ChapterPDF Available

Tectonics of Sedimentary Basins: Recent Advances

Authors:

Abstract and Figures

One of Earth's greatest mountain-building episodes, the Grenvillian orogeny, occurred with the assembly of the supercontinent Rodinia at the end of the Mesoproterozoic era, about 1.2-1.0 billion years ago. Weathering and erosion of the Grenvillian mountain chain, the roots of which can be traced today for nearly 12,000 km, produced huge volumes of sedimentary detritus that were dispersed by an enormous system of braided rivers. Erosion, denudation, and sediment throughput were enhanced by a lack of vegetation and vigorous continental weathering under a climate that favored strong chemical alteration.
Geological sketch map of Precambrian North America to show the location of Proterozoic sedimentary basins and orogenic belts in the southeastern part of the continent. The orogenic zones contributed sedimentary detritus to huge river systems that crossed the continent, spreading northwesterly and forming the thick sandstone successions preserved in the Thelon, Athabasca and other basins. The Grenvillian orogenic belt is interpreted as the roots of a deeply eroded mountain chain, resulting from multi-stage continental accretion that formed from about 1.2 to 1.0 billion years ago. Sediments from this orogen have been recognized by westward-directed paleocurrents (black arrows) and by common occurrence of detrital zircon grains yielding ages that are characteristic of the Grenville Province (see Fig. 29.3), several thousand kilometers away, on the other side of the continent. The long blue dashed arrows show the suggested transport direction of the Paleoproterozoic sediments, shed from the Yapavai and Trans-Hudson orogens (shown in green and pink, respectively, on the map) and the dashed red arrows illustrate transport directions proposed for Neoproterozoic sediments deposited in the, Grenvillian hinterland basins, located along the western margin of Laurentia (e.g., Amundsen Basin, Mackenzie Mountains, Uinta Mountains Group [UMG], and Grand Canyon Supergroup [GCS]). In North America, the proximal part of the Grenvillian depositional system is represented by deposits located in the Middle Run Basin (subsurface) and Midcontinent Rift-MCR. The Grenville Front is the approximate position of the thrust-fault marking the inboard edge of Grenvillian deformation. The position and extent of the Grenvillian mountain belt in the supercontinent Rodinia, is shown in orange on Figure 29.1. Source : Simplified from Whitmeyer and Karlstrom (2007).
… 
Content may be subject to copyright.
Chapter 29
The great Grenvillian sedimentation episode: record
of supercontinent Rodinia’s assembly
ROBERT RAINBIRD, PETER CAWOOD
and GEORGE GEHRELS
z
Geological Survey of Canada, Ottawa, Canada
School of Earth and Environment, University of Western Australia, Crawley, Australia and
Department of Geography and Geosciences, University of St. Andrews, St. Andrews, UK
z
Department of Geosciences, University of Arizona, Tucson, USA
ABSTRACT
One of Earth’s greatest mountain-building episodes, the Grenvillian orogeny, occurred
with the assembly of the supercontinent Rodinia at the end of the Mesoproterozoic era,
about 1.2–1.0 billion years ago. Weathering and erosion of the Grenvillian mountain
chain, the roots of which can be traced today for nearly 12,000 km, produced huge
volumes of sedimentary detritus that were dispersed by an enormous system of braided
rivers. Erosion, denudation, and sediment throughput were enhanced by a lack of
vegetation and vigorous continental weathering under a climate that favored strong
chemical alteration.
The enormity of the erosional episode and broad extent of river system that drained the
Grenvillian Mountains was first recognized with the advent of detrital zircon geochro-
nology as a tool of provenance analysis. Initially, zircon grains of Grenvillian age were
recovered from early Neoproterozoic sedimentary basins located in northwestern
Canada, more than 3000 km away from the nearest probable sources in the Grenville
Province of eastern Laurentia. Paleocurrents derived from cross-bedding in thick fluvial
deposits preserved in these basins showed regionally consistent west-northwesterly
transport, lending support to the paleogeographic model. Correlative strata, located
thousands of kilometers to the south, in the Canadian and US Cordillera, exhibit similar
detrital zircon age distributions providing further support for the large-scale river
system. These data also indicate that the fluvial system was laterally extensive and
likely originated from multiple sources along the great length of the Grenvillian
mountain front.
Deposits representing the proximal parts of the system have now been recognized in
the subsurface of the central US, where they comprise several stratigraphic sequences
that can be tied to the various stages of tectonic evolution of the Grenvillian orogeny.
The sequences correlate well with outcrop exposures preserved in the Midcontinent
Rift system and Great Lakes region to the north. Among these are syn-collisional rift
deposits and post-collisional foreland basin deposits displaying features such as axial
flow patterns that indicate deposition by trunk rivers flowing parallel to the mountain
front. Similar stratigraphic successions are preserved around the North Atlantic in
Scotland, Shetland, East Greenland, Svalbard, and Norway. Detrital zircon grains from
these successions are dominated by late Paleoproterozoic and late Mesoproterozoic
ages inferred to have been derived from source terranes of the Grenville Province in
eastern Laurentia.
Detrital zircon geochronology indicates that Grenville-age detritus was reworked into
numerous Phanerozoic successions around the globe. Some of this detritus was derived
directly from uplift and erosion of Grenville Province rocks or recycling of detritus from
Tectonics of Sedimentary Basins: Recent Advances, First Edition. Edited by Cathy Busby and Antonio Azor.
Ó2012 Blackwell Publishing Ltd. Published 2012 by Blackwell Publishing Ltd.
583
Grenvillian foreland basin deposits during Appalachian-Hercynian orogenesis and
assembly of the Pangea supercontinent. Late Mesoproterozoic detritus has been con-
tinually recycled into younger stratigraphic successions and remains a significant
component of modern river sediments.
Keywords: detrital zircon geochronology; neoproterozoic; paleogeography; Rodinia;
fluvial
INTRODUCTION
Periods of global-scale orogeny have punctuated
Earth’s history and are a key component of the
evidence for a supercontinent cycle (e.g., Murphy
et al., 2009; Nance et al., 1988). When superconti-
nents formed, continents collided and great por-
tions of the crust were uplifted and exhumed along
lengthy collisional plate boundaries, forming
extensive chains of mountains. The mountains
were subjected to vigorous weathering and erosion,
producing large volumes of sedimentary detritus
that were dispersed by enormous riversystems. The
remnants of these sedimentation events are pre-
served as thick and extensive sedimentary deposits
(e.g., Archer and Greb, 1995; Cawood et al., 2007a;
Squire et al., 2006; Veevers, 2004). Compilations of
zircon crystallization age data from sediments from
modern river systems around the globe show peaks
in crystallization ages at around 2.7 Ga, 1.8 Ga,
1.1 Ga, 0.5 Ga, and 0.3 Ga that correspond to the
inferred cycles of supercontinent assembly (e.g.,
Campbell and Allen, 2008; Hawkesworth et al.,
2009; Rino et al., 2008).
