Вы находитесь на странице: 1из 18

Palaeogeography, Palaeoclimatology, Palaeoecology 210 (2004) 367 – 384

www.elsevier.com/locate/palaeo

Cherty carbonate facies of the Montoya Group, southern New


Mexico and western Texas and its regional correlatives: a record of
Late Ordovician paleoceanography on southern Laurentia
Michael C. Pope *
Department of Geology, Washington State University, Pullman, WA 99164-2812, USA
Received 25 November 2002; accepted 23 February 2004

Abstract

The Upper Ordovician Montoya Group in southern New Mexico and westernmost Texas records predominantly subtidal
deposition on a gently dipping carbonate ramp that was subsequently nearly entirely dolomitized. The medial unit of the
Montoya Group, the Aleman Formation is unique because it contains abundant chert (10 – 70% by volume). The chert occurs as:
(1) thin continuous beds of sponge spicules within mudstone or calcisiltite; (2) discontinuous, lenses or nodules within skeletal
wackestones and packstones; or (3) as a replacement of skeletal grains and burrows. Coeval skeletal grainstones and muddy
peritidal facies contain little chert. Phosphate (up to 5 wt.%) occurs within the underlying Upham Formation and the Aleman
Formation as replacement of fossils and peloids. The abundance of chert and phosphate in these subtidal facies indicates they
formed within a region of strong upwelling. Regional correlation with Upper Ordovician cherty units along the periphery of
southern Laurentia and other low latitude continents suggests that upwelling was widespread and long-lived during the Late
Ordovician. The upwelling is interpreted to record vigorous oceanic circulation produced by the onset of glaciation on
Gondwana during this period. Late Ordovician relative sea-level curves around the periphery of Laurentia indicate correlative
third-order (1 – 3 my duration) fluctuations that may provide a means for high-resolution global correlations. However, the
mechanism(s) that produced these long-term fluctuations are unclear.
D 2004 Elsevier B.V. All rights reserved.

Keywords: Upwelling; Glaciation; Chert; Phosphate; Eustasy

1. Introduction house conditions in which atmospheric pCO2 was an


estimated 10 – 18 times greater than present (Berner,
Dropstones, diamictites, and striated pavements in 1994). Glaciation was likely triggered by the migra-
northern Africa record a Late Ordovician glacial event tion of Gondwana across the South Pole (Crowley and
(Crowell, 1999). This glaciation is enigmatic because Baum, 1991, 1995), possibly in conjunction with a
it formed during a prolonged period of global green- global decrease in atmospheric temperature driven by
intense silicate weathering of Taconic orogenic high-
lands (Kump et al., 1999). A 3– 7x positive excur-
* Fax: +1-509-335-5989. sion in both d13 C and d18 O stable isotopes of
E-mail address: mcpope@wsu.edu (M.C. Pope). carbonates, and a record of widespread sea-level

0031-0182/$ - see front matter D 2004 Elsevier B.V. All rights reserved.
doi:10.1016/j.palaeo.2004.02.035
368 M.C. Pope / Palaeogeography, Palaeoclimatology, Palaeoecology 210 (2004) 367–384

Fig. 1. Late Ordovician global map modified from Scotese (1997). L = Laurentia, B = Baltica, S = Siberia, and C = China. Montoya Group
locality marked by black circle. The location of cherty carbonates is approximate, see references in Pope and Steffen (2003). Paleoceanographic
currents modified from Pope and Read (1998).

Fig. 2. Location map of Montoya Group outcrops in southern New Mexico and western Texas. Generalized Montoya Group sequence
stratigraphic cross-section (Fig. 5) occurs along B – BV.
M.C. Pope / Palaeogeography, Palaeoclimatology, Palaeoecology 210 (2004) 367–384 369

drawdown from three continents during the Hirnan- Long-term relative sea-level fluctuations recorded in
tian Stage is interpreted as evidence for a short-lived, the Montoya Group are correlative with similar events
possibly < 1 million-year-long glaciation (Brenchley around Laurentia and may provide a means for high-
et al., 1994; Marshall et al., 1997). A longer glacial resolution stratigraphic correlation. The mechanism(s)
event is indicated by Late Middle – Late Ordovician that produced these long-term relative sea-level fluc-
glaciogenic sedimentary rocks in Africa (Barnes, tuations are unclear.
1986; Frakes et al., 1992; Theron, 1994; Crowell,
1999). However, the physical and paleontologic re-
cord for the duration of this glaciation is equivocal. 2. Regional stratigraphic framework
High-frequency, moderate amplitude (20 –30 m/20–
100 ky) sea-level fluctuations in the Upper Ordovician Upper Ordovician Montoya Group strata in New
Lexington Limestone of the Appalachian Basin were Mexico and western Texas are up to 180 m thick and
interpreted to be indirect proxy evidence for a pro- formed on a gently sloping carbonate ramp on the
longed Ordovician glaciation (Pope and Read, 1997a, Laurentian passive margin (Measures, 1985a,b; LeM-
1998). This paper discusses new sedimentologic and one, 1988). Paleomagnetic reconstructions indicate
stratigraphic data from the Upper Ordovician Mon- this passive margin (Fig. 1) was positioned between
toya Group of Texas and southern New Mexico and 5j and 20j south latitude, aligned approximately
their regional equivalents that indicate the southern east –west and faced the Panthalassic Ocean (Scotese,
margin of Laurentia was the locus of intensive up- 1997; Mac Niocaill et al., 1997).
welling. Enhanced thermohaline circulation caused by The Montoya Group outcrops in uplifted, block-
the Gondwana glaciation is postulated to have pro- faulted mountain ranges throughout southern New
duced this long-lived, widespread upwelling zone. Mexico and westernmost Texas that are surrounded

Fig. 3. Stratigraphic chart of Montoya Group and its correlatives along the southern margin of Laurentia. Biostratigraphic references for each
area given in the text. Dark gray stippling indicates cherty carbonate units. P indicates units containing abundant phosphate. The duration of
upwelling around North America is indicated by the extent of cherty carbonate. The Turinian – Chatfieldian boundary occurs at f 454 based on
U – Pb dating of the Diecke bentonite in the US Midcontinent (Tucker et al., 1990), whereas the Ordovician – Silurian boundary is estimated at
442 Ma (Tucker and McKerrow, 1995).
370 M.C. Pope / Palaeogeography, Palaeoclimatology, Palaeoecology 210 (2004) 367–384

by flat-floored valleys (Fig. 2). These mountains were The Montoya Group in southern New Mexico and
produced by Cenozoic basin and range extension and westernmost Texas is subdivided into three formations
they trend north – northwesterly. Forty-two full and (Fig. 3) given in ascending order, Upham, Aleman,
partial outcrop measured sections (Fig. 2) and over and Cutter. All of these units are extensively dolomi-
300 standard thin sections were used for analysis of tized with limestone occurring locally only in the
these rocks. Upham and basal Aleman. The Upham Formation
Extensive studies of the Montoya Group by strat- (Fig. 4) conformably overlies a basal, thin (up to 16 m
igraphers and paleontologists (e.g. Kelley and Silver, thick), calcareous-cemented sandstone and gravel
1952; Pray, 1958; Hill, 1959; Howe, 1959; Flower, conglomerate (Cable Canyon Sandstone) that rests
1961, 1969; Pratt and Jones, 1961; Kottlowski, 1963; unconformably above the Lower Ordovician El Paso
LeMone, 1969; Hayes, 1975; Sweet, 1979) produced Group. The Upham Formation typically is a light to
the gross regional lithostratigraphy (Fig. 3) that dark colored, bioturbated wackestone to grainstone
remains in use today. These studies indicate that the approximately 13 – 42 m thick. This unit has a grada-
Montoya Group was deposited during the Cincinna- tional contact with the underlying sandstone or lies
tian Series (Late Ordovician) and that the latest disconformably on the El Paso Group. The Aleman
Ordovician Hirnantian Stage was a time of erosion Formation (16 – 85 m thick) conformably overlies the
or non-deposition in this area (Fig. 3). Upham Formation and generally consists of a lower,

