10A) (Thomas et
al., 2017). The same age range is similarly abundant in the Gondwanan Coahuila accreted terrane
(Fig. 10D) along the Marathon orogen, representing
the Sunsás province of Amazonia. Detrital grains of
this age group are abundant in synorogenic sandstones in the Marathon foreland (Fig. 10E) (Thomas
et al., 2019). This age group does not provide a clear
distinction between an Appalachian Grenville or a
Gondwanan Sunsás provenance.
1963–1202 Ma
This multimodal (generally four to six modes)
age group (Table 3) is represented in all of the
samples. With some variations, the abundance
of grains generally decreases with increasing age.
The complete range of ages spans the Penokean,
Trans-Hudson, Central Plains, Yavapai, Mazatzal,
and Granite-Rhyolite provinces of the Laurentian
craton and also includes the Ventuari-Tapajós,
Rio Negro–Juruena, and Rondonia–San Ignacio
provinces of Amazonia (Fig. 6). The Central Plains–
Yavapai–Mazatzal subgroup (1800–1600 Ma) is
represented by intermediate modes in the detrital zircons, suggesting a possible source in the
Nemaha basement uplift (Xie et al., 2016, 2018a);
however, Appalachian samples have similar
modes in this age range (Fig. 10A), interpreted
to be recycled from synrift sedimentary rocks
along the Iapetan margin (Thomas et al., 2017).
Ages of the Granite-Rhyolite province are particularly abundant in the passive-margin off-shelf
sandstones that were deposited around the
Ouachita embayment and subsequently imbricated in the Ouachita thrust belt (McGuire, 2017),
providing a potential source for recycling from
the Ouachita thrust belt to the synorogenic sandstones (Fig. 10F). Because of the matches of ages
with both Laurentian and Amazonian/Gondwanan
provinces, as well as the incorporation of Laurentian detritus in the Ouachita thrust belt, these
detrital zircons provide no clear distinction of the
provenance.
2224–1945 Ma
Most samples have small numbers of grains
in this age group (Table 3). The age group, which
has no counterpart in Laurentia, is distinctive of
the Moroni-Itacaiúnas province and the Trans-​
Amazonian belt of Amazonia and the Eburnian
province of West Africa (Fig. 6). Sandstones in
the Appalachian basin have small numbers of
grains of these ages, indicating a provenance
in Gondwanan accreted terranes in the Appalachian orogen (Fig. 10A) (Thomas et al., 2017).
This age group is clearly linked to Gondwana,
but the numbers of detrital zircons are less than
those of the Pan-African/Brasiliano age group
(Fig. 10).
3036–2387 Ma
Most samples have a minor mode in this
age group (Table 3), which overlaps the ages of
the Superior province of Laurentia and the Central Amazonia province of Amazonia/Gondwana
(Cordani and Teixeira, 2007; Cardona et al., 2010).
Grains with ages of the Superior province of
Laurentia are relatively abundant in Ordovician
passive-margin off-shelf sandstones, which were
deposited along the rifted margin of Laurentia
(Fig. 2, box A1; Figs. 4 and 10F) (Gleason et al.,
2002; McGuire, 2017). The Ordovician sandstones
are in the Ouachita thrust belt and were available
for recycling of Laurentian Superior detritus from
sources in the Ouachita orogen.
Thomas et al. | Detrital zircons and late Paleozoic sediment dispersal, Ouachita orogen
1235
Research Paper
Provenance and Dispersal of Late Paleozoic
Detritus in the Ouachita Foreland
Fort Worth Basin
Similarities through all the analyzed samples from the Fort Worth basin (Fig. 8) indicate a
consistent ultimate provenance through Middle
Pennsylvanian to Early Permian time. Most of the
age groups of detrital zircons in the Fort Worth
basin are equally available from Gondwanan
and Appalachian sources, making identity of the
provenance problematic and leaving only a few
age groups that are clear discriminants (Table 3).
One of the most dominant modes (616–532 Ma)
in the Fort Worth basin (Fig. 8) matches Gondwanan Pan-African/Brasiliano ages, which are not
abundant in samples from Mississippian–Permian sandstones of the Alleghanian Appalachian
basin (Fig. 10A). In some of the Fort Worth samples (TX-2-PB and TX-5-CC; Fig. 8), grains with
Neoproterozoic ages (898–500 Ma) are especially
abundant in proportion to grains with Mesoproterozoic ages (1256–914 Ma) (ratios of 0.570 and
0.983, respectively). Minor modes between 2092
and 1984 Ma match the ages of the Moroni-Itacaiúnas province and Trans-Amazonian belt of
Amazonia/Gondwana (Table 3); this age group is
rare in Mississippian–Permian sandstones of the
Alleghanian Appalachian basin (Fig. 10A). The
mix of inferred Laurentian and Gondwanan ages
in the Fort Worth basin has been used to support an interpretation that detritus from sources
in the Appalachian orogen of eastern Laurentia
was mixed with clearly non-Appalachian, more
proximal detritus from the accreted Sabine terrane in the Ouachita orogen (Alsalem et al., 2018).
The detrital-zircon ages that correspond to Appalachian age groups, however, are also available
from Gondwanan accreted terranes (Table 3)
along the Ouachita-Marathon orogen, which could
have supplied all of the detritus in the Fort Worth
basin (Fig. 12). Chert-clast conglomerates in Middle Pennsylvanian sandstones in the Fort Worth
basin have been interpreted to record dispersal
from exposures of Ordovician Bigfork Chert and/
or Devonian Arkansas Novaculite in the Ouachita
GEOSPHERE | Volume 17 | Number 4
Downloaded from http://pubs.geoscienceworld.org/gsa/geosphere/article-pdf/17/4/1214/5362417/1214.pdf
by Idaho State University user
sedimentary thrust belt (Brown et al., 1973), further
indicating sediment dispersal from the Ouachita
orogen to the foreland (Fig. 12C). In summary,
the consistent age distributions throughout the
stratigraphic succession and into the distal Fort
Worth basin indicate persistent dispersal of clastic
sediment from the orogenic source, as well as no
mixing of sediment from cratonic or more distal
sources into the distal foreland (Fig. 12).
The detrital-zircon age distributions in Pennsylvanian–Permian sandstones of the Fort Worth
basin are remarkably similar to those in the Marathon foreland (cf. Figs. 10I and 10E). In an MDS
plot, the sandstones in the Fort Worth basin are
near the coeval sandstones in the Marathon foreland (Fig. 13), indicating that detritus along the
Ouachita-​Marathon foreland in the Fort Worth
basin and Marathon foreland came from the same
or similar sources. In the MDS plot, the separation between the Middle Pennsylvanian Haymond
Formation and the Upper Pennsylvanian Gaptank
Formation and Middle Permian Road Canyon Formation in the Marathon foreland (Fig. 13) suggests
a change through time in the provenance, possibly
because of erosional unroofing. The distributions
of ages of detrital zircons in Middle Pennsylvanian
to Middle Permian sandstones in the Marathon
foreland are similar to those in the Coahuila terrane (Figs. 10D and 10E), and the age groups in the
Marathon foreland have potential sources in the
Coahuila terrane (Thomas et al., 2019). The MDS
plot shows that the Fort Worth samples are closer
to the Marathon and Coahuila data and are more
distant from the Appalachian data (Fig. 13B). The
Hf data have substantial overlap; however, the distribution of values from the Fort Worth basin in the
age range of 850–350 Ma is nearly identical to that
of the Marathon samples but only partially overlaps with the more restricted range of Appalachian
values (Fig. 11B). If the Sabine terrane is similar in
composition to the other components (Coahuila,
Sierra Madre, and Maya) of the Oaxaquia assemblage, the Sabine terrane could have supplied all
of the age groups of detrital zircons in the sandstones of the Fort Worth basin (Table 3), similar to
the supply from Coahuila to the Marathon foreland
(Fig. 12) (Thomas et al., 2019).
Anadarko Basin and Other Intracratonic Basins
Along the southern margin of the Anadarko
basin, the Lower Permian Post Oak Conglomerate
unconformably overlies the Cambrian synrift Wichita Granite Group in the Wichita uplift (Figs. 1B, 5B,
and 12D). The Post Oak Conglomerate has detrital
zircons almost exclusively near the crystallization
age (539–530 Ma) of the Wichita Granite Group, and
distinctively positive εHft values between +10.1 and
+4.7 (Fig. 11C) indicate juvenile magmas (Thomas
et al., 2016).
Farther east, the sandstone matrix of the Upper
Pennsylvanian Vanoss Conglomerate, which overlies an angular unconformity on lower Paleozoic
strata above the Precambrian basement in the
Arbuckle uplift (Figs. 1B, 5B, and 12D), is dominated
by Superior-age (2860–2620 Ma) zircons (Thomas
et al., 2016). The Ordovician limestone succession beneath the unconformity includes units of
quartzose sandstone that contain Superior-age
detrital zircons (Pickell, 2012). The combination of
limestone clasts from proximal strata beneath the
unconformity and the lack of Superior-age basement rocks in the region strongly argues that the
detrital zircons in the Vanoss Conglomerate were
recycled from the sandstones within the Ordovician
limestone succession (Thomas et al., 2016).
Stratigraphically above the Post Oak and
Vanoss Conglomerates, the Lower Permian Wellington Formation (Fig. 5A) in the lower part of a
succession of widespread Lower–Middle Permian
red beds and evaporites (Johnson et al., 1988) has
detrital-zircon age distributions distinctly different
from those of the conglomerates (Thomas et al.,
2016). A subset (673–504 Ma) of a dominant age
group (857–504 Ma) in the Wellington Formation
(Fig. 9) overlaps the age of detrital zircons from the
Cambrian synrift igneous rocks in the Post Oak Conglomerate. Although some grains in the Wellington
have εHft values (+6.5) similar to those in the Post
Oak Conglomerate, most Wellington grains (+6.5
to −32.2) are significantly more negative (Figs. 9
and 11C), indicating that the Wellington sands are
from different, more evolved sources than the
juvenile sources of the detritus in the Post Oak Conglomerate (Thomas et al., 2016). The abundance
Thomas et al. | Detrital zircons and late Paleozoic sediment dispersal, Ouachita orogen
1236
Research Paper
A – 326 Ma
105W
MICHIGAN
90W
95W
100W
75W
80W
Late Mississippian
35N
uth
ern
KO
Ok
lah
om
a
INO
IS
ILL
A-
oc
e
flo an
or
ST
no
a
30N
c
ar
ic
n
INE
lca
vo SAB
Ou
TN
tuff
h
nc
tre
y
ar
n
io
et m
cr ris
c
p
a
in
fore
arc bas
ILA
AHU
O
C
105W
105W
N
YA
MA
MICHIGAN
90W
RM
30N
INO
TY
ILL
ST
CI
TW
RE
om
LIN
CM
AT
AT
AT
a
IA
Ozark
RO
SP
O
EE
ANN
SUW
30N
INE
B
SA
105W
ILA
AHU
CO
RA E
ER R
SI AD
M
accretionary
prism
Ou
a
DP foredeep
35N
A-
BS
no
Lla
PE
lah
Ben
d
N
IA
Ok
KO
40N
CA
ern
SM
FO
GW
uth
AR
75W
80W
IS
SALIN
A
aha
Nem
on
mar
Cim
So
35N
AD
Cent. Okla.
AN
600 km
80W
early Middle Pennsylvanian
40N
400
85W
95W
100W
200
0
95W
100W
B – 313 Ma
35N
tuff O
HS
ST
Lla
30N
Ozark
BW
B en
I
RM
PE
AR
d
AN
AD
SG
HB
CA
RO
LIN
IA
So
40N
SALIN
A
N em
aha
FO
RE
ST
CI
TY
AN
Cent. Okla.
