Elrick & Snider, 2002 - Deep-Water Stratigraphic Cyclicity and Carbonate Mud Mound Development (Middle Cambrian)

Download as pdf or txt
Download as pdf or txt
You are on page 1of 27

Sedimentology (2002) 49, 1021–1047

Deep-water stratigraphic cyclicity and carbonate mud mound


development in the Middle Cambrian Marjum Formation,
House Range, Utah, USA
MAYA ELRICK* and ANNA C. SNIDER
*Earth and Planetary Sciences, University of New Mexico, Albuquerque, NM 87131, USA
(E-mail: [email protected])
US Department of Energy, Carlsbad, NM 88220, USA

ABSTRACT

In mid-Middle Cambrian time, shallow-water sedimentation along the


Cordilleran passive margin was abruptly interrupted by the development of
the deep-water House Range embayment across Nevada and Utah. The Marjum
Formation (330 m) in the central House Range represents deposition in the
deepest part of the embayment and is composed of five deep-water facies:
limestone–argillaceous limestone rhythmites; shale; thin carbonate mud
mounds; bioturbated limestone; and cross-bedded limestone. These facies
are cyclically arranged into 1Æ5 to 30 m thick parasequences that include
rhythmite–mound, rhythmite–shale, rhythmite–bioturbated limestone and
rhythmite–cross-bedded limestone parasequences. Using biostratigraphically
constrained sediment accumulation rates, the parasequences range in duration
from 14 to 270 kyr. The mud mounds are thin (<2 m), closely spaced,
laterally linked, symmetrical domes composed of massive, fenestral, peloidal
to clotted microspar with sparse unoriented, poorly sorted skeletal material,
calcitized bacterial(?) filaments/tubes and abundant fenestrae and stroma-
tactoid structures. These petrographic and sedimentological features suggest
that the microspar, peloids/clots and syndepositional micritic cement were
precipitated in situ from the activity of benthic microbial communities.
Concentrated growth of the microbial communities occurred during periods of
decreased input of fine detrital carbonate transported offshore from the
adjacent shallow-water carbonate platform. In the neighbouring Wah Wah
Range and throughout the southern Great Basin, coeval mid-Middle Cambrian
shallow-water carbonates are composed of abundant metre-scale, upward-
shallowing parasequences that record high-frequency (104)105 years) eustatic
sea-level changes. Given this regional stratigraphic relationship, the Marjum
Formation parasequences probably formed in response to high-frequency sea-
level fluctuations that controlled the amount of detrital carbonate input into
the deeper water embayment. During high-frequency sea-level rise and early
highstand, detrital carbonate input into the embayment decreased as a result of
carbonate factory retrogradation, resulting in the deposition of shale (base of
rhythmite–shale parasequences) or thin nodular rhythmites, followed by in situ
precipitated mud mounds (lower portion of rhythmite–mound parasequences).
During the ensuing high-frequency sea-level fall/lowstand, detrital carbonate
influx into the embayment increased on account of carbonate factory pro-
gradation towards the embayment, resulting in deposition of rhythmites
(upper part of rhythmite–mound parasequences), reworking of rhythmites by a
lowered storm wave base (cross-bedded limestone deposition) or bioturbation
of rhythmites by a weakened/lowered O2-minimum zone (bioturbated lime-
stone deposition). This interpreted sea-level control on offshore carbonate
Ó 2002 International Association of Sedimentologists 1021
13653091, 2002, 5, Downloaded from https://onlinelibrary.wiley.com/doi/10.1046/j.1365-3091.2002.00488.x by Korea University Library, Wiley Online Library on [07/05/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
1022 M. Elrick and A. C. Snider

sedimentation patterns is unique to Palaeozoic and earliest Mesozoic deep-


water sediments. After the evolution of calcareous plankton in the Jurassic,
the presence or absence of deeper water carbonates was influenced by a variety
of chemical and physical oceanographic factors, rather than just physical
transport of carbonate muds.
Keywords Cambrian, cyclicity, deep water, mud mounds, sea level.

INTRODUCTION GEOLOGICAL AND STRATIGRAPHIC


SETTING
Much of the understanding of Palaeozoic eustasy
comes from the study of relatively shallow- During latest Proterozoic to Late Devonian time,
marine cratonic deposits. This emphasis is be- 4–7Æ5 km of carbonate and siliciclastic marine
cause shallow-marine deposits are most sensitive deposits accumulated along the western edge of
to eustatic change, and deeper water deposits North America (Stewart & Poole, 1974; Levy &
lying outboard of the craton are commonly buried Christie-Blick, 1991). This westward-deepening,
or deformed by subsequent tectonic events. passive margin extended from north-western
Where correlative deeper water deposits are Canada to Mexico and was bordered on its
preserved, however, they provide an opportunity landward side (east) by the Transcontinental
to evaluate potential linkages between various Arch (Stewart, 1991). During the Cambrian, both
scales of sea-level change and associated sea-water eastern and western margins of North America
chemistry and physical oceanographic changes. were characterized by persistent, but temporally
These are best recorded in deeper water deposits shifting, carbonate and siliciclastic facies belts
because they represent a more continuous record (Robison, 1960; Aitken, 1978, 1997). The inner
of sedimentation. detrital facies belt lay closest to the craton and is
An instructive example of deeper water cratonic composed of shale, siltstone and sandstone; the
deposits is found in the Cambrian House Range middle carbonate belt (shallow-water carbonates)
embayment of western Utah and Nevada (Fig. 1). was situated seaward of the inner detrital belt.
This fault-generated basin accumulated nearly The outer detrital belt lay furthest offshore and
1000 m of deeper water Middle to Upper Cambrian is composed of shale and argillaceous carbonates
carbonates and fine siliciclastics. The majority of of deeper water origin. In the western USA, the
these deposits display sedimentary cyclicity at early Middle Cambrian shoreline (and associated
various scales including centimetre-scale (rhyth- inner detrital facies belt) trended approximately
mites), metre-scale (parasequences) and 100 m north–south and was located along the Utah–
scale depositional sequences. Of particular Colorado border and central Arizona (Fig. 1A).
interest is that many of the parasequences are In the mid-Middle Cambrian, north–south-
defined by the presence of thin, laterally linked, trending, passive-margin sedimentation patterns
carbonate mud mounds that can be correlated were disrupted by the development of a deeper
across the deepest water portion of the embay- water cratonic embayment that extended across
ment. In addition, coeval shallow-water deposits the former shallow-water carbonate platform in
across most of the Great Basin also display prom- the vicinity of central Nevada and western Utah
inent cyclicity at a similar scale to that observed in (House Range embayment; Fig. 1; Robison, 1964;
the embayment (e.g. Bond et al., 1991; Montañez & Rees, 1986). The abrupt development of the House
Osleger, 1993), thus permitting a comparison Range embayment resulted from vertical tectonic
between coeval shallow- and deep-water deposits. movements along a south-west–north-east-trend-
The main objectives of this paper are to: (1) ing normal fault, with the central and northern
describe and interpret the origin of deeper water House Range of western Utah occupying the
carbonate mud mounds in the upper Marjum downthrown block (Fig. 1; Kepper, 1976; Brady
Formation; (2) describe metre-scale parasequences & Koepnick, 1979; Rees, 1986). At its largest, the
(which include the mud mounds) occurring House Range embayment extended over 120 km
throughout the Marjum Formation; and (3) integ- from north to south and 400 km from east to west
rate local sedimentological and regional strati- (non-palinspastic distances; Rees, 1986). The
graphic data to interpret the origin of the northern margin maintained a ramp geometry
parasequences. throughout the embayment’s lifespan; the geom-
Ó 2002 International Association of Sedimentologists, Sedimentology, 49, 1021–1047
13653091, 2002, 5, Downloaded from https://onlinelibrary.wiley.com/doi/10.1046/j.1365-3091.2002.00488.x by Korea University Library, Wiley Online Library on [07/05/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Deep-water stratigraphic cyclicity and carbonate mud mound development 1023

A 116° 112° 108°


42°

NEVADA UTAH

WYOMING
COLORADO

N
FS

nHR
HOUSE RANGE
cHR
EMBAYMENT sHR

38° WW

0 100 km

study area

ARIZONA NEW MEXICO

DEEPER-WATER SHALLOW-WATER
DEPOSITS CARBONATES

B N S
FS nHR cHR sHR WW
SL

<<1 °
20 km

early Bolaspidella

FS nHR cHR sHR WW


SL

<1°

late Bolaspidella

Fig. 1. (A) Palaeogeographic map (non-palinspastic) of the House Range embayment during the Middle Cambrian
late Bolaspidella trilobite zone (late in the embayment’s lifespan). The southern embayment boundary is interpreted
as a normal fault. FS, Fish Springs Range; nHR, northern House Range (Swasey Peak); cHR, central House Range
(near Marjum Pass); sHR, southern House Range (Steamboat Pass); WW, northern Wah Wah Range, summit area
(from Robison, 1964). Approximate position of Middle Cambrian shoreline shown by dashed line (from Stewart &
Suczek, 1977). (B) Schematic north–south depositional profiles across the House Range embayment during early and
late Bolaspidella time. Note that the northern margin remains gentle throughout the embayment’s lifespan.

etry of the southern margin is not well constrained time, the embayment was completely filled,
because of the lack of Middle Cambrian outcrops and shallow-marine deposition resumed along a
in the southern House Range. By latest Cambrian north-trending passive margin (Fig. 1B).
Ó 2002 International Association of Sedimentologists, Sedimentology, 49, 1021–1047
13653091, 2002, 5, Downloaded from https://onlinelibrary.wiley.com/doi/10.1046/j.1365-3091.2002.00488.x by Korea University Library, Wiley Online Library on [07/05/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
1024 M. Elrick and A. C. Snider

The Middle Cambrian Marjum Formation con- (1) limestone–argillaceous limestone rhythmites;
formably overlies the deep-water Wheeler Shale (2) carbonate mud mounds; (3) shale; (4) biotur-
and is conformably overlain by the deep-water bated limestone; and (5) cross-bedded limestone
Weeks Limestone (Figs 2 and 3). The Marjum facies. Facies interpretations are given after the
Formation spans the majority of the Bolaspidella geochemical analysis discussion.
zone (polymerid trilobites) or part of the Ptych-
agnostus atavus, Ptychagnostus punctuosus and
Rhythmite facies
lower Lejopyge laevigata zones (agnostid trilo-
bites; Fig. 2; Robison, 1976, 1984). The thin, interbedded limestone and argillaceous
limestone beds of the rhythmite facies are strik-
ingly uniform in thickness and lateral continuity
FACIES DESCRIPTION (Fig. 4A). Dark grey limestone beds are 2–10 cm
thick and are characterized by laminated to
In the central House Range study area, the massive pelleted microspar (originally micrite)
Marjum Formation is 330 m thick and accumu- with sparse sponge spicules and burrows, rare
lated within the deepest part of the House Range agnostid trilobites and rare ripple cross-laminae
embayment. It is composed predominantly of (with eastward or onshore palaeoflow directions).
deep-water ‘rhythmites’ (thin, interbedded lime- The majority of grains are less than 50 lm
stone and argillaceous limestone beds) and sub- (medium to coarse silt size), oval, moderately
ordinate shales (Fig. 3). The upper 70 m are well sorted, medium to dark brown pellets com-
characterized by deep-water rhythmites interbed- posed of microspar that float within the microspar
ded with thin (<2 m) carbonate mud mounds. In matrix, forming a pellet wackestone to packstone
the study area, there are five depositional facies texture (Fig. 4B). The term pellet is used to
recognized in the Marjum Formation including: differentiate oval to spherical, moderately well-

TRILOBITE ZONES S. NEVADA


OD

SE CENTRAL WAH WAH &


RI

Polymerid/ HOUSE RANGE S. HOUSE RANGE


PE

Agnostids CALIFORNIA
Nopah Fm
Aphelaspis
UPPER CAMBRIAN

Orr Fm Orr Fm

Crepicephalus
and
Cedaria Weeks Wah Wah
Formation

Limestone Summit Fm

Lejopyge
CAMBRIAN

calva
Banded Trippe
Limestone
King

Mountain
Bolaspidella
CAMBRIAN

Ptychagnostus Marjum
Member
punctuosus
Formation
Pierson Cove
Bonanza

Formation
MIDDLE

P. atavatus
Wheeler Eye of the Needle
P. gibbus Shale Limestone Fig. 2. Chronostratigraphic diagram
Ehmaniella

Swasey Lst Swasey Lst of Middle and Upper Cambrian


P. praecurrens deposits across the House, Fish
Whirlwind Fm Whirlwind Fm Springs and Wah Wah Ranges
(from Hintze & Robison, 1975).