Fig. 4. Cross-section diagrams showing the distribution of facies across the Montoya Group carbonate ramp during the second-order TST and
HST. During the TST siliciclastics of the inner part of the ramp are progressively overlain by mid-ramp burrowed skeletal packstone or
grainstone then deep ramp interbedded calcisiltite or mudstone and spiculitic chert. The HST is a shallowing-upward succession characterized
by the initial deep ramp calcisiltite and mudstone interbedded with spiculitic chert grading upward into mid-ramp skeletal wackestone and
packstone then skeletal packstone and grainstone. Burrowed skeletal wackestone and mudstone, laminated mudstone and fenestral mudstone
were deposited updip, behind the skeletal grainstone and packstone.
M.C. Pope / Palaeogeography, Palaeoclimatology, Palaeoecology 210 (2004) 367–384 371

intercalated dark brown-gray, thin-bedded carbonate follows the hierarchy of Weber et al. (1995) for
and chert unit that grades up into nodular, cherty naming sequence stratigraphic depositional units.
dolostone (Fig. 4). The Cutter Formation (30 – 60 m The generalized sequence stratigraphic framework
thick), composed of light colored, fine-grained dolo- of the Montoya Group sequence relies on Line B –
mite with little chert, is the uppermost unit of the BV (Fig. 2). This generalized sequence stratigraphic
Montoya Group (Fig. 4). A regional unconformity analysis indicates the Montoya Group is a single
separates the Montoya Group from the overlying second-order (8 – 10 my duration) transgressive to
Silurian Fusselman Dolomite or younger units. regressive supersequence composed of six regionally
correlative third-order (1 –3 my duration) sequences
(Pope et al., 2001). The third-order sequences are
3. Generalized sequence stratigraphy numbered sequentially upward (0 –5) from the base.
Sequences 0 and 1 have quartz, sandy bases that are
The Montoya Group measured sections were ana- overlain by subtidal, mid- to shallow-ramp biotur-
lyzed using sequence stratigraphic principles outlined bated skeletal wackestone/packstone. Updip, these
in Harris et al. (1999). The Montoya Group was sequences are capped by high-energy skeletal grain-
deposited on a carbonate ramp (sensu Ahr, 1973) stone. Sequence 2 has chert-rich, deep ramp carbon-
because shallow to deep water facies pass laterally ate at its base and its upper part is marked by
into one another (Fig. 4) without evidence for a progradation of a high-energy shallow ramp skeletal
substantial change in depositional slope (Read, 1985). grainstone with abundant colonial corals. Muddy
A generalized outline of the Montoya Group peritidal carbonates were deposited behind the skel-
sequence stratigraphy (Fig. 5) is presented here to etal shoal. Sequence 3 is comprised of an aggrada-
facilitate the following discussions of: (1) the region- tional stack of skeletal packstone with colonial corals
al stratigraphic position and distribution of cherty that separates muddy inner ramp peritidal facies from
facies on this ramp; and (2) the correlative third-order deeper ramp cherty carbonates. Sequence 4 is marked
(1– 3 my) sea-level changes recorded in these units. by a pronounced basinward shift of peritidal facies
The sequence stratigraphic terminology in this paper into the basin. Sequence 5 is only preserved in

Fig. 5. Generalized sequence stratigraphy along Line B – BV from Fig. 2. The second-order systems tracts are separated by a regional maximum
flooding zone of cherty carbonate (Pope et al., 2001). The cherty carbonates of the Aleman Formation occur within third-order sequences 2 and 3.
372 M.C. Pope / Palaeogeography, Palaeoclimatology, Palaeoecology 210 (2004) 367–384

downdip locations and is comprised of coarse skel-


etal grainstone.

4. Cherty facies of the Aleman Formation

The cherty facies of the Aleman Formation are


shown graphically in Fig. 4. The Aleman Formation
is a complex subtidal carbonate unit containing
abundant chert (10 – 70% by volume). The Aleman
Formation commonly is subdivided into upper and
lower cherty units that are separated by a widespread
medial packstone/grainstone marker unit (Fig. 5).
The depositional environments represented by the
Aleman Formation range from shallow, high-energy
shoals to deep-water settings, below storm wave base
(Fig. 4).

4.1. Even-bedded laminated calcisiltite or mudstone


and spiculitic chert

Even-bedded laminated calcisiltite and spiculitic


chert (Fig. 6A) is the basal unit of the Aleman
Formation. The calcisiltite generally is horizontally
laminated but locally shows small hummocky cross-
lamination. The chert is composed almost entirely
(>90%) of sponge spicules that were cemented by
later silica or carbonate. Locally the spiculitic chert
is interbedded with massive mudstone and this facies
is most common in the base of the lower part of the
upper Aleman above the medial packstone/grain-
stone marker. The even-bedded calcisiltite or mud-
stone commonly contains 1 –5 wt.% phosphate as
pellets, with rare phosphatic sponge spicules occur- Fig. 6. (A) Photo of even-bedded calcisiltite and spiculitic chert,
ring in the interbedded spiculitic chert. Locally lower Aleman Formation, Cooks Range, New Mexico. Ten-
within the even-bedded calcisiltite and spiculitic centimeter black scale bar in lower right center of photo. (B) Photo
chert are beds 1 – 3 m thick of cross-laminated of nodular chert in burrowed skeletal wackestone and packstone,
north Franklin Mountains, Texas. (C) Silicified brachiopods in
spiculite up to 3 m thick.
skeletal packstone, Caballos Mountains, New Mexico. Five-
The even bedded calcisiltite or mudstone inter- centimeter scale bar.
bedded with spiculitic chert is interpreted to repre-
sent deposition in deep waters commonly below
storm wave base. The calcisiltite and mudstone higher frequency climate or oceanographic changes
likely represent the background sedimentation on (e.g. Elrick et al., 1991). The abundance of sponge
this ramp, whereas the spiculitic chert formed as the spicules and absence of any other fauna, save small
disarticulated sponge spicules were redistributed by brachiopods, indicates this facies may have formed
storms or currents and accumulated. It is unclear in cool or oxygen-poor waters (for example, James,
what caused the alternation of carbonate and silica, 1997). The hummocky beds within this facies
the periods of abundant silica deposition may record indicate that storm wave base did sometimes im-
M.C. Pope / Palaeogeography, Palaeoclimatology, Palaeoecology 210 (2004) 367–384 373