Cim
mar
on
40N
GEOSPHERE | Volume 17 | Number 4
100W
YA
MA
N
95W
200
0
85W
400
Figure 12. Sequence of regional maps showing interpreted sediment dispersal pathways
(yellow lines and arrows) into and across the
Ouachita foreland and southern Midcontinent.
Iapetan rifted margin of Laurentia and intracratonic basins are from Figure 1. Tectonic evolution
through the sequence of maps is adapted from
Mack et al. (1983), Houseknecht (1986), Ross
(1986), Thomas (1989, 2011), Viele and Thomas
(1989), and Mueller et al. (2014). Correlations of
all named stratigraphic units and approximate
stratigraphic level of each map are shown in
Figure 5B. A-O—Alabama-Oklahoma transform
fault; Oua—Ouachita rift. (A) Late Mississippian, 326 Ma. Southward subduction beneath
continental-margin volcanic arcs on the leading
edges of the accreted Coahuila (subduction initiated 355 Ma) and Sabine (subduction initiated
338 Ma) terranes, and rapid accumulation of
deep-water muddy turbidites along with eruptions of ash-flow tuffs (ST—Stanley, including
tuffs with ages of 328–320 Ma; TN—Tesnus). Fluvial dispersal of synorogenic detritus through
the Appalachian basin (SG—Stony Gap, Mauch
Chunk–Pottsville clastic wedge) via the southern
Illinois basin (HB—Hardinsburg) to the northern shelf of the Arkoma basin (BW—Batesville
and Wedington) and over the shelf margin into
deep water (HS—Hot Springs). (B) Early Middle Pennsylvanian, 313 Ma. Fort Worth basin:
filling of the foredeep and progradation of synorogenic clastic sediment onto the distal shelf
(BS—Big Saline). Marathon foreland: restricted
clastic sediment from the Coahuila terrane and
deposition of carbonate detritus from the shelf
(DP—Dimple). Arkoma basin: rapid and deep
flexural subsidence in response to emplacement of the tectonic load of the accretionary
prism onto the edge of continental crust and
deposition of very thick synorogenic turbidites
(AT—Atoka), dispersal from the Appalachian basin (CM—Cross Mountain, for example) via the
southern Illinois basin (TW—Tradewater) to the
northern shelf of the Arkoma basin (AT—Atoka)
and over the shelf margin (SP—Spiro) into deep
water (AT—Atoka) to mix with proximal sediment from the Sabine terrane. Appalachian
basin: Pennington-Lee clastic wedge (CM—
Cross Mountain) and Mauch Chunk–Pottsville
clastic wedge (SM—Sharp Mountain) prograde
to distal foreland. Anadarko basin: deposition of
proximal detritus (GW—Granite Wash) eroded
from Arbuckle and Wichita basement uplifts.
(Continued on following two pages.)
600 km
80W
Thomas et al. | Detrital zircons and late Paleozoic sediment dispersal, Ouachita orogen
Downloaded from http://pubs.geoscienceworld.org/gsa/geosphere/article-pdf/17/4/1214/5362417/1214.pdf
by Idaho State University user
1237
Research Paper
C – 309 Ma
105W
MICHIGAN
90W
95W
100W
75W
80W
late Middle Pennsylvanian
uth
ern
Ok
lah
a
INO
IS
ILL
PR
35N
A
d
-O
successor
basin
NEE
AN
SUW
E
BIN
no
fore
HM
om
Lla
P
Ozark
SV
B en
I
30N
KO
SN
AN
M
ER
AR
CD
CA
RO
LIN
IA
35N
AD
GW
So
40N
SALIN
A
N em
aha
FO
RE
ST
CI
TY
AN
Cent. Okla.
Cim
mar
on
40N
SA
dee
Ou
a
p
30N
ILA
AHU
CO
YA
MA
RA E
ER R
SI AD
M
N
105W
D – 300 Ma
105W
MICHIGAN
90W
95W
100W
Ok
lah
a
no
Lla
GT
80W
N
AC
HIA
40N
PA
L
ILL
GR
AP
IS
INO
TY
CI
IA
Ozark
LIN
ITA
75W
80W
RO
OUACH
VN
om
CS
RM
35N
A-
O
successor
basin
E
BIN
SA
EE
ANN
SUW
30N
Ou
a
30N
KO
PO
Ben
d
N
IA
AR
600 km
CA
ern
HON
MARAT
PE
FO
AD
uth
ST
Nem
on
mar
Cim
AN
Cent. Okla.
So
35N
RE
SALIN
A
aha
Pennsylvanian–Permian
40N
400
85W
95W
100W
200
0
Figure 12 (continued ). (C) Late Middle Pennsylvanian, 309 Ma. Final accretion of Sabine terrane
and last stage of Ouachita thrusting (blue lines
from Fig. 1). Marathon foreland: filling of the
foredeep with turbidites (HM—Haymond). Fort
Worth basin: progradation of deltaic deposits
from the overfilled foredeep to the distal shelf
(SN—Strawn). Arkoma basin: progradation of
fluvial to deltaic deposits from the Appalachian
basin (PR—Princess, Pennington-Lee clastic
wedge) via the southern Illinois basin (CD—
Carbondale) to southward prograding deltaic
deposits on the northern Arkoma shelf to mix
with minor deltaic deposits (SV—Savanna, for
example) from the overfilled Arkoma foredeep;
initial shallow-marine deposits unconformably
overlie deformed Atokan strata in a successor
basin in the Ouachita orogenic belt. Anadarko
basin: deposition of proximal detritus (GW—
Granite Wash) eroded from Arbuckle and Wichita
basement uplifts. (D) Pennsylvanian–Permian, 300 Ma. Final transpressional accretion of
Suwannee terrane and latest stages of Appalachian thrusting (blue lines from Fig. 1). Marathon
foreland: filling of foredeep and progradation
of deltaic to shallow-marine deposits onto the
distal shelf (GT—Gaptank). Fort Worth basin:
progradation of deltaic deposits to the distal
shelf (CS—Cisco). Arkoma basin: successor-basin deposits of Desmoinesian–Guadalupian age
cover deformed Ouachita sedimentary rocks
(Atoka) at an angular unconformity. Anadarko
basin: Wichita and Arbuckle basement uplifts
at southern edge of basin; onlap of conglomerates (PO—Post Oak, VN—Vanoss) above
angular unconformity on the basement uplifts.
Appalachian basin: northward dispersal of clastic sediment of Pennington-Lee clastic wedge
(GR—Greene, for example).
ILA
AHU
CO
MAYA
A
RR
E
SI
N
M
100W
E
GEOSPHERE | Volume 17 | Number 4
R
AD
105W
95W
200
0
85W
400
600 km
80W
Thomas et al. | Detrital zircons and late Paleozoic sediment dispersal, Ouachita orogen
Downloaded from http://pubs.geoscienceworld.org/gsa/geosphere/article-pdf/17/4/1214/5362417/1214.pdf
by Idaho State University user
1238
Research Paper
E – 270 Ma
105W
MICHIGAN
90W
95W
100W
HON
MARAT
Ok
lah
a
PA
LA
CH
IA N
Figure 12 (continued ). (E) Middle Permian,
270 Ma. Final accretion of Coahuila terrane and
last stage of Marathon thrusting at 295 Ma (blue
lines from Fig. 1). Marathon foreland: above
multiple unconformities, progradation of deltaic to shallow-marine deposits (RC—Road
Canyon) onto southern craton, to merge with
detritus from the overfilled Fort Worth basin,
and possibly continuing to the Permian basin
(DM—Delaware Mountain). Anadarko basin: progradation of detritus from Marathon foreland
and/or Fort Worth basin to cover eroded Wichita
and Arbuckle uplifts and the overfilled Anadarko
basin (RS—Rush Springs); dispersal northward
on the craton to the southwest of the Forest City
basin (WH—Whitehorse); detritus reworked by
south-southwest winds (orange arrowheads)
over northern part of Anadarko basin and possibly to Permian basin. Arkoma basin: filling of
successor basin.
35N
A-
O
successor
basin
Lla
DM
RC
EE
ANN
SUW
E
BIN
no
SA
30N
Ou
a
30N
om
B en
MI
R
PE
AN
KO
RS
Ozark
ITA
OUACH
d
ern
AR
CA
RO
LIN
IA
uth
Cent. Okla.
So
AD
40N
AP
WH
AN
35N
ILL
Cim
mar
on
40N
IN O
IS
SALIN
A
N em
ah a
FO
RE
ST
CI
TY
Middle Permian
75W
80W
ILA
AHU
CO
MAYA
RA
ER
SI
N
200
400
600 km
RE
AD
M
0
105W
0.3
0.2
Dimension II
0.6
A
0.1
SM2
0.5
TCP2
NR03
PC02
OK-9-RS2
Road Canyon
AR-2-HC
PC01
OK-4-WL
Atoka WS
Spiro
OK-5-GB2
TX-2-PB
AR-6-SV
Atoka NE Big Saline
Strawn-Canyon(S4-8)
Bloyd
-0.1
passive-margin Wedington
Gaptank
off-shelf
Stanley
-0.4
-0.4
Haymond(Z4)
-0.3
-0.2
Arkoma basin
-0.1
0
0.1
Dimension I
Fort Worth basin
Marathon foreland
0.2
0.1
0
-0.2
-0.3
Haymond(Z1)
0.3
0.4
0.5
A7
0.2
-0.1
TX-5-CC
Stanley tuffs (>340 Ma)
-0.3
A9
Haymond(Z4)
0.3
0
-0.2
B
0.4
NR02
PC04
PC03
AR-1-BV
Whitehorse
Hot Springs
Dimension II
0.4
80W
85W
95W
100W
A6
Stanley tuffs A17
A25
A11
A23 A27
A26
PC04
SM2
NR03
TCP2
Stanley
A24
NR02
PC01
A3 PC02
A21
Road Canyon
Whitehorse
TX-2-PB
A22
A13 A5
OK-9-RS2
OK-4-WL
A16
A28
A15
Atoka NE
Strawn-Canyon
Atoka WS
PC03
Big Saline
TX-5-CC
AR-6-SV
Gaptank
passive- Wedington
A1
AR-1-BV
Bloyd
margin
A2 OK-5-GB2
off-shelf Hot Springs
AR-2-HC
A12
Spiro
Haymond(Z1)
A8
A14
A10
A20
-0.4
-0.4
A19 A4
-0.3
-0.2
Anadarko basin
Coahuila terrane
A18
A1: Stony Gap
A2: Hinton
A3: Mauch Chunk
A4: Princeton
A5: Bluestone
A6: Raccoon Mtn
A7: Pocahontas
-0.1
0
0.1
0.2
Dimension I
0.3
Appalachian basin
A8: Lee
A15: Montevallo
A9: TumblingRun A16: SharonN
A10: Sewanee
A17: SharonS
A11: BottomCreek A18: Pottsville
A12: Corbin
A19: Norton
A13: LRaleigh
A20: CrossMtn
A14: URaleigh
A21: SharpMtn
0.4
0.5
0.6
A22: Princess
A23: Conemaugh
A24: LMonongahela
A25: UMonongahela
A26: Washington
A27: Greene
A28: Proctor
Figure 13. Two-dimensional multidimensional scaling (MDS) plots for detrital-zircon U-Pb age distributions of Mississippian–Permian sandstones in the Arkoma (Fig. 7), Fort Worth (Fig. 8), and Anadarko (Fig. 9) basins in comparison with distributions of samples from the Appalachian basin (from Thomas et al.,
2017, and references cited therein; Fig. 10A), from the Whitehorse Group in the distal Anadarko basin (KS-5-WH, from Thomas et al., 2020; Fig. 10G), from the
Marathon foreland (from Thomas et al., 2019, and Gleason et al., 2007; Fig. 10E), from the Coahuila terrane (only samples with n >45; from Thomas et al., 2019,
and Lawton and Molina-Garza, 2014; Fig. 10D), and in a composite from Ouachita passive-margin off-shelf sandstones (from McGuire, 2017, and Gleason et al.,
2002; Fig. 10F). Points that plot closer together have greater correspondence (are more similar) for detrital-zircon ages. The MDS plots (Vermeesch, 2013) were
constructed using the program DZmds (github.com/kurtsundell/DZmds) (Saylor et al., 2018) using Mismatch of age-probability density plots (Amidon et al.,
2005) to calculate dissimilarity. Axis scales are calculated dissimilarities between samples. (A) Samples from Arkoma, Fort Worth, and Anadarko basins; Marathon foreland; and Coahuila terrane. (B) Comparison of samples plotted in A with samples from Appalachian basin. See text for discussion of interpretations.