Ó 2002 International Association of Sedimentologists, Sedimentology, 49, 1021–1047


13653091, 2002, 5, Downloaded from https://onlinelibrary.wiley.com/doi/10.1046/j.1365-3091.2002.00488.x by Korea University Library, Wiley Online Library on [07/05/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Deep-water stratigraphic cyclicity and carbonate mud mound development 1025

WEEKS
LIMESTONE FACIES
Limestone-dominated
rhythmite facies

Argillaceous limestone-
dominated rhythmite facies

Mud mound facies

Shale facies

Shallow-water limestones

MARJUM
FORMATION PARASEQUENCE TYPE

Rhythmite-mound

Rhythmite-shale

Rhythmite-bioturbated
limestone

Rhythmite-crossbedded
limestone

Approximate location of
detailed stratigraphic columns
shown in Figure 10

Fig. 3. Generalized stratigraphic WHEELER


column of the Wheeler Shale and SHALE

100 metres
Marjum Formation at Marjum Pass,
central House Range, showing
depositional facies and stratigraphic
distribution of parasequence types.
See Fig. 10 for detailed partial
stratigraphic columns at positions SWASEY
shown by arrows. LIMESTONE

sorted, silt-size carbonate grains from peloids, monly accentuated by compaction and pressure
which are larger (fine to medium sand size), more solution; however, gradations (over <0Æ25 cm)
irregular in shape and more poorly sorted. Lam- between the two rock types also occur. Laminae
inated argillaceous limestone layers (0Æ25–10 cm in both rock types are defined by submillimetre-
thick) are composed of tan to reddish, dolomitic, thick graded pellet and microspar layers. The
argillaceous pelleted microspar with sparse majority of individual limestone and argillaceous
sponge spicules and agnostid trilobites. The argil- limestone beds are strikingly uniform in thick-
laceous material is composed of subangular ness and lateral continuity (i.e. they form parallel-
quartz silt and illite, and ranges in abundance bedded layers) (Fig. 4A). A consistent exception
between 10 and 70 weight per cent (wt%) vs. 5–25 to this occurs in the upper Marjum Formation;
wt% in limestone layers. The fine skeletal mater- 10–60 cm below each carbonate mud mound
ial in both layers is abraded, and trilobite carap- layer, individual limestone layers decrease in
aces generally lie with their concave side up, thickness and become nodular, whereas underly-
suggesting deposition from suspension settling. ing and overlying argillaceous limestone layers
Individual limestone–argillaceous limestone maintain similar thicknesses (Fig. 5). This dec-
couplets range in thickness from 3 to 20 cm and rease in limestone bed thickness occurs beneath
average 5 cm (Fig. 4A). Contacts between lime- carbonate mud mounds as well as thinner inter-
stone and argillaceous limestone beds are com- mound deposits (Fig. 5), suggesting that thinning
Ó 2002 International Association of Sedimentologists, Sedimentology, 49, 1021–1047
13653091, 2002, 5, Downloaded from https://onlinelibrary.wiley.com/doi/10.1046/j.1365-3091.2002.00488.x by Korea University Library, Wiley Online Library on [07/05/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
1026 M. Elrick and A. C. Snider

Fig. 4. (A) Limestone-dominated


rhythmite facies in upper Marjum
Formation at Section A. Limestone
layers are dark grey; light grey layers
are argillaceous limestone. (B) Pho-
tomicrograph (plane-polarized light)
of rhythmite limestone from upper
Marjum Formation at Section C.
Fine laminae are composed of gra-
ded pellets and microspar. Photo-
graph and sample locations shown
in Fig. 9.

Ó 2002 International Association of Sedimentologists, Sedimentology, 49, 1021–1047


13653091, 2002, 5, Downloaded from https://onlinelibrary.wiley.com/doi/10.1046/j.1365-3091.2002.00488.x by Korea University Library, Wiley Online Library on [07/05/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Deep-water stratigraphic cyclicity and carbonate mud mound development 1027

Fig. 5. (A) Carbonate mud mound


layer (Mound 2, Section B); top and
bottom contact partially outlined
with white lines. Note the nodular
character of rhythmites directly
below the mounds. Scale bar is 1 m.
(B) Carbonate mud mound layer
(Mound 5, Section B). Note that
limestone layers within rhythmite
couplets thin directly below the
mud mound and show nodular
geometry particularly below inter-
mound deposits. Scale bar is 0Æ5 m.
Photograph locations shown on
Fig. 9.

during burial compaction was not the cause; if it limestone-dominated vs. argillaceous limestone-
were, the limestone beds would thin more di- dominated stratigraphic intervals are gradational
rectly beneath the carbonate mud mounds than in over tens of centimetres.
intermound areas.
The proportion of limestone to argillaceous
Carbonate mud mound facies
limestone in individual rhythmite couplets varies
throughout the formation. Intervals of argilla- The Marjum Formation carbonate mud mounds
ceous limestone-dominated rhythmites (intervals are carbonate buildups with depositional relief
in which argillaceous limestone beds are thicker above the sea floor and are composed dominantly
than interbedded limestone beds) vary in thick- of carbonate mud and peloidal mud, i.e. they are
ness from several metres to decimetres and not framework supported (Bosence & Bridges,
predominate in the lower and middle Marjum 1995). Marjum Formation mud mounds are small
Formation (Fig. 3). Intervals of limestone-domin- symmetrical, dome-shaped buildups that are 0Æ4–
ated rhythmites (individual limestone beds thicker 2 m thick and between 0Æ5 and 2 m in diameter.
than interbedded argillaceous limestone beds) They are closely spaced (0Æ5–3 m) and laterally
dominate the middle and parts of the upper connected to adjacent mounds by deposits that
Marjum Formation (Fig. 3). Contacts between are similar in all aspects to the mud mound
Ó 2002 International Association of Sedimentologists, Sedimentology, 49, 1021–1047
13653091, 2002, 5, Downloaded from https://onlinelibrary.wiley.com/doi/10.1046/j.1365-3091.2002.00488.x by Korea University Library, Wiley Online Library on [07/05/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
1028 M. Elrick and A. C. Snider

facies, except that the intermound beds lack bedding) in shape and range from a few millime-
topographic relief (Fig. 5). The mud mound tres to 1 cm in length and height (Fig. 7B). They are
slopes range from 10° to 50°. Within each mud filled with coarse, cloudy calcite cement, which is
mound-bearing bed, individual mud mounds irregularly replaced by microcrystalline quartz or,
were spaced laterally in a three-dimensional array less commonly, by coarse, equant quartz crystals.
along the sea floor (Fig. 6). Between 11 and 14 Stromatactoid cavities are most abundant and
mud mound layers occur across the field area; the largest in the cores of mud mounds; however,
greatest number of mud mound layers occurs in they also occur within intermound deposits.
the most western stratigraphic Section I (situated Stromatactoid cavities are between 0Æ5 and 7 cm
in a more offshore location). thick and a few centimetres to tens of centimetres
Mud mounds are composed of medium to dark long; most are oriented subparallel to bedding,
grey, clotted to peloidal microspar with abundant but vertical to near-vertical cavities commonly
fenestrae and stromatactoid structures, sponge connect adjacent horizontal cavities (Fig. 8B).
spicules, and rare trilobites (Fig. 7). The micro- Cavity roofs and floors vary from smooth to
spar (originally micrite) grades from distinct oval irregular and are filled by a combination of fine
to spherical peloids (60–100 lm; very fine sand) geopetal sediment and/or coarse calcite cement. If
to vague clotted textures surrounded by light geopetal sediment is present, it lies directly on
brown to clear microspar (Fig. 7B). Unabraded the cavity floor and does not alternate with the
monaxon and tetraxon sponge spicules are re- cements. The unfossiliferous geopetal sediment
placed by calcite and float within the microspar; ranges from microspar to peloidal microspar and
trilobite bioclasts are poorly sorted, unoriented is massive or rarely laminated. The first gen-
and show little evidence of abrasion. Scanning eration of stromatactoid cement is composed
electron microscope (SEM) analysis of the micro- dominantly of inclusion-rich, isopachous bladed
spar, peloids and clots reveals submicrometre- calcite (< 1 mm-long crystals; clear in hand speci-
wide, calcitized filaments and tubes in voids men); the second generation is characterized
between calcite crystals (Fig. 8A). by inclusion-rich, equant, very coarse calcite
Fenestrae are abundant in all mud mounds (3–8 mm; white in hand specimen). Stroma-
(locally comprising up to 30% of the rock) and tactoid structures are always associated with
are slightly less abundant in intermound deposits. fenestrae; in particular, fenestrae are more con-
Fenestrae are irregular to elongate (subparallel to centrated directly above stromatactoid roofs.

20 km

SHALLOW-WATER CARBONATES DEEP-WATER DEPOSITS


and MUD MOUNDS
Fig. 6. Schematic block diagram illustrating the three-dimensional growth array of mud mound layers within the
House Range embayment and the adjacent shallow-water carbonate platform.

Ó 2002 International Association of Sedimentologists, Sedimentology, 49, 1021–1047


13653091, 2002, 5, Downloaded from https://onlinelibrary.wiley.com/doi/10.1046/j.1365-3091.2002.00488.x by Korea University Library, Wiley Online Library on [07/05/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Deep-water stratigraphic cyclicity and carbonate mud mound development 1029

Fig. 7. Photomicrographs (plane-


polarized light) of mud mound
facies. (A) Fine skeletal wackestone
with replaced sponge spicules (S),
trilobites (T), miscellaneous skeletal
debris and small fenestrae (F). (B)
Peloidal (P) to clotted microspar and
spar-filled fenestrae (F). Location of
samples shown in Fig. 9.

study area and reported graded laminae, rare


Shale facies
soft-sediment deformation folds and low-angle
The shale facies is most common in the lower truncation surfaces. Fine current-oriented agnos-
and middle Marjum Formation and occurs as 0Æ5 tid trilobites, fragmented filamentous algae and
to 11 m thick intervals (Fig. 3). The shale is rare tool and sole marks were also recorded.
composed of dark grey, fissile to platy, calcar-
eous clayshale to mudshale with sparse agnostid
Bioturbated limestones
trilobites and rare bioturbation. Near contacts
with associated rhythmites, the shale contains In the lower Marjum Formation, rhythmite facies
centimetre-thick, laterally discontinuous lenses are interbedded with thin (0Æ3–2 m thick), biotur-
of fine pelletal microspar (Fig. 10). Contacts bated to slightly nodular-bedded limestones
with rhythmite facies are gradational over tens of (Fig. 10). Limestones are composed of bioturba-
centimetres. Rogers (1984) studied shales within ted, pelleted microspar with sparse skeletal
the lower Marjum Formation 5 km north of the material (polymerid trilobites, small thin-walled
Ó 2002 International Association of Sedimentologists, Sedimentology, 49, 1021–1047
13653091, 2002, 5, Downloaded from https://onlinelibrary.wiley.com/doi/10.1046/j.1365-3091.2002.00488.x by Korea University Library, Wiley Online Library on [07/05/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
1030 M. Elrick and A. C. Snider

Fig. 8. (A) SEM image of clotted


microspar in mud mounds. White
arrows point to abundant submi-
crometre, calcitized bacteria(?) fila-
ments or tubes linking calcite
crystals. (B) Stromatactoid struc-
tures near core of mud mound 2.
White arrows point to geopetal
sediment (S) and cement (C). Geop-
etal sediment has been partially
dolomitized, imparting a lighter
colour than the host limestone. Note
horizontal to vertical orientation of
cavities. Tip of finger for scale at
bottom left. Location of samples
shown in Fig. 9.

brachiopods and sponge spicules) and rare lam- forms indicates eastward-directed (onshore) dune
inae. Contacts with parallel-bedded rhythmite migration.
facies are sharp or gradational over a few centi-
metres.
SEDIMENTARY CYCLICITY
Cross-bedded limestones
Cyclic repetition of the five depositional facies
Five cross-bedded limestone beds (10–30 cm defines metre- to decimetre-scale packages ter-
thick) are present in the middle Marjum Forma- med parasequences. Four parasequence types
tion. They are composed of pelletal microspar (Fig. 10) are recognized in the Marjum Formation
with sparse bioturbation (Fig. 10). These beds including rhythmite–mound, rhythmite–shale,
contain well-preserved dune forms between 7 rhythmite–bioturbated limestone and rhythmite–
and 15 cm high, with 0Æ3–1 m wavelengths and cross-bedded limestone parasequences. Parase-
nearly symmetrical, north–south-trending crests. quence origins are discussed after presentation of
Slight but consistent asymmetry of dune the geochemical and regional stratigraphic data.
Ó 2002 International Association of Sedimentologists, Sedimentology, 49, 1021–1047
13653091, 2002, 5, Downloaded from https://onlinelibrary.wiley.com/doi/10.1046/j.1365-3091.2002.00488.x by Korea University Library, Wiley Online Library on [07/05/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Deep-water stratigraphic cyclicity and carbonate mud mound development 1031