Perhaps the greatest orogenic event in Earth’s
history, the Grenvillian orogeny, occurred with the
assembly of the supercontinent Rodinia at the end
of the Mesoproterozoic era (e.g., Hoffman, 1991; Li
et al., 2008). The roots of the ancient Grenvillian
mountain chain are preserved today in an almost
contiguous belt that stretches nearly 12,000 km
from what is now southern Mexico north-eastward
through eastern North America, across to the Brit-
ish Isles and on into Scandinavia and eastern
Russia (Fig. 29.1). Elements of it and similar-age
orogenic belts are preserved on many other cratons
worldwide. Evidence of the enormity of the ero-
sional episode and broad extent of river systems
that drained the Grenvillian mountains was first
recognized with the advent of detrital zircon geo-
chronology as a tool of provenance analysis. U-Pb
dating of detrital zircon was used to test a proposal
Fig. 29.1. One possible configuration of the supercon-
tinent Rodinia formed as a result of the continental
collisions that occurred during the Grenvillian orogeny,
between about 1.2 and 1.0 billion years ago (simplified
from Li et al., 2008). Himalayan-style mountains formed
in the areas show in orange (the Grenvillian orogenic
belt) and shed sediments over a huge area of the super-
continent, as suggested by the red arrows (river systems
may have been more extensive than what is suggested,
but paleogeographic information is unavailable for
these areas). The North American area is shown in
more detail in Figure 29.2. Possible components of
the Grenvillian belt as discussed in the text include
the Sveco-Norwegian orogen of Baltica (S-N), the
Suns
as orogen on the western margin of Amazonia
(S), the Wilkes province in the Mawson block (W),
the Albany-Fraser (A-F) and Musgrave (M) orogens in
Australia, the Eastern Ghats belt of India (EG), the
Namaqua-Natal province of the Zimbabwe craton
(N-N), and the Kibaran and Irumide belts of the
Congo craton (K-I).
584 Part 5: Plate Interior Basins and Basin Types
inspired by the ideas of Potter (1978) concerning
“big river” systems, that Proterozoic sedimentary
basins in northwestern Canada were the remnants
of enormous river systems that emerged when
major orogenies were taking place in areas to the
east of the sedimentary basins (Young, 1978, 1979).
Paleocurrents determined from measurement of
cross-bedding in thick fluvial deposits preserved
in these basins showed consistent west-northwest-
erly transport (e.g., Young, 1979; Fig. 29.2). Initial
studies of early Neoproterozoic sandstones from
the Amundsen basin revealed that nearly half of
the zircon analyzed was of an age unique to the
Grenville Province of eastern Laurentia and thus
must have been transported more than 3000 kilo-
meters to their final resting place on the other side
of the continent (Rainbird et al., 1992; Fig. 29.3). A
follow-up study on correlative deposits from the
northern Canadian Cordillera 1000 km away (Rain-
bird et al., 1997), and from Paleozoic strata along
the western margin of North America (Gehrels
et al., 1995), documented the widespread occur-
rence of Grenville-age detritus thereby lending
support to the findings of the initial study. The
advent of ion microprobe and laser-ablation ICP-
MS analytical technology has facilitated numerous
detrital zircon provenance studies of similar-age
rocks, younger rocks, and modern sediments
and has revealed that the Grenvillian sedimenta-
tion episode was voluminous and widespread
(e.g., Becker et al., 2005; Cawood et al., 2007;
Dickinson, 2008; Eriksson et al., 2003; Mueller
et al., 2007).
RODINIA AND THE GRENVILLIAN
OROGENY
Rodinia was a supercontinent composed of most of
the Earth’s continental blocks that amalgamated at
the end of the Mesoproterozoic era (e.g., Hoff-
man, 1991; Fig. 29.1). It probably formed between
1200 and 1000 million years ago and began to
break-up about 800 million years ago. Many recon-
structions have been proposed for Rodinia but
most are based on correlation and linkage of so-
called Grenvillian-age orogenic belts that occur
within each of the blocks (Dalziel, 1991; Moores,
1991). Though the configuration of the interior
cratons of Rodinia is now reasonably well estab-
lished, recent reconstructions still differ consider-
ably in detail. Most reconstructions portray a
supercontinent with Laurentia as the central or
keystone block surrounded in the southeast by
Baltica, Amazonia, and the West African craton;
in the south by the Rio de la Plata and S~
ao Fran-
cisco cratons; in the southwest with the Congo and
Kalahari cratons; and in the northeast by Australia,
India, and east Antarctica (Fig. 29.1). The positions
of Siberia and North and South China, north of
Laurentia, differ strongly depending on the recon-
struction (Li et al., 2008; Pisarevsky et al., 2003).
Assembly of Rodinia was marked by global col-
lisional orogenesis, the remnants of which are
preserved today as an extensive belt of moderate
to high-grade metamorphic rocks that constitute
the Grenville orogen. Within the Grenville orogen,
the common occurrence of high-grade metamor-
phic rocks at the Earth’s surface today indicates
uplift and erosion of tens of kilometers of crust
since the bulk of mountain building was com-
pleted, over a billion years ago (Jamieson and
Beaumont, 1989; Jamieson et al., 2007). The record
of this prolonged period of exhumation and denu-
dation should be preserved in the huge volumes of
detritus that would have been delivered to sedi-
mentary basins located across the interior of Rodi-
nia by a pan-continental drainage system.
Remnants of some of these interior basins are
now preserved on continental margins that devel-
oped when Rodinia, and the successor Pangea
supercontinent, rifted apart.
The Grenvillian orogeny is best documented in
its type area in eastern Canada and the State of
New York where it records major, northwest-
directed, crustal contraction during the mid- to
late Mesoproterozoic (Gower et al., 2003, 2008).
Most of its rocks are severely deformed and meta-
morphosed from upper amphibolite to granulite
facies with a wide range of rock types and crystal-
lization ages. From work in the Grenville Province
(that part of the orogen exposed in the Canadian
Shield), summarized by (Hoffman, 1989) and
(Davidson, 1995, 1998, 2008), several discrete
orogenic episodes are recognized. Moore and
Thompson (1980) proposed a two-stage Grenvillian
orogenic cycle comprising the Elzevirian and
Ottawan orogenies. Rivers (1997) identified three
events that followed Elzevirian accretion and
amalgamation (1.29–1.19 Ga): Shawinigan (1.19–
1.14 Ga) for collisional and post-collisional mag-
matism, and Ottawan (1.08–1.02 Ga) and Rigolet
(1.0–0.98 Ga) for two younger phases of Grenvillian
orogeny. The ages of these events are constrained by
syn- to post-orogenic granitic plutons that mainly
represent continental arcs, which developed with
accretion of the various components of the orogen.
The Great Grenvillian Sedimentation Event 585
Fig. 29.2. Geological sketch map of Precambrian North America to show the location of Proterozoic sedimentary basins and
orogenic belts in the southeastern part of the continent. The orogenic zones contributed sedimentary detritus to huge river
systems that crossed the continent, spreading northwesterly and forming the thick sandstone successions preserved in the
Thelon, Athabasca and other basins. The Grenvillian orogenic belt is interpreted as the roots of a deeply eroded mountain
chain, resulting from multi-stage continental accretion that formed from about 1.2 to 1.0 billion years ago. Sediments from this
orogen have been recognized by westward-directed paleocurrents (black arrows) and by common occurrence of detrital
zircon grains yielding ages that are characteristic of the Grenville Province (see Fig. 29.3), several thousand kilometers away,
on the other side of the continent. The long blue dashed arrows show the suggested transport direction of the Paleoproterozoic
sediments, shed from the Yapavai and Trans-Hudson orogens (shown in green and pink, respectively, on the map) and the
dashed red arrows illustrate transport directions proposed for Neoproterozoic sediments deposited in the, Grenvillian
hinterland basins, located along the western margin of Laurentia (e.g., Amundsen Basin, Mackenzie Mountains, Uinta
Mountains Group [UMG], and Grand Canyon Supergroup [GCS]). In North America, the proximal part of the Grenvillian
depositional system is represented by deposits located in the Middle Run Basin (subsurface) and Midcontinent Rift-MCR.