GEOSPHERE | Volume 17 | Number 4
Thomas et al. | Detrital zircons and late Paleozoic sediment dispersal, Ouachita orogen
Downloaded from http://pubs.geoscienceworld.org/gsa/geosphere/article-pdf/17/4/1214/5362417/1214.pdf
by Idaho State University user
1239
Research Paper
of recycled Superior-age zircons (2860–2620 Ma)
in the Vanoss Conglomerate is only weakly represented in the detrital zircons in the Wellington
Formation (Fig. 9). The mix of zircon ages in the
Wellington Formation (Fig. 9), along with the generally more negative εHft values, indicates that the
sands were derived neither from the local synrift
igneous rocks in the Wichita and Arbuckle uplifts
nor from recycling from the conglomerates that lap
onto the uplifts. Instead, the age distributions suggest a source separate and distal from the Arbuckle
and Wichita uplifts. In a broader sense, these data
show that the crystalline rocks of the Wichita and
Arbuckle uplifts were rapidly covered in the Early
Permian and were not available later as a source
of regionally dispersed sediment (Soreghan and
Soreghan, 2013; Thomas et al., 2016).
Along with the Wellington Formation, the
Garber and Rush Springs Sandstones include a
distinct Pan-African/Brasiliano component (857–
502 Ma), the usual abundance of Mesoproterozoic
grains (1228–924 Ma, Sunsás = Grenville age), and
moderate abundance of zircons in the age range of
2200–1200 Ma (Fig. 9). Similarities of detrital-zircon distributions in Permian sandstones in and
above the Wellington Formation in the intracratonic Anadarko basin (Figs. 9, 10H, 11C, and 13)
to those in the Marathon foreland (Figs. 10E, 11C,
and 13) and Fort Worth basin (Figs. 8, 10I, 11, and
13) suggest the same or a similar provenance.
Dispersal of sediment through either or both the
Marathon foreland and Fort Worth foreland basin
requires fluvial systems that prograded onto the
craton and into the Anadarko basin (Fig. 12E).
The Rush Springs Sandstone consists primarily
of eolian deposits in an erg system with southwest-directed wind currents (Poland and Simms,
2012), which are compatible with overall north to
northwest drainage from the Ouachita-Marathon
orogen onto the craton and into the Anadarko
intracratonic basin with episodic wind reworking
back over the fluvial system (Fig. 12E) (Soreghan
and Soreghan, 2013).
A sample from the Middle Permian Whitehorse
Group (KS-5-WH; Fig. 10G; Thomas et al., 2020)
on the structural divide between the Anadarko
basin and the Forest City basin on the north
GEOSPHERE | Volume 17 | Number 4
Downloaded from http://pubs.geoscienceworld.org/gsa/geosphere/article-pdf/17/4/1214/5362417/1214.pdf
by Idaho State University user
(Fig. 1) suggests distant transport from the Ouachita-Marathon orogen onto the craton (Fig. 12E). The
detrital-zircon distributions are generally similar to
those in the stratigraphically lower Pennsylvanian
and Mississippian sandstones in the Forest City
basin, which are characteristic of an Appalachian
provenance (Kissock et al., 2018; Chapman and
Laskowski, 2019; Thomas et al., 2020). In contrast,
the Middle Permian Whitehorse sample includes
a secondary group with ages of 638–519 Ma and a
minor mode of 576 Ma (Fig. 10G); in addition, four
grains have ages of 2196–2021 Ma. These distinctly
Pan-African/Brasiliano and Moroni-Itacaiúnas/
Trans-Amazonian/Eburnian ages suggest a Gondwanan provenance. Another, distinctively young
secondary distribution with ages of 335–287 Ma has
a mode of 319 Ma (Fig. 10G), corresponding to the
age of the Alleghanian orogeny in the Appalachians,
but contrasting with the general lack of Alleghanian-age grains in the synorogenic dispersal
systems of the Appalachian basin (Table 3; Thomas
et al., 2017). The secondary mode with Alleghanian
ages in the Whitehorse sample is similar in age to
distributions in coeval sandstones in the Anadarko
basin (Fig. 9) and Marathon foreland (Fig. 10E; fig.
4 in Thomas et al., 2019), as well as a prominent
mode in Jurassic–Cretaceous sandstones in faultbounded basins in the Coahuila terrane (Fig. 10D).
The association of these detrital-zircon age distributions with Permian red beds forms a distinctive
assemblage, implying long-distance fluvial transport onto the southern craton from Gondwanan
accreted terranes, such as Coahuila (Fig. 12E).
Detrital-zircon data from Lower and Middle
Permian sandstones in the Permian (Delaware and
Midland basins) basin (Figs. 1A, 5B, and 10C) suggest widespread sediment dispersal from sources
like those for the Marathon foreland, as well as the
Fort Worth and Anadarko basins (Soreghan and
Soreghan, 2013; Xie et al., 2018b; Liu and Stockli,
2019). Most samples from the Permian basin have
abundant detrital-zircon ages in the age ranges of
1975–1211, 1208–930, and 500–335 Ma (Fig. 10C),
all consistent with either Appalachian or Gondwanan sources (Table 3). The samples from the
Permian basin have a dominant age distribution
of 852–500 (mostly 685–500) Ma, as well as a lesser
age distribution of 2203–1975 Ma (Fig. 10C), both
distinctly from Gondwanan sources (Table 3). Most
samples from the Permian basin have a secondary distribution with ages of 334–277 Ma similar
to those in the Marathon and Anadarko samples
(Fig. 10). The data lead to an interpretation of fluvial
dispersal of detritus from the Ouachita-Marathon
orogen and accreted Gondwanan terranes across
the broad foreland and southern craton (Fig. 12E).
Arkoma Basin
The tuff beds in muddy turbidites of the Mississippian–Pennsylvanian Stanley Shale in the lower
part of the proximal synorogenic clastic wedge in
the Arkoma basin provide a clear link to a source in
an arc system south of the foredeep (Niem, 1977),
consistent with a continental-margin arc on the
leading edge of the Sabine terrane (Fig. 12A), as
well as a definitive age (328–320 Ma; Fig. 6) of arc
volcanism (Shaulis et al., 2012). Detrital zircons
with ages equal to the crystallization ages of the
tuffs are lacking in the foreland detritus. Although
dominated by synmagmatic grains, the tuffs contain zircon grains older than the crystallization age
(Fig. 6), which have been interpreted to be detrital
grains from the muddy turbidites (Shaulis et al.,
2012), but which, alternatively, may be xenocrysts
from the magma chamber in the Sabine terrane.
Although the proportions differ somewhat, most
of the age modes of the prevolcanic zircons in the
tuffs of the Stanley Shale are similar to those of the
Coahuila terrane, as well as those in the Marathon
foreland and Fort Worth basin (cf. Figs. 6 and 10D,
10E, and 10I, respectively); however, the tuffs differ
primarily in having a noticeably greater abundance
of zircon grains in the age range of 1577–1283 Ma
(Fig. 6). If the prevolcanic grains in the tuffs are detrital from Laurentian/Appalachian sources (Shaulis et
al., 2012), the source of the 1577–1283 Ma grains is
problematic, because grains of this age range are
only secondary components of the Appalachian signature (cf. Figs. 6 and 10A). Detrital grains with ages
of the Granite-Rhyolite province (1500–1320 Ma)
of Laurentia are relatively abundant in the Ordovician–Silurian Ouachita passive-margin off-shelf
Thomas et al. | Detrital zircons and late Paleozoic sediment dispersal, Ouachita orogen
1240
Research Paper
strata (Fig. 10F) and might have been available for
deltaic systems modified by marine reworking
recycling, but grains of those ages are rare in the
and interaction with carbonate deposition on the
Stanley Shale (Fig. 7). If the prevolcanic grains in
shallow-marine shelf (Garner, 1967; Ogren, 1968;
the tuffs are xenocrystic, the composition of the
Glick, 1979; Sutherland, in Johnson et al., 1988; Xie
basement of the Sabine terrane may differ from
et al., 2016). The two most dominant detrital-zirthat of the Coahuila terrane in having a locally more
con age groups have ages of 1210–937 Ma and of
substantial Rondonia–San Ignacio (1550–1300 Ma)
496–339 Ma (Fig. 7). Ages are scattered between
component along with the Sunsás rocks, as sug- 1907 and 1210 Ma, and an older group has ages
gested for northwestern Amazonia (e.g., Cardona
of 3036–2466 Ma. The age distributions have relaet al., 2010), or, alternatively, the magma may have
tively few grains with distinctive Gondwanan ages
been contaminated by subducted sediment from
of 2186–1945 Ma (Moroni-Itacaiúnas/Trans-Amathe Ouachita margin. Future research may provide
zonian/Eburnian) or of 762–501 Ma (Pan-African/
answers to these questions, as well as better res- Brasiliano). The detrital-zircon age distributions
olution of the composition of the Sabine terrane.
(Fig. 7) of the Batesville and Wedington Sandstones,
A sandstone bed in the upper part of the
as well as of the Hot Springs Sandstone, are similar
proximal deposits of the Stanley Shale has a dom- to the Appalachian signature (Fig. 10A). The Batesinant group of detrital-zircon ages of 874–501 Ma
ville, Wedington, and Hot Springs data, as well as
(McGuire, 2017), corresponding in age to the
data from coeval sandstones in the southern Illinois
Pan-African/Brasiliano assemblage of Gondwa- basin and in the Appalachian basin, are all close
nan terranes (Fig. 7; Table 3). Another important
together in an MDS plot (Fig. 14A), indicating a
age group of 1197–1018 Ma is equivalent in age to
similar provenance. Values of εHft from the Batesthe Amazonian Sunsás and Laurentian Grenville
ville Sandstone have a remarkably close match to
provinces (Fig. 7; Table 3). A secondary age group
those of Mississippian sandstones in the Appalaof 2185–1953 Ma matches the distinctive Moroni-​ chian foreland and in the southern Illinois basin
Itacaiúnas/Trans-Amazonian ages of Amazonia
(Fig. 15A) (Thomas et al., 2017, 2020). An ultimate
(Table 3). The overall distribution of detrital-​zircon
source in the Appalachians with dispersal from the
ages indicates a Gondwanan provenance (Fig. 12A), Appalachian basin through the southern part of the
consistent with the composition of the Coahuila
Illinois basin may have supplied the Batesville and
terrane and of the Sabine terrane as suggested by
Wedington deltaic sand to the cratonic fringe of
the prevolcanic age distributions in the tuff beds
the Arkoma basin, as well as the Hot Springs sand
(Figs. 6, 7, and 10D; Table 3). In contrast, the Hot
(within the muddy turbidites of the Stanley Shale)
Springs Sandstone Member of the Stanley Shale
to the adjacent continental slope, during the Late
has dominant age groups of 1195–994 Ma and
Mississippian (Fig. 12A).