<20 cm. Ten rhythmite–mound parasequences


Parasequences
are correlated across the study area (Fig. 9).
Rhythmite–mound parasequences
Rhythmite–shale parasequences
Rhythmite–mound parasequences (2–24 m thick)
are present in the upper 70 m of the Marjum Ten rhythmite–shale parasequences (6–30 m
Formation (Fig. 3). A typical parasequence con- thick) are present in the lower and middle
sists of a 10–50 cm base of nodular-bedded Marjum Formation (Fig. 3). They are character-
rhythmites overlain by mud mound facies, which ized by a thin, basal interval of shale interbedded
are sharply overlain by parallel-bedded rhyth- with rare limestone lenses, gradationally overlain
mites (Figs 5 and 10). The contacts between by uniform shale and gradationally capped by
stacked parasequences are gradational over rhythmite facies (Fig. 10). The contact between

W E
OFFSHORE
W Ekm
N

E
10

NG
RA
E
I

US
HO
L
RA
10
A

NT
CE
9 B c Marjum Pass
A
8b I Road
8a
?
B
8
C
10
9
8
7

Fig. 4a

6
5
7

4b
6
? 5
4a
Fig. 5b
Figs 7, 8a
4 4
3 3
2a
?
2 2
Fig. 9. Correlation of upper Marjum 1
Formation mud mounds 1 to 11 1
across the study area near Marjum Figs 5a, 8b
Pass, central House Range. A greater Mud mounds
number of mud mounds are present
in the offshore Section I. Inset map 10 m Rhythmites
shows the location of measured
sections near Marjum Pass Road;
grey stippled pattern refers to Weeks Limestone
Palaeozoic outcrop. 2 km
Ó 2002 International Association of Sedimentologists, Sedimentology, 49, 1021–1047
13653091, 2002, 5, Downloaded from https://onlinelibrary.wiley.com/doi/10.1046/j.1365-3091.2002.00488.x by Korea University Library, Wiley Online Library on [07/05/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
1032 M. Elrick and A. C. Snider

Rhythmite-mound Rhythmite-shale Rhythmite-bioturbated Rhythmite-crossbedded


parasequence parasequence limestone parasequence limestone parasequence

Rhythmites F Fenestrae
F

Nodular rhythmites Limestone lenses

Bioturbated T Trilobites
3m limestones

F
Shale Bioturbation

Mud mounds Skeletal debris

Parasequence
0 Stromatactoid
top
Cross bedded
limestone

Fig. 10. Partial stratigraphic columns from Section A (see Fig. 3 for stratigraphic locations) illustrating typical
parasequence types. Rhythmite–mound parasequences occur in the upper Marjum, rhythmite–shale parasequences
in the lower and middle Marjum, and rhythmite–bioturbated limestone and rhythmite–cross-bedded limestone
parasequences occur in the middle Marjum Formation.

successive parasequences is commonly grada- Parasequence duration


tional over <0Æ5 m.
The average duration of parasequences is estima-
ted using parasequence thickness divided by
Rhythmite–bioturbated limestone
average sediment accumulation rates. Accumula-
parasequences
tion rates were calculated using the total strati-
Twelve stacked rhythmite–bioturbated limestone graphic thickness of Middle Cambrian deposits in
parasequences (1Æ5–6Æ5 m thick) occur in the the central House Range (thickness of 1020 m
middle Marjum Formation (Fig. 3) and are char- defined by trilobite zones; Robison, 1984) divided
acterized by basal nodular-bedded rhythmite by the duration of the Middle Cambrian. The
facies gradationally overlain by bioturbated lime- duration of the Middle Cambrian varies widely
stone facies (Fig. 10). Contacts between parase- depending on which time scale is used (see for
quences are sharp. example Elrick & Hinnov, 1996). Recently, Bowr-
ing & Erwin (1998) reported an isotopic age for the
Rhythmite–cross-bedded limestone top of the Middle Cambrian at 500 Myr (no
parasequences uncertainties cited), and the base of the Middle
Cambrian is reported at 509 Myr (Young & Laurie,
Three of these parasequences occur in the
1996; no uncertainties cited). Using this 9 Myr
middle Marjum Formation (1Æ5–6 m thick;
duration for the Middle Cambrian results in an
Fig. 3). They are composed of rhythmite facies
average sediment accumulation rate within the
abruptly overlain by cross-bedded limestone
study area of 11 cm kyr)1. This rate lies within
facies (Fig. 10). Contacts between parasequences
the range determined for the central House Range
are sharp.
Ó 2002 International Association of Sedimentologists, Sedimentology, 49, 1021–1047
13653091, 2002, 5, Downloaded from https://onlinelibrary.wiley.com/doi/10.1046/j.1365-3091.2002.00488.x by Korea University Library, Wiley Online Library on [07/05/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Deep-water stratigraphic cyclicity and carbonate mud mound development 1033

stratigraphic thickness using a variety of pub- MAT Delta E isotope ratio mass spectrometer at
lished Cambrian time scales (4–21 cm kyr)1; the University of New Mexico. Reproducibility on
Elrick & Hinnov, 1996). Given the 11 cm kyr)1 replicated samples and the standard was better
average sediment accumulation rate calculated in than 0Æ1% for both carbon and oxygen isotopes.
this study, the Marjum Formation parasequences All isotope data are reported using the conven-
range in duration from between 14 and 270 kyr. tional d notation relative to VPDB.

GEOCHEMICAL ANALYSIS Results


Insoluble residues
To aid in interpreting the origin of the mud
mounds and the controls on their formation, the The wt% IRs of mud mounds ranged between 3
Marjum Formation was sampled for weight per and 10 wt% (average 6Æ6 wt%). Weight per cent IR
cent insoluble residues (wt% IR), total organic for rhythmite limestone ranges between 5 and 24
carbon content (TOC) and stable isotope analysis. wt% (average 14Æ3 wt%) (Fig. 11A, Table 1). The
lower wt% IR in mud mounds indicates that,
during mud mound deposition, either less silic-
Methods iclastic material reached the basinal study area
and/or carbonate accumulation and early cemen-
Insoluble residues
tation related to mud mound formation diluted
At selected stratigraphic sections, rhythmite lime- the background siliciclastic influx.
stone layers were collected every 1Æ5 m, then
every 10 cm directly below mud mounds. In Total organic carbon
addition, various mud mounds at each measured
TOC values range from below detection limits (0Æ1
section were collected, producing a total of 115
wt%) to 0Æ85 wt%, with shale facies containing
samples. Samples were powdered, weighed, then
the highest TOC values and rhythmite facies
dissolved in 0Æ5 m HCl for 24 h. Insoluble resi-
typically containing below 0Æ3 wt%. No TOC
dues were filtered and dried overnight, then
trends were observed between rhythmite and
reweighed to determine the wt% IR.
mud mound facies.
Total organic carbon
Stable isotopes
A smaller subset of samples from rhythmite
Carbon isotope values from rhythmite limestones
limestone, argillaceous limestone, mud mound
range between )1Æ53& and +1Æ35&, with an aver-
and shale facies were collected for TOC analysis
age of +0Æ52&, whereas d13C values of mud mounds
(44 samples). TOC values were determined using a
are between )1Æ92& and +0Æ83&, with an average
colorimetric method for carbon determination
of )0Æ08& (Fig. 11B; Table 2). The majority of d18O
(Metson, 1961). Organic carbon was extracted by
values range between )8% and )12&, with a mean
dissolution with HCl, followed by organic matter
of )9Æ8&. Approximately 20% of mud mound
oxidation with chromium trioxide. Measurements
samples have d18O values that are less than the
of total chromium trioxide used for oxidation were
Middle Cambrian range reported by Veizer et al.,
determined by spectrophotometric comparison
1999), whereas 40% of the rhythmite limestone
with a sucrose standard (Blakemore et al., 1987).
samples have values that are less than the Middle
Cambrian range (Fig. 11B). This trend may reflect
Stable isotopes
syndepositional to very early diagenetic marine
The same suite of samples collected for insoluble cementation of mud mound limestones and a more
residues was analysed for carbon and oxygen closed pore fluid system during diagenesis. Rhyth-
isotopes (except Section C; Fig. 9); in addition, mite limestones display little evidence of synde-
two samples of stromatactoid cement were ana- position/early marine cementation, and thus their
lysed, producing a total of 80 samples. Analyses oxygen isotope values were more likely to be reset
were conducted on bulk rock (pelleted and peloi- to lower values during burial diagenesis. The two
dal microspar) and individual crystals of stroma- stromatactoid cement samples have the highest
tactoid cement. Carbonate samples were reacted d18O values and some of the lowest d13C values
under vacuum with purified orthophosphoric (Fig. 11B). The low d13C values may reflect some
acid (McCrea, 1950) and measured on a Finnigan influence of organic matter diagenesis.
Ó 2002 International Association of Sedimentologists, Sedimentology, 49, 1021–1047
13653091, 2002, 5, Downloaded from https://onlinelibrary.wiley.com/doi/10.1046/j.1365-3091.2002.00488.x by Korea University Library, Wiley Online Library on [07/05/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
1034 M. Elrick and A. C. Snider

12
A
Mounds
10
Rhythmites
FREQUENCY

0
0 5 10 15 20 25 30

INSOLUBLE RESIDUE (wt%)

2
B 1.5

Middle Cambrian 1

0.5
13
δ C
0
( o/oo VPDB)

Rhythmites -0.5
1
Mounds -1

2 Stromatactoid -1.5
cement
-2

-16 -14 -12 -10 -8 -6 -4 -2 0

18
δ O ( o/ooVPDB)
Fig. 11. (A) Histogram of wt% insoluble residues in limestone rhythmites and mud mound facies. (B) Carbon and
oxygen isotope cross-plot of rhythmite limestone, mud mounds and stromatactoid cement (1, earliest clear calcite
cement; 2, later white calcite cement). Field labelled Middle Cambrian is range of Middle Cambrian isotope values
reported by Veizer et al. (1999) (d18O and d13C) and Montañez et al. (2000) (d13C).

Unlike oxygen, d13C values are only slightly range (Fig. 11B and Table 2), and both lie within
affected by diagenesis in carbonate rocks because the range of values reported for other Middle
the carbon composition of diagenetic fluids is Cambrian deposits of the Great Basin (e.g. Braiser,
buffered by the host carbonates (Marshall, 1992). 1993; Montañez et al., 1996, 2000; Saltzman et al.,
Marjum Formation d13C values are interpreted as 1998) and globally (Veizer et al., 1999). The
primary, reflecting the composition of the fluids similarity of d13C values for both facies and their
that precipitated the carbonate. The d13C values of near-zero mean values indicate that the fluids
mud mounds and rhythmites generally overlap in responsible for mud mound carbonate precipita-
Ó 2002 International Association of Sedimentologists, Sedimentology, 49, 1021–1047
13653091, 2002, 5, Downloaded from https://onlinelibrary.wiley.com/doi/10.1046/j.1365-3091.2002.00488.x by Korea University Library, Wiley Online Library on [07/05/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Deep-water stratigraphic cyclicity and carbonate mud mound development 1035
Table 1. Insoluble residue contents.