The Grenville Front is the approximate position of the thrust-fault marking the inboard edge of Grenvillian deformation. The
position and extent of the Grenvillian mountain belt in the supercontinent Rodinia, is shown in orange on Figure 29.1.
Source: Simplified from Whitmeyer and Karlstrom (2007).
586 Part 5: Plate Interior Basins and Basin Types
It has been proposed that many of the plutons were
particularly fertile with respect to the production of
zircon (Moecher and Samson, 2006). Earlier Meso-
proterozoic events (Tucker and Gower, 1994) and
the 1.65 Ga Labradorian orogeny (Sch
arer and
Gower, 1988; Gower et al., 2003) may have played
an important role in the evolution of the southeast
margin of Laurentia that was later incorporated in
the Grenville orogen.
Other prominent late Mesoproterozoic moun-
tain belts on other cratons include the Sveco-
Norwegian orogen of Baltica (Bingen et al., 2008a,
2008b), which is essentially the northeastward
extension of the Grenville Province, the Suns
as
orogen that lies along the western margin of Ama-
zonia (Santos et al., 2008), the Maud, Rayner, and
Wilkes provinces of East Antarctica (Mawson Cra-
ton; Fitzsimons, 2000), with possible links to the
Albany-Fraser and Musgrave orogens in southern
and central Australia (Black et al.,1992; Cawood and
Korsch, 2008; Clark et al., 2000), the Eastern Ghats
belt of India (Mezger and Cosca, 1999) and the
Namaqua-Natal province, located along the south-
west margin of the Zimbabwe craton (Fitzsimons,
2000; Fig. 29.1). The Kibaran and Irumide belts of
the Congo craton are prominent Grenville-age
orogens that likely formed during the amalgam-
ation of Rodinia (Kokonyangi et al., 2006; De
Waele et al., 2008).
THEORY: BIG RIVER SYSTEMS IN
THE PROTEROZOIC
Thick, mature, cratonic sheet sandstones
Thick, quartz-rich sandstone deposits are most
commonly preserved in rocks that formed between
about 2500 and 400 million years ago, a period
representing nearly half of Earth history (e.g., Soe-
gaard and Eriksson, 1989). Some of the thickest and
most extensive of these dominantly terrestrial
deposits occur above regionally extensive uncon-
formities that are developed on the crystalline
rocks of broad and stable Precambrian cratons
(Eriksson and Donaldson, 1986). This period is
marked by Earth’s earliest examples of thick (some-
times exceeding 1000 m), widespread deposits of
mature sandstone (mainly quartzarenite), many of
which show evidence of deposition in broad,
braided river channels (e.g., Athabasca Group of
northern Saskatchewan, Ramaekers and Catu-
neanu, 2004), Dubawnt Supergroup of Thelon
and Baker basins (Rainbird et al., 2003); Baraboo
interval of north-central United States (Medaris
et al., 2003, 2007); Roraima Supergroup of south-
ern Venezuela (Santos et al., 2003); and the Huron
Supergroup of south-central Canada (Young
et al., 2001), to name a just a few (North American
examples located on Figure 29.2). These deposits
Fig. 29.3. Probability density diagram with superimposed histogram of detrital zircon U-Pb ages from early Neoproterozoic
sandstones of the Mackenzie Mountains and Amundsen basins of northwestern Canada. ID-TIMS data. Source: From
Rainbird et al. (1992, 1997).
The Great Grenvillian Sedimentation Event 587
contain a relatively high proportion of coarse, bed-
load material and there is little evidence of the
existence of deep channels. These characteristics
indicate a very different hydrologic regime from
that observed in modern humid regions and may be
ascribed to the lack of sediment-stabilizing conti-
nental vegetation (Schumm, 1968). High mineral-
ogical maturity of sandstones and the chemical
composition of interbedded mudstones (they are
enriched in alumina and depleted in alkali and
alkaline earth elements) indicate a regime of vig-
orous continental weathering under a climate that
favored strong chemical alteration (Nesbitt
et al., 1996).
Our geochronological work and that of others
shows that the ages of some of the deposits
described above range from roughly the same as
that of major episodes of global-scale orogenesis up
to about 200 million years younger (Campbell and
Allen, 2008; Rino et al., 2008). We think that their
ages and their paleo-transport directions, as
obtained from studying structures such as cross-
bedding, support the idea that they were the pro-
ducts of vigorous weathering and erosion of vast
mountain ranges developed during these periods
of supercontinent amalgamation. The source
regions are, in many cases, quite remote from
their eventual sites of deposition. This is in appar-
ent contrast to at least some modern large river
systems, for example the Amazon and Mississippi
(Iizuka et al., 2005; Mapes et al., 2004), which
contain detritus mainly derived from local sources.
The difference in provenance signature may be due
to the lack of vegetation during Proterozoic time,
resulting in higher rates of erosion, sediment yield
and more efficient sediment throughput.
Stratigraphic correlations and paleogeography
The idea that several large intracontinental sedi-
mentary basins of late Paleoproterozoic to Meso-
proterozoic age might be remnants of a formerly
continuous blanket of sand that was deposited
unconformably on crystalline rocks of the Cana-
dian Shield was first proposed by Fraser
et al. (1970). The basins are isolated now because,
after burial and lithification, they were gradually
uplifted and eroded so that only dish- to trough-
shaped depressions are preserved where the thick-
est parts of the blanket were deposited (Fig. 29.2).
Evidence for their formerly greater extent includes
the presence of small outliers of sandstone, well
beyond the present-day basin margins, and
numerous exposures of regolith that show evi-
dence of reworking by formerly overlying fluvial
deposits (Hadlari et al., 2004). The basins record a
composite history and are localized at the sites of
earlier, syn-orogenic extension and related subsi-
dence and sedimentation events (Aspler
et al., 2004; Rainbird et al., 2003; Ramaekers
et al., 2007). The sequence stratigraphy, sedimen-
tology and depositional age of the basinsfill are
similar and can be correlated between depocenters
that are thousands of kilometers apart. Sedimen-
tology shows that the bulk of the sandstones in the
post-orogenic sequences were deposited by west-
erly flowing braided streams. A similar paleocur-
rent pattern emerges from each of the basins
(Campbell, 1979; Rainbird et al., 2007; Ramaekers
et al., 2007; Ross, 1983; Fig. 29.2), supporting the
notion that they are remnants of a regional sand
blanket deposited by a series of large river systems.
These observations and those from younger Pro-
terozoic basins led to the suggestion that the pre-
served sedimentary basins were the remnants of
enormous river systems that emerged when major
orogenies were taking place in areas to the east of
the sedimentary basins (Young, 1978, 1979; Rain-
bird et al., 1992). Such conditions and relation-
ships existed not once but at least three times
during the evolution of the ancient North Ameri-
can craton. Examples include the Paleoproterozoic
Hudsonian orogeny and sedimentary rocks in the
Athabasca, Thelon, Hornby Bay, and Elu basins;
the Mesoproterozoic Grenvillian orogeny and the
contemporaneous Middle Run, Torridonian, and
Amundsen basins (Fig. 29.2; see below); the Paleo-
zoic Appalachian-Hercynian orogeny and the Cen-
tral Appalachian (eastern Kentucky-northern
Tennessee), Eastern Interior (Illinois-western Indi-
ana) basins, and southwestern US basins. In each
case, the orogenies can be linked to the formation of
supercontinents (Nuna, Rodinia, and Pangea,
respectively; e.g., Rainbird and Young, 2009).