484–373 Ma (McGuire, 2017), respectively, correPennsylvanian sandstones (Cane Hill Memsponding to the Grenville and Taconic age groups
ber of the Hale Formation, Bloyd Formation, and
in the Appalachians and overlapping with Gondwa- Savanna Sandstone; Figs. 2 and 5A) on the distal
nan age groups (Fig. 7; Table 3). The Hot Springs
shelf of the Arkoma basin generally are interpreted
sample, however, lacks distinctive Gondwanan
to be fluvial to deltaic deposits with southward to
components, thereby indicating a provenance
westward directed sediment dispersal (Sutherland,
different from the other sample from the Stanley
in Johnson et al., 1988; Xie et al., 2018a). The detriShale, most likely the Appalachians (Fig. 12A).
tal-zircon distributions are generally similar through
In the most distal shelf of the Arkoma basin
the Upper Mississippian to Middle Pennsylvanian
in northern Arkansas, distribution patterns and
succession (Fig. 7), suggesting persistent sedipaleocurrents of Mississippian sandstones
ment dispersal. The most prominent age groups of
(Batesville Sandstone and Wedington Sandstone
detrital zircons have ages of 1265–935 Ma and 495–
Member of the Fayetteville Shale; Fig. 2, box B1;
374 Ma (Fig. 7). Ages are spread broadly from 1950
Figs. 5A and 12A) indicate southward prograding
to 1265 Ma with a strong secondary mode between
GEOSPHERE | Volume 17 | Number 4
Downloaded from http://pubs.geoscienceworld.org/gsa/geosphere/article-pdf/17/4/1214/5362417/1214.pdf
by Idaho State University user
1635 and 1618 Ma (Fig. 7). An older age group has
modes between 2820 and 2569 Ma. The groups of
ages in these samples generally match the Appalachian signature (Figs. 10A, 13, and 14B); however,
these age groups are also available from Gondwanan terranes (Table 3). The sandstones (Fig. 7)
have minor groups with ages of 2190–1950 Ma
(Moroni-Itacaiúnas/Trans-Amazonian/Eburnian)
and of 682–528 Ma (Pan-African/Brasiliano) that
suggest a possible Gondwanan provenance; however, the sandstones in the Appalachian basin have
similar minor groups of these ages (Thomas et al.,
2017). Age data from the Corbin Sandstone in the
Appalachian basin, the Caseyville Sandstone in
the southern Illinois basin, and the Pennsylvanian
sandstones in the northern shelf of the Arkoma
basin are very close together in an MDS plot
(Fig. 14B), indicating a common provenance and
dispersal system. Distributions of εHft values from
the Pennsylvanian Cane Hill and Savanna sandstones in the Arkoma basin are nearly identical
to those from the Appalachian foreland and from
the southern Illinois basin (Fig. 15B) (Thomas et al.,
2017, 2020). The detrital-zircon U-Pb ages and εHft
values strongly indicate an Appalachian source and
dispersal through the southern part of the Illinois
basin to the northern shelf of the Arkoma basin
(Fig. 12C). One possible exception is the Savanna
Sandstone, which has a somewhat greater proportion of grains with Pan-African/Brasiliano ages and
a minor mode of 589 Ma (Fig. 7). The Pan-African/
Brasiliano grains in the Savanna Sandstone, along
with the chert pebbles in the stratigraphically higher
Thurman Sandstone (Fig. 2, box C1) (Sutherland,
in Johnson et al., 1988), show that some sediment
spread from the Ouachita orogenic belt into the
distal foreland to mix with the extrabasinal sediment (Fig. 12C).
To document the provenance of the synorogenic turbidites in the Ouachita foredeep, detrital
zircons were analyzed from the Middle Pennsylvanian Atoka Formation in the Ouachita thrust belt
from sample sites with northeast-directed, west-directed, and south-directed (Spiro) paleocurrents
(Sharrah, 2006). All three sets of samples have a
dominance of Mesoproterozoic (1205–933 Ma) zircons, a range of older zircons between 1947 and
Thomas et al. | Detrital zircons and late Paleozoic sediment dispersal, Ouachita orogen
1241
Research Paper
0.5
Mauch Chunk
0.2
0.1
Aux Vases (IB-AV)
Hardinsburg(IB-HB)
Stony Gap
Hinton
Tar Springs
-0.1
Stanley
Aux Vases (C&L)
-0.3
-0.2
-0.1
0
SharonS
SharpMtn
0.2
0.1
0
UMonongahela
Conemaugh
Pocahontas
Greene
0.1
Washington
0.3
0.4
0.5
0.6
Sewanee
Dixon SharonN
Atoka NE
Princess Atoka WS
Caseyville(KY-15)
Norton
AR-6-SV
Proctor
AR-2-HC
Caseyville(KY-4)
Corbin
LRaleigh
Spiro
Bloyd
-0.4
-0.4
URaleigh
Montevallo
-0.3
0.2
CrossMtn
BottomCreek
LMonongahela
RaccoonMtn
-0.2
AR-1-BV
-0.2
Pottsville
-0.1
Hardinsburg(KY-1-HB)
Cypress
TumblingRun
0.3
Wedington
0
B
0.4
Stanley tuffs >340 Ma
Hot Springs
Dimension II
0.6
Bluestone
Princeton
A
Dimension II
0.3
Lee
-0.3
-0.2
-0.1
Dimension I
0
0.1
0.2
0.3
0.4
Dimension I
Arkoma basin
Illinois basin
Appalachian basin
Figure 14. Two-dimensional multidimensional scaling (MDS) plots for detrital-zircon U-Pb age distributions for (A) Mississippian and (B) Pennsylvanian sandstones in the Arkoma basin in comparison with distributions in the southern Illinois basin and the Appalachian basin (B includes Lower Permian samples)
along the inferred dispersal pathway from the Appalachians via the southern Illinois basin to the distal Arkoma basin. Data sources for panel A: Arkoma basin
from original data herein and references cited in Figure 7; southern Illinois basin from Thomas et al. (2020) and sample Aux Vases (C&L) from Chapman and
Laskowski (2019); Appalachian basin from Park et al. (2010) and Thomas et al. (2017). Data sources for panel B: Arkoma basin from original data herein and
references cited in Figure 7; southern Illinois basin from Thomas et al. (2020); Appalachian basin from references cited for Figure 10A. Specifications for the
MDS plots are described in the caption for Figure 13. See text for discussion of interpretations.
1210 Ma, and a scattering of zircons older than
2340 Ma, all of which have counterparts in both
Amazonia/Gondwana and Laurentia (Fig. 7; Table 3).
All of the samples have a group of younger zircons
(492–336 Ma; Fig. 7), suggesting possible recycling
from Appalachian sources (Sharrah, 2006); however, the same age range of zircons is available
from the Coahuila terrane (Thomas et al., 2019) and
may be from the Sabine terrane. The samples from
sandstones with northeast-directed paleocurrents
have a secondary group of 34 grains with ages of
799–505 Ma and a mode of 632 Ma, as well as 15
grains with ages of 2218–1962 Ma (Fig. 7), both
distinctly Gondwanan (Table 3). The other two
sets of Atoka samples have fewer grains in these
age groups (Fig. 7). For these distinctive Gondwanan age groups in the northeast-dispersed Atoka
GEOSPHERE | Volume 17 | Number 4
Downloaded from http://pubs.geoscienceworld.org/gsa/geosphere/article-pdf/17/4/1214/5362417/1214.pdf
by Idaho State University user
sandstones (Table 3), the interpreted provenance is
the west end of the accreted Sabine terrane, which
is inferred to have collided with the Laurentian rift
margin near the present location of the Waco uplift
(Fig. 12B); the Atoka west-dispersed and south-dispersed (Spiro) sands are interpreted to be from
Appalachian sources (Sharrah, 2006). An MDS plot
shows that the northeast-dispersed, west-dispersed,
and south-dispersed (Spiro) Atoka sandstones form
a tight cluster that also includes a sample of the
Corbin Sandstone from the Appalachian basin
and two samples of Caseyville Sandstone from
the southern Illinois basin (Fig. 14B), suggesting
a common provenance for all of these sandstones.
The Gondwanan grains in the northeast-dispersed
sandstones suggest, however, that dispersal from
the Sabine terrane was mixed with Appalachian
detritus that prograded across the northern distal
shelf into the distal side of the foredeep, similar
to the dispersal of sand in the Spiro Sandstone
(Fig. 12B).
In summary, some of the most proximal Mississippian–Pennsylvanian sediments in the Arkoma
basin have detrital-zircon age distributions that are
generally similar to those in the Fort Worth and
Marathon foreland basins, suggesting the same
or similar provenance within the accreted Gondwanan terranes of the Ouachita orogen (Fig. 12A).
The orogenic source is compatible with longitudinal paleocurrents in the Ouachita foredeep. The
Mississippian Hot Springs Sandstone, which lacks
distinctive Gondwanan ages of detrital zircons, may
represent southward dispersal across the northern distal shelf and down the cratonic side of the
Thomas et al. | Detrital zircons and late Paleozoic sediment dispersal, Ouachita orogen
1242
Research Paper
A) Arkoma basin – Mississippian
20
20
Appalachian
basin
15
10
5
5
εHft
εHft
DM
CHUR
0
-5
-5
-10
Arkoma basin
-10
southern Illinois
l
basin
sta
-15
400
800
1200
1600
2000
Detrital Zircon Age (Ma)
2400
Arkoma basin southern Illinois
basin
t al
us
-15
cru tion
olu
ev
-20
-25
10
DM
CHUR
0
Appalachian
basin
15
n
cr
tio
olu
ev
-20
2800
-25
400
800
1200
1600
2000
Detrital Zircon Age (Ma)
2400
2800
B) Arkoma basin – Pennsylvanian
20
20
Appalachian
basin
15
10
5
5
southern Illinois
basin
-15
l
sta
cru tion
lu
o
ev
Arkoma basin
-20
400
800
1200
1600
2000
Detrital Zircon Age (Ma)
2400
Density contours
68% (1σ)
95% (2σ)
Ouachita foredeep, mixing with the synorogenic
muddy turbidites (Fig. 12A). In the distal fringe
of the Arkoma basin along the southern margin
of the Laurentian craton, depositional systems
indicate dominant sediment dispersal southward
from the craton into the subsiding Arkoma foreland basin (Figs. 12B and 12C). Although all of
the components of these sandstones could have
come from the Sabine terrane, the paucity of truly
distinctive Gondwanan components suggests
a different source, most likely the Appalachians
with dispersal through the southern part of the
GEOSPHERE | Volume 17 | Number 4
Downloaded from http://pubs.geoscienceworld.org/gsa/geosphere/article-pdf/17/4/1214/5362417/1214.pdf
by Idaho State University user
-5
-10
Arkoma basin
-15
southern Illinois
basin
l
sta
cru tion
u
l
o
ev
-20
2800
-25
DM
CHUR
0
εHft
εHft
-5
-10
-25
10
DM
CHUR
0
Appalachian
basin
15
Figure 15. Comparisons of εHft data from
the Arkoma basin (Fig. 7) with respect
to the southern Illinois basin and Appalachian basin (from Thomas et al., 2017,
2020): (A) Mississippian; (B) Pennsylvanian (Appalachian data include Lower
Permian). CHUR—chondritic uniform
reservoir; DM—depleted mantle. Specifications for the plots are described in
the caption for Figure 11. See text for
discussion of interpretations.
400
800
1200
1600
2000
Detrital Zircon Age (Ma)
Minimum Intensity
2400
2800
Maximum Intensity
Illinois basin to the distal fringe of the Arkoma
basin (Figs. 12B and 12C). The chert pebbles and
some detrital zircons, however, indicate at least
some mixing from Ouachita orogenic sources to
the most distal foreland (Fig. 12C). In this regard,
the Arkoma foreland basin differs from the Fort
Worth basin, where even the most distal fringe has
a distinctive detrital-zircon age distribution from
accreted Gondwanan terranes. The Arkoma basin
evidently marks the southwestward limit of dispersal through the Appalachian foreland, and the
basins in the Ouachita-Marathon foreland farther
southwest have distinct Gondwanan provenances
in the accreted terranes (Fig. 12).