Facies Sample wt%IR Facies Sample wt%IR Facies Sample wt%IR


Rhythmite Rhythmite Rhythmite
limestone E-23 7Æ9 limestone B-133 13Æ0 limestone C-995 10Æ7
E-21 11Æ6 B-101 8Æ5 C-996 12Æ0
E-19 8Æ9 B-103 9Æ8 C-997 7Æ7
E-17 17Æ6 B-104 15Æ7 C-998 9Æ0
E-15 17Æ1 B-105 16Æ81 C-999 12Æ2
E-13 20Æ4 B-105b 6Æ2
E-11 15Æ2 B-106 12Æ6
E-10 23Æ4 B-107 14Æ9 Mud
J-19 11Æ0 B-108 10Æ2 mound A-38 6Æ1
J-18 12Æ0 B-109 10Æ8 A-36 4Æ4
J-17 11Æ1 B-110 11Æ9 A-5 4Æ4
J-16 12Æ2 B-111 16Æ0 J-20 5Æ4
J-15 13Æ9 B-112 12Æ3 J-5 4Æ7
J-14 13Æ7 B-113 9Æ0 B2-T 10Æ8
J-13 10Æ3 B-114 13Æ1 B2-C 4Æ1
J-12 14Æ1 C-9 10Æ1 B3-C1 4Æ7
J-11 15Æ3 C-2 8Æ8 B3-T2 4Æ1
J-10 11Æ5 C-3 12Æ6 B2-C2 5Æ2
J-9 13Æ0 C-4 17Æ6 B-31 9Æ0
J-8 15Æ0 C-7 13Æ6 B-30 3Æ4
J-7 13Æ5 C-8 23Æ9 B-29 5Æ2
J-6 15Æ0 C-101 28Æ1 B-28 3Æ8
A-35 12Æ4 C-200 18Æ6 B-27 2Æ0
A-34 9Æ4 C-300 10Æ8 B-26 5Æ8
A-4 17Æ1 C-400 17Æ8 B-116 7Æ4
A-1 6Æ6 C-500 21Æ4 B-134 6Æ8
B-1 7Æ6 C-600 18Æ5 B-100 4Æ7
B-117 13Æ3 C-700 17Æ5 B-115 10Æ7
B-118 8Æ2 C-800 22Æ6 C-100 21Æ7
B-119 15Æ0 C-900 32Æ0 C-10 10Æ3
B-120 14Æ1 C-910 18Æ9 C-11 10Æ1
B-121 13Æ5 C-920 14Æ8 C-12 5Æ6
B-122 15Æ8 C-930 16Æ7 C-13 9Æ4
B-123 10Æ8 C-940 14Æ5 C-1 3Æ6
B-124 14Æ8 C-950 16Æ6 I2-T 4Æ3
B-125 20Æ4 C-960 16Æ5 I3-C2 6Æ4
B-126 22Æ1 C-970 20Æ5
B-127 11Æ3 C-980 16Æ8
B-128 24Æ9 C-990 14Æ3
B-129 9Æ8 C-991 15Æ9
B-130 9Æ3 C-992 13Æ1
B-131 9Æ8 C-993 9Æ2
B-132 10Æ1 C-994 14Æ7

tion were not sourced from: (1) methane-bearing & Chafetz, 2002). Instead, we interpret the carbon
or magmatic fluids (e.g. Beauchamp & Savard, isotope values of the mud mounds to reflect
1992; Mounji et al., 1998); (2) 12C-enriched fluids precipitation from deeper (substorm wave base)
affected by organic matter diagenesis; or (3) fluids marine or near-marine sea water.
generated during bacterial fermentation (Wu &
Chafetz, 2002). In the first two cases, the d13C
values would be expected to be significantly lower
FACIES INTERPRETATIONS
than the observed range, whereas bacterial fer-
mentation strongly fractionates carbon isotopes
Rhythmite facies
with the CO2 (and consequently the bicarbonate)
generated having distinctly positive d13C values Limestone and argillaceous limestone beds con-
(Irwin et al., 1977; Anderson & Arthur, 1983; Wu tain laminae, rare oriented skeletal fossils and
Ó 2002 International Association of Sedimentologists, Sedimentology, 49, 1021–1047
13653091, 2002, 5, Downloaded from https://onlinelibrary.wiley.com/doi/10.1046/j.1365-3091.2002.00488.x by Korea University Library, Wiley Online Library on [07/05/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
1036 M. Elrick and A. C. Snider
Table 2. Stable isotope data (& VPDB).

Facies Sample d18O d13C Facies Sample d18O d13C


Rhythmite Rhythmite
limestone A-39 )11Æ07 0Æ52 limestone B-121 )8Æ74 0Æ51
E-28 )10Æ76 1Æ35 B-122 )9Æ38 0Æ57
E-21 )10Æ39 1Æ35 B-123 )9Æ38 0Æ68
E-19 )12Æ28 0Æ98 B-124 )10Æ23 0Æ24
E-17 )12Æ11 0Æ58 B-125 )10Æ23 0Æ14
E-15 )10Æ45 0Æ90 B-126 )11Æ00 0Æ27
E-13 )10Æ56 0Æ97 B-127 )9Æ88 0Æ35
E-10 )10Æ46 1Æ07 B-128 )9Æ29 0Æ20
E-11 )9Æ97 1Æ22 B-129 )9Æ22 0Æ61
J-19 )13Æ16 0Æ38 B-130 )8Æ70 0Æ55
J-18 )12Æ33 0Æ45 B-131 )8Æ75 0Æ04
J-17 )10Æ70 0Æ44 B-132 )9Æ01 )0Æ23
J-16 )12Æ56 0Æ56 C-200 )8Æ49 0Æ70
J-15 )10Æ07 0Æ42 C-300 )8Æ64 0Æ58
J-14 )11Æ56 0Æ75 C-800 )8Æ34 0Æ79
J-13 )10Æ38 0Æ52 C-940 )8Æ47 0Æ71
J-12 )10Æ55 0Æ05 C-995 )9Æ28 1Æ06
J-11 )9Æ86 0Æ53 C-990 )8Æ03 0Æ63
J-10 )10Æ74 0Æ46
J-9 )10Æ55 0Æ54 Mud mound A-38 )9Æ41 )0Æ08
J-8 )10Æ33 0Æ14 J-20 )10Æ89 0Æ36
J-7 )10Æ60 0Æ70 J-5 )9Æ77 )0Æ75
J-6 )9Æ90 0Æ44 B2-T )9Æ09 0Æ00
B-101 )10Æ27 0Æ68 B2-C )9Æ20 )0Æ02
B-103 )9Æ81 0Æ82 B2-C2 )8Æ57 0Æ19
B-104 )9Æ10 0Æ85 B3-C1 )9Æ02 )0Æ02
B-105 )8Æ98 0Æ72 B3-T )9Æ07 0Æ83
B-105b )8Æ52 0Æ94 B3-T2 )11Æ25 0Æ15
B-106 )9Æ68 0Æ61 I3-C2 )9Æ19 0Æ71
B-107 )9Æ51 0Æ55 B-134 )9Æ50 )0Æ33
B-108 )8Æ83 0Æ75 B-116 )8Æ83 0Æ08
B-109 )8Æ75 0Æ52 B-115 )9Æ00 0Æ60
B-110 )8Æ75 0Æ57 B-100 )9Æ21 0Æ33
B-111 )9Æ01 0Æ37 C-100 )9Æ04 0Æ30
B-112 )10Æ07 )0Æ30 C-13 )13Æ48 )1Æ23
B-113 )9Æ18 )0Æ43 C-12 )9Æ12 )1Æ92
B-114 )9Æ36 )1Æ53 C-11 )8Æ90 )0Æ58
B-117 )10Æ81 0Æ37 C-10 )9Æ69 )0Æ04
B-118 )10Æ54 0Æ29
B-119 )8Æ87 0Æ48 Stromatactoid c-1 )7Æ57 )0Æ67
B-120 )9Æ79 0Æ59 cement c-2 )7Æ79 )1Æ51

rare ripple cross-laminae, suggesting deposition zoic, so the majority of fine carbonate is most
in quiet waters below and near storm wave base; likely detrital in origin and was probably gener-
the paucity of bioturbation suggests low bottom- ated in shallow-water regions and transported
water oxygen levels related to an O2-minimum offshore by seaward-directed currents. Some of
zone. Individual submillimetre-thick graded lam- the fine carbonate also includes micritic cement.
inae are interpreted to represent suspension- Spectral analysis of the Marjum limestone–
settling deposition from a single storm event or argillaceous limestone couplets reveals peaks
single dilute density current; therefore, each with wavelengths of between 4 and 9 cm and,
limestone and argillaceous limestone bed con- using a range of biostratigraphically controlled
tains many tens to hundreds of events. Calcareous sediment accumulation rates (4–21 cm kyr)1;
plankton such as planktonic foraminifera and Elrick & Hinnov, 1996), this suggests that the
coccolithophorids did not evolve until the Meso- individual limestone–argillaceous limestone cou-

Ó 2002 International Association of Sedimentologists, Sedimentology, 49, 1021–1047


13653091, 2002, 5, Downloaded from https://onlinelibrary.wiley.com/doi/10.1046/j.1365-3091.2002.00488.x by Korea University Library, Wiley Online Library on [07/05/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Deep-water stratigraphic cyclicity and carbonate mud mound development 1037

plets represent between 200 and 2100 years and poorly sorted skeletal material; these latter
(millennial scale; Elrick & Hinnov, 1996). Sedi- features are inconsistent with deposition by nor-
mentological and spectral results indicate that the mal hydraulic processes. On the other hand,
rhythmic interbedding represents very short-term rhythmite limestones contain laminae of silt-size
fluctuations in: (1) fluvial and/or aeolian material pellets and microspar, with rare oriented and
derived from the exposed craton (Transconti- abraded fine skeletal material; these features sug-
nental Arch); or (2) the intensity or location of gest deposition from discrete, successive deposi-
storm-generated or dilute density currents that tional episodes (suspension-settling events).
transported fine detrital carbonate offshore into In addition to an in situ origin for the fine
the deep-water embayment. Both these might be carbonate, several lines of evidence suggest that
controlled by short-term tectonic, eustatic or the mud mounds were bound by micritic cement
climate changes. Given the stratigraphic and and/or were biologically stabilized during or very
lateral persistence of the rhythmites within the soon after deposition: (1) the lack of slump or
study area, the occurrence of millennial-scale gravity flow deposits along mound flanks, despite
records in other Palaeozoic–Mesozoic deeper the fact that many mound slopes are greater 35°;
water successions (Anderson, 1982; Burchell (2) little or no evidence for internal mound
et al., 1990; Elrick et al., 1991) and the strong compaction; and (3) abundant stromatactoid cav-
millennial-scale signal in many Pleistocene– ities filled with fine geopetal sediment and spar-
Holocene records (e.g. Pisias et al., 1973; Grootes filled fenestrae.
et al., 1993; Bond & Lotti, 1995; Raymo et al., Apart from the calcitized filaments and tubes,
1998; Bond et al., 2001), it is likely that the there is no direct evidence for the presence of a
Marjum rhythmites represent climatically con- lime mud-producing organism, but a combination
trolled changes in the fine terrigenous and/or of circumstantial evidence best fits a microbial
detrital carbonate fluxes. origin. It is now well documented that certain
ancient and modern types of bacteria produced
carbonate mud and peloids (e.g. Morita, 1980;
Mud mound facies
Chafetz, 1986; Monty, 1995; Reid & Macintyre,
Carbonate mud mounds occur in deep-water 2000; Riding, 2000). They produce carbonate in
deposits ranging in age from Proterozoic to Cre- two main ways: (1) by secreting an extracellular
taceous, but are most notable in the Palaeozoic polymeric material (slime and filaments combi-
(Table 3). The origin of the mud and peloidal ning to form a mucilaginous biofilm) that scav-
mud in Palaeozoic mud mounds is derived from enges Ca2+, Mg2+ and HCO32– from the ambient
either: (1) in situ organic production by microbial- waters and kinetically favours CaCO3 formation;
mediated precipitation and/or degradation of and/or (2) by having a cell coat that acts as a
delicate skeletons; or (2) external sources and template for adsorbing Ca2+, Mg2+ and HCO32– and
hydrodynamically concentrated to form mud forms CaCO3. The abundant peloids/clotted mi-
mounds (Monty, 1995 and references therein). crospar are interpreted to represent fine-grained
For the Marjum mud mounds, a combination of carbonate precipitated within and around clumps
petrographic, sedimentological and stratigraphic of bacteria (Chafetz, 1986). The active upper
observations indicates an in situ origin for the portions of the bacterial biofilms are responsible
carbonate mud (now microspar) and peloids. for the generation of the fine-grained carbonate
Most importantly, their three-dimensional growth and micritic cement. Where present, the mucila-
array on the sea floor (Fig. 6) and their symmetric ginous character of the biofilm would have
shape argue against hydrodynamic concentration stabilized the mound surface, prevented erosion
of fine detrital carbonate by bottom currents. In and discouraged settlement by many infaunal and
addition, it is unlikely that settling of fine detrital epifaunal organisms, thus reducing competition
carbonate or mud from whitings could have for the microbial communities. The extracellular
generated the buildups. The grain size, fossil size biofilm and bacterial cell walls can continue to
and orientation and extent of grain-size sorting is produce fine-grained carbonate even after cell
distinct between mud mound and associated death (Monty, 1995), meaning that carbonate
rhythmite facies, suggesting different sediment precipitation probably continued into shallow
sources and transport histories. In particular, burial and provided additional mud and cement.
massive mud mounds are composed of fenestral, The occurrence of abundant fenestrae throughout
peloidal microspar with calcitized bacterial(?) the mud mound and intermound facies is inter-
filaments/tubes, sparse unoriented, unabraded preted to represent the decay of buried bacterial
Ó 2002 International Association of Sedimentologists, Sedimentology, 49, 1021–1047
13653091, 2002, 5, Downloaded from https://onlinelibrary.wiley.com/doi/10.1046/j.1365-3091.2002.00488.x by Korea University Library, Wiley Online Library on [07/05/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
1038 M. Elrick and A. C. Snider
Table 3. General characteristics of substorm wave base Palaeozoic mud mounds dominated by lime mud to peloidal
mud textures.