Testing the theory: detrital zircon geochronology
Provenance analysis attempts to identify source
areas from which sedimentary detritus was
derived and thereby glean important information
about the origin and evolution of a sedimentary
basin. The method has been revolutionized by the
introduction of detrital zircon geochronology, a
technique whereby zircon grains are separated
from clastic sedimentary rocks and analyzed for
their U-Pb isotopic age (e.g., Fedo et al., 2004). Like
588 Part 5: Plate Interior Basins and Basin Types
the rest of the grains in a sandstone, the zircon
grains derive from weathering of an older igneous,
metamorphic or sedimentary rock, so by compar-
ing the age of a particular zircon grain with the ages
of potential source rocks, matches can be made to
constrain its provenance. In the past 10–15 years,
the field of detrital zircon geochronology has been
modernized by the development of microbeam
analytical techniques that can sample and analyze
sub-regions of individual zircon grains rapidly and
with considerable precision.
THE GRENVILLIAN SEDIMENTATION
EVENT AND PALEOGEOGRAPHY OF
RODINIA
The amalgamation of the Rodinia supercontinent
was associated with deposition of large volumes of
clastic sediment produced by erosion of a broad
and extensive chain of mountains, the Grenvillian
orogeny. We consider that the sediments were
transported away from the orogeny by a pan-con-
tinental river system, the proximal and distal ele-
ments of which are preserved in sedimentary
basins mainly in North America, Greenland, and
the United Kingdom (see Rainbird et al., 1992,
1997). Since the publication of the big river
model, numerous studies, all of them employing
detrital zircon geochronology, have added addi-
tional pieces to the puzzle of Rodinian paleogeog-
raphy (Cawood et al., 2004; Kirkland et al., 2008;
Krabbendam et al., 2008; Mueller et al., 2007;
Santos et al., 2002; Dehler et al., 2010).
Grenvillian Foreland basins of Laurentia
One of the major breakthroughs in testing the
model of a pan-continental river system draining
the Grenville mountains was the discovery of sed-
imentary deposits representing the proximal parts
of the drainage system in Laurentia. In North
America, a survey of existing geological maps
and studies of the Grenville thrust front, where it
is exposed in eastern Canada and US (Grenville
Front; Figs. 29.2 and 29.4), shows high grade meta-
morphic rocks of the Grenville Province juxta-
posed against reworked crust of the Superior,
Southern, and Nain provinces with no vestige of
any syn-orogenic sedimentary basin in the fore-
land (e.g., Wheeler et al., 1997; Fig. 29.2). The
deposits were first recognized in the COCORP
(Consortium for Continental Reflection Profiling)
OH-1 seismic reflection profile and other strati-
graphic drilling tests from the subsurface of west-
ern Ohio (Shrake et al., 1991; Drahovzal et al., 1992;
Fig. 29.4). These data outlined an asymmetric,
westward-tapering wedge of coarse, immature
clastic red-beds of interpreted Neoproterozoic
age straddling the Grenville Front and referred to
as the Middle Run Formation. The seismic profiles
also reveal a wide, well-defined zone of east-dip-
ping reflectors inferred to represent thrust struc-
tures of the Grenville front tectonic zone
(Hauser, 1993). In the subsurface, southwest of
OH-1, is a shallow, east-dipping sequence of sed-
imentary strata, similar to Middle Run Formation
elsewhere. Together these rocks were interpreted
as the molasse phase of a previously unrecognized
foreland basin to the Grenville orogen (Middle Run
Basin, Figs. 29.2 and 29.4; Hauser, 1993). This
interpretation is supported by the detrital zircon
geochronology of drill core samples from the Mid-
dle Run Formation (see Fig. 29.5), which reveals
that a very high percentage of the basin fill was
derived from erosion of the adjacent Grenville
Province (Santos et al., 2002). A recent re-interpre-
tation based on reprocessing of the COCORP seis-
mic reflection data and limited well-core data
recognized four, unconformity-bounded, Protero-
zoic sedimentary sequences, which are interpreted
to represent episodic sedimentation correspond-
ing to alternating stages of extension and contrac-
tion during the Grenvillian orogeny (Baranoski
et al., 2009; cf. Rivers, 1997). The older sequences
are interpreted to record deposition into a system
of NW-trending, normal fault-bounded basins (e.
g., Fort Wayne and East Continent rifts; Fig. 29.4),
which were uplifted, eroded and subsequently
buried by a westward-prograding clastic wedge
(e.g., Middle Run Basin) marking foreland basin
development toward the end of the Grenvillian
orogeny (Baranoski et al., 2009). The older rift
basins could have been formed by terrane collision
at the beginning of the Grenvillian orogeny
(Donaldson and Irving, 1972) and thus may be
analogous to “impactogens” like the Rhine graben
(Seng
or et al., 1978) or tectonic-escape basins in
the hinterland of the Himalayan orogen (Morley
et al., 2001). Alternatively, they were initiated by
mantle plume-related extension (Hauser, 1996).
The general stratigraphic succession, limited
age constraints and inferred depositional and tec-
tonic settings of the sequences described by Bar-
anoski et al. (2009) allow correlation with the
Keweenawan Supergroup, exposed at surface in
The Great Grenvillian Sedimentation Event 589
the Midcontinent Rift system to the north
(Figs. 29.2 and 29.4). It comprises a lower rift-
associated subaerial volcanic and volcano-sedi-
mentary succession ranging in age from 1109 to
1087 Ma (see references in Hollings et al., 2007)
and an overlying sedimentary succession that
comprises two to three unconformity-bounded
sequences as summarized by Ojakangas et al.
(2001). The lower part of the sedimentary succes-
sion is typified by coarse, lithic-rich, sedimentary
rocks deposited adjacent to rift margins on alluvial
fans with rapid facies changes to cyclicly depos-
ited fluvial and lacustrine facies as exemplified by
the Oronto Group of Wisconsin and northern
Michigan (Morey and Ojakangas, 1992). Uncon-
formably overlying braided stream deposits of the
Bayfield Group are more texturally and composi-
tionally mature. Correlative rocks include the
Fond du Lac formation, an up to 2100 m-thick
deposit of conglomerates and arkoses deposited
by an east-flowing fluvial system (Morey and
Ojakangas, 1992). Detrital zircon geochronology
Fig. 29.4. Neoproterozoic sedimentary basins of the Grenvillian foreland of North America. Most of the deposits, except for
those of the Keweenawan Supergroup in the Great Lakes region, are inferred in the subsurface from seismic profiles and well-
core information. Source: Adapted from Baranoski et al. (2009), Figure 29.1.
590 Part 5: Plate Interior Basins and Basin Types
of the Fond du Lac and overlying Hinckley sand-
stone show prominent Mesoproterozoic age peaks
reflecting derivation from underlying volcanic
rocks but also from the adjacent Grenville Province
and Superior Province, which was reworked by the
Grenvillian orogeny (Finley-Blasi et al., 2006;
Wirth et al., 2006), supporting the paleogeographic
model that these rocks may be part of the late
Grenvillian foreland basin described above. Paleo-
currents from the Bayfield Group and equivalents
show localized flow at basin margins but overall
east-northeasterly flow, which is consistent with
trunk rivers flowing through the rift and parallel to
the Grenvillian orogenic front. These strata
also correlate with the Jacobsville sandstone
(Ojakangas et al., 2001), a >900 m-thick succession
of subarkosic to sublithic arenites, conglomerates
and siltstones, which are exposed mainly in north-
ern Michigan’s Keweenaw Peninsula and eastward
along the south shore of Lake Superior into Ontario
(Kalliokoski, 1982). Tectonic uplift and provenance
from the south is indicated by petrologic data,
paleocurrent analysis (Kalliokoski, 1982; Hedg-
man, 1992) and detrital zircon ages (J. Craddock
and K. Wirth, personal communication, 2009).