■■ CONCLUSIONS
Thomas et al. | Detrital zircons and late Paleozoic sediment dispersal, Ouachita orogen
The age distributions of detrital zircons in Middle Pennsylvanian to Middle Permian sandstones in
the Marathon foreland indicate a sediment supply
from the accreted Coahuila (Gondwanan) terrane in
the interior of the Marathon orogenic belt (Thomas
et al., 2019). All of the ages of the Marathon detritus
1243
Research Paper
could have been derived from the Coahuila terrane,
requiring no other contribution from mixed sources
such as the Appalachians. The documented composition of the Coahuila terrane may serve as a model
for other probable, less-well-documented, accreted
Gondwanan terranes, such as the Sabine terrane.
The age distributions of detrital zircons in Middle Pennsylvanian to Lower Permian sandstones in
the Fort Worth basin in the Ouachita foreland are
similar to those in the Marathon foreland, indicating the same source or another source with similar
composition. The Sabine terrane along the interior
part of the Ouachita orogen is the likely source for
the foreland detritus in the Fort Worth basin. The
stratigraphically highest and most distal sandstones in the Fort Worth basin have the Gondwanan
detrital-zircon signature, consistent with progradation of synorogenic clastic sediment to the cratonic
side of the basin throughout the history of basin
filling as the basin became overfilled.
The age distributions of detrital zircons in Lower
and Middle Permian sandstones in the Anadarko
basin indicate dispersal of detritus onto the adjacent southern craton from the same sources that
supplied the Marathon and Fort Worth foreland
detritus. Published results from the intracratonic
Permian basin have been interpreted similarly.
Sediment dispersal from orogenic sources adjacent to either or both the Marathon and Fort Worth
forelands spread onto the craton into intracratonic basins.
The age distributions of detrital zircons in Mississippian to Middle Pennsylvanian sandstones
differ both laterally and vertically within the Arkoma
basin. Some sandstones in the most proximal part
of the basin have a Gondwanan signature; whereas
those in the most distal basin generally do not have
distinctive Gondwanan components, suggesting a
provenance in the Appalachian orogen and sediment dispersal through the Appalachian foreland
and southern Illinois basin to the cratonic side
of the Arkoma foreland basin. Some sandstones
indicate intermittent mixing from opposite sides
of the basin.
The combined data from the Fort Worth and
Arkoma foreland basins and from the Anadarko
intracratonic basin suggest a broad pattern of
GEOSPHERE | Volume 17 | Number 4
Downloaded from http://pubs.geoscienceworld.org/gsa/geosphere/article-pdf/17/4/1214/5362417/1214.pdf
by Idaho State University user
sediment dispersal from the Ouachita-Marathon
orogen (largely accreted Gondwanan terranes)
onto the southern craton during the Late Mississippian through Early Permian and at least into
the Middle Permian. Dispersal from the Appalachians via the Illinois basin brought sediment to
the cratonic side of the Arkoma basin and distal
foreland, and the Arkoma basin evidently marks
the southwestern limit of dispersal of recognizable
Appalachian detritus along the Ouachita-Marathon
foreland.
ACKNOWLEDGMENTS
Sample collection and analyses were funded by National Science Foundation (NSF) award EAR-1304980 to Gehrels and
Thomas. Thomas collected the samples with the assistance of
Beth Welke. In the Anadarko basin, we were guided through the
stratigraphy and depositional systems and to collecting sites by
Neil Suneson, Jim Puckette, and Zach Poland. Majie Fan led us
through the outcrops in the Fort Worth basin and also provided
an instructive review of an early draft of the manuscript. In the
Arkoma basin, Doy Zachry provided an overview of stratigraphy
and locations for collecting; after our field season was frozen
to an end by an ice storm, Doy along with Peggy Gucchione
collected the sample from the Savanna Sandstone. All analyses
were conducted at the Arizona LaserChron Center with the support of NSF awards EAR-1338583 and EAR-1649254. Tim Lawton
and Ed Osborne provided helpful reviews and sharpened the
discussion in a draft manuscript. We appreciate the efforts of
reviewer Ryan Leary, an anonymous reviewer, Associate Editor
Todd LaMaskin, and Editor David Fastovsky in seeing this article
through the Geosphere review process.
REFERENCES CITED
Alsalem, O.B., Fan, M., Samora, J., Xie, X., and Griffin, W.R., 2018,
Paleozoic sediment dispersal before and during the collision between Laurentia and Gondwana in the Fort Worth
basin, USA: Geosphere, v. 14, p. 325–342, https://​doi​.org​
/10​.1130​/GES01480​.1.
Amidon, W.H., Burbank, D.W., and Gehrels, G.E., 2005, U-Pb
zircon ages as a sediment mixing tracer in the Nepal Himalaya: Earth and Planetary Science Letters, v. 235, no. 1–2,
p. 244–260, https://​doi​.org​/10​.1016​/j​.epsl​.2005​.03​.019.
Arbenz, J.K., 1989a, The Ouachita system, in Bally, A.W., and
Palmer, A.R., eds., The Geology of North America—An Overview: Boulder, Colorado, Geological Society of America, The
Geology of North America, v. A, p. 371–396, https://​doi​.org​
/10​.1130​/DNAG​-GNA​-A​.371.
Arbenz, J.K., 1989b, Ouachita thrust belt and Arkoma basin, in
Hatcher, R.D., Jr., Thomas, W.A., and Viele, G.W., eds., The
Appalachian-Ouachita Orogen in the United States: Boulder,
Colorado, Geological Society of America, The Geology of
North America, v. F-2, p. 621–634, https://​doi​.org​/10​.1130​
/DNAG​-GNA​-F2​.621.
Arbenz, J.K., 2008, Structural Framework of the Ouachita Mountains: Oklahoma Geological Survey Circular 112A, 40 p.
Arbenz, J.K., Muehlberger, W.R., Nicholas, R.L., Tauvers, P.R.,
and Viele, G.W., 1989, Ouachita system—cross sections, in
Hatcher, R.D., Jr., Thomas, W.A., and Viele, G.W., eds., The
Appalachian-Ouachita Orogen in the United States: Boulder,
Colorado, Geological Society of America, The Geology of
North America, v. F-2, Plate 12.
Becker, T.P., Thomas, W.A., Samson, S.D., and Gehrels, G.E., 2005,
Detrital zircon evidence of Laurentian crustal dominance in
the lower Pennsylvanian deposits of the Alleghanian clastic
wedge in eastern North America: Sedimentary Geology,
v. 182, p. 59–86, https://​doi​.org​/10​.1016​/j​.sedgeo​.2005​.07​.014.
Becker, T.P., Thomas, W.A., and Gehrels, G.E., 2006, Linking late
Paleozoic sedimentary provenance in the Appalachian basin
to the history of Alleghanian deformation: American Journal of Science, v. 306, p. 777–798, https://​doi​.org​/10​.2475​
/10​.2006​.01.
Bouvier, A., Vervoort, J.D., and Patchett, J.D., 2008, The Lu-Hf
and Sm-Nd isotopic composition of CHUR: Constraints
from unequilibrated chondrites and implications for the
bulk composition of terrestrial planets: Earth and Planetary Science Letters, v. 273, p. 48–57, https://​doi​.org​/10​.1016​
/j​.epsl​.2008​.06​.010.
Bowring, S., 1984, U-Pb zircon ages of granitic boulders in the
Ordovician Blakely Sandstone, Arkansas, and implications
for their provenance, in Stone, C.G., and Haley, B.R., eds.,
A Guidebook to the Geology of the Central and Southern
Ouachita Mountains, Arkansas: Arkansas Geological Commission Guidebook 84–2, p. 123.
Brown, L.F., Jr., Cleaves, A.W., II, and Erxleben, A.W., 1973, Pennsylvanian Depositional Systems in North-Central Texas:
Texas Bureau of Economic Geology Guidebook 14, 122 p.
Cameron, K.L., López, R., Ortega-Gutiérrez, F., Solari, L.A., Keppie,
J.D., and Schulze, C., 2004, U-Pb geochronology and Pb
isotope compositions of leached feldspars: Constraints on
the origin and evolution of Grenvillian rocks from eastern
and southern Mexico, in Tollo, R.P., Corriveau, L., McLelland,
J., and Bartholomew, M.J., eds., Proterozoic Tectonic Evolution of the Grenville Orogen in North America: Geological
Society of America Memoir 197, p. 755–769, https://​doi​.org​
/10​.1130​/0​-8137​-1197​-5​.755.
Cardona, A., Chew, D., Valencia, V.A., Bayona, G., Miskovic, A.,
and Ibañez-Mejía, M., 2010, Grenvillian remnants in the
Northern Andes: Rodinian and Phanerozoic paleogeographic
perspectives: Journal of South American Earth Sciences,
v. 29, p. 92–104, https://​doi​.org​/10​.1016​/j​.jsames​.2009​.07​.011.
Cecil, R., Gehrels, G., Patchett, J., and Ducea, M., 2011, U-Pb-Hf
characterization of the central Coast Mountains batholith:
Implications for petrogenesis and crustal architecture:
Lithosphere, v. 3, p. 247–260, https://​doi​.org​/10​.1130​/L134​.1.
Centeno-García, E., 2005, Review of Upper Paleozoic and Lower
Mesozoic stratigraphy and depositional environments of
central and west Mexico: Constraints on terrane analysis
and paleogeography, in Anderson, T.H., Nourse, J.A., McKee,
J.W., and Steiner, M.B., eds., The Mojave-Sonora Megashear
Hypothesis: Development, Assessment, and Alternatives:
Geological Society of America Special Paper 393, p. 233–258,
https://​doi​.org​/10​.1130​/0​-8137​-2393​-0​.233.
Chapman, A.D., and Laskowski, A.K., 2019, Detrital zircon U-Pb
data reveal a Mississippian sediment dispersal network
Thomas et al. | Detrital zircons and late Paleozoic sediment dispersal, Ouachita orogen
1244
Research Paper
originating in the Appalachian orogen, traversing North
America along its southern shelf, and reaching as far as
the southwest United States: Lithosphere, v. 11, p. 581–587,
https://​doi​.org​/10​.1130​/L1068​.1.
Clift, P.D., Heinrich, P., Dunn, D., Jacobus, A., and Blusztajn, J.,
2018, The Sabine block, Gulf of Mexico: Promontory on the
North American margin?: Geology, v. 46, p. 15–18, https://​
doi​.org​/10​.1130​/G39592​.1.
Cohen, K.M., Finney, S.C., Gibbard, P.L., and Fan, J.-X., 2013,
updated, The ICS [International Commission on Stratigraphy] International Chronostratigraphic Chart: Episodes, v. 36,
p. 199–204, https://​doi​.org​/10​.18814​/epiiugs​/2013​/v36i3​/002.
Cordani, U.G., and Teixeira, W., 2007, Proterozoic accretionary
belts in the Amazonian craton, in Hatcher, R.D., Jr., Carlson, M.P., McBride, J.H., and Martínez Catalán, J.R., eds.,
4-D Framework of Continental Crust: Geological Society of
America Memoir 200, p. 297–320, https://​doi​.org​/10​.1130​
/2007​.1200​(14).
Crosby, E.J., and Mapel, W.J., 1975, Central and west Texas, in
McKee, E.D., and Crosby, E.J., coordinators, Paleotectonic
Investigations of the Pennsylvanian System in the United
States, Part I. Introduction and Regional Analyses of the
Pennsylvanian System: U.S. Geological Survey Professional
Paper 853, p. 197–232.
Culotta, R., Latham, T., Sydow, M., Oliver, J., Brown, L., and
Kaufman, S., 1992, Deep structure of the Texas Gulf passive
margin and its Ouachita-Precambrian basement: Results of
the COCORP San Marcos arch survey: American Association
of Petroleum Geologists Bulletin, v. 76, p. 270–283.
Denison, R.E., 1982, Geologic Cross Section from the Arbuckle
Mountains to the Muenster Arch, Southern Oklahoma and
Texas: Geological Society of America Map and Chart Series
MC-28R, 8 p.