pinge upon the seafloor during deposition of this 4.3. Skeletal packstone/grainstone
facies.
Skeletal packstone/grainstone including common
4.2. Skeletal wackestone to packstone with irregular, crinoids, bryozoans, brachiopods and rare rugose
discontinuous bedded to nodular chert corals and stromatoporoids is interbedded with thin
beds of colonial tabulate coral bafflestone comprises
Skeletal wackestone to packstone containing irreg- widespread marker unit in the middle of the Aleman
ular and discontinuously bedded chert up to a few Formation (Howe, 1959). This unit developed be-
meters wide and a few cm thick on outcrop grades tween upramp peritidal mudstone and downramp
laterally and vertically into skeletal wackestone/pack- subtidal carbonates with abundant chert (Fig. 5). This
stone with nodular bedded chert (Fig. 6B). This facies facies commonly contains little silica, except where
occurs within both the lower and upper Aleman the colonial corals commonly are replaced by chert.
Formation, seaward of the grainstone shoal complex Low-angle tangential cross-bedding locally is com-
and above the interbedded calcisiltite or mudstone and mon within this facies. Hardground surfaces also are
spiculitic chert (Fig. 4). The abundance of chert in this common within this facies and these locally are
facies varies from 5– 60%. The chert nodules range encrusted by phosphate.
from a few cm’s to tens of cm’s in diameter. The chert The skeletal packstone/grainstone and coral baffle-
margins vary from smooth to sharp and irregular. stone is interpreted to represent a high-energy skeletal
Some chert nodules contain carbonate within their shoal or coral thicket. Widespread cross-bedding in
centers giving them a ‘‘hollow’’ appearance (see this unit indicates high-energy currents during depo-
Howe, 1959). Primary laminations within the skeletal sition of this unit. The abundance of corals in this
wackestone/packstone are rare, but where they occur facies indicates that they were deposited in warm,
with the chert, the laminations commonly are bent normal marine waters.
around the nodules. The nodular chert occurs primarily
in skeletal wackestone to skeletal packstone. Sponge
spicules occur in many of the silicified nodules but 5. Types of chert in Montoya Group
their abundance commonly is much less than in the
interbedded calcisiltite and spiculite facies. Phosphate The chert in the Montoya Group is divided into
occurs throughout this facies as peloids, coatings on three types: primary, early diagenetic and late diage-
hardgrounds and as a replacement of skeletal grains, netic. The nature of contacts between the chert and
most commonly of bryozoans. Additionally, there are carbonate within an individual bed varies greatly from
many silicified burrows and fossils (Fig. 6C) within gradational to sharp. Many of the chert beds contain
the nodular cherty wackestone/packstone. Brachio- small euhedral dolomite crystals formed after forma-
pods are the main skeletal grains in this facies with tion of the chert.
lesser amounts of bryozoan and crinoid fragments.
These cherty nodular carbonate facies formed on 5.1. Primary depositional chert
an open marine ramp. The variety of chert abundance
and morphologies reflects both original depositional Centimeter-thick beds of chert interbedded with
features and subsequent early diagenetic silica enrich- calcisiltite or mudstone and elongate discontinuous
ment. The abundance of brachiopods and lack of other chert lenses in calcisiltite or mudstone are considered
warm-water organisms (corals, green algae) suggest primary depositional because petrography and etching
this facies formed in cool waters. The lack of bedding of samples with HF reveals that they are composed
and nodular appearance of chert suggests this facies almost entirely of sponge spicules. Similarly, cross-
was intensely bioturbated. Deformed lamination sur- bedding in these spiculites indicate either storm re-
rounding chert nodules, silicification of burrows and working or downslope current transportation of abun-
unflattened skeletal fragments indicate that much of dant spicules. These chert beds formed from an
the silica in this facies formed prior to burial and accumulation of sponge spicules on the seafloor, and
compaction. the lack of abundant storm features (i.e. graded
374 M.C. Pope / Palaeogeography, Palaeoclimatology, Palaeoecology 210 (2004) 367–384

bedding, hummocky cross stratification) suggests they conformably overlain by Cincinnatian shaly cool-
commonly formed below storm wave base. water carbonates (Holland, 1993; Pope and Read,
1997b; Holland and Patzkowsky, 1997). As the Ap-
5.2. Early diagenetic chert palachian Basin filled during the Cincinnatian, fine
siliciclastics prograded progressively farther to the
Almost all the nodular and irregular chert nodules west and south as that basin was filled (Kolata et
in the Aleman Formation are considered early dia- al., 2001).
genetic because laminae in the carbonate sediment The Viola Group of Oklahoma has two parts, a
are bent around the chert. Also, many whole, uncom- lower Trenton Group (Upper Ordovician) equivalent
pacted fossils and undeformed burrows in the Ale- that is unconformably overlain by Cincinnatian strata
man Formation are silicified suggesting this chert (O’Brien and Derby, 1997). The Viola Group consists
formed early on the sea floor, or immediately below of interbedded carbonate and chert deposited on a
the sediment – water interface, prior to complete steep ramp that includes contourites and turbidites
lithification. (Brown and Sentfle, 1997). The basal part of the Viola
Group is particularly siliceous (up to 70% chert),
5.3. Late diagenetic chert including both biogenic (sponge spicules) and sec-
ondary silica (Galvin, 1983; Candelaria and Roux,
Late diagenetic chert commonly occurs in three 1997). The lower cherty carbonate of the Viola Group
forms. (1) White to light gray nodules that cross-cut is comprised of nodular-bedded skeletal wackestone/
bedding and occur primarily in tidal flat facies are packstone with light brown-gray chert nodules (Fig.
interpreted to be a replacement of evaporite nodules. 7). The nodular bedding was likely produced by a
(2) Gray to white nodules within subtidal facies cross- combination of burrowing and uneven marine precip-
cut bedding of dark gray carbonate containing early itation, followed by subsequent uneven compaction.
diagenetic chert (e.g. Geeslin and Chafetz, 1982) The chert is interpreted to be early diagenetic since
occur primarily updip and are also interpreted to be bedding in this facies is deformed around the chert.
a replacement of evaporites. However, these evapor- Karst within the Viola Group in Oklahoma formed
ites likely formed from burial brines during subse- prior to the Richmondian (Sykes et al., 1997) and may
quent exposure of the Montoya carbonate platform. be correlative to shallowing and progradation of
(3) Elongate veins, or tabular beds that cross cut or grainstone within the Aleman Formation. The top part
parallel bedding. of carbonate part of the Viola Group is a crinoidal
grainstone (equivalent to Fernvale, Welling) that is
conformably overlain by the Richmondian Sylvan
6. Regional correlations Shale. This grainstone may be equivalent to the
grainstone that prograded near the top of Sequence
The Montoya Group is regionally correlative with 2 in the Montoya Group. Structural differentiation of
Upper Ordovician cherty and phosphatic carbonates basins and uplifts segmented the southern Oklahoma
of the southern Midcontinent and Appalachian Basin part of the Ouachita basin to produce distinctive
(Fig. 3). These correlations are described here to lithologies in each area (Denison, 1997). The basins
determine the paleoceanographic significance of these commonly are muddier and chert-rich, whereas the
units. uplifts are grain-rich, preferentially dolomitized, and
The Trenton Group (Upper Ordovician) and its contain abundant hardgrounds, more karstic surfaces,
equivalents (Galena Group, Lexington Limestone) in and less chert. Locally the Viola Group contains
the Appalachian Basin and northern midcontinent are sandstone beds such as First Wilcox Sand that were
shallow-water carbonates that were deposited in cool, shed off subtle highs in Arkansas and Oklahoma
phosphate-rich waters (Brookfield, 1988; Lavoie, during high-frequency sea-level falls (see O’Brien
1995; Patzkowsky and Holland, 1996; Pope and and Derby, 1997; Denison, 1997). Organic-rich shale
Read, 1997a, 1998; Holland and Patzkowsky, 1997; and chert in the Big Fork Chert and Polk Creek Shale
Kolata et al., 2001). These cool-water carbonates are units (Finney, 1986) occurs outboard of the Viola
M.C. Pope / Palaeogeography, Palaeoclimatology, Palaeoecology 210 (2004) 367–384 375

Fig. 7. Photo of nodular carbonate and light-colored chert in lower Viola Group, I-35, Ardmore County, Oklahoma. Ten-centimeter scale bar in
lower middle of photo.

Group cherty-carbonate facies and indicates this area grained, organic-rich chert and carbonate deposited
of the margin received abundant dissolved silica in a deep subtidal setting (McBride, 1969, 1970,
throughout the Late Ordovician. 1989). The Maravillas Formation was deposited
The Richmondian Sylvan Shale is a poorly ex- during the Late Ordovician (Goldman et al., 1995).
posed unit that overlies the top of the Viola Group and The lower part of the Maravillas Formation is a
underlies the latest Ordovician (Gamachian – Hirnan- succession, up to 130 m thick of interbedded thin to
tian) Keel Oolite. The Sylvan Shale records a single thick beds of carbonate and chert (Fig. 8A). The
shallowing-upward trend of graptolitic black shale carbonate beds (5 cm to 2 m thick) are calcisiltite and
passing upward into unfossiliferous gray and green carbonate-clast conglomerate which commonly fine
shale (Finney, 1988). The water depths of the Sylvan upward (Fig. 8B). The carbonate units were deposited
are unknown, but it is not entirely a deepwater unit, by carbonate turbidity currents (McBride, 1969,
rather it reflects muddying of the waters which poi- 1970) whose presence suggests the Maravillas For-
soned the carbonate system shutting down the car- mation was deposited in a tectonically active setting,
bonate factory (Amsden, 1988). This scenario is or below a break in the ramp slope, possibly on a
compatible with the onset of a gradual long-term distally steepened ramp. The chert morphologies in
sea-level drop that begins in the Cincinnatian and the Maravillas Formation are quite varied from even,
induced progradation of fine siliciclastics to the west uniformly dark homogeneous beds to undulatory beds
and south from the Appalachian Basin (Kolata et al., and less common chert nodules. The Maravillas
2001). Formation chert contains abundant sponge spicules,
The Hirnantian Keel Oolite is part of a thin, locally phosphate and glauconite (McBride, 1969, 1989).
developed succession of iron and aragonitic ooids that The chert of the Maravillas Formation is interpreted
occur discontinuously in the US Midcontinent and to have formed during deposition and early diagene-
southern Canada (see Sharma and Dix, this volume). sis. The upper 20 – 30 m of the Maravillas Formation
These units formed during the sea-level drawdown consists of cherty shale or shale (McBride, 1969,
likely associated with the maximum extent of Late 1970) that may be correlative with the Sylvan Shale
Ordovician glaciation (Brenchley et al., 1994). in Oklahoma. The Maravillas Formation was thrusted
The Maravillas Formation in the Marathon Uplift northward to its present position during the Late
of south-central Texas consists of interbedded fine- Paleozoic, indicating the Ordovician carbonate ramp
376 M.C. Pope / Palaeogeography, Palaeoclimatology, Palaeoecology 210 (2004) 367–384