Paleocurrents from fluvial units in the main part
of the basin suggest axial (NE and SW) transport,
similar to the Bayfield Group, and perhaps related
to development of a fluvial trunk system, as is
commonly observed in foreland basins.
Fig. 29.5. Location and probability density diagrams of detrital zircon U-Pb ages from Neoproterozoic sandstones of
metasedimentary units within the Caledonian orogen in Scotland, eastern Laurentia, and Greenland. Source: SHRIMP
data sources for the Middle Run Formation are Santos et al. (2002); Sleat Group, Kinnaird et al. (2007); Stoer Group and
Torridon Group, Rainbird et al. (2001); Moine (Cawood et al., 2004; Friend et al., 2003); for the Grampian Group, Cawood
et al. (2003); and for the upper Dalradian Group, Cawood et al. (2003). SHRIMP data sources for Greenland successions in
Cawood et al. (2007b). Map modified from Cawood et al. (2007b).
The Great Grenvillian Sedimentation Event 591
Correlative basins of the North Atlantic region
Paleomagnetic data suggest that Baltica underwent
95clockwise rotation with respect to Laurentia at
the end of the Mesoproterozoic, sometime after
1265 Ma (and probably 1120 Ma) and prior to
1000 Ma, synchronous with formation of the Gren-
ville mountain chain (Cawood et al., 2010; Pisar-
evsky et al., 2003). Baltica moved from a position in
which its northern margin (current coordinates)
abutted East Greenland to one in which the Scan-
dinavian margin faced Scotland, the Rockall Bank
and southeast Greenland (Fig. 29.6). This oroclinal
bending of Baltica with respect to Laurentia
resulted in formation of a triangular ocean basin
(sphenochasm) to the north of the Grenville-
Sveconorwegian orogen. Remnants of this basin
occur in end Mesoproterozoic to Neoproterozoic
successions preserved around the margins of the
North Atlantic. The successions are divisible into
two cycles of accumulation at around 1000 Ma and
900 Ma that were deformed and stabilized during
tectonothermal episodes involving crustal thick-
ening and igneous activity associated at 980–
910 Ma and 840–710 Ma (Cawood et al., 2010).
The orientation of preserved sedimentary succes-
sions is generally parallel to the Laurentian margin
suggesting that the geometry of the margin con-
trolled sedimentation. This same margin geometry
was reactivated during contractional deformation
that stabilized the successions. The region
remained a site of lithospheric weakness during
the Caledonian cycle, initiated by the separation of
Baltica and opening of the northern Iapetus ocean
at the end of the Neoproterozoic (Cawood and
Nemchin, 2001).
Remnants of the first sedimentary cycle are pre-
served in Scotland, Shetland, East Greenland, East
Svalbard, and northern Norway (Cawood et al.,
2007; Gee and Tebenkov, 2004; Kalsbeek et al.,
2000; Kirkland et al., 2008; Watt and Thrane, 2001).
These early successions accumulated between
1030 and 980 Ma based on constraints from youn-
gest detrital zircon grains and overprinting by
tectonothermal events. Sedimentological studies
in the least deformed sequences in Scotland
indicate accumulation in high energy fluvial
braid-plain environments (Friend et al., 2003;
Krabbendam et al., 2008; Nicholson, 1993;
Stewart, 2002; Williams, 2001). Depositional
settings of other units that accumulated during
this late Mesoproterozoic to early Neoprotero-
zoic sedimentary cycle cannot be confidently
resolved due to subsequent deformation.
The second depositional cycle includes succes-
sions in Scotland and northern Norway (Cawood
et al., 2003, 2007; Kirkland et al., 2007). These
successions are generally spatially discrete from
first cycle units as the two have not been recog-
nized in stratigraphic contact. Constraints from
youngest detrital zircon indicate deposition after
920 Ma with an upper age limit provided by tecto-
nothermal events at around 840 Ma or younger.
Siliciclastic units of inferred similar age also
occur in the SW terranes of Svalbard but lack
evidence for an extensive mid-Neoproterozoic
tectonothermal event (Gee and Tebenkov, 2004).
Fig. 29.6. Reconstruction of eastern Laurentia, Baltica, and
northern Amazonia for late Mesoproterozoic to Neoproter-
ozoic. Baltica is shown in both its pre-1265 Ma (grey shad-
ing) and post-1000 Ma (coloured) positions with respect to
Laurentia (after Cawood et al., 2010). The rotation of Baltica
created a triangular ocean basin, the Aesir Sea. The disrup-
tion of Svalbard by a series of Caledonian (mid-Paleozoic)
strike-slip faults into Western, Central and Eastern terranes
follows Harland (1997). The South Western and Nordaus-
tlandet terranes of Gee and Tebenkov (2004) lie within the
Western and Eastern terranes of Harland (1997), respec-
tively. Scotland is also disrupted by Caledonian faults
into the Grampian terrane, Moine succession and Hebridean
foreland. Red coloured blocks correspond to sedimentary
sequences associated with first cycle successions and yel-
low with the second cycle. Abbreviations: GT – Grampian
terrane of Scotland; M – Moine succession of Scotland; Hf –
Hebridean foreland of Scotland; K – Krummedal succession
of East Greenland; N – Nordaustlandet terrane of Svalbard;
Rb – Rockall Bank; Sa/So – Svaerholt and Sørøy successions
of northern Norway; Sh – Shetland Islands; Sn – Sveconor-
wegian orogen; Ss – Sunsas orogen; Sv – central Svalbard;
and Sw – Southwestern terranes of Svalbard.
592 Part 5: Plate Interior Basins and Basin Types
Detrital zircon grains from the two successions
are dominated by late Paleoproterozoic and late
Mesoproterozoic ages inferred to have been
derived from source terranes exposed within east-
ern Laurentia (e.g., Labradoran, Makkovikian-
Ketilidian) and the Grenville-Sveconorwegian
mountain belts (Fig. 29.5). They lack significant
Archean detritus, which typically is exposed adja-
cent to the sites of sediment accumulation or is
a component of underlying successions (e.g.,
Rainbird et al., 2001), suggesting subdued relief
in these regions (Cawood et al., 2003, 2007, and
references therein).
These late Mesoproterozoic to Neoproterozoic
successions are unconformably overlain by late
Neoproterozoic (<750 Ma) to early Paleozoic
(Cambro-Ordovician) siliciclastic- and carbonate-
dominated successions that accumulated during
rifting and subsequent development of passive
margins associated with the breakup of Rodinia
and opening of the Iapetus Ocean at the start
of the Appalachian-Caledonian orogenic cycle
(Cawood and Nemchin, 2001). Detrital zircon
data from these units indicate continued input
from the Grenvillian mountains and, at least
locally, increased input from the Archean cratons
(Fig. 29.5, Cawood and Nemchin, 2001; Cawood
et al., 2003, 2007a, 2007b; Dhuime et al., 2007).
In the Appalachian Orogen, along-strike, correl-
ative late Neoproterozoic to early Paleozoic suc-
cessions, occur to the south of the Grenville-
Sveconorwegian deformation front and may
directly overlie basement rocks of the Grenville
Province. Detrital zircon ages from these units are
almost exclusively of Mesoproterozoic age and
older detritus is absent suggesting that the Grenville
orogen was acting as a topographic and drainage
divide that prevented input from Archean and
Paleoproterozoic terranes in the interior of Laur-
entia (Cawood et al., 2007, and references therein).