Denison, R.E., 1989, Foreland structure adjacent to the Ouachita
foldbelt, in Hatcher, R.D., Jr., Thomas, W.A., and Viele, G.W.,
eds., The Appalachian-Ouachita Orogen in the United
States: Boulder, Colorado, Geological Society of America,
The Geology of North America, v. F-2, p. 681–688, https://​
doi​.org​/10​.1130​/DNAG​-GNA​-F2​.681.
Dickinson, W.R., and Lawton, T.F., 2001, Carboniferous to Cretaceous assembly and fragmentation of Mexico: Geological
Society of America Bulletin, v. 113, p. 1142–1160, https://​
doi​.org​/10​.1130​/0016-​ 7606(​ 2001)113<
​ 1142:C
​ TCAAF>2.​ 0.​ CO;2.
Dodson, S.A., 2008, Petrographic and Geochronologic Provenance Analysis of Upper Pennsylvanian Fluvial Sandstones
of the Conemaugh and Monongahela Groups, Athens County,
Ohio [M.S. thesis]: Athens, Ohio, Ohio University, 86 p.
Dunn, D.P., 2009, Arkansas crustal xenoliths: Implications for
basement rocks of the northern Gulf Coast, USA: Lithosphere, v. 1, p. 60–64, https://​doi​.org​/10​.1130​/L10​.1.
Eriksson, K.A., Campbell, I.H., Palin, J.M., Allen, C.M., and Bock,
B., 2004, Evidence for multiple recycling in Neoproterozoic
through Pennsylvanian sedimentary rocks of the central
Appalachian basin: The Journal of Geology, v. 112, p. 261–
276, https://​doi​.org​/10​.1086​/382758.
Flawn, P.T., Goldstein, A., Jr., King, P.B., and Weaver, C.E., 1961,
The Ouachita System: University of Texas, Bureau of Economic Geology Publication 6120, 401 p.
Galloway, W.E., and Brown, L.F., Jr., 1973, Depositional
systems and shelf-slope relations on cratonic basin margin, uppermost Pennsylvanian of north-central Texas:
GEOSPHERE | Volume 17 | Number 4
Downloaded from http://pubs.geoscienceworld.org/gsa/geosphere/article-pdf/17/4/1214/5362417/1214.pdf
by Idaho State University user
American Association of Petroleum Geologists Bulletin,
v. 57, p. 1185–1218.
Garner, H.F., 1967, Moorefield–Batesville stratigraphy and sedimentation in Arkansas: Geological Society of America
Bulletin, v. 78, p. 1233–1245, https://​doi​.org​/10​.1130​/0016​
-7606​(1967)78​[1233:​MSASIA]2​.0​.CO;2.
Gehrels, G.E., 2000, Introduction to detrital zircon studies of
Paleozoic and Triassic strata in western Nevada and northern California, in Soreghan, M.J., and Gehrels, G.E., eds.,
Paleozoic and Triassic Paleogeography and Tectonics of
Western Nevada and Northern California: Geological Society of America Special Paper 347, p. 1–17, https://​doi​.org​/10​
.1130​/0​-8137​-2347​-7​.1.
Gehrels, G.E., and Pecha, M.E., 2014, Detrital zircon U-Pb geochronology and Hf isotope geochemistry of Paleozoic and Triassic
passive margin strata of western North America: Geosphere,
v. 10, p. 49–65, https://​doi​.org​/10​.1130​/GES00889​.1.
Gehrels, G.E., Valencia, V., and Ruiz, J., 2008, Enhanced precision,
accuracy, efficiency, and spatial resolution of U-Pb ages by
laser ablation–multicollector–inductively coupled plasma–
mass spectrometry: Geochemistry Geophysics Geosystems,
v. 9, Q03017, https://​doi​.org​/10​.1029​/2007GC001805.
Gleason, J.D., Finney, S.C., and Gehrels, G.E., 2002, Paleotectonic implications of a Mid- to Late-Ordovician provenance
shift, as recorded in sedimentary strata of the Ouachita and
southern Appalachian Mountains: The Journal of Geology,
v. 110, p. 291–304, https://​doi​.org​/10​.1086​/339533.
Gleason, J.D., Gehrels, G.E., Dickinson, W.R., Patchett, P.J.,
and Kring, D.A., 2007, Sources of detrital zircon grains in
turbidite and deltaic sandstones of the Lower to Middle
Pennsylvanian Haymond Formation, Marathon assemblage,
Texas: Journal of Sedimentary Research, v. 77, p. 888–900,
https://​doi​.org​/10​.2110​/jsr​.2007​.084.
Glick, E.E., 1979, Arkansas, in Craig, L.C., and Connor, C.W.,
coordinators, Paleotectonic Investigations of the Mississippian System in the United States; Part I. Introduction
and Regional Analyses of the Mississippian System: U.S.
Geological Survey Professional Paper 1010, p. 125–145.
Gradstein, F.M., Ogg, J.G., and Smith, A.G., 2004, A Geologic
Time Scale 2004: Cambridge, UK, Cambridge University
Press, 589 p., https://​doi​.org​/10​.4095​/215638.
Graham, S.A., Ingersoll, R.V., and Dickinson, W.R., 1976, Common
provenance for lithic grains in Carboniferous sandstones
from Ouachita Mountains and Black Warrior basin: Journal
of Sedimentary Petrology, v. 46, p. 620–632.
Gray, M.B., and Zeitler, P.K., 1997, Comparison of clastic wedge
provenance in the Appalachian foreland using U/Pb ages
of detrital zircons: Tectonics, v. 16, p. 151–160, https://​doi​
.org​/10​.1029​/96TC02911.
Grimm, R.P., Eriksson, K., and Carbaugh, J., 2013, Tectono-sedimentary evolution of Early Pennsylvanian alluvial systems
at the onset of the Alleghanian orogeny, Pocahontas basin,
Virginia: Basin Research, v. 25, p. 450–470, https://​doi​.org​
/10​.1111​/bre​.12008.
Handford, C.R., 1986, Facies and bedding sequences in shelfstorm–deposited carbonates—Fayetteville Shale and Pitkin
Limestone (Mississippian), Arkansas: Journal of Sedimentary Petrology, v. 56, p. 123–137, https://​doi​.org​/10​.1306​
/212F88A0​-2B24​-11D7​-8648000102C1865D.
Handford, C.R., 1995, Baselap patterns and the recognition of
lowstand exposure and drowning—A Mississippian-ramp
example and its seismic signature: Journal of Sedimentary
Research, v. B65, p. 323–337.
Hanson, R.E., Puckett, R.E., Jr., Keller, G.R., Brueseke, M.E.,
Bulen, C.L., Mertzman, S.A., Finegan, S.A., and McCleery,
D.A., 2013, Intraplate magmatism related to opening of
the southern Iapetus Ocean: Cambrian Wichita igneous
province in the Southern Oklahoma rift zone: Lithos, v. 174,
p. 57–70, https://​doi​.org​/10​.1016​/j​.lithos​.2012​.06​.003.
Harry, D.L., Londono, J., and Huerta, A., 2003, Early Paleozoic
transform-margin structure beneath the Mississippi coastal
plain, southeast United States: Geology, v. 31, p. 969–972,
https://​doi​.org​/10​.1130​/G19787​.1.
Hatcher, R.D., Jr., 2010, The Appalachian orogen: A brief summary,
in Tollo, R.P., Bartholomew, M.J., Hibbard, J.P., and Karabinos,
P.M., eds., From Rodinia to Pangea: The Lithotectonic Record
of the Appalachian Region: Geological Society of America
Memoir 206, p. 1–19, https://doi.org/10.1130/2010.1206(01).
Hibbard, J.P., van Staal, C.R., and Miller, B.V., 2007, Links among
Carolinia, Avalonia, and Ganderia in the Appalachian
peri-Gondwanan realm, in Sears, J.W., Harms, T.A., and
Evenchick, C.A., eds., Whence the Mountains? Inquiries into
the Evolution of Orogenic Systems: A Volume in Honor of
Raymond A. Price: Geological Society of America Special
Paper 433, p. 291–311, https://​doi​.org​/10​.1130​/2007​.2433​(14).
Hogan, J.P., and Gilbert, M.C., 1998, The Southern Oklahoma
aulacogen: A Cambrian analog for mid-Proterozoic AMCG
(anorthosite-mangerite-charnockite-granite) complexes?, in
Hogan, J.P., and Gilbert, M.C., eds., Central North America
and Other Regions: Proceedings of the Twelfth International
Conference on Basement Tectonics: Dordrecht, Netherlands,
Kluwer, p. 39–78.
Houseknecht, D.W., 1986, Evolution from passive margin to
foreland basin: The Atoka Formation of the Arkoma basin,
south-central U.S.A., in Allen, P.A., and Homewood, P., eds.,
Foreland Basins: International Association of Sedimentologists Special Publication 8, p. 327–345.
Johnson, K.S., Amsden, T.W., Denison, R.E., Dutton, S.P., Goldstein, A.G., Rascoe, B., Jr., Sutherland, P.K., and Thompson,
D.M., 1988, Southern Midcontinent region, in Sloss, L.L., ed.,
Sedimentary Cover—North American Craton: U.S.: Boulder,
Colorado, Geological Society of America, The Geology of
North America, v. D-2, p. 307–359, https://​doi​.org​/10​.1130​
/DNAG​-GNA​-D2​.307.
Jusczuk, S.J., 2002, How Do the Structures of the Late Paleozoic
Ouachita Thrust Belt Relate to the Structures of the Southern Oklahoma Aulacogen [Ph.D. dissertation]: Lexington,
Kentucky, University of Kentucky, 339 p.
Keller, G.R., Braile, L.W., McMechan, G.A., Thomas, W.A., Harder,
S.H., Chang, W.-F., and Jardine, W.G., 1989, Paleozoic continent-ocean transition in the Ouachita Mountains imaged
from PASSCAL wide-angle seismic reflection-refraction
data: Geology, v. 17, p. 119–122, https://d
​ oi.​ org/​ 10.​ 1130/​ 0091​
-7613​(1989)017​<0119:​PCOTIT>2​.3​.CO;2.
Kier, R.S., Brown, L.F., Jr., and McBride, E.F., 1979, The Mississippian and Pennsylvanian (Carboniferous) Systems in the
United States—Texas: U.S. Geological Survey Professional
Paper 1110-S, p. S1–S45.
Kissock, J.K., Finzel, E.S., Malone, D.H., and Craddock, J.P.,
2018, Lower–​Middle Pennsylvanian strata in the North
American midcontinent record the interplay between erosional unroofing of the Appalachians and eustatic sea-level
Thomas et al. | Detrital zircons and late Paleozoic sediment dispersal, Ouachita orogen
1245
Research Paper
rise: Geosphere, v. 14, p. 141–161, https://​doi​.org​/10​.1130​
/GES01512​.1.
Konstantinou, A., Wirth, K.R., Vervoort, J.D., Malone, D.H., Davidson, C., and Craddock, J.P., 2014, Provenance of quartz arenites
of the early Paleozoic midcontinent region, USA: The Journal
of Geology, v. 122, p. 201–216, https://​doi​.org​/10​.1086​/675327.
Krogh, T.E., Kamo, S.L., Sharpton, V.L., Marin, L.E., and Hilde­
brand, A.R., 1993, U-Pb ages of single shocked zircons
linking distal K/T ejecta to the Chicxulub crater: Nature,
v. 366, p. 731–734, https://​doi​.org​/10​.1038​/366731a0.
Lawlor, P.J., Ortega-Gutierrez, F., Cameron, K.L., Ochoa-Camarillo,
H., Lopez, R., and Sampson, D.E., 1999, U-Pb geochronology,
geochemistry, and provenance of the Grenvillian Huiznopala Gneiss of eastern Mexico: Precambrian Research, v. 94,
p. 73–99, https://​doi​.org​/10​.1016​/S0301​-9268​(98)00108​-9.