Fig. 8. (A) Photo of interbedded carbonate and chert the Maravillas Formation along Texas State Highway 385, approximately 15 km south of
Marathon, Texas. Person for scale is approximately 1.7 m tall. (B) Close-up of normal graded carbonate conglomerate between black chert and
light gray lime mudstone in Maravillas Formation.

upon which it formed was much wider than its Anderson, 1974; Anderson and Silver, 1979) does not
current configuration. negate the correlation, just its relative position.
Biostratigraphic correlation indicates the Montoya Upper Ordovician cherty carbonates are not re-
Group, Maravillas Formation, and Viola Springs stricted to southern Laurentia, but also occur discon-
Group formed a laterally continuous zone of cherty tinuously in the North American Cordillera from the
carbonates across the southern Laurentia margin dur- Great Basin to northern Canada and Alaska (Figs. 1
ing the Late Ordovician (Fig. 9). Similarly, Upper and 9). In these areas, interbedded chert and grapto-
Ordovician shelf carbonates in Sonora, Mexico pass litic shale were deposited in deep water seaward of
downramp into interbedded chert and carbonate then cherty carbonates that pass upramp into peritidal
into basinal chert and shale (Poole et al., 1995a,b), carbonates (Ross, 1976; Miller, 1975, 1976).
suggesting this belt of cherty carbonates was contin- Coeval interbedded chert and shale or cherty shale
uous from Oklahoma into northern Mexico. The units were also deposited in low-latitude settings on
possibility of many hundreds of kilometers of Meso- Baltica, Siberia, New Zealand and Australia (Fig. 1).
zoic left-lateral offset along this margin (Silver and The shale in these units commonly is organic-rich,
M.C. Pope / Palaeogeography, Palaeoclimatology, Palaeoecology 210 (2004) 367–384 377

Fig. 9. Map showing evidence for late Middle to Late Ordovician upwelling around the US. Upwelling is marked in the Appalachian Basin by
cool-water carbonates (Brookfield, 1988; Patzkowsky and Holland, 1993; Lavoie, 1995; Pope and Read, 1997a, 1998) containing abundant
phosphate (up to 10 wt.%) along the Cincinnati Arch (Holland and Patzkowsky, 1997; Pope and Read, 1997a, b). The Sebree Trough funneled
cool, deep oceanic waters onto the US Midcontinent (Kolata et al., 2001). Warm water tidal flats only developed in the Late Ordovician, updip
along the western and northern limit of outcrops. Cherty carbonates described in this paper occur along the southern margin of Laurentia from
Oklahoma to Mexico. Chert-rich carbonate facies occur in the subsurface data from the Permian basin, Texas. Outboard of the cherty carbonates
are cherty shales. Cherty carbonates and cherty shale also occur in the southern Great Basin and Idaho.

and may contain abundant phosphate and numerous Formation indicates that these rocks also formed in
graptolites. The chert in these units commonly occurs warm oceanic waters. However, the deeper subtidal
as nodules containing radiolarians or sponge spicules. portions of the Aleman are dominated by siliceous
Silicified nodules within the shaly units are interpreted sponge spicules and brachiopods indicating this is a
to have formed when sea-level was high (e.g. Loi and heterozoan fauna (James, 1997). The brachiopods in
Dabard, 2002), similar to the model for formation of the Aleman Formation and deeper subtidal portions of
Aleman Formation chert nodules proposed below. the Cutter Formation (Howe, 1959) contain many of
the genuses (Hebertella, Platystrophia, Rafinesquina)
interpreted to define a cool water biofacies in the
7. Discussion Appalachian Basin (Patzkowsky and Holland, 1999).
Additionally, the lack of ooids, peloids, hardgrounds
7.1. Upwelling in the Late Ordovician and calcareous algae in these rocks also may indicate
they formed in cooler oceanic waters (cf. Patzkowsky
The abundance of stromatoporoids, corals, and and Holland, 1993; Pope and Read, 1997a,b). The
receptacularid algae (LeMone, 1969, 1988) in the mudcracked and bioturbated dolomudsone with occa-
Upham Formation indicates that this unit was depos- sional corals of the peritidal facies in the Cutter
ited in warm oceanic waters. Similarly, the occurrence Formation suggest the updip part of the Montoya
of corals in the shallow subtidal facies of the Aleman Group was deposited in warm oceanic waters. Thus,
378 M.C. Pope / Palaeogeography, Palaeoclimatology, Palaeoecology 210 (2004) 367–384

it appears the upper part of the Montoya Group may North American midcontinent are anomalous in the
have been deposited on a temperature-stratified ramp Paleozoic because they represent the influx of cool
with cool, deeper subtidal waters passing updip into oceanic waters, with abundant silica and phosphate,
warm, shallow subtidal and peritidal waters. This hundreds of kilometers onto the interior of this con-
depositional model is very similar to one put forth tinent (Lavoie, 1995; Patzkowsky and Holland, 1996;
to explain the coeval juxtaposition of subtidal cool- Pope and Read, 1998; Kolata et al., 2001). Interbed-
water and peritidal warm-water Late Ordovician car- ded chert and shale commonly were deposited ocean-
bonates in the US Midcontinent (Kolata et al., 2001) ward of the cherty carbonates (e.g. Big Fork chert,
and the Appalachian Basin (Pope and Read, 1998). upper Maravillas Formation shaly chert, unnamed
The abundance of spiculitic chert, early diagenetic units in Mexico). Graptolitic-rich shales may mark
chert and phosphate (1 – 5 wt.%) in the Montoya the position of upwelling along the shelf margin (e.g.
Group indicates these formed in an upwelling zone Finney and Berry, 1997) and coeval cherty carbonates
(e.g., Parrish, 1998). Direct correlations and similar- inboard and graptolitic shale and chert outboard are
ities between the Montoya Group, Maravillas Forma- consistent with laterally extensive (up to a few hun-
tion, Viola Group and unnamed Upper Ordovician dred kilometers wide) upwelling across the southern
units in Mexico indicates the upwelling zone was Laurentia margin. The development of this upwelling
widespread along the southern margin of Laurentia zone corresponds with a northward expansion of cool
(Fig. 9). Global sea-level was at, or near, its Paleozoic water trilobite faunas (Shaw, 1991) and a pronounced
maximum during the Late Ordovician and most of shift to cooler-water benthic faunas across eastern
North America was submerged and could not have North America (Patzkowsky and Holland, 1993).
been a substantial source of silica. Thus, the silica and Similarly, the abundance of silica-replaced fossils
phosphate in the Upper Ordovician cherty carbonates and bedded chert throughout the Late Ordovician
of southern Laurentia were brought into this area by (Kidder and Erwin, 2001; Pope and Steffen, 2003)
upwelling of cool, deeper oceanic waters. Biostratig- suggests a substantial global increase in upwelling
raphy (Fig. 3) suggests the upwelling lasted up to 10 – during this period. The abundance of Late Ordovician
12 my (454 – 442 Ma). cherty carbonate and cherty shale units and the dearth
In modern oceans, coastal upwelling occurs pri- of these units in the Middle Ordovician and during the
marily along the western margins of low – middle Early Silurian suggests a global climatic or oceano-
latitude continents where surface winds blowing par- graphic origin for these deposits.
allel to the coast, and slightly offshore push surface Glacially mixed oceans commonly are marked by
waters through Ekman transport perpendicular to the the development of widespread equatorial upwelling
shoreline. This offshore movement of water is bal- zones, whereas greenhouse oceans generally lack
anced through replacement from below by cool, widespread upwelling zones (Hay, 1988). Similarly,
nutrient rich waters. These upwelling waters generally the occurrence and abundance of widespread phos-
come from the upper few hundred meters of the phorites commonly indicates well-mixed oceans (e.g.
ocean. These upwelling zones are characterized by Berner, 1996). Thus, the coupled origination of geo-
abundant phosphate, silica, and organic-rich shale graphically widespread equatorial or near-equatorial
(Parrish, 1998). The appearance of widespread and upwelling and phosphogenesis in the Late Ordovician
anomalously cherty and phosphatic carbonates in the may reflect the initiation of Gondwana glaciation. The
Late Ordovician of Laurentia indicates that strong initiation of this upwelling in the Late Ordovician
equatorial upwelling began during this period. Unlike corresponds with cool (13 – 19 jC) surface waters in
modern upwelling the distribution of cool water, the Appalachian Basin (Railsback et al., 1990), and
phosphate-, and silica-rich carbonates occurs on mar- the approximate initiation of glaciogenic deposits in
gins that were perpendicular to prevailing winds (Figs. Africa (Theron, 1994). The presence of geographical-
1 and 9), especially on the US Midcontinent, and ly widespread upwelling zones developed hundreds of
possibly along southern Laurentia. kilometers onto Laurentia over a 10 – 12 my period
Upper Ordovician bedded carbonate, chert, and also is in accord in accord with computer modeling,
phosphate deposits of southern Laurentia and the which indicates that there was increased poleward
M.C. Pope / Palaeogeography, Palaeoclimatology, Palaeoecology 210 (2004) 367–384 379