Mound Mound width Dominant


Geographic thickness (isolated mound
Age location Setting and (no.)* or linked) lithology Reference
Cambrian Western Utah, Basinal 0Æ4–2 m 0Æ5–2 m Stromatactid 1a
USA (this study) (100s) (linked) mudstone–wackestone
South Australia Shelf 60 m 120 m Stromatactid 1b
margin (few) (isolated) mudstone–wackestone
Northwest Slope 1–25 m 120 m Stromatactid microbial 1c
Territories (few) (isolated) boundstone
Ordovician Virginia, USA Down- 20–250 m 1–65 km Stromatactid 2a
slope (7) (isolated) mudstone–grainstone
Nevada, USA Deep 80 m 300 m Stromatactid 2b
ramp (3) (isolated) peloidal mudstone
Alberta, Deep 2m 15 m Microbial 2c
Canada shelf (several) (isolated) boundstone
Silurian Canadian Slope and 310 m <1 km Microbial 3a
Arctic downslope (1) (isolated) boundstone
1140 m 15–25 km Stromatactid
(4) (isolated) mudstone–wackestone
Northern Slope >100 m 100s of m Stromatactid 3b
Greenland (6+) (isolated) mudstone–wackestone
Michigan Intracra- >10–60 m 10s of m– Stromatactid 3c
Basin, USA tonic (>5) <2Æ5 km mudstone/
basin (isolated) wackestone/boundstone
Devonian Morocco, Deep Up to 50 m 150 m Stromatactid 4a
Spanish shelf (several 10s) (isolated) mudstone–wackestone
Sahara,
Southern Deep Up to 70 m Several 100 m Stromatactid 4b
France, shelf (several) (isolated) mudstone–wackestone
Belgium Open Up to 30 m 10s of m Stromatactid 4c
shelf (several) (isolated) mudstone–wackestone
Mississippian Montana, USA Deep ramp- 10–25 m 75–185 m Mudstone–wackestone 5a
basinal (several) (isolated)
New Slope Up to 40 m 10s)100s of m Stromatactid mudstone– 5b
Mexico, USA (10s) (isolated to grainstone (microbial)
linked?)
Belgium, Deep ramp m)100s of m 10s)1000s Stromatactid 5c
England, and slope (several) of m mudstone–packstone
Ireland (microbial)
Kentucky, Deep 15 m 500 m Green shale, 5d
USA shelf (>12) (isolated) mudstone–packstone
Pennsylvanian Canadian Slope to Up to 550 m 100s of m Stromatactid mudstone– 6a
Arctic basinal (many) (isolated) packstone (microbial)
Permian Hubei and Platform 50–120 m No data Sponge 7a
Sichuan, China margin (several) provided wackestone–packstone

* Number refers to the approximate number of mud mounds reported from the studied stratigraphic succession.
1a, this study; 1b, James & Gravestock (1990); 1c, Pratt (1989a); 2a, Read (1982); 2b, Ross et al. (1975); 2c, Pratt (1989b,
1995); 3a, De Freitas & Dixon (1995) (mounds A and C, Fig. 2); 3b, Sonderholm & Harland (1989); 3c, Textoris &
Carozzi (1964); Lehmann & Simo (1989); Pratt (1995); 4a, Brachert et al. (1992); Dumestre & Illing (1967); 4b, Flajs &
Hussner (1993); 4c, Dreesen et al. (1985); Pratt (1995); 5a, Stone (1972); Elrick & Read (1991); 5b, Pray (1961); Jeffery
et al. (1996); 5c, Lees & Miller (1995); Bridges et al. (1995); Pratt (1995); Meyer et al. (1995); 6a, Davies et al. (1989);
7a, Liu et al. (1991).

communities/mats and trapping of gas within the implies that, during their development, the entire
cohesive deposits. sea floor within the study area was colonized by
The three-dimensional spacing of mud mounds microbial communities generating micrite and
separated by thinner intermound deposits peloids. Microbial activity and therefore mud
Ó 2002 International Association of Sedimentologists, Sedimentology, 49, 1021–1047
13653091, 2002, 5, Downloaded from https://onlinelibrary.wiley.com/doi/10.1046/j.1365-3091.2002.00488.x by Korea University Library, Wiley Online Library on [07/05/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Deep-water stratigraphic cyclicity and carbonate mud mound development 1039

and peloid production was locally concentrated What can be interpreted from the Marjum
within regularly spaced, metre-wide regions that stromatactoid cavities is that they formed before
eventually formed the small mud mounds significant burial and compaction because the
(Fig. 6). It is not understood what originally cavities are commonly filled with marine geopetal
controlled this regular spacing. sediment that could not have infiltrated into the
cavities once they were buried more than a few
Stromatactoid origin tens of centimetres below the sea floor. In addi-
tion, the preservation of large cavities indicates
The mechanism of stromatactis (or stromatactoid)
that at least the first generation of internal
cavity formation in mud mounds remains uncer-
isopachous marine cement precipitated before
tain. The most current ideas include: (1) decay of
compaction.
siliceous sponge bodies (Bourque & Gignac, 1983;
Bourque & Boulvain, 1993); (2) patchy hard-
ground formation or microbial growth and subse- Shale facies
quent submarine erosion of surrounding unbound
The fine grain size, the preservation of suspen-
sediment (Bathurst, 1980, 1982; Pratt, 1982, 1995
sion laminae and the rare occurrence of unidi-
respectively); or (3) winnowing of unbound sedi-
rectional current-oriented fossils and sole and
ment by through-flowing pore waters below the
tool marks (Rogers, 1984) suggest deposition from
sediment–water interface (Wallace, 1987) or by
siliciclastic-rich, dilute density currents in relat-
winnowing on the sea floor (Matyszkiewicz,
ively quiet, poorly oxygenated waters below
1993). Observations from the Marjum stromatac-
storm wave base (within the O2-minimum zone).
toid structures do not support an origin related to
Palaeoflow and palaeoslope indicators from fos-
siliceous sponge body decay because: (1) geopetal
sils, soft-sediment folds and truncation surfaces
sediments within stromatactoid structures lack
indicate a south-west to west-dipping palaeo-
sponge spicules (at least some spicules would be
slope and turbidity current flow towards the
expected if the cavities were related to sponge
south-west (Rogers, 1984). However, palaeoflow
collapse and decay); (2) unlike examples provi-
directions from interbedded cross-bedded lime-
ded by Bourque & Boulvain (1993), there are no
stone facies and rare rippled rhythmite facies indi-
mud mounds that display transitions between
cate periodic eastward (onshore) current flows.
well-preserved, recognizable sponge bodies to
Previous workers have suggested that the fine
poorly preserved sponges indicated only by
siliciclastics in the Marjum and underlying
variably abundant sponge spicules; instead, the
Wheeler Formations were derived from the adja-
Marjum mounds are consistently composed of
cent craton via bypass across the adjacent car-
peloidal microspar with variably abundant
bonate platform (Hintze & Robison, 1975; Rogers,
sponge spicules; and (3) modern and previously
1984; Rees, 1986) or by aeolian transport
documented fossil sponges do not attain growth
(Dalrymple et al., 1985). However, coeval, shal-
morphologies with the geometry and scale
low-water deposits adjacent to the central House
implied by the model of Bourque & Boulvain
Range are composed of very pure carbonates. In
(1993; for additional arguments, see Pratt, 1986).
addition, no localized siliciclastic bypass zone
If bottom currents or escaping pore fluids
deposits are known within Middle Cambrian
removed unbound or non-cohesive sediments
platform carbonates in Utah, southern Idaho or
and formed the cavities, it seems likely that some
southern Nevada. This suggests that the fine
additional evidence for sediment reworking or
siliciclastic material in the Marjum and Wheeler
winnowing (scours, ripples, reworked skeletal
Formations was not transported offshore (west-
material) would be apparent within the mud
ward) from the adjacent craton. Instead, it is
mound deposits, particularly given the current
proposed here that the fine siliciclastics were
velocities and persistence required to remove the
transported from north of the study area by
volume of sediment represented by the cavities.
deeper water (geostrophic?) currents. This inter-
No evidence for current reworking or winnowing
pretation agrees with regional studies by Aitken
is observed in the Marjum mud mounds. At the
(1978, 1997) in the Canadian Rocky Mountains,
same time, it is clear from the abundant early
which indicate a persistent far northern source
cement within stromatactoid cavities and fenest-
(near the British Columbia–North-West Territor-
rae that large amounts of fluids passed through
ies border) for Middle Cambrian quartz and clay
the sediment; the origin of these fluids is not well
deposited in western Canada and as far south as
constrained.
northern Utah. A northern siliciclastic sediment
Ó 2002 International Association of Sedimentologists, Sedimentology, 49, 1021–1047
13653091, 2002, 5, Downloaded from https://onlinelibrary.wiley.com/doi/10.1046/j.1365-3091.2002.00488.x by Korea University Library, Wiley Online Library on [07/05/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
1040 M. Elrick and A. C. Snider

source for the study area is supported by the 104 and 105 years (given the parasequence dur-
palaeogeographic position of western North ation estimations discussed above). This indi-
America during the Middle Cambrian (rotated cates that, within the study area, deposition in
90° clockwise and less than 10–15°N of the quiet, poorly oxygenated waters below storm
palaeoequator; Scotese & Golonka, 1992; Scotese, wave base (conditions during rhythmite depos-
1997), such that north-east palaeotrade winds ition) was interrupted every 104)105 years by
probably controlled the general wind patterns in environmental conditions that resulted in mud
the study area. mound, shale, bioturbated limestone or cross-
bedded limestone facies deposition. From the
facies interpretations discussed above, these
Bioturbated limestone facies
environmental changes involved: (1) a decrease
This facies is interpreted as having been depo- to near absence of detrital carbonate input into
sited in quiet, moderately oxygenated, substorm the embayment (mud mounds and underlying
wave base waters based on the occurrence of the thin nodular rhythmites and shales respectively);
fine-grained texture, sparse trace fossils and ben- (2) an increase in bottom-water O2 levels (biotur-
thic skeletal fossils and facies associations. The bated limestone); or (3) a lowering of storm wave
slight increase in bottom-water O2 levels, with base concurrent with a slight increase in bottom-
respect to associated rhythmite facies, may have water O2 levels (cross-bedded limestone).
resulted from a weakening or lowering of the To evaluate what mechanism(s) might control
O2-minimum zone, which may be related to: all three of these high-frequency palaeoenviron-
(1) a decrease in surface water productivity and mental changes, the stratigraphic record of time-
an associated decrease in O2 demand for organic equivalent, shallow-water deposits was examined
matter decomposition; or (2) a relative sea-level in other parts of the Great Basin. Approximately
fall and an associated fall in the O2-minimum 60 km south of the study area in the Wah Wah
zone. Range (Figs 1 and 2), Middle Cambrian (Bolas-
pidella zone) shallow-water carbonates of the Eye
of the Needle Limestone, Pierson Cove Formation
Cross-bedded limestone facies
and lower Trippe Limestone are characterized by
Dune-crest orientation, asymmetry and stoss and stacked metre-scale, upward-shallowing subtidal
lee preservation suggest deposition from oscilla- and peritidal cycles (or parasequences). The
tory to eastward-directed, relatively energetic cyclicity of these formations has been described
traction currents during times of relatively high by Kepper (1972, 1976), and a combined field and
sediment accumulation. The associated rhyth- statistical analysis was undertaken by Bond et al.
mites have few if any wave- or current-generated (1991). The parasequences average between 3 and
sedimentary structures or textures, indicating that, 5 m in thickness and are characterized by a
during dune formation, storm wave base lowered repetition of facies formed in environments vary-
in response to more intense storms (stronger ing from shallow subtidal (burrow-mottled lime
winds) or a relative sea-level fall. The occurrence mudstone with thrombolites) to high-energy bars
of some bioturbation suggests an increase in (oolitic grainstones) and to stromatolitic tidal flats
bottom-water O2 levels (compared with interbed- (laminated dolomudstone to lime mudstone).
ded rhythmites) concurrent with the lower storm Spectral analysis and the ‘gamma method’ age
wave base; as with the bioturbated limestone model indicate significant spectral peaks in the
facies, the increase in O2 levels may indicate a Trippe Limestone and Pierson Cove Formation,
weakening/lowering of the O2-minimum zone or, which correspond to Cambrian orbital eccentri-
more likely (given the lowered storm wave base), city (127 kyr) and climatic precession (15 kyr)
as a result of relative sea-level fall. models (Bond et al., 1991), i.e. preservation of
high-frequency, sea-level fluctuations in these
shallow-water carbonates.
PARASEQUENCE ORIGINS Across the southern Great Basin, Middle and
AND DISCUSSION Upper Cambrian shallow-water carbonates are
represented by the Bonanza King Formation
Although the type of parasequence varies strati- (Fig. 2). The Banded Mountain Member spans
graphically within the Marjum Formation, rhyth- the Bolaspidella, Cedaria and Crepicephalus
mite facies occur in each type, and the average zones and is characterized by 150–250, upward-
duration of each parasequence type is between shallowing, subtidal and peritidal parasequences
Ó 2002 International Association of Sedimentologists, Sedimentology, 49, 1021–1047
13653091, 2002, 5, Downloaded from https://onlinelibrary.wiley.com/doi/10.1046/j.1365-3091.2002.00488.x by Korea University Library, Wiley Online Library on [07/05/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Deep-water stratigraphic cyclicity and carbonate mud mound development 1041