Lawton, T.F., and Molina-Garza, R.S., 2014, U-Pb geochronology of the type Nazas Formation and superjacent strata,
northeastern Durango, Mexico: Implications of a Jurassic
age for continental-arc magmatism in north-central Mexico:
Geological Society of America Bulletin, v. 126, p. 1181–1199,
https://​doi​.org​/10​.1130​/B30827​.1.
Lillie, R.J., Nelson, K.D., de Voogd, B., Brewer, J.A., Oliver, J.E.,
Brown, L.D., Kaufman, S., and Viele, G.W., 1983, Crustal
structure of Ouachita Mountains, Arkansas: A model based
on integration of COCORP reflection profiles and regional
geophysical data: American Association of Petroleum Geologists Bulletin, v. 67, p. 907–931.
Liu, L., and Stockli, D.F., 2019, U-Pb ages of detrital zircons in
lower Permian sandstone and siltstone of the Permian Basin,
west Texas, USA: Evidence of dominant Gondwanan and
peri-Gondwanan sediment input to Laurentia: Geological
Society of America Bulletin, v. 132, p. 245–262, https://​doi​
.org​/10​.1130​/B35119​.1.
Loomis, J., Weaver, B., and Blatt, H., 1994, Geochemistry of
Mississippian tuffs from the Ouachita Mountains, and implications for the tectonics of the Ouachita orogen, Oklahoma
and Arkansas: Geological Society of America Bulletin, v. 106,
p. 1158–1171, https://​doi​.org​/10​.1130​/0016​-7606​(1994)106​
<1158:​GOMTFT>2​.3​.CO;2.
Lopez, R., 1997, High-Mg Andesites from the Gila Bend Mountains, Southwestern Arizona: Evidence for Hydrous Melting
of Lithosphere during Miocene Extension and the Pre-Jurassic Geotectonic Evolution of the Coahuila Terrane,
Northwestern Mexico: Grenville Basement, a Late Paleozoic Arc, Triassic Plutonism, and the Events South of the
Ouachita Suture [Ph.D. thesis]: Santa Cruz, California, University of California–Santa Cruz, 147 p.
Lopez, R., Cameron, K.L., and Jones, N.W., 2001, Evidence for
Paleoproterozoic, Grenvillian, and Pan-African age Gondwanan crust beneath northeastern Mexico: Precambrian
Research, v. 107, p. 195–214, https://​doi​.org​/10​.1016​/S0301​
-9268​(00)00140​-6.
Mack, G.H., Thomas, W.A., and Horsey, C.A., 1983, Composition
of Carboniferous sandstones and tectonic framework of
southern Appalachian-Ouachita orogen: Journal of Sedimentary Petrology, v. 53, p. 931–946.
Mapel, W.J., Johnson, R.B., Bachman, G.O., and Varnes, K.L.,
1979, Southern Midcontinent and Southern Rocky Mountains
Region: Paleotectonic Investigations of the Mississippian
System in the United States: U.S. Geological Survey Professional Paper 1010-J, p. 161–187.
GEOSPHERE | Volume 17 | Number 4
Downloaded from http://pubs.geoscienceworld.org/gsa/geosphere/article-pdf/17/4/1214/5362417/1214.pdf
by Idaho State University user
Martens, U., Weber, B., and Valencia, V.A., 2010, U/Pb geochronology of Devonian and older Paleozoic beds in the
southeastern Maya block, Central America: Its affinity with
peri-Gondwanan terranes: Geological Society of America
Bulletin, v. 122, p. 815–829, https://d
​ oi.​ org/​ 10.​ 1130/​ B26405.​ 1.
McGuire, P.R., 2017, U-Pb Detrital Zircon Signature of the
Ouachita Orogenic Belt [M.S. thesis]: Fort Worth, Texas,
Texas Christian University, 78 p.
Mickus, K.L., and Keller, G.R., 1992, Lithospheric structure of the
south-central United States: Geology, v. 20, p. 335–338, https://​
doi​.org​/10​.1130​/0091​-7613​(1992)020​<0335:​LSOTSC>2​.3​.CO;2.
Morris, E.M., and Stone, C.G., 1986, A preliminary report on the
metagabbros of the Ouachita core, in Stone, C.G., Howard,
J.M., and Haley, B.R., eds., Sedimentary and Igneous Rocks
of the Ouachita Mountains of Arkansas, Part 1: Arkansas
Geological Commission Guidebook 86–3, p. 87–90.
Morris, R.C., 1989, Stratigraphy and sedimentary history of post–
Arkansas Novaculite Carboniferous rocks of the Ouachita
Mountains, in Hatcher, R.D., Jr., Thomas, W.A., and Viele,
G.W., eds., The Appalachian-Ouachita Orogen in the United
States: Boulder, Colorado, Geological Society of America,
The Geology of North America, v. F-2, p. 591–602, https://​
doi​.org​/10​.1130​/DNAG​-GNA​-F2​.591.
Mueller, P.A., Heatherington, A.L., Foster, D.A., Thomas, W.A.,
and Wooden, J.L., 2014, The Suwannee suture: Significance
for Gondwana-Laurentia terrane transfer and formation of
Pangaea: Gondwana Research, v. 26, p. 365–373, https://d
​ oi​
.org​/10​.1016​/j​.gr​.2013​.06​.018.
Nelson, K.D., Lillie, R.J., de Voogd, B., Brewer, J.A., Oliver, J.E.,
Kaufman, S., Brown, L.D., and Viele, G.W., 1982, COCORP seismic reflection profiling in the Ouachita Mountains of western
Arkansas: Geometry and geologic interpretation: Tectonics,
v. 1, p. 413–430, https://​doi​.org​/10​.1029​/TC001i005p00413.
Nicholas, R.L., and Rozendal, R.A., 1975, Subsurface positive
elements within Ouachita foldbelt in Texas and their relation to Paleozoic cratonic margin: American Association of
Petroleum Geologists Bulletin, v. 59, p. 193–216.
Nicholas, R.L., and Waddell, D.E., 1989, The Ouachita system
in the subsurface of Texas, Arkansas, and Louisiana, in
Hatcher, R.D., Jr., Thomas, W.A., and Viele, G.W., eds., The
Appalachian-Ouachita Orogen in the United States: Boulder,
Colorado, Geological Society of America, The Geology of
North America, v. F-2, p. 661–672, https://doi.org/10.1130​
/DNAG-GNA-F2.661.
Nielsen, K.C., Viele, G.W., and Zimmerman, J., 1989, Structural setting of the Benton-Broken Bow uplifts, in Hatcher,
R.D., Jr., Thomas, W.A., and Viele, G.W., eds., The Appalachian-Ouachita Orogen in the United States: Boulder, Colorado,
Geological Society of America, The Geology of North America,
v. F-2, p. 635–660, https://​doi​.org​/10​.1130​/DNAG​-GNA​-F2​.635.
Niem, A.R., 1977, Mississippian pyroclastic flow and ash-fall
deposits in the deep-marine Ouachita flysch basin, Oklahoma and Arkansas: Geological Society of America Bulletin,
v. 88, p. 49–61, https://​doi​.org​/10​.1130​/0016​-7606​(1977)88​
<49:​MPFAAD>2​.0​.CO;2.
Ogren, D.E., 1968, Stratigraphy of Upper Mississippian rocks
of northern Arkansas: American Association of Petroleum
Geologists Bulletin, v. 52, p. 282–294.
Ortega-Gutierrez, F., Ruiz, J., and Centeno-Garcia, E., 1995,
Oaxaquia, a Proterozoic microcontinent accreted to
North America during the late Paleozoic: Geology, v. 23,
Thomas et al. | Detrital zircons and late Paleozoic sediment dispersal, Ouachita orogen
p. 1127–​1130, https://​doi​.org​/10​.1130​/0091​-7613​(1995)023​
<1127:​OAPMAT>2​.3​.CO;2.
Park, H., Barbeau, D.L., Jr., Rickenbaker, A., Bachmann-Krug,
D., and Gehrels, G., 2010, Application of foreland basin
detrital-zircon geochronology to the reconstruction of the
southern and central Appalachian orogen: The Journal of
Geology, v. 118, p. 23–44, https://​doi​.org​/10​.1086​/648400.
Perry, W.J., Jr., 1989, Tectonic Evolution of the Anadarko Basin
Region, Oklahoma: U.S. Geological Survey Bulletin 1866A,
19 p.
Pickell, J.M., 2012, Detrital Zircon Geochronology of Middle
Ordovician Siliciclastic Sediment on the Southern Laurentian Shelf [M.S. thesis]: College Station, Texas, Texas A&M
University, 120 p.
Pindell, J.L., and Kennan, L., 2009, Tectonic evolution of the Gulf
of Mexico, Caribbean and northern South America in the
mantle reference frame: An update, in James, K.H., Lorente,
M.A., and Pindell, J.L., eds., The Origin and Evolution of
the Caribbean Plate: Geological Society [London] Special
Publication 328, p. 1–55, https://​doi​.org​/10​.1144​/SP328​.1.
Poland, Z.A., and Simms, A.R., 2012, Sedimentology of an erg
to an erg-margin depositional system, the Rush Springs
Sandstone of western Oklahoma, U.S.A.: Implications for
paleowinds across northwestern Pangea during the Guadalupian (Middle Permian): Journal of Sedimentary Research,
v. 82, p. 345–363, https://​doi​.org​/10​.2110​/jsr​.2012​.32.
Poole, F.G., Perry, W.J., Jr., Madrid, R.J., and Amaya-Martínez, R.,
2005, Tectonic synthesis of the Ouachita-Marathon-Sonora
orogenic margin of southern Laurentia: Stratigraphic and
structural implications for timing of deformational events
and plate-tectonic model, in Anderson, T.H., Nourse, J.A.,
McKee, J.W., and Steiner, M.B., eds., The Mohave-Sonora
Megashear Hypothesis: Development, Assessment, and
Alternatives: Geological Society of America Special Paper
393, p. 543–596, https://​doi​.org​/10​.1130​/0​-8137​-2393​-0​.543.
Ross, C.A., 1986, Paleozoic evolution of southern margin of
Permian basin: Geological Society of America Bulletin, v. 97,
p. 536–554, https://​doi​.org​/10​.1130​/0016​-7606​(1986)97​<536:​
PEOSMO>2​.0​.CO;2.
Rozendal, R.A., and Erskine, W.S., 1971, Deep test in the Ouachita
structural belt of central Texas: American Association of
Petroleum Geologists Bulletin, v. 55, p. 2008–2017.
Saylor, J.E., Jordan, J.C., Sundell, K.E., Wang, X., Wang, S., and
Deng, T., 2018, Topographic growth of the Jishi Shan and
its impact on basin and hydrology evolution, NE Tibetan
Plateau: Basin Research, v. 30, no. 3, p. 544–563, https://​doi​
.org​/10​.1111​/bre​.12264.
Schuchert, C., 1923, Sites and nature of the North American
geosynclines [1922 Geological Society of America Presidential Address]: Geological Society of America Bulletin, v. 34,
p. 151–229, https://​doi​.org​/10​.1130​/GSAB​-34​-151.
Sharrah, K.L., 2006, Comparative Study of the Sedimentology
and Provenance of the Atoka Formation in the Frontal
Ouachita Thrust Belt, Oklahoma [Ph.D. thesis]: Tulsa, Oklahoma, University of Tulsa, 252 p.
Shaulis, B.J., Lapen, T.J., Casey, J.F., and Reid, D.R., 2012, Timing and rates of flysch sedimentation in the Stanley Group,
Ouachita Mountains, Oklahoma and Arkansas, U.S.A.: Constraints from U-Pb zircon ages of subaqueous ash-flow tuffs:
Journal of Sedimentary Research, v. 82, p. 833–840, https://​
doi​.org​/10​.2110​/jsr​.2012​.68.