heat flow during the Late Ordovician glaciation (Chatfieldian – Hirnantian; 454 – 442 Ma) transgres-
(Poussart et al., 1999). sive to regressive sequence during the Late Ordovi-
cian containing 6– 12 superimposed sea-level rises
7.2. Relative sea-level record from North America and falls each 1 –3 my apiece (Fig. 10). The Chat-
fieldian record is poorly resolved but it appears to
The Late Ordovician is a time of high sea-level consist of two to five sea level rises and falls. The
around North America following a global sea-level Edenian rise and fall appears to be the longest
low that occurred during the Middle Ordovician duration ( f 3 Ma) but may locally record at least
(Schutter, 1992). The six, regionally correlative de- two separate events. There are two long-term sea
positional sequences in the Montoya Group record level rise and fall events in both the Maysvillian
long-term (1– 3 my) changes in relative water-depth and Richmondian and a single long-term rise and fall
in this area during the later Late Ordovician (Fig. 10). recorded in the Hirnantian. A very detailed record of
The North American data indicates a long-term multiple high-frequency high-amplitude fluctuations

Fig. 10. Correlation diagram of published late Middle to Late Ordovician sea-level, coastal onlap and water depth curves from the periphery of
North America. Maximum water depth, marked with a black circle, approximates the local maximum flooding surface in each study area. The
Montoya curve shows six third-order (1 – 3 my duration) relative sea-level rises and falls corresponding to the six depositional sequences. These
relative sea-level rises are roughly correlative with events from the Appalachian Basin (Holland and Patzkowsky, 1997; Pope and Read, 1998)
and Great Basin (Harris et al., 1996) suggesting there were four to seven third-order rises and falls in sea-level in the Late Ordovician. The
similarity of these curves suggests the rise and falls in relative sea-level were likely produced by glacio-eustatic fluctuations. The largest sea-
level fall in the Late Ordovician occurs during the Hirnantian indicating the maximum extent of Late Ordovician icesheets during this time.
Differences in the timing of maximum water depth in any individual location were likely produced by local variations in subsidence and
sediment supply.
380 M.C. Pope / Palaeogeography, Palaeoclimatology, Palaeoecology 210 (2004) 367–384

in the very latest Ordovician (Fig. 10) occurs on (e.g. Holland, 1993; Jennette and Pryor, 1993; Pope
Anticosti Island (Long, 1993). and Read, 1997b; Brett and Algeo, 2001). However,
The similarity in timing of these long-term events, determining the magnitude and duration of sea-level
their duration, and distribution around Laurentia sug- fluctuations that produced this cyclicity is controver-
gests there is likely a long-term eustatic driver to these sial (cf. Jennette and Pryor, 1993; Holland et al., 1997
relative sea-level fluctuations, but variations in subsi- for contrasting views).
dence, sediment distribution or tectonics are likely Alternatively, the nearly synchronous long-term
obscuring the true eustatic component of these curves. Late Ordovician fluctuations delineated here may be
For example, in the Appalachian Basin, the deepest recording nonlinear oscillations that are not directly
water facies in the Late Ordovician occurs during the related to any eustatic forcing mechanisms (e.g. Gaf-
Chatfieldian (Diecchio and Broderson, 1994) or the fin, 1992). More detailed sequence stratigraphic
earliest Edenian (Holland and Patzkowsky, 1997; records through Late Ordovician successions are nec-
Pope and Read, 1998). However, the deepest water essary to determine the mechanism(s) that produced
facies of the Montoya Group occur during the Mays- these long-term relative sea-level fluctuations. Re-
villian (Fig. 10), whereas they occur in the Richmon- gardless of the mechanism(s), these long-term relative
dian in Great Basin rocks (Harris and Sheehan, 1996, sea level fluctuations they may provide a means for
1997). Although, the absolute magnitude of these regional and global correlations that are below the
long-term sea-level events is uncertain, it is clear resolution of biostratigraphy.
sea-level fell the farthest in the Hirnantian because
that lowstand led to widespread subaerial exposure of
Laurentia and elsewhere (cf. Brenchley et al., 1994), 8. Conclusions
probably during the maximum advance of Late Ordo-
vician icesheets. The Late Ordovician glacial event was unique
The mechanism(s) creating these long-term relative because it formed during a time of elevated atmo-
sea-level fluctuations around Laurentia during the spheric CO2. Determining the duration of this glacial
Late Ordovician is unclear. The duration of these episode is essential to determining how glaciation
fluctuations, 1 –3 my each, is generally below the began, how waxing and waning of the glacial ice-
accepted record of long-term changes in seafloor sheets affected coeval sedimentation, and how and
spreading rates that may cause eustatic variations why this glaciation ceased. The abundance of cherty
(Pitman, 1978). Similarly, these fluctuations are much carbonate and phosphate in the Montoya Group of
longer than the short-term Milankovitch climate forc- New Mexico and Texas indicates these rocks formed
ing mechanisms, precession, obliquity and eccentric- in a widespread upwelling zone. The duration of this
ity, that have frequencies of approximately 21– 23, 41, upwelling and correlation with other Late Ordovician
and 100 or 400 ky, respectively. These fluctuations upwelling zones around the periphery of Laurentia
may reflect long-term changes in ice volume, possibly suggest that Gondwana glaciation may have been
as a long-term (1.3 or 2.0 Ma) harmonic of eccentric- occurring throughout this period. The regionally cor-
ity (De Boer and Smith, 1994). However, if long-term relative third-order depositional sequences (1 –3 my
changes in ice volume are being recorded in the Late duration) in the Montoya Group refine a long-term
Ordovician why is the record of short-term ice volume Late Ordovician relative sea-level curve.
changes, which should be represented by high-fre-
quency, high-amplitude depositional cycles be so
scarce? Moderate to high-amplitude, high-frequency Acknowledgements
meter-scale cycles developed during the Chatfieldian
(Pope and Read, 1998) and the Hirnantian (Long, This research supported by ACS-PRF #35837-G8.
1993), but unequivocable evidence for such cylicity in Bob Myers, Range Geologist at White Sands Missile
the Cinncinnatian is not evident. Meter- and decame- Range provided invaluable help in accessing sections
ter-scale depositional cycles are ubiquitous in the in the San Andres range. Dan Hunter, Bryn Clark,
Cincinnatian successions of the Appalachian Basin Luke LeMond, Steve Turpin and John Bengelsdorf all
M.C. Pope / Palaeogeography, Palaeoclimatology, Palaeoecology 210 (2004) 367–384 381