Distal remnants of the Grenvillian river system
The distal parts of the Grenvillian river system are
preserved in the Amundsen and Mackenzie basins,
covering a broad region of northwestern Canada,
(Fig. 29.2) and were the first indication of the
potentially enormous breadth and width of the
Grenvillian sedimentation event (Rainbird et al.,
1997). These deposits are mature quartzarenite
sandstones bodies, up to 1.8 km thick in the case
of the Katherine Group in the Mackenzie Moun-
tains (see, Long et al., 2008). The sandstones are
mostly braided fluvial deposits (Fig. 29.7), with
consistent west-northwesterly paleocurrent indi-
cators (Fig. 29.2). They are interbedded with del-
taic deposits and shallow marine carbonates and
evaporites laid down on the edge of an epicratonic
basin that subsequently rifted apart with the
breakup of Rodinia (Rainbird et al., 1996).
Detrital zircon geochronology of Neoproterozoic
sedimentary rocks from sedimentary basins
located elsewhere along the western margin of
Laurentia, and now exposed mainly in the Cordil-
lera of California, Nevada, Utah, and northern
Mexico, yielded detrital zircon with late Mesopro-
terozoic ages (e.g., Stewart et al., 2001). Although
they attribute some of these ages to provenance
Fig. 29.7. Distal end of the Grenvillian
river system represented by braided-type
deposits of the early Neoproterozoic Sha-
ler Supergroup (Nelson Head Formation),
Amundsen Basin, Northwest Territories,
Canada (see Rainbird et al., 1996). Alter-
nating, thick, tabular channel sandstones
(light tones) and thin, tabular to broadly
lenticular flood-plain deposits composed
of fine sandstone and siltstone. Exposed
section has a true thickness of about 35 m.
The Great Grenvillian Sedimentation Event 593
from local volcanic centres, much of the material is
considered to have derived from the Grenville
Province of eastern Laurentia, elements of which
extend into west Texas and northern Mexico (Bick-
ford et al., 2000; Barth, 2001). One candidate for a
distal component of the Grenvillian river system is
the Nankoweap Formation, a 100 m thick section of
red crossbedded sandstones and subordinate
mudrocks sandwiched unconformably between
the ca. 1250–1100 Ma Unkar Group (Timmons
et al., 2005) and the ca 800–742 Ma Chuar Group
(Dehler et al., 2001) in the Grand Canyon Super-
group of the southwestern United States (Fig. 29.2).
The detrital zircon U-Pb age profile for the Nanko-
weap Fm. (Timmons et al., 2005, Figure 29.8a)
matches so-called bar codes established for the
Amundsen and Mackenzie basins and those of
more proximal deposits (cf. Hinckley sandstone;
Finley-Blasi et al., 2006). Another potential correl-
ative and component of the Grenvillian river sys-
tem is the early Neoproterozoic Uinta Mountains
Group and Big Cottonwood Formation of eastern
Utah (Fig. 29.2; Link et al., 1993). A series of prov-
enance studies of these rocks (Condie et al., 2001;
Mueller et al., 2007; Dehler et al., 2010) yielded
detrital zircon U-Pb age profiles similar to those
described above (Fig. 29.8b), and support the idea
that this is yet another remnantof the vast drainage
network that delivered detritus westward across
Laurentia to an epicratonic basin or basins that
Fig. 29.8. Probability density diagrams with superimposed histograms of LA-ICPMS detrital zircon U-Pb ages from early
Neoproterozoic sandstones from the (a) Nankoweap Formation (Grand Canyon Supergroup; data from Timmons et al., 2005)
and the (b) Big Cottonwood Formation, Uinta Mountains Group. Source: Data from Dehler et al., 2010.
594 Part 5: Plate Interior Basins and Basin Types
existed along its western and northern margin
(Figs. 29.1 and 29.2).
Beyond Laurentia
A possible example of a Grenvillian foreland basin
in Siberia is the Mayamkan Formation, a 1 km-
thick upward-coarsening succession of immature
alluvial red-beds of Neoproterozoic age located in
the Uchur-Maya depression of eastern Russia
(Rainbird et al., 1998). Detrital zircon grains from
this unit are dominated by Mesoproterozoic ages,
unlike any known ages of potential source terranes
from the Siberian craton to the west. Paleogeo-
graphic information indicates that the basin
received its detritus from an eastern source, pos-
tulated to be a Grenvillian orogenic belt that rifted
away with the break-up of Rodinia and eventual
opening of the proto-Pacific ocean.
Neoproterozoic basins in southern Africa,
largely of Pan-African age and related to assembly
of Gondwana, record input from late Mesoproter-
ozoic orogenic belts associated with formation of
the supercontinent Rodinia. A provenance study
of Neoproterozoic siliciclastic successions in the
Gariep Belt of southwestern Africa and equivalents
in the Dom Feliciano Belt in South America are
dominated by zircon grains with ages of 1200–
1000 Ma with a further peak between 2000 and
1700 Ma (Basei et al., 2005). These ages compare
well to the pre-Gariep basement geology of south-
ern Africa with the Mesoproterozoic (Grenvillian)
detritus likely derived from the Namaqua-Natal
Belt (Fig. 29.1) and the Paleoproterozoic detritus
from the Richtersveld terrane, an Andean type arc
source. In Zambia and the Democratic Republic of
Congo, Neoproterozoic successions host major Cu-
Co deposits within the Central African Copperbelt
(Master et al., 2005). Detrital zircon ages are mainly
Paleoproterozoic (2080–1835 Ma) and probably
derived from the Lufubu Metamorphic Complex
with only a minor Mesoproterozoic component
which was derived from the Kibaran Belt
(Fig. 29.1).
EVIDENCE FOR RECYCLING OF
GRENVILLIAN BASINS INTO
YOUNGER SEDIMENTARY DEPOSITS
Following accumulation of continent-scale clas-
tic wedges during late Mesoproterozoic to early
Neoproterozoic time, Grenville-age detritus was
reworked into younger stratigraphic successions.
In western North America, Grenville-age detrital
zircon comprises a significant proportion of
most sedimentary successions, including late
Neoproterozoic successions of the southwestern
US (Apache Group: Stewart et al., 2001), Cordil-
leran miogeoclinal strata (Gehrels et al., 1995) and
off-shelf basinal strata (Gehrels et al., 2000) of
Cambrian through Triassic age, Jurassic strata of
the Colorado Plateau (Dickinson and Geh-
rels, 2009), and Cretaceous foreland-basin strata
(Dickinson, 2008). A well-preserved analogue of
the Rodinian paleogeographic model comes from
the supercontinent Pangea, which amalgamated
during the Appalachian-Hercynian orogeny from
the Pennsylvanian to Triassic. The paleogeogra-
phy of the Laurasia block of Pangea was domi-
nated by a series of coalescing, alluvial-deltaic
wedges and axial braided rivers that filled fore-
land basins formed by flexural loading along the
Alleghenian-Appalachian thrust front (Absaroka
Sequence; Sloss, 1988). The Central Appalachian
Basinisanexamplewhoseinllhasbeeninter-
preted to represent an Amazon-scale drainage
system (Archer and Greb, 1995). As the foreland
basins filled, excess detritus was transported
westward, across the craton, by fluvial and
eolian processes, eventually reaching the west-
ern margin of Laurentia (e.g., Dickinson and
Gehrels, 2003). Provenance studies of these
deposits reveal that the majority of their detrital
zircon records ages characteristic of the Grenville
Province (Becker et al., 2005; Gray and Zeitler,
1997; McLennan et al., 2001; Thomas et al., 2004;
Dickinson and Gehrels, 2003). This detritus
was derived directly from uplift and erosion of
Grenville Province rocks during Appalachian
orogenesis or recycling of detritus from Grenvillan
foreland basin deposits such as those described
above.