(1–7 m thick) (Montañez & Osleger, 1993; Mont- When the House Range embayment was dom-
añez et al., 1996). The average duration of the inated by deeper water carbonates rather than
parasequences is estimated to range between 30 shale, rhythmite–mound, rhythmite–bioturbated
and 150 kyr, and these were interpreted by limestone and rhythmite–cross-bedded limestone
Montañez & Osleger (1993) and Montañez et al. parasequences developed (Figs 3 and 12A). Dur-
(1996) to record high-frequency sea-level oscilla- ing initial high-frequency sea-level rises, the
tions. carbonate factory retrograded northward, result-
If the parasequences within the Bolaspidella ing in a decrease in detrital carbonate transported
zone of the Wah Wah Range and southern Great to the embayment (Fig. 12A); the initial detrital
Basin record the effects of high-frequency, sea- carbonate decrease is recorded by the thinning
level changes, this implies that the carbonate of limestone layers within rhythmites directly
platform surrounding the House Range embay- below each mud mound layer. In addition to
ment was also affected by these oscillations. limestone layer thinning, the nodular fabric of
Although the magnitude of the oscillations was limestone layers directly below mud mounds is
likely to be low (given the peritidal facies-dom- probably a result of differential early cementa-
inated nature of the shallow-water parasequences tion, compaction and later pressure solution,
and the paucity of evidence for Middle Cambrian which may be related to bioturbation (Scoffin,
continental ice sheets), deeper water environ- 1987); this bioturbation may indicate a slight
ments may have been affected by eustatic- increase in bottom-water O2 levels concomitant
induced fluctuations in carbonate platform size with initial transgression. At maximum flooding,
(size and location of carbonate factory with when the carbonate factory lay furthest from the
progradation and retrogradation), depth to wave embayment and the detrital carbonate input was
base and O2-minimum zone and offshore marine lowest, deeper water microbial communities col-
circulation patterns related to shoreline positions. onized the sea floor, leading to the in situ
Based on stratigraphic interpretations from this precipitation of fine-grained carbonate to form
study and previous studies of coeval shallow- the mud mounds (Fig. 12A). The microbially
marine carbonates, it is likely that the Marjum induced precipitation of lime mud, peloids/clots
Formation parasequences developed in response and micritic cement would have diluted the
to high-frequency (104)105 years) sea-level fluc- amount of terrigenous material within the mud
tuations that controlled: (1) the amount of fine mound facies and may explain why the insoluble
detrital carbonate reaching the House Range residue contents are lower in mud mounds than
embayment; (2) the depth/intensity of the O2- in the associated rhythmite limestones. During
minimum zone; and (3) the depth to storm wave high-frequency sea-level falls, the carbonate fac-
base. The following depositional model is pro- tory prograded towards the embayment, resulting
posed. During time intervals when the House in a greater input of detrital carbonate to the basin
Range embayment was flooded by northern-deri- and deposition of rhythmite facies on top of the
ved, fine-grained siliciclastics (Figs 3 and 12B), mud mounds (Fig. 12A).
parasequences are composed of abundant shale Rhythmite–bioturbated limestone and rhyth-
(rhythmite–shale parasequences). Shale was mite–cross-bedded limestone parasequences are
deposited during high-frequency sea-level rise few in number and stratigraphically restricted to a
and early highstand, when the shallow subtidal carbonate-dominated interval within the middle
carbonate factory along the northern House Range Marjum Formation (Fig. 3). This stratigraphic
embayment ramp retrograded (northward) away distribution is interpreted to be the result of
from the study area. This would have increased high-frequency sea-level oscillations superim-
the factory-to-basin distance, thus decreasing the posed on an interval of longer term (third order
amount of detrital carbonate reaching the basin or 1–5 Myr-scale) sea-level fall/lowstand. The
from offshore-directed turbidity currents or accommodation space loss during longer term
storm-generated currents (Fig. 12B). During the sea-level fall/lowstand would diminish the
ensuing high-frequency sea-level falls, the car- effects of carbonate factory retrogradation during
bonate factory prograded towards the embay- high-frequency sea-level rises (cf. Goldhammer
ment, decreasing the factory-to-basin distance et al., 1990); thus, little or no change in detrital
and thus increasing the detrital carbonate input carbonate input was recorded, and rhythmites
to the embayment from offshore-directed cur- accumulated even during the high-frequency
rents; this resulted in the deposition of rhythmite rises. The combined accommodation space loss
facies over basal shales (Fig. 12B). during high-frequency and longer term sea-level
Ó 2002 International Association of Sedimentologists, Sedimentology, 49, 1021–1047
13653091, 2002, 5, Downloaded from https://onlinelibrary.wiley.com/doi/10.1046/j.1365-3091.2002.00488.x by Korea University Library, Wiley Online Library on [07/05/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
1042 M. Elrick and A. C. Snider

ED ORY . .
. . . . . . . . . . . . . . . . . . .
AD T . . N
GR FAC
UPPER MARJUM FORMATION A TR
O
RE ATE . .
. .

ON . . .
B
R . .
CA . .
.
. .

SL 20 km
KEY
. . . . . . . . . . . . . . . . . Shale facies
. . . . . . . . . . . . . . . . . .
D .
DE . RY . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
RA . CTO. . .
.
G . . . .
O . .F .A . . Rhythmite facies
PR .
. .
. ATE. .
.
. ON. .
.
. RB. .
. . CA. .
. Mud mound facies

SL Thin nodular rhythmites

Shallow-water limestones
LOWER & MIDDLE MARJUM FORMATION

D Y . . . . . . . . . . . . . . . . . . . .
B DE OR . . N ....
GRA FACT . . ....
Shallow subtidal
O . carbonate factory
TR E .
RE NAT . .
BO . . .
R
CA . . Exposed carbonate platform
. .
. .

Shoreline position
SL
20 km
Offshore detrital carbonate
transport (line thickness
NORTHERN-DERIVED
indicates relative abundance)
SILICICLASTICS
. . .. .. .. .. .. .. .. .. .. .. . . . . . . . . . . . . .
. . . . . . . . . . . . . . . . .
ED . ORY. . . . . . . . . . . . . . .
.
R AD .
. CT . . . . . .
OG .
. . FA. .
PR .
. TE . .
. A .
.
. ON . .
. ARB. .
.
.
. . C. .

SL

Fig. 12. Schematic block diagram across the House Range embayment illustrating sea-level-induced control on
offshore sedimentation patterns and Marjum parasequence development. (A) Sedimentation patterns associated with
high-frequency sea-level change during deposition of the upper Marjum Formation (limestone rich). (B) Sedimen-
tation patterns associated with high-frequency sea-level oscillations during deposition of lower and parts of the
middle Marjum Formation (shale rich).

fall/lowstand was great enough to: (1) lower storm low carbonate platforms are flooded and carbon-
wave base and rework the rhythmites to form ate production rates are high, resulting in
cross-bedded limestones; and (2) lower the depth/ abundant offshore carbonate sediment transport
intensity of the O2-minimum zone, permit limited (highstand shedding). In Schlager’s model, dur-
bioturbation of the rhythmites and form biotur- ing late highstand and lowstand, the shallow
bated limestones. carbonate platform is exposed, resulting in a
In this interpretation, shale, mud mound and significant decrease in size or total loss of the
underlying thin, nodular rhythmites are trans- carbonate factory, possible fluvial incising of the
gressive deposits, whereas rhythmite facies are exposed platform and a resultant increase in
regressive (highstand) and/or lowstand deposits. siliciclastic sediment deposition within the basin.
Bioturbated limestone and cross-bedded lime- This highstand shedding model appears to be
stone facies are also regressive/lowstand deposits. appropriate for flat-topped platforms; however,
This interpretation is opposite to the highstand for ramp margin geometries, the carbonate factory
shedding model of Schlager (1993). who sugges- migrates along the gently dipping slope, but is
ted that offshore carbonate deposition peaks never fully lost during sea-level falls/lowstands
during transgression/early highstand when shal- (Burchette & Wright, 1992).
Ó 2002 International Association of Sedimentologists, Sedimentology, 49, 1021–1047
13653091, 2002, 5, Downloaded from https://onlinelibrary.wiley.com/doi/10.1046/j.1365-3091.2002.00488.x by Korea University Library, Wiley Online Library on [07/05/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Deep-water stratigraphic cyclicity and carbonate mud mound development 1043

The House Range embayment’s sensitivity to storm wave base. In the lower and middle Mar-
varying detrital carbonate input is supported by jum Formation, the rhythmites are cyclically
the fact that the northern embayment margin was interbedded with deep-water shale and, less
a ramp throughout its lifetime (Figs 1 and 12) commonly, with bioturbated limestone and cross-
and, given Cambrian palaeogeography and gen- bedded limestone to form metre- to decimetre-
eral atmospheric circulation patterns, it is likely scale, rhythmite–shale and rhythmite–bioturbat-
that this ramp margin contributed the most ed limestone, and rhythmite–cross-bedded lime-
detrital carbonate material to the basin. For slopes stone parasequences. In the upper Marjum
of less than a few degrees, even slight changes in Formation, rhythmites are punctuated by thin, car-
sea level will dramatically shift the position of bonate mud mounds to form rhythmite–mound
the carbonate factory; the transport distances parasequences. The Marjum parasequences have
between the migrating factory and the deeper an average duration of 14–270 kyr using bio-
water embayment would be expected to vary stratigraphically controlled average sediment
significantly. However, the geometry of the accumulation rates and parasequence thicknesses.
southern embayment margin is not well under- 2 The mud mounds are composed of fenestral,
stood because of the lack of Middle Cambrian stromatactoid-bearing peloidal microspar with
outcrop between the central and southern House sparse unoriented, unabraded and poorly sorted
Range. skeletal material. SEM observations indicate
This interpreted sea-level control on detrital common submicrometre, calcitized bacterial(?)
carbonate and siliciclastic sedimentation patterns filaments/tubes. This combination of features
and subsequent parasequence development is along with the three-dimensional spacing of mud
similar to that invoked by many authors (Aitken, mounds on the sea floor and their symmetry
1978, 1997; Markello & Read, 1981; Chow & suggests that the carbonate mud/peloids were
James, 1987; Bond et al., 1989; Westhrop, 1989) precipitated and cemented in situ as a result of
for the development of larger scale, shale–car- substorm wave base benthic microbial activity.
bonate sequences of Cambrian age (grand cycles 3 The concentrated growth of microbial com-
of Aitken, 1978, 1997). In each of these grand munities across the study area is interpreted to
cycle depositional models, deeper water, shale- have been controlled by periodic decreases in the
rich intervals accumulated during long-term amount of fine detrital carbonate transported off-
(third-order) sea-level rises (or during periods of shore from the adjacent carbonate platform. In
increased sea-level rise rates) when the rate of contrast, rhythmites, bioturbated limestones and
accommodation space gain outpaced that of car- cross-bedded limestones were deposited during
bonate production (lower part of the grand cycle). periods of renewed detrital carbonate influx,
During ensuing long-term sea-level fall (or during lowering/weakening of the O2-minimum zone
decreased rates of sea-level rise), carbonate pro- and/or lowering of storm wave base. Deposition
duction kept pace with or outpaced accommoda- of shale at the base of some parasequences also
tion space gains, and shallow-water carbonates records periods of little or no detrital carbonate
were deposited to form the upper part of the influx.
grand cycles. This pattern of shale-rich facies 4 Coeval Middle Cambrian shallow-water car-
accumulating during third-order relative sea-level bonates (specifically within the Bolaspidella
rise and carbonate-rich facies deposited on the zone) throughout the Great Basin are character-
fall is particularly well supported by Bond et al. ized by stacked, metre-scale, upward-shallowing
(1989), who documented repeated 2–6 Myr sea- parasequences that have been interpreted to
level events that have similar timing in the record high-frequency (104)105 years) sea-level
southern Canadian Rockies, the Great Basin and oscillations. The formation of these shallow-wa-
the southern Appalachians, suggesting a eustatic ter parasequences coincident with the deposition
control on Cambrian third-order sequences. of Marjum Formation parasequences, which have
similar estimated durations, suggests that high-
frequency sea-level oscillations also affected
CONCLUSIONS deep-water sedimentation patterns in the House
Range embayment. During high-frequency sea-
1 The Middle Cambrian Marjum Formation is level rises, the carbonate factory along the nor-
dominated by thin, interbedded limestone and thern margin of the House Range embayment
argillaceous limestone rhythmites, which were retrograded northwards, increasing the transport
deposited in quiet, oxygen-poor waters below distance between the deep-water study area and
Ó 2002 International Association of Sedimentologists, Sedimentology, 49, 1021–1047
13653091, 2002, 5, Downloaded from https://onlinelibrary.wiley.com/doi/10.1046/j.1365-3091.2002.00488.x by Korea University Library, Wiley Online Library on [07/05/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
1044 M. Elrick and A. C. Snider