provided invaluable field assistance. Steven Holland Arbuckle Mountians, Oklahoma. In: Johnson, K.S. (Ed.),
and Carlton Brett provided insightful comments on a Simpson and Viola Groups in the Southern Midcontinent,
1994 Symposium. Oklahoma Geological Survey Circular,
previous version of this manuscript. Jessica Steffen vol. 99, p. 102.
initiated many useful discussions about the Montoya Candelaria, M.P., Roux, B.P., 1997. Reservior analysis of a hor-
Group and especially the significance of its chert. izontal-well completion in the Viola Limestone ‘‘chocolate
Kate Giles provided access to a digital camera, advice brown zone’’, Marietta Basin, Oklahoma. In: Johnson, K.S.
(Ed.), Simpson and Viola Groups in the Southern Midconti-
on logistics and acted as a sounding board for many of
nent, 1994 Symposium. Oklahoma Geological Survey Circular,
the ideas in this paper. This paper was approved for vol. 99, pp. 183 – 193.
public release by White Sands Missile Range; Crowell, J.C., 1999. Pre-Mesozoic Ice Ages: Their Bearing on
distribution unlimited. OPSEC review completed on Understanding the Climate System. Geological Society of
August 19, 2002. America Memoir, vol. 192. Geological Society of America,
Boulder, CO.
Crowley, T.J., Baum, S.K., 1991. Toward reconciliation of Late
Ordovician ( f 440 Ma) glaciation with very high CO2 levels.
References Journal of Geophysical Research 96, 22597 – 22610.
Crowley, T.J., Baum, S.K., 1995. Reconciling Late Ordovician (440
Ahr, W., 1973. The carbonate ramp: an alternative to the shelf Ma) glaciation with very high (14  ) CO2 levels. Journal of
model. Transactions of the Gulf Coast Association of Geological Geophysical Research 100, 1093 – 1101.
Societies 23, 221 – 225. De Boer, P.L., Smith, D.G., 1994. Orbital forcing and cyclic sequen-
Amsden, T.W., 1988. Depositional and post-depositional history of ces. In: De Boer, P.L., Smith, D.G. (Eds.), Orbital Forcing and
middle Paleozoic (Late Ordovician through Early Devonian) Cyclic Sequences. International Association of Sedimentologists
strata in the ancestral Anadarko Basin. In: Johnson, K.S. Special Publication, vol. 19. Blackwell, Oxford, pp. 1 – 14.
(Ed.), Anadarko Basin Symposium, 1988. Oklahoma Geological Denison, R.E., 1997. Contrasting sedimentation inside and outside
Survey Circular, vol. 90. University of Oklahoma, Norman, OK, the southern Oklahoma aulacogen during Middle and Late Or-
pp. 143 – 146. dovician. In: Johnson, K.S. (Ed.), Simpson and Viola Groups in
Anderson, T.H., Silver, L.T., 1979. The role of the Mojave – Sonora the Southern Midcontinent, 1994 Symposium. Oklahoma Geo-
Megashear in the tectonic evolution of northern Sonora. In: logical Survey Circular, vol. 99, pp. 39 – 47.
Anderson, T.H., Roldan-Quintana, J., (Eds.), Geology of north- Diecchio, R., Broderson, B.T., 1994. Recognition of regional (eu-
ern Sonora (Geological Society of America Annual Meeting static?) and local (tectonic) relative sea-level events in outcrop
Trip 27). University of Pittsburgh, PA and Hermosillo, Son., and gamma-ray logs, Ordovician, West Virginia. In: Dennison,
Instituto de Geologia, U.N.A.M., pp. 59 – 68. J., Ettensohn, F.R. (Eds.), Tectonic and Eustatic Controls On
Barnes, C.R., 1986. The faunal extinction event near the Ordovi- Sedimentary Cycles: Society of Economic Paleontologists and
cian – Silurian boundary: a climatically induced crisis. In: Wall- Mineralogists Concepts in Sedimentology and Paleontology,
iser, O.H. (Ed.), Global Bioevents, Springer-Verlag Lecture vol. 4, pp. 170 – 180.
Notes in Earth Science, vol. 8, 121 – 126. Elrick, M., Read, J.F., Coruh, C., 1991. Short-term paleoclimatic
Berner, R.A., 1994. GEOCARB II, A revised model of atmospheric fluctuations expressed in lower Mississippian ramp-slope depos-
CO2 over Phanerozoic time. American Journal of Science 294, its, southwestern Montana. Geology 19, 799 – 802.
56 – 91. Finney, S.C., 1986. Graptolite biofacies and correlation of eustatic,
Berner, R.A., 1996. A new look at the long-term carbon cycle. GSA subsidence and tectonic events in the Middle – Upper Ordovi-
Today 9, 1 – 6. cian of North America. Palaios 1, 435 – 461.
Brenchley, P.J., Marshall, J.D., Carden, G.A.F., Robertson, D.B.R., Finney, S.C., 1988. Middle Ordovician strata of the Arbuckle and
Meidla, T., Hints, L., Anderson, T.F., 1994. Bathymetric and Ouachita Mountains, Oklahoma: contrasting lithofacies and
isotopic evidence for short-lived Late Ordovician glaciation in biofacies deposited in the southern Oklahoma aulacogen and
a greenhouse period. Geology 22, 295 – 298. Ouachita geosyncline. In: Hayward, O.T. (Ed.), Centennial Field
Brett, C.E, Algeo, T., 2001. Sequence stratigraphy of Upper Ordo- Guide 4, South-Central Section. Geological Society of America,
vician and Lower Silurian strata of the Cincinnati Arch region. Boulder, CO, pp. 171 – 176.
Field Trip Guidebook for the 1999 Field Conference of the Finney, S.C., Berry, W.B.N., 1997. New perspectives on graptolite
Great Lakes Section of the Society for Sedimentary Geologists distributions and their use as indicators of platform margin dy-
and Kentucky Society of Professional Geologists. SEPM, Soci- namics. Geology 25, 919 – 922.
ety for Sedimentary Geology, pp. 34 – 46. Flower, R.H., 1961. Montoya and related colonial corals. New
Brookfield, M.E., 1988. A mid-Ordovician temperate carbonate Mexico Bureau of Mines and Mineral Resources Memoir 7,
shelf – – The Black River and Trenton Limestone Groups of 229.
southern Ontario, Canada. Sedimentary Geology 60, 137 – 154. Flower, R.H., 1969. Early Paleozoic of New Mexico and the El
Brown, A.A., Sentfle, J.T., 1997. Source potential of the Viola Paso region. The Ordovician Symposium. El Paso Geological
Springs Formation, southern limb of the Arbuckle anticline, Society Annual Fieldtrip, vol. #3, pp. 32 – 101.
382 M.C. Pope / Palaeogeography, Palaeoclimatology, Palaeoecology 210 (2004) 367–384