1246
Research Paper
Silverman, B.W., 1986, Density Estimation for Statistics and Data
Analysis: London, Chapman and Hall, 175 p., https://​doi​.org​
/10​.1007​/978​-1​-4899​-3324​-9.
Soreghan, G.S., and Soreghan, M.J., 2013, Tracing clastic delivery to the Permian Delaware basin, U.S.A.: Implications for
paleogeography and circulation in westernmost equatorial
Pangea: Journal of Sedimentary Research, v. 83, p. 786–802,
https://​doi​.org​/10​.2110​/jsr​.2013​.63.
Steiner, M.B., and Walker, J.D., 1996, Late Silurian plutons in
Yucatan: Journal of Geophysical Research, v. 101, p. 17727–
17735, https://​doi​.org​/10​.1029​/96JB00174.
Stewart, J.H., Blodgett, R.B., Boucot, A.J., Carter, J.L., and Lopez,
R., 1999, Exotic Paleozoic strata of Gondwanan provenance
near Ciudad Victoria, Tamaulipas, Mexico, in Ramos, V.A.,
and Keppie, J.D., eds., Laurentia–Gondwana Connections
before Pangea: Geological Society of America Special Paper
336, p. 227–252, https://​doi​.org​/10​.1130​/0​-8137​-2336​-1​.227.
Stone, C.G., and Haley, B.R., 1977, The occurrence and origin of
the granite–meta-arkose erratics in the Ordovician Blakely
Sandstone, Arkansas, in Stone, C.G., ed., Symposium on the
Geology of the Ouachita Mountains, Stratigraphy, Sedimentology, Petrography, Tectonics, and Paleontology: Little Rock,
Arkansas, Arkansas Geological Commission, v. 1, p. 107–111.
Sundell, K., Saylor, J.E., and Pecha, M., 2019, Provenance and
recycling of detrital zircons from Cenozoic Altiplano strata
and the crustal evolution of western South America from
combined U-Pb and Lu-Hf isotopic analysis, in Horton, B.,
and Folguera, A., eds., Andean Tectonics: Amsterdam, Netherlands, Elsevier, p. 363–397, https://​doi​.org​/10​.1016​/B978​-0​
-12​-816009​-1​.00014​-9.
Sutherland, P.K., 1988, Late Mississippian and Pennsylvanian
depositional history in the Arkoma basin area, Oklahoma
and Arkansas: Geological Society of America Bulletin, v. 100,
p. 1787–1802, https://​doi​.org​/10​.1130​/0016​-7606​(1988)100​
<1787:​LMAPDH>2​.3​.CO;2.
Thomas, W.A., 1977, Evolution of Appalachian-Ouachita salients
and recesses from reentrants and promontories in the
continental margin: American Journal of Science, v. 277,
p. 1233–1278, https://​doi​.org​/10​.2475​/ajs​.277​.10​.1233.
Thomas, W.A., 1989, The Appalachian-Ouachita orogen beneath
the Gulf Coastal Plain between the outcrops in the Appalachian and Ouachita Mountains, in Hatcher, R.D., Jr., Thomas,
W.A., and Viele, G.W., eds., The Appalachian-Ouachita Orogen in the United States: Boulder, Colorado, Geological
Society of America, The Geology of North America, v. F-2,
p. 537–553, https://​doi​.org​/10​.1130​/DNAG​-GNA​-F2​.537.
Thomas, W.A., 2006, Tectonic inheritance at a continental margin
[2005 Geological Society of America Presidential Address]:
GSA Today, v. 16, no. 2, p. 4–11, https://​doi​.org​/10​.1130​/1052​
-5173​(2006)016​[4:​TIAACM]2​.0​.CO;2.
Thomas, W.A., 2011, The Iapetan rifted margin of southern Laurentia: Geosphere, v. 7, p. 97–120, https://​doi​.org​/10​.1130​
/GES00574​.1.
Thomas, W.A., 2014, A mechanism for tectonic inheritance
at transform faults of the Iapetan margin of Laurentia:
GEOSPHERE | Volume 17 | Number 4
Downloaded from http://pubs.geoscienceworld.org/gsa/geosphere/article-pdf/17/4/1214/5362417/1214.pdf
by Idaho State University user
Geoscience Canada, v. 41, p. 321–344, https://​doi​.org​/10​
.12789​/geocanj​.2014​.41​.048.
Thomas, W.A., 2019, Tectonic inheritance at multiple scales
during more than two complete Wilson cycles recorded in
eastern North America, in Wilson, R.W., Houseman, G.A.,
McCaffrey, K.J.W., Doré, A.G., and Buiter, S.J.H., eds., Fifty
Years of the Wilson Cycle Concept in Plate Tectonics: Geological Society [London] Special Publication 470, p. 337–352,
https://​doi​.org​/10​.1144​/SP470​.4.
Thomas, W.A., and Astini, R.A., 1996, The Argentine Precordillera: A traveler from the Ouachita embayment of North
American Laurentia: Science, v. 273, p. 752–757, https://​doi​
.org​/10​.1126​/science​.273​.5276​.752.
Thomas, W.A., and Astini, R.A., 1999, Simple-shear conjugate
rift margins of the Argentine Precordillera and the Ouachita
embayment of Laurentia: Geological Society of America
Bulletin, v. 111, p. 1069–1079, https://​doi​.org​/10​.1130​/0016​
-7606​(1999)111​<1069:​SSCRMO>2​.3​.CO;2.
Thomas, W.A., Viele, G.W., Arbenz, J.K., Nicholas, R.L., Denison,
R.E., Muehlberger, W.R., and Tauvers, P.R., 1989, Tectonic
map of the Ouachita orogen, in Hatcher, R.D., Jr., Thomas,
W.A., and Viele, G.W., eds., The Appalachian-Ouachita Orogen in the United States: Boulder, Colorado, Geological
Society of America, The Geology of North America, v. F-2,
Plate 9.
Thomas, W.A., Becker, T.P., Samson, S.D., and Hamilton, M.A.,
2004, Detrital zircon evidence of a recycled orogenic foreland provenance for Alleghanian clastic-wedge sandstones:
The Journal of Geology, v. 112, p. 23–37, https://​doi​.org​/10​
.1086​/379690.
Thomas, W.A., Tucker, R.D., Astini, R.A., and Denison, R.E.,
2012, Ages of pre-rift basement and synrift rocks along
the conjugate rift and transform margins of the Argentine
Precordillera and Laurentia: Geosphere, v. 8, p. 1366–1383,
https://​doi​.org​/10​.1130​/GES00800​.1.
Thomas, W.A., Gehrels, G.E., and Romero, M.C., 2016, Detrital zircons from crystalline rocks along the Southern
Oklahoma fault system, Wichita and Arbuckle Mountains,
USA: Geosphere, v. 12, p. 1224–1234, https://d
​ oi.​ org/​ 10.​ 1130​
/GES01316​.1.
Thomas, W.A., Gehrels, G.E., Greb, S.F., Nadon, G.C., Satkoski,
A.M., and Romero, M.C., 2017, Detrital zircons and sediment
dispersal in the Appalachian foreland: Geosphere, v. 13,
p. 2206–2230, https://​doi​.org​/10​.1130​/GES01525​.1.
Thomas, W.A., Gehrels, G.E., Lawton, T.F., Satterfield, J.I.,
Romero, M.C., and Sundell, K.E., 2019, Detrital zircons and
sediment dispersal from the Coahuila terrane of northern
Mexico into the Marathon foreland of the southern Midcontinent: Geosphere, v. 15, p. 1102–1127, https://​doi​.org​
/10​.1130​/GES02033​.1.
Thomas, W.A., Gehrels, G.E., Sundell, K.E., Greb, S.F., Finzel,
E.S., Clark, R.J., Malone, D.H., Hampton, B.A., and Romero,
M.C., 2020, Detrital zircons and sediment dispersal in the
eastern Midcontinent of North America: Geosphere, v. 16,
p. 817–843, https://​doi​.org​/10​.1130​/GES02152​.1.
Thompson, D.M., 1982, The Atoka Group (Lower-Middle Pennsylvanian), Northern Fort Worth Basin, Texas; Terrigenous
Depositional Systems, Diagenesis, Reservoir Distribution,
and Quality: University of Texas at Austin, Bureau of Economic Geology Report of Investigations 125, 62 p.
Van Avendonk, H.J.A., and Dalziel, I.W.D., 2018, Comment: The
Sabine block, Gulf of Mexico: Promontory on the North
American margin?: Geology, v. 46, p. e440, https://​doi​.org​
/10​.1130​/G40053C​.1.
Van Schmus, W.R., et al., 1993, Transcontinental Proterozoic
provinces, in Reed, J.C., Jr., Bickford, M.E., Houston, R.S.,
Link, P.K., Rankin, D.W., Sims, P.K., and Van Schmus, W.R.,
eds., Precambrian: Conterminous U.S.: Boulder, Colorado,
Geological Society of America, The Geology of North America, v. C-2, p. 171–334, https://​doi​.org​/10​.1130​/DNAG​-GNA​
-C2​.171.
Vermeesch, P., 2013, Multi-sample comparison of detrital age
distributions: Chemical Geology, v. 341, p. 140–146, https://​
doi​.org​/10​.1016​/j​.chemgeo​.2013​.01​.010.
Vervoort, J.D., and Patchett, P.J., 1996, Behavior of hafnium
and neodymium isotopes in the crust: Constraints from
crustally derived granites: Geochimica et Cosmochimica
Acta, v. 60, no. 19, p. 3717–3733, https://d
​ oi.​ org/​ 10.​ 1016/​ 0016​
-7037​(96)00201​-3.
Vervoort, J.D., Patchett, P.J., Blichert-Toft, J., and Albarede, F.,
1999, Relationships between Lu-Hf and Sm-Nd isotopic
systems in the global sedimentary system: Earth and Planetary Science Letters, v. 168, p. 79–99, https://​doi​.org​/10​.1016​
/S0012​-821X​(99)00047​-3.
Viele, G.W., and Thomas, W.A., 1989, Tectonic synthesis of the
Ouachita orogenic belt, in Hatcher, R.D., Jr., Thomas, W.A.,
and Viele, G.W., eds., The Appalachian-Ouachita Orogen in
the United States: Boulder, Colorado, Geological Society of
America, The Geology of North America, v. F-2, p. 695–728,
https://​doi​.org​/10​.1130​/DNAG​-GNA​-F2​.695.
Wright, J.E., Hogan, J.P., and Gilbert, M.C., 1996, The Southern
Oklahoma aulacogen: Not just another B.L.I.P.: Eos (Washington, D.C.), v. 77, no. 46, p. F845.
Xie, X., Cains, W., and Manger, W.L., 2016, U-Pb detrital zircon
evidence of transcontinental sediment dispersal: Provenance of Late Mississippian Wedington Sandstone member,
NW Arkansas: International Geology Review, v. 58, no. 15,
p. 1951–1966, https://d
​ oi.​ org/​ 10.​ 1080/​ 00206814.​ 2016.​ 1193775.
Xie, X., Buratowski, G., Manger, W.L., and Zachry, D., 2018a,
U-Pb detrital-zircon geochronology of the middle Bloyd
sandstone (Morrowan) of northern Arkansas (U.S.A.): Implications for Early Pennsylvanian sediment dispersal in the
Laurentian foreland: Journal of Sedimentary Research, v. 88,
p. 795–810, https://​doi​.org​/10​.2110​/jsr​.2018​.47.
Xie, X., Anthony, J.M., and Busbey, A.B., 2018b, Provenance of
Permian Delaware Mountain Group, central and southern
Delaware basin, and implications of sediment dispersal
pathway near the southwestern terminus of Pangea: International Geology Review, v. 61, no. 3, p. 361–380, https://​
doi​.org​/10​.1080​/00206814​.2018​.1425925.
Thomas et al. | Detrital zircons and late Paleozoic sediment dispersal, Ouachita orogen
1247
Download