the factory, which resulted in less detrital car- deep-water environments. Once calcareous
bonate being transported offshore to the embay- microfossils evolved in the Jurassic, the presence
ment. The decrease in detrital carbonate input or absence of deeper water carbonates was influ-
permitted the concentrated growth of benthic enced by a variety of chemical and physical
microbial communities and in situ carbonate oceanographic factors (nutrient supply, tempera-
precipitation in the form of micrite, peloids/clots ture, upwelling, chemistry, circulation) rather
and micritic cement (base of rhythmite–mound than just physical transport of carbonate muds.
parasequences), or in the absence of microbial
activity, shale was deposited to form the base of
rhythmite–shale parasequences. During high-fre- ACKNOWLEDGEMENTS
quency sea-level falls and lowstands, the car-
bonate factory prograded southwards and lay Funding for this research was provided by NSF
closer to the study area, providing more detrital grants EAR-9210462 and EAR-9405734. Valuable
carbonate to the deep-water embayment, which field assistance was provided by Margaret Rees,
resulted in the deposition of rhythmites. During Sheila Hutcherson, Erin Knox and Laura Snider.
superimposed third-order and high-frequency The manuscript benefited from reviews by Viorel
falls/lowstands, bioturbated limestone or cross- Atudorei, Linda Hinnov, Carol Dehler, Brian
bedded limestone was deposited in response to a Pratt, Carl Drummond and Isabel Montañez;
sea-level-induced lowering (or weakening) of the special thanks go to Mike Pope, Eric Mountjoy
O2-minimum zone and the lowering of storm and editor Ian Jarvis for their essential and
wave base respectively. constructive reviews.
5 When compared with many other deep-water
Palaeozoic mud mounds, the Marjum mounds are
conspicuously thinner and smaller in diameter, REFERENCES
more closely spaced and laterally linked, occur in
significantly greater abundance along single Aitken, J.D. (1978) Revised models for depositional grand
stratigraphic horizons and occur repeatedly (up to cycles, Cambrian of the southern Rocky Mountains. Can.
Bull. Petrol. Geol., 26, 515–542.
14 times) throughout the deep-water succession Aitken, J.D. (1997) Stratigraphy of the Middle Cambrian plat-
(Table 3). In addition, each mud mound-bearing formal succession, southern Rocky Mountains. Geol. Surv.
layer can be correlated over > 5 km distances, Can. Bull., 398, 322 pp.
implying that they did not grow on local palaeo- Anderson, R.Y. (1982) A long geoclimatic record from the
topographic features or sea-floor vents; rather, Permian. J. Geophys. Res., 87, 7285–7294.
Anderson, T.F. and Arthur, M.A. (1983) Stable isotopes of
their growth within the embayment was con- oxygen and carbon and their application to sedimentologic
trolled by more regional palaeoenvironmental and paleoenvironmental problems. In: Stable Isotopes in
conditions. Despite these geometric and strati- Sedimentary Geology. SEPM Short Course, 10, 1–151.
graphic differences, the Marjum mud mounds Bathurst, R.G.C. (1980) Stromatactis – origin related to sub-
have features common to other Palaeozoic mud marine-cemented crusts in Paleozoic mud mounds. Geol-
ogy, 8, 131–134.
mounds including: (a) textures dominated by Bathurst, R.G.C. (1982) Genesis of stromatactis cavities
microbial lime mud and peloidal mud; (b) com- between submarine crusts in Palaeozoic carbonate mud
mon stromatactoid (or stromatactis) structures; (c) buildups. J. Geol. Soc. London, 139, 165–181.
evidence of syndepositional cementation/stabil- Beauchamp, B. and Savard, M.M. (1992) Cretaceous chemo-
ization; and (d) a paucity of skeletal material synthetic carbonate mounds in the Canadian Arctic. Palai-
os, 7, 434–450.
(Table 3). Blakemore, L.C., Searle, P.L. and Daly, B.K. (1987) Methods
6 Results from this study imply that Middle for the chemical analysis of soils. NZ Soil Bur. Sci. Rep., 80.
Cambrian sea-level changes affected shallow- Bond, G.C. and Lotti, R. (1995) Iceberg discharges into the
marine sedimentation patterns through changes North Atlantic on millennial time scales during the last
in water depth, as well as controlling the type glaciation. Science, 267, 1005–1010.
Bond, G.C., Kominz, M.A., Steckler, M.S. and Grotziner, J.P.
and amount of offshore sediment delivery via (1989) Role of thermal subsidence, flexure, and eustasy in
carbonate factory progradation/retrogradation the evolution of early Paleozoic passive-margin carbonate
patterns. This interpreted sea-level control on platforms. In: Controls on Carbonate Platform and Basin
offshore carbonate sedimentation patterns is Development (Eds P.D. Crevello, J.L. Wilson, J.F. Sarg and
particularly unique to Palaeozoic deep-water J.F. Read), SEPM Spec. Publ., 44, 39–61.
Bond, G.C., Kominz, M.A. and Beavan, J. (1991) Evidence for
sediments because calcareous microfossils orbital forcing of Middle Cambrian peritidal cycles: Wah
(planktonic foraminifera and coccolithophorids) Wah range, south-central Utah. In: Sedimentary Modeling:
had not yet evolved to contribute their shells to
Ó 2002 International Association of Sedimentologists, Sedimentology, 49, 1021–1047
13653091, 2002, 5, Downloaded from https://onlinelibrary.wiley.com/doi/10.1046/j.1365-3091.2002.00488.x by Korea University Library, Wiley Online Library on [07/05/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Deep-water stratigraphic cyclicity and carbonate mud mound development 1045
Computer Simulations and Methods for Improved Param- genetic history of the ‘mabre rouge a crinoides de Baelen’
eter Definition (Eds E.K. Franseen, W.L. Watney, C.G.StC. (Late Upper Devonian, Verviers Synclinorium, eastern
Kendall and W. Ross), Kan. Geol. Surv. Bull., 233, 293–317. Belgium). Ann. Soc. Géol. Belg., 108, 311–359.
Bond, G.C., Kromer, B., Juerg, B., Muscheler, R., Evans, M.N., Dumestre, A. and Illing, L.V. (1967) Middle Devonian reefs. I.
Showers, W., Hoffmann, S., Lotti-Bond, R., Hajadas, I. and Spanish Sahara. In: International Symposium on the Devo-
Bonani, G. (2001) Persistent solar influence on North nian System (Ed. D.H. Oswald), Vol. II, pp. 333–350. Alberta
Atlantic climate during the Holocene. Science, 294, 2130– Society for Petroleum Geology, Calgary.
2136. Elrick, M. and Hinnov, L.A. (1996) Millennial-scale climate
Bosence, D.W.J. and Bridges, P.H. (1995) A review of the ori- origins for stratification in Cambrian and Devonian deep-
gin and evolution of carbonate mud-mounds. In: Carbonate water rhythmites, western USA Palaeogeogr. Palaeoclima-
Mud-Mounds Their Origin and Evolution (Eds C.L.V. tol. Palaeoecol., 123, 353–372.
Monty, D.W.J. Bosence, P.H. Bridges and B.R. Pratt), Int. Elrick, M. and Read, J.F. (1991) Cyclic ramp-to-basin carbon-
Assoc. Sedimentol. Spec. Publ., 23, 3–9. ate deposits, Lower Mississippian, Wyoming and Montana:
Bourque, P.-A. and Boulvain, F. (1993) A model for the origin a combined field and computer modelling study. J. Sed.
and petrogenesis of the red stromatactis limestone of Petrol., 61, 1194–1224.
Paleozoic carbonate mounds. J. Sed. Petrol., 63, 607–619. Elrick, M., Read, J.F. and Coruh, C. (1991) Short-term paleo-
Bourque, P.-A. and Gignac, H. (1983) Sponge-constructed climatic fluctuations expressed in Lower Mississippian
stromatactis mud mounds, Silurian of Gaspé, Québec. ramp-slope deposits, southwestern Montana. Geology, 19,
J. Sed. Petrol., 53, 521–532. 799–802.
Bowring, S.A. and Erwin, D.H. (1998) A new look at evolu- Flajs, G. and Hussner, H. (1993) A microbial model for the
tionary rates in deep time: uniting paleontology and high- Lower Devonian stromatactis mud mounds of the Montagne
precision geochronology. GSA Today, 8, 1–8. Noire (France). Facies, 29, 179–194.
Brachert, T.C., Buggisch, W., Flugel, E., Hussner, H.M., Goldhammer, R.K., Dunn, P.A. and Hardie, L.A. (1990)
Joachmiski, M.M., Tourner, F. and Walliser, O.H. (1992) Depositional cycles, composite sea-level changes, cycle
Controls of mud mound formation: the Early Devonian kess- stacking patterns, and the hierarchy of stratigraphic forcing:
kess carbonates of Hamar Laghdad, Anti-Atlas, Morocco. examples from Alpine Triassic platform carbonate. Bull.
Geol. Rundsch., 81, 15–44. Geol. Soc. Am., 102, 535–562.
Brady, M.J. and Koepnick, R.B. (1979) A Middle Cambrian Grootes, P.M., Stuiver, M., White, J.W.C., Johnsen, S. and
platform-to-basin transition, House Range, west central Jouzel, J. (1993) Comparison of oxygen isotope records for the
Utah. Brigham Young Univ. Geol. Studies, 26, 1–17. GISP2 and GRIP Greenland ice cores. Nature, 366, 552–554.
Braiser, M.D. (1993) Towards a carbon isotope stratigraphy of Hintze, L.F. and Robison, R.A. (1975) Middle Cambrian stra-
the Cambrian system: potential of the Great Basin succes- tigraphy of the House, Wah Wah, and adjacent ranges in
sion. In: High Resolution Stratigraphy (Eds E.A. Hailwood western Utah. Bull. Geol. Soc. Am., 86, 881–891.
and R.B. Kidd), Geol. Soc. Am. Spec. Publ., 70, 341–347. Irwin, H., Curtis, C. and Coleman, M. (1977) Isotopic evidence
Bridges, P.H., Gutteridge, P. and Pickard, H.A.H. (1995) The for source of diagenetic carbonates formed during burial of
environmental setting of Early Carboniferous mud-mounds. organic-rich sediments. Nature, 269, 209–213.
In: Carbonate Mud-Mounds. Their Origin and Evolution James, N.P. and Gravestock, D.I. (1990) Lower Cambrian shelf
(Eds C.L.V. Monty, D.W.J. Bosence, P.H. Bridges and B.R. and shelf margin buildups, Flinders Ranges, South Austra-
Pratt), Int. Assoc. Sedimentol. Spec. Publ., 23, 171–190. lia. Sedimentology, 37, 455–480.
Burchell, M.T., Stefani, M. and Masetti, D. (1990) Cyclic Jeffery, D.L. and Stanton, R.J., Jr (1996) Growth history of
sedimentation in the Southern Alpine Rhaetic: the import- Lower Mississippian Waulsortian mounds; distribution,
ance of climate and eustasy in controlling platform–basin stratal patterns, and geometries, New Mexico. Facies, 35,
interactions. Sedimentology, 37, 795–815. 29–58.
Burchette, T.P. and Wright, V.P. (1992) Carbonate ramp Kepper, J.C. (1972) Paleoenvironmental patterns in middle to
depositional systems. Sed. Geol., 79, 3–57. lower Upper Cambrian interval in eastern Great Basin.
Chafetz, H.S. (1986) Marine peloids: a product of bacterially AAPG Bull., 56, 503–527.
induced precipitation of calcite. J. Sed. Petrol., 56, 812–817. Kepper, J.C. (1976) Stratigraphic relationships and deposi-
Chow, N. and James, N.P. (1987) Cambrian Grand Cycles: a tional facies in a portion of the Middle Cambrian of the
northern Appalachian perspective. Bull. Geol. Soc. Am., Basin and Range. Brigham Young Univ. Geol. Studies, 23,
105, 1576–1590. 75–91.
Dalrymple, R.W., Narbonne, G.M. and Smith, L. (1985) Eolian Lees, A. and Miller, J. (1995) Waulsortian banks. In: Carbonate
action and the distribution of Cambrian shales in North Mud-Mounds. Their Origin and Evolution (Eds C.L.V.
America. Geology, 13, 607–610. Monty, D.W.J. Bosence, P.H. Bridges and B.R. Pratt), Int.
Davis, G.R., Nassichuk, W.W. and Beauchamp, B. (1989) Assoc. Sedimentol. Spec. Publ., 23, 191–271.
Upper Carboniferous ‘Waulsortian’ Reefs, Canadian Arctic Lehmann, P.J. and Simo, A. (1989) Depositional facies and
Archipelago. In: Reefs, Canada and Adjacent Areas (Eds diagenesis of the Pipe Creek Jr. Reef, Silurian, Great Lakes
H.H.J. Geldsetzer, N.P. James and G.E. Tebbutt), Mem. Can. Region, Indiana. In: Reefs, Canada and Adjacent Areas (Ed.
Soc. Petrol. Geol., 13, 658–666. H.H.J. Geldsetzer, N.P. James and G.E. Tebbutt), Mem. Can.
De Freitas, T.A. and Dixon, O.A. (1995) Silurian microbial Soc. Petrol. Geol., 13, 319–329.
buildups of the Canadian. Arctic. In: Carbonate Mud- Levy, M. and Christie-Blick, N. (1991) Tectonic subsidence of
Mounds. Their Origin and Evolution (Eds C.L.V. Monty, the early Paleozoic passive continental margin in eastern
D.W.J. Bosence, P.H. Bridges and B.R. Pratt), Int. Assoc. California and southern Nevada. Bull. Geol. Soc. Am., 103,
Sedimentol. Spec. Publ., 23, 151–170. 1590–1606.
Dressen, R., Bless, M.J.M., Conil, R., Flajs, G. and Laschet, C. Liu, H., Rigby, J.K., Li, G., Xia, K. and Liu, L. (1991) Upper
(1985) Depositional environment, paleoecology and dia- Permian carbonate buildups and associated lithofacies,