Frakes, L.A., Francis, J.E., Syktus, J.I., 1992. Climatic modes of the James, N.P., 1997. The cool-water carbonate depositional realm. In:
Phanerozoic. Cambridge Univ. Press, Cambridge. James, N.P., Clarke, J.A.D. (Eds.), Cool-Water Carbonates.
Gaffin, S.R., 1992. Unforced oscillations in a freeboard and basin SEPM (Society for Sedimentary Geology) Special Publication,
model: analogue to glacial/climate oscillators? The Journal of vol. 56. Tulsa, OK, pp. 1 – 20.
Geology 100, 717 – 729. Jennette, D.C., Pryor, W.A., 1993. Cyclic alternation of proximal
Galvin, P.K., 1983. Deep to shallow carbonate ramp transition in and distal storm facies: Kope and Fairview Formations (Up-
Viola Limestone (Ordovician), southwest Arbuckle Mountains, per Ordovician), Ohio and Kentucky. Journal of Sedimentary
Oklahoma. American Association of Petroleum Geologists Bul- Petrology 63, 183 – 203.
letin 63, 466 – 467. Kelley, V.C., Silver, C., 1952. Geology of the Caballo Mountains.
Geeslin, J.H., Chafetz, H.S., 1982. Silicification prior to carbonate University of New Mexico Publications in Geology 4, 286.
lithification. Journal of Sedimentary Petrology 52, 1283 – 1293. Kidder, D.L., Erwin, D.H., 2001. Secular distribution of biogenic
Goldman, D., Bergstrom, S.M., Mitchell, C.E., 1995. Revision of silica through the Phanerozoic: comparison of silica-replaced
the Zone 13 graptolite biostratigraphy in the Marathon, Texas, fossils and bedded cherts at the series level. The Journal of
standard succession and its bearing on Upper Ordovician bio- Geology 109, 509 – 522.
geography. Lethaia 28, 115 – 128. Kolata, D.R., Huff, W.D., Bergstrom, S.M., 2001. The Ordovi-
Harris, M.T., Sheehan, P.M., 1996. Upper Ordovician – Lower Si- cian Sebree Trough: an oceanic passage to the Midcontinent
lurian depositional sequences determined from middle shelf sec- United States. Geological Society of America Bulletin 113,
tions, Barn Hills and Lakeside Mountains, eastern Great Basin. 1067 – 1078.
In: Witzke, B.J., Ludvigson, G.A., Day, J. (Eds.), Paleozoic Kottlowski, F.E., 1963. Paleozoic and Mesozoic strata in southwest-
sequence stratigraphy; views from the North American Craton. ern and south-central New Mexico. New Mexico Bureau of
Geological Society of America Special Paper, vol. 306. Geolog- Mines and Mineral Resources Bulletin 79, 100.
ical Society of America, Boulder, CO, pp. 161 – 176. Kump, L.R., Arthur, M.A., Patzkowsky, M.E., Gibbs, M.T., Pinkus,
Harris, M.T., Sheehan, P.M., 1997. Carbonate sequences and fossil D.S., Sheehan, P.M., 1999. A weathering hypothesis for glacia-
communities from the Upper Ordovician – Lower Silurian of the tion at high atmospheric pCO2 during the Late Ordovician.
Eastern Great Basin. In: Link, P., Kowallis, B.J. (Eds.), Prote- Palaeogeography, Palaeoclimatology, and Palaeoecology 152,
rozoic to Recent stratigraphy, tectonics and vocanology, Utah, 173 – 187.
Nevada, southern Idaho and central Montana. Brigham Young Lavoie, D., 1995. A Late Ordovician high-energy temperate-water
Geology Studies, vol. 42, Pt. I, pp. 105 – 128. carbonate ramp, southern Quebec, Canada: implications for Late
Harris, P.M., Saller, A.H., Simo, J.A. (Eds.), 1999. Advances in Ordovician oceanography. Sedimentology 42, 95 – 116.
carbonate sequence stratigraphy: Application to Reservoirs, out- LeMone, D.V., 1969. Cambrian and Ordovician in the El Paso
crops and models. SEPM (Society for Sedimentary Geology) border region. In: LeMone, D.V. (Ed.), The Ordovician Sym-
Special Publication, vol. 63. Society for Sedimentary Geology, posium: El Paso Geological Society, 3rd Annual Field Trip,
Tulsa, OK. pp. 145 – 161.
Hay, W.H., 1988. Paleoceanography: a review for the GSA LeMone, D.V., 1988. Precambrian and Paleozoic stratigraphy;
Centennial. Geological Society of America Bulletin 100, Franklin Mountains, west Texas. In: Hayward, O.T. (Ed.), Cen-
1934 – 1956. tennial Field Guide. South-Central section, vol. 4. Geological
Hayes, P.T., 1975. Cambrian and Ordovician rocks of Arizona, New Society of America, Boulder, CO, pp. 387 – 394.
Mexico, Texas. United States Geological Survey Professional Loi, A., Dabard, M., 2002. Controls of sea-level fluctuations on the
Paper, vol. 873. United States Geological Survey, Denver, CO. formation of Ordovician siliceous nodules in terrigenous off-
Hill, D., 1959. Some Ordovician corals from New Mexico, Arizona shore environments. Sedimentary Geology 153, 65 – 84.
and Texas. New Mexico Bureau of Mines and Mineral Resour- Long, D.G.F., 1993. Limits on late Ordovician eustatic sea-level
ces Bulletin 64, 25. change from carbonate shelf sequences: an example from Anti-
Holland, S.M., 1993. Sequence stratigraphy of a carbonate-clastic costi Island, Quebec. In: Posamentier, H.W., Summerhayes,
ramp: the Cincinnatian Series (Upper Ordovician) in its type C.P., Haq, B.U., Allen, G.P. (Eds.), Sequence Stratigraphy
area. Geological Society of America Bulletin 105, 306 – 322. And Facies Associations. Special Publication of the Internation-
Holland, S.M., Patzkowsky, M.E., 1997. Distal orogenic effects on al Association of Sedimentologists, vol. 18. Blackwell, Oxford,
peripheral bulge sedimentation, Middle and Upper Ordovician pp. 487 – 499.
of the Nashville dome. Journal of Sedimentary Research 67, Mac Niocaill, C., van der Pluijm, B.V., van der Voo, R., 1997.
250 – 263. Ordovician paleogeography and the evolution of the Iapetus
Holland, S.M., Miller, A.I., Dattilo, B.F., Meyer, D.L., Diekmeyer, Ocean. Geology 25, 159 – 162.
S.L., 1997. Cycle anatomy and variability in the storm-domi- Marshall, J.D., Brenchley, P.J., Mason, P., Wolff, G.A., Astini,
nated type Cincinnatian (Upper Ordovician); coming to grips R.A., Hints, L., Meidla, T., 1997. Global carbon isotopic
with cycle delineation and genesis. Journal of Geology 105, events associated with mass extinction and glaciation in the late
135 – 152. Ordovician. Palaeogeography, Palaeoclimatology, and Palaeoe-
Howe, H.J., 1959. Montoya Group stratigraphy (Ordovician) of cology 132, 195 – 210.
Trans-Peco Texas. American Association of Petroleum Geolo- McBride, E.F., 1969. Stratigraphy, sedimentary structures and origin
gists Bulletin 43, 2285 – 2333. of flysch and pre-flysch rocks of the Marathon Basin, Texas.
M.C. Pope / Palaeogeography, Palaeoclimatology, Palaeoecology 210 (2004) 367–384 383