Late Mesoproterozoic detritus has been contin-
ually recycled into younger stratigraphic succes-
sions and remains a significant component of
modern river sediments. Detrital zircon ranging
in age from 1.2 to 1.0 Ga represents a significant
component of the detritus in modern rivers, such
as the Colorado (Grove and Kimbrough, 2008),
Mississippi, several Appalachian rivers (Eriksson
et al., 2003, 2004) and are recognized in numerous
other river systems around the world (Campbell
and Allen, 2008). Eriksson et al. (2004) noted a
discrepancy between the discrete events defined
by geochronological studies of Grenvillian rocks in
The Great Grenvillian Sedimentation Event 595
eastern Laurentia and a more continuous record of
magmatism seen in the detrital zircon ages from
Appalachian rivers. They suggested that the age
of the Grenville Province and orogenic episodes
of the Grenvillian orogeny are perhaps better
reflected by the broader sampling base provided
by zircon from modern Appalachian rivers. In
addition to present-day Grenvillian basement
exposures, modern river zircon grains are recycled
from sedimentary rocks which were derived, in
part, from Grenvillian crystalline rocks that either
have been removed by erosion or are presently
overlain by younger cover.
In considering the relative abundance of Gren-
ville-age detritus, it is important to understand that
Grenville-age rocks are considered to contain a
higher abundance of zircon than rocks in other
igneous assemblages (Dickinson, 2008; Moecher
and Samson, 2006). Grenville-age zircon may be
accordingly over-represented in most sedimentary
assemblages, which should be considered when
attempting to reconstruct volumes of sediment
generated. This does not significantly impact our
conclusions concerning the breadth of dispersal of
Grenville-age detritus, however, which are based
largely on the presence of 1.2–1.0 Ga grains in
deposits more than 3000 km away from their
inferred source rather than their absolute or rela-
tive abundance.
SUMMARY
The Grenvillian tectonomagmatic event was con-
tinuous between 1250 and 950 Ma and produced
massive and extensive mountain belts and a huge
volume of sedimentary detritus that has been
recycled through the sedimentary record for the
past 1.0 billion years. Sediment yield and through-
put was enhanced initially by vigorous chemical
weathering and a lack of sediment-stabilizing veg-
etation until the establishment of land plants in the
Silurian. Evidence for the Great Grenvillian sedi-
mentation event is documented here mainly from
studies based in North America and adjacent con-
tinental components of Laurentia but must doubt-
lessly be recorded in sedimentary deposits
preserved on other continental blocks that were
part of the supercontinent Rodinia. Ongoing and
future provenance studies of these other sedimen-
tary sequences will provide a rigorous evaluation
of the extent and significance of Grenville-derived
clastic strata.
ACKNOWLEDGMENTS
Rob Rainbird would like to thank Grant Young for
his vision and guidance and Charlie Jefferson for
support of this research. He would also like to
thank Bill Davis, Larry Heaman, Vicki McNicoll,
and Richard Stern for help with collection of
detrital zircon data. Mark Baranoski reviewed
the section on Grenvillian foreland basins and
provided a copy of Figure 29.1 from his paper.
The chapter benefitted from reviews by Stan
Finney, Scott Samson, and Thomas Hadlari.
Thank you to editor Cathy Busby for her encour-
agement and a particularly thorough critique of
the manuscript. This is Geological Survey of
Canada contribution # 20090432.
REFERENCES
Archer, A.W., and Greb, S.F. (1995) An Amazon-scale
drainage system in the early Pennsylvanian of central
North America. Journal of Geology, 103, 611–628.
Aspler, L.B., Chiarenzelli, J.R., and Cousens, B.L. (2004)
Fluvial, lacustrine and volcanic sedimentation in the
Angikuni sub-basin, and the initiation of 1.84 Ga
Baker Lake Basin, western Churchill Province, Nunavut,
Canada. Precambrian Research, 129, 225–250.
Baranoski, M.T., Dean, S.L., Wicks, J.L., and Brown, V.M.
(2009) Unconformity bounded seismic reflection
sequences define Grenville-age rift system and foreland
basins beneath the Phanerozoic in Ohio. Geosphere, 5,
140–151.
Barth, A.P., Wooden, J.L., and Coleman, D.S. (2001)
SHRIMP-RG U-Pb zircon geochronology of Mesoproter-
ozoic metamorphism and plutonism in the southwest-
ernmost United States. Journal of Geology, 109, 319–327.
Basei, M.A.S., Frimmel, H.E., Nutman, A.P., Preciozzi, F.,
and Jacob, J. (2005) A connection between the Neopro-
terozoic Dom Feliciano (Brazil/Uruguay) and Gariep
(Namibia/South Africa) orogenic belts: evidence from
a reconnaissance provenance study. Precambrian
Research, 139, 195–221.
Becker, T.P., Thomas, W.A., Samson, S.D., and Gehrels, G.E.
(2005) Detrital zircon evidence of Laurentian crustal
dominance in the lower Pennsylvanian deposits of the
Alleghanian clastic wedge in eastern North America.
Sedimentary Geology, 182, 59–86.
Bickford, M.E., Soegaard, K., Nielsen, K.C., and McLelland,
J.M. (2000) Geology and geochronology of Grenville-age
rocks in the Van Horn and Franklin Mountains area,
West Texas; implications for the tectonic evolution of
Laurentia during the Grenville. Geological Society of
America Bulletin, 112, 1134–1148.
Bingen, B., Anderson, J., S
oderlund, U., and M
oller, C.
(2008a) The Mesoproterozoic in the Nordic countries.
Episodes, 31, 29–34.
Bingen, B., Nordgulen, Ø., and Viola, G. (2008b) A four-
phase model for the Sveconorwegian orogeny, SW Scan-
dinavia. Norwegian Journal of Geology, 88, 43–72.
596 Part 5: Plate Interior Basins and Basin Types
Black, L.P., Harris, L.B., and Delor, C.P. (1992) Reworking of
Archean and Early Proterozoic components during a
progressive, middle Proterozoic tectonothermal event
in the Albany Mobile Belt, Western Australia. Precam-
brian Research, 59, 95–123.
Campbell, F.H.A. (1979) Stratigraphy and sedimentation in
the Helikian Elu Basin and Hiukitak Platform, Bathurst
Inlet-Melville Sound, Northwest Territories, Geological
Survey of Canada Paper 79, 19.
Campbell, I.H., and Allen, C.M. (2008) Formation of super-
continents linked to increases in atmospheric oxygen.
Nature Geoscience, 1, 554–558.
Cawood, P.A., and Korsch, R.J. (2008) Assembling Austra-
lia: Proterozoic building of a continent. Precambrian
Research, 166, 1–35.
Cawood, P.A., and Nemchin, A.A. (2001) Paleogeographic
development of the eastern Laurentian margin: con-
straints from U-Pb dating of detrital zircons from the
northern Appalachians. Geological Society of America
Bulletin, 113, 1234–1246.
Cawood, P.A., Nemchin, A.A., Smith, M., and Loewy, S.
(2003) Source of the Dalradian Supergroup constrained
by U-Pb dating of detrital zircon and implications for the
East Laurentian margin. Journal of the Geological Soci-
ety, London, 160, 231–246.
Cawood, P.A., Nemchin, A.A., and Strachan, R. (2007a)
Provenance record of Laurentian passive-margin strata
in the northern Caledonides: implications for paleodrai-
nageand paleogeography. Geol Soc Am Bull, 119,
993–1003.
Cawood, P.A., Nemchin, A.A., Strachan, R., Prave, T., and
Krabbendam, M. (2007b) Sedimentary basin and detrital
zircon record along East Laurentia and Baltica during
assembly and breakup of Rodinia. Journal of the Geolog-
ical Society, 164, 257–275.