Ó 2002 International Association of Sedimentologists, Sedimentology, 49, 1021–1047


13653091, 2002, 5, Downloaded from https://onlinelibrary.wiley.com/doi/10.1046/j.1365-3091.2002.00488.x by Korea University Library, Wiley Online Library on [07/05/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
1046 M. Elrick and A. C. Snider
western Hubei–eastern Sichuan Provinces, China. AAPG Pratt, B.R. (1989b) Early Ordovician cryptalgal-sponge reefs,
Bull., 75, 1447–1467. Survey Peak Formation, Rocky Mountains, Alberta. In:
McCrea, J.M. (1950) The isotope chemistry of carbonates and a Reefs, Canada and Adjacent Areas (Ed. H.H.J. Geldsetzer,
paleotemperature scale. J. Chem. Phys., 18, 849–857. N.P. James and G.E. Tebbutt), Mem. Can. Soc. Petrol. Geol.,
Markello, J.R. and Read, J.F. (1981) Carbonate ramp-to-deeper 13, 213–217.
shale shelf transitions of an Upper Cambrian intrashelf Pratt, B.R. (1995) The origin, biota and evolution of deep-
basin, Nolichucky Formation, southwest Virginia Appala- water mud-mounds. In: Carbonate Mud-Mounds. Their
chians. Sedimentology, 28, 573–597. Origin and Evolution (Eds C.L.V. Monty, D.W.J. Bosence,
Marshall, J.D. (1992) Climatic and oceanographic isotopic P.H. Bridges and B.R. Pratt), Int. Assoc. Sedimentol. Spec.
signals from the carbonate rock record and their preserva- Publ., 23, 49–123.
tion. Geol. Mag., 129, 143–160. Pray, L.C. (1961) Geology of the Sacramento Mountains
Matyszkiewicz, J. (1993) Genesis of stromatactis in an Upper escarpment, Otero Co., New. Mexico. Bull. New Mex. Bur.
Jurassic carbonate buildup (Mlynka, Cracow region, south- Mines Mineral Resour., 35, 133 pp.
ern Poland): internal reworking and erosion of organic Raymo, M.E., Ganley, K., Carter, S., Oppo, D.W. and
growth cavities. Facies, 28, 87–96. McManus, J. (1998) Millennial-scale climate instability
Metson, A.J. (1961) Methods of chemical analysis for soil during the early Pleistocene epoch. Nature, 392, 699–702.
survey samples. NZ Dept Sci. Ind. Res., Soil Bureau Bull., Read, J.F. (1982) Geometry, facies, and development of Middle
12, 133–145. Ordovician carbonate buildups, Virginia Appalachians.
Meyer, D.L., Ausich, W.I., Bohl, D.T., Norris, W.A. and Potter, AAPG Bull., 66, 189–209.
P.E. (1995) Carbonate mud-mounds in the Fort Payne For- Rees, M.N. (1986) A fault-controlled trough through a
mation, (lower Carboniferous), Cumberland Saddle region, carbonate platform: the Middle Cambrian House Range
Kentucky and Tennessee, USA. In: Carbonate Mud-Mounds. embayment. Bull. Geol. Soc. Am., 97, 1054–1069.
Their Origin and Evolution (Eds C.L.V. Monty, D.W.J. Reid, R.P. and Macintyre, I.G. (2000) Microboring versus re-
Bosence, P.H. Bridges and B.R. Pratt), Int. Assoc. Sedi- crystallization: further insight into the micritization pro-
mentol. Spec. Publ., 23, 273–287. cess. J. Sed. Petrol., 70, 24–28.
Montañez, I.P. and Osleger, D.A. (1993) Parasequences Riding, R. (2000) Microbial carbonates: the geological record
stacking patterns, third-order accommodation events, and of calcified bacterial–algal mats and biofilms. Sedimentol-
sequence stratigraphy of Middle to Upper Cambrian plat- ogy, 47 (Suppl. 1), 179–214.
form carbonates, Bonanza King Formation, southern Great Robison, R.A. (1960) Lower and Middle Cambrian stratigraphy
Basin. In: Recent Advances and Applications of Carbonate of the eastern Great. Basin. In: Geology of East-Central
Sequence Stratigraphy (Eds B. Loucks and J.F. Sarg), AAPG Nevada. Intermountain Assoc. Petrol. Geol., 11th Ann. Field
Mem., 57, 305–326. Conference, pp. 43–52.
Montañez, I.P., Banner, J.L., Osleger, D.A., Borg. L.E. and Robison, R.A. (1964) Upper Middle Cambrian stratigraphy of
Bosserman, P.J. (1996) Integrated Sr isotope stratigraphy western Utah. Bull. Geol. Soc. Am., 75, 995–1010.
and relative sea-level history in Middle Cambrian platform Robison, R.A. (1976) Middle Cambrian biostratigraphy of the
carbonates. Geology, 24, 917–920. Great Basin. Brigham Young Univ. Geol. Studies, 23, 93–
Montañez, I.P., Osleger, D.A., Banner, J.L., Mack, L.E. 109.
and Musgrove, M. (2000) Evolution of the Sr and C Robison, R.A. (1984) Cambrian Agnostida from North America
isotope composition of Cambrian oceans. GSA Today, 10, and Greenland: Part 1. Ptychagnostidae. Univ. Kansas
1–7. Paleontol. Contrib., 109, 28 pp.
Monty, C.L.V. (1995) The rise and nature of carbonate mud- Rogers, J.C. (1984) Depositional environments and paleoecol-
mounds: an introductory actualistic approach. In: Carbon- ogy of two quarry sites in the Middle Cambrian Marjum and
ate Mud-Mounds. Their Origin and Evolution (Ed. C.L.V. Wheeler Formations, House Range, Utah. Brigham Young
Monty, D.W.J. Bosence, P.H. Bridges and B.R. Pratt), Int. Univ. Geol. Studies, 31, 97–115.
Assoc. Sedimentol. Spec. Publ., 23, 11–48. Ross, R.J., Jaanusson, V. and Friedman, I. (1975) Lithology
Morita, R.Y. (1980) Calcite precipitation by marine bacteria. and origin of Middle Ordovician calcareous mudmound at
Geomicrobiol. J., 2, 63–82. Meiklejohn Peak, southern Nevada. US Geol. Surv. Prof.
Mounji, D., Bourque, P.A. and Savard, M.M. (1998) Hydro- Paper, 871, 48 pp.
thermal origin of Devonian conical mounds (kess-kess) of Saltzman, M.R., Runnegar, B. and Lohmann, K.C. (1998)
Hamar Lakhdad, Anti-Atlas, Morocco. Geology, 26, 1123– Carbon isotope stratigraphy of Upper Cambrian (Steptoean
1126. Stage) sequences of the eastern Great Basin: record of a
Pisias, N.G., Dauphin, J.P. and Sancetta, C. (1973) Spectral global oceanographic event. Bull. Geol. Soc. Am., 110, 285–
analysis of Late Pleistocene–Holocene sediments. Quatern. 297.
Res., 3, 3–9. Schlager, W. (1993) Accommodation and supply – a dual
Pratt, B.R. (1982) Stromatolitic framework of carbonate mud- control on stratigraphic sequences. Sed. Geol., 86, 111–136.
mounds. J. Sed. Petrol., 52, 1203–1227. Scoffin, T.P. (1987) An Introduction to Carbonate Sediments
Pratt, B.R. (1986) Sponge-constructed stromatactis mud and Rocks. Chapman & Hall, New York, 274 pp.
mounds, Silurian of Gaspé, Quebec – discussion. J. Sed. Scotese, C.R. (1997) PALEOMAP Paleogeographic Atlas.
Petrol., 56, 459–460. PALEOMAP Progess Report 90-0497. Department of Geol-
Pratt, B.R. (1989a) Deep-water Girvanella–Epiphyton reef on a ogy, University of Texas, Arlington.
mid-Cambrian continental slope, Rockslide Formation, Scotese, C.R. and Golonka, J. (1992) PALEOMAP Paleogeo-
Mackenzie Mountains, Northwest Territories. In: Reefs, graphic Atlas. PALEOMAP Progess Report 20. Department
Canada and Adjacent Areas (Ed. H.H.J. Geldsetzer, N.P. of Geology, University of Texas, Arlington.
James and G.E. Tebbutt), Mem. Can. Soc. Petrol. Geol., 13, Sonderholm, M. and Harland, T.L. (1989) Franklinian reef
161–164. belt, Silurian, north Greenland. In: Reefs, Canada and

Ó 2002 International Association of Sedimentologists, Sedimentology, 49, 1021–1047


13653091, 2002, 5, Downloaded from https://onlinelibrary.wiley.com/doi/10.1046/j.1365-3091.2002.00488.x by Korea University Library, Wiley Online Library on [07/05/2024]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Deep-water stratigraphic cyclicity and carbonate mud mound development 1047
Adjacent Areas (Eds H.H.J. Geldsetzer, N.P. James and G.E. Veizer, J., Ala, D., Azmy, K., Bruckschen, P., Buhl, D., Bruhn,
Tebbutt), Mem. Can. Soc. Petrol. Geol., 13, 356–366. F., Carden, G., Diener, A., Ebneth, S., Godderis, Y., Jasper,
Stewart, J.H. (1991) Latest Proterozoic and Cambrian rocks of T., Korte, C., Pawellek, F., Podlaha, O. and Strauss, H.
the western United States – an overview. In: Paleozoic (1999) 87Sr/86Sr, d13C, and d18O evolution of Phanerozoic
Paleogeography of the Western United States (Eds J.D. seawater. Chem. Geol., 161, 59–88.
Cooper and C.H. Stevens), Pacific Section SEPM, pp. 13–38. Wallace, M.V. (1987) The role of internal erosion and sedi-
Stewart, J.H. and Poole, F.G. (1974) Lower Paleozoic and mentation in the formation of stromatactis mudstones and
uppermost Precambrian Cordilleran miogeocline, Great associated lithologies. J. Sed. Petrol., 57, 695–700.
Basin, western United States. In: Tectonics and Sedimen- Westhrop, S.R. (1989) Facies anatomy of an Upper Cambrian
tation (Ed. W.R. Dickinson), SEPM Spec. Publ., 22, 27–57. grand cycle: Bison Creek and Mistaya Formations, southern
Stewart, J.H. and Suczek, C.A. (1977) Cambrian and Latest Alberta. Can. J. Earth Sci., 26, 2292–2304.
Precambrian paleogeography and tectonics in the western Wu, Y. and Chafetz, H.S. (2002) 13C-encriched carbonate in
United States. In: Paleozoic Paleogeography of the Western Mississippian mud mounds: Alamogordo Member, Lake
United States (Ed. J.H. Stewart, C.H. Stevens and A.E. Valley Formation, Sacramento Mountains, New Mexico,
Fritsche), Pacific Section SEPM, pp. 1–18. USA. J. Sed. Petrol., 72, 138–145.
Stone, R.A. (1972) Waulsortian-type bioherms (reefs) of Mis- Young, G.C. and Laurie, J.R. (1996) An Australian Phanero-
sissippian age, central Bridger Range, Montana. Montana zoic Timescale. Oxford University Press, Melbourne.
Geol. Soc. Guidebook, pp. 37–55.
Textoris, D.A. and Carozzi, A.V. (1964) Petrography and evo- Manuscript received 25 July 2001;
lution of Niagaran (Silurian) reefs, Indiana. AAPG Bull., 48,
revision accepted 8 May 2002.
397–426.

Ó 2002 International Association of Sedimentologists, Sedimentology, 49, 1021–1047

You might also like