Guidebook for the American Association of Petroleum Geolo- In: Cooper, J.D., Droser, M.L., Finney, S.C. (Eds.), Ordovician
gists and Society of Economic Paleontologists and Mineralogists Odyssey: Pacific Section of the Society for Sedimentary Geol-
Annual Meeting. Dallas Geological Society, Dallas, p. 104. ogy, pp. 277 – 284.
McBride, E.F., 1970. Stratigraphy and origin of Maravillas Forma- Pope, M.C., Read, J.F., 1997a. High-frequency cyclicity of the
tion (upper Ordovician), west Texas. American Association of Lexington Limestone (Middle Ordovician), a cool-water carbon-
Petroleum Geologists Bulletin 54, 1719 – 1745. ate clastic ramp in an active foreland basin. In: James, N.P.,
McBride, E.F., 1989. Stratigraphy and sedimentary history of Pre- Clarke, J.P. (Eds.), Cool-Water Carbonates. Society of Econom-
Permian Paleozoic rocks of the Marathon Uplift. In: Hatcher Jr., ic Paleontologists and Mineralogists Special Publication, vol.
R.D., Thomas, W.A., Viele, G.W. (Eds.), The Geology of North 56. Tulsa, OK, pp. 411 – 429.
America. The Appalachian – Ouachita Orogen in the United Pope, M.C., Read, J.F., 1997b. High-resolution surface and subsur-
States, vol. F-2. Geological Society of America, Boulder, CO, face sequence stratigraphy of Middle to Late Ordovician (Late
pp. 603 – 620. Mohawkian to Cincinnatian) foreland basin rocks, Kentucky
Miller, R.H., 1975. Late Ordovician – Early Silurian conodont bio- and Virginia. American Association of Petroleum Geologists
stratigraphy, Inyo Mountains, California. Geological Society of Bulletin 81, 1866 – 1893.
America Bulletin 86, 159 – 162. Pope, M.C., Read, J.F., 1998. Ordovician metre-scale cycles: impli-
Miller, R.H., 1976. Revision of Upper Ordovician, Silurian and cations for Ordovician climate and eustatic fluctuations in the
Lower Devonian stratigraphy, southwestern Great Basin. Geo- central Appalachian Basin, USA. Palaeoclimatology, Palaeo-
logical Society of America Bulletin 87, 961 – 968. geography, and Palaeoecology 138, 27 – 42.
Measures, E.A., 1985a. Carbonate facies of the Montoya Group— Pope, M.C., Steffen, J.B., 2003. Widespread, prolonged Late Mid-
description of a shoaling-upward ramp, Part I. West Texas Geo- dle to Late Ordovician upwelling in North America: a proxy
logical Society Bulletin 25 (2), 4 – 8. record of glaciation? Geology 31, 63 – 66.
Measures, E.A., 1985b. Carbonate facies of the Montoya Group— Pope, M.C., Hunter, D.M., Steffen, J., Clark, B., 2001. 2nd-order
description of a shoaling-upward ramp: Part II. West Texas Sequence Stratigraphy of Late Middle to Late Ordovician Mon-
Geological Society Bulletin 25 (3), 4 – 8. toya Group, southern New Mexico and West Texas. American
O’Brien, J.E., Derby, J.R., 1997. Progress report on Simpson and Association of Petroleum Geologists. Annual Meeting Program
Viola correlations from the Arbuckles to the Ozarks. In: John- with Abstracts, p. A160.
son, K.S. (Ed.), Simpson and Viola Groups in the Southern Poussart, P.F., Weaver, A.J., Barnes, C.R., 1999. Late Ordovician
Midcontinent, 1994 Symposium. Oklahoma Geological Survey glaciation under high atmospheric CO2: a coupled model anal-
Circular, vol. 99, pp. 260 – 266. ysis. Paleoceanography 14, 542 – 558.
Parrish, J.T., 1998. Interpreting pre-Quaternary climate from the Pratt, W.P., Jones, W.R., 1961. Montoya Dolomite and Fusselman
rock record. Columbia Univ. Press, New York, p. 338. Dolomite in Silver City region, New Mexico. American Asso-
Patzkowsky, M.E., Holland, S.M., 1993. Biotic response to a Mid- ciation of Petroleum Geologists Bulletin 37, 1894 – 1918.
dle Ordovician paleoceanographic event in eastern North Amer- Pray, L.C., 1958. Stratigraphic section, Montoya Group and Fussel-
ica. Geology 21, 619 – 622. man Formation, Franklin Mountains, Texas. West Texas Geo-
Patzkowsky, M.E., Holland, S.M., 1996. Extinction, invasion and logical Society, 30 – 42.
sequence stratigraphy; patterns of faunal change in the Middle Railsback, L.B., Acherly, S.C., Anderson, T.F., Cisne, J.L., 1990.
and Upper Ordovician of the Eastern United States. In: Witzke, Palaeontological and isotope evidence for warm saline deep
B.J., Ludvigson, G.A., Day, J. (Eds.), Paleozoic Sequence Stra- waters in Ordovician oceans. Nature 343, 156 – 159.
tigraphy; Views from the North American Craton. Geological Read, J.F., 1985. Carbonate platform facies models. American As-
Society of America Special Paper, vol. 306. Geological Society sociation of Petroleum Geologists Bulletin 69, 1 – 21.
of America, Boulder, CO, pp. 131 – 142. Ross Jr., R.J., 1976. Ordovician sedimentation in the western United
Patzkowsky, M.E., Holland, S.M., 1999. Biofacies replacement in States. Rocky Mountain Association of Geologists—1976 Sym-
a sequence stratigraphic framework: Middle and Upper Ordo- posium, pp. 109 – 133.
vician of the Nashville Dome, Tennessee, USA. Palaios 14, Schutter, S.R., 1992. Ordovician hydrocarbon distribution in North
301 – 323. America and its relationship to eustatic cycles. In: Webby, B.D.,
Pitman III, W.C., 1978. Relationship between eustasy and strati- Laurie, J.R. (Eds.), Global Perspectives on Ordovician Geology.
graphic sequences of passive margins. Geological Society of Balkema, Rotterdam, pp. 421 – 432.
America Bulletin 89, 1389 – 1403. Scotese, C.R., 1997. Continental Drift, 7th Edition. Paleomap Proj-
Poole, F.G., Stewart, J.H., Repetski, J.E., Harris, A.G., Ross Jr., ect, Arlington, TX, p. 79.
R.J., Ketner, K.B., Amaya-Martinez, R., Morales-Martinez, Shaw, F.C., 1991. Viola Group (Ordovician, Oklahoma) cryptoli-
J.M., 1995a. Ordovician carbonate-shelf rocks of Sonora, Mex- thinid trilobites: biogeography and taxonomy. Journal of Pale-
ico. In: Cooper, J.D., Droser, M.L., Finney, S.C. (Eds.), Ordovi- ontology 65, 919 – 935.
cian Odyssey: Pacific Section of the Society for Sedimentary Silver, L.T., Anderson, T.H., 1974. Possible left-lateral early to
Geology, pp. 267 – 275. middle Mesozoic disruption of the southwestern North Ameri-
Poole, F.G., Stewart, J.H., Berry, W.B.N., Harris, A.G., Repetski, can margin. Geological Society of America Abstracts with Pro-
J.E., Madrid, R.J., Ketner, K.B., Carter, C., Morales-Martinez, grams 6, 955 – 956.
J.M., 1995b. Ordovician ocean-basin rocks of Sonora, Mexico. Sweet, W.C., 1979. Late Ordovician conodonts and biostratigraphy
384 M.C. Pope / Palaeogeography, Palaeoclimatology, Palaeoecology 210 (2004) 367–384

of the western Midcontinent province. Brigham Young Univer- Tucker, R.D., Krogh, T.E., Ross Jr., R.J., Williams, S.H., 1990.
sity Geological Studies 26, 45 – 85. Time-scale calibration by high-precision U – Pb zircon dating
Sykes, M., Puckette, J., Abdolla, A., Al-Shaieb, Z., 1997. Karst of interstratified volcanic ashes in the Ordovician and Lower
Development in the Viola Limestone in Southern Oklahoma. Silurian stratotypes of Britain. Earth and Planetary Science Let-
In: Johnson, K.S. (Ed.), Simpson and Viola Groups in the ters 100, 51 – 58.
Southern Midcontinent, 1994 Symposium. Oklahoma Geologi- Weber, L.J., Sarg, J.F., Wright, F.M., 1995. Sequence stratigraphy
cal Survey Circular, vol. 99, pp. 66 – 75. and reservoir delineation of the Middle Pennsylvanian (Desmoi-
Theron, J.N., 1994. The Ordovician System in South Africa; cor- nesian), Paradox basin and Aneth field, southwestern USA. In:
relation chart and explanatory notes. In: Williams, S.H. (Ed.), Read, J.F., Kerans, C., Weber, L.J., Sarg, J.F., Wright, F.M.
The Ordovician System in Greenland and South Africa. Inter- (Eds.), Milankovitch Sea-Level Changes, Cycles, and Reser-
national Union of Geological Sciences, vol. 29, pp. 1 – 5. voirs on Carbonate Platforms in Greenhouse and Ice-House
Tucker, R.D., McKerrow, W.S., 1995. Early Paleozoic chronology: Worlds: Tulsa. SEPM (Society for Sedimentary Geology),
a review in light of new U – Pb zircon ages from Newfoundland Tulsa, OK. Short Course Notes No. 35, 81.
and Britain. Canadian Journal of Earth Science 32, 368 – 379.

Вам также может понравиться