Cawood, P.A., Nemchin, A.A., Strachan, R.A., Kinny, P.D.,
and Loewy, S. (2004) Laurentian provenance and an
intracratonic tectonic setting for the Moine Supergroup,
Scotland, constrained by detrital zircons from the Loch
Eil and Glen Urquhart successions. Journal of the Geo-
logical Society, London, 161, 861–874.
Cawood, P.A., Strachan, R., Cutts, K., Kinny, P.D., Hand, M.,
and Pisarevsky, S. (2010) Neoproterozoic orogeny along
the margin of Rodinia: Valhalla orogen, North Atlantic.
Geology, 38, 99–102.
Clark, D.J., Hensen, B.J., and Kinny, P.D. (2000) Geochro-
nological constraints for a two-stage history of the
Albany-Fraser Orogen, Western Australia. Precambrian
Research, 102, 155–183.
Condie, K.C., Lee, D., and Farmer, G.L. (2001) Tectonic
setting and provenance of the Neoproterozoic Uinta
Mountain and Big Cottonwood groups, Northern Utah:
constraints from geochemistry, Nd isotopes, and detrital
modes. Sedimentary Geology, 141–142, 443–464.
Dalziel, I.W.D. (1991) Pacific margins of Laurentia and East
Antarctica-Australia as a conjugate rift pair: evidence
and implications for an Eocambrian supercontinent.
Geology, 19, 598–601.
Davidson, A. (1995) A review of the Grenville orogen in its
North American type area. AGSO Journal of Australian
Geology and Geophysics, 16, 3–24.
Davidson, A. (1998) An overviewof Grenville Province
geology, in Lucas, S.B., and St-Onge, M. R., eds., The
Geology of North America C-1: geology of the Precam-
brian Superior and Grenville Provinces and Precambrian
fossils in North America. Geological Society of America,
205–270.
Davidson, A. (2008) Late Paleoproterozoic to mid-Neopro-
terozoic history of northern Laurentia: an overview of
central Rodinia. Precambrian Research, 160, 5–22.
Dehler, C.M., Elrick, M., Karlstrom, K.E., Smith, G.A.,
Crossey, L.J., Timmons, J.M., Eriksson, P.G., Catuneanu,
O., Aspler, L.B., Chiarenzelli, J.R., and Martins, N.M.A.
(2001) Neoproterozoic Chuar Group (approximately
800–742 Ma) Grand Canyon; a record of cyclic marine
deposition during global cooling and supercontinent
rifting. Sedimentary Geology, 141–142, 465–499.
Dehler, C.M., Fanning, C.M., Link, P.K., Kingsbury, E.M.,
and Rybczynski, D., 2010, Maximum depositional age
and provenance of the Uinta Mountain Group and Big
Cottonwood Formation, northern Utah: paleogeography
of rifting western Laurentia. Geological Society of Amer-
ica Bulletin, 122, 1686–1699.
De Waele, B., Johnson, S.P., and Pisarevsky, S.A. (2008)
Palaeoproterozoic to Neoproterozoic growth and evolu-
tion of the eastern Congo Craton: its role in the Rodinia
puzzle. Precambrian Research, 160, 127–141.
Dhuime, B., Bosch, D., Bruguier, O., Caby, R., and Pourtales,
S. (2007) Age, provenance and post-deposition metamor-
phic overprint of detrital zircons from the Nathorst Land
group (NE Greenland)–A LA-ICP-MS and SIMS study.
Precambrian Research, 155, 24–46.
Dickinson, W.R. (2008) Impact of differential zircon fertility
of granitoid basement rocks in North America on age
populations of detrital zircons and implications for gran-
ite petrogenesis. Earth and Planetary Science Letters,
275, 80–92.
Dickinson, W.R., and Gehrels, G.E. (2003) U-Pb ages of
detrital zircons from Permian and Jurassic eolianite
sandstones of the Colorado Plateau, USA: paleogeo-
graphic implications. Sedimentary Geology, 163, 29–66.
Dickinson, W.R., and Gehrels, G.E. (2008) Sediment deliv-
ery to the Cordilleran foreland basin: insights from U-Pb
ages of detrital zircons in Upper Jurassic and Cretaceous
strata of the Colorado Plateau. American Journal of Sci-
ence, 308, 1041–1082.
Dickinson, W.R., and Gehrels, G.E. (2009) U-Pb ages of
detrital zircons in Jurassic eolian and associated sand-
stones of the Colorado Plateau: evidence for transconti-
nental dispersal and intraregional recycling of sediment.
Geological Society of America Bulletin, 121, 408–433.
Donaldson, J.A., and Irving, E. (1972) Grenville Front and
rifting of the Canadian Shield. Nature, 273, 139–140.
Drahovzal, J.A., Harris, D.C., Wickstrom, L.H., Walker, D.,
Baranoski, M.T., Keith, B., and Furer, L.C. (1992) The
East Continent Rift Basin: a new discovery. Ohio Divi-
sion of Geological Survey, Information Circular 57, 25 p.
Eriksson, K.A., Campbell, I.H., Palin, J.M., and Allen, C.M.
(2003) Predominance of Grenvillian magmatism
recorded in detrital zircons from modern Appalachian
rivers. Journal of Geology, 111, 707–717.
Eriksson, K.A., Campbell, I.H., Palin, J.M., Allen, C.M., and
Bock, B. (2004) Evidence for multiple recycling in Neo-
proterozoic through Pennsylvanian sedimentary rocks of
the central Appalachian basin. Journal of Geology, 112,
261–276.
The Great Grenvillian Sedimentation Event 597
Eriksson, K.A., and Donaldson, J.A. (1986) Basinal and shelf
sedimentation in relation to the Archean-Proterozoic
boundary. Precambrian Research, 33, 103–121.
Fedo, C.M., Sircombe, K.N., and Rainbird, R.H. (2004)
Detrital zircon analysis of the sedimentary record, Han-
char, J.M., and Hoskin, P. O., eds., Zircon: experiments,
isotopes, and trace element investigation. Mineralogical
Society of America, Reviews in Mineralogy and Geo-
chemistry, 53, 277–303.
Finley-Blasi, L., Davidson, C., Wirth, K., Craddock, J., and
Vervoort, J. (2006) U-Pb dating of Detrital Zircon from the
Neoproterozoic (?) Fond Du Lac and Hinckley Sandstone
Formations near Duluth, Minnesota. Abstracts with Pro-
grams, 38. Boulder, CO, Geological Society of America,
505.
Fitzsimons, I.C.W. (2000) Grenville-age basement provinces
in East Antarctica: evidence for three separate collisional
orogens. Geology, 28, 879–882.
Fraser, J.A., Donaldson, J.A., Fahrig, W.F., and Tremblay, L.
P. (1970) Helikian basins and geosynclines of the
northwestern Canadian Shield, in Baer, A.J., ed., Sym-
posium on Basins and Geosynclines of the Canadian
Shield, Geological Survey of Canada Paper 70–40,
213–238.
Friend, C.R.L., Strachan, R.A., Kinny, P., and Watt, G.R.
(2003) Provenance of the Moine Supergroup of NW Scot-
land: evidence from geochronology of detrital and inher-
ited zircons from (meta)sedimentary rocks granites and
migmatites. Journal of the Geological Society, London,
160, 247–257.
Gee, D.G., and Tebenkov, A.M. (2004) Svalbard: a fragment
of the Laurentian margin, in Gee, D.G., and Pease, V.L.,
eds., The Neoproterozoic Timanide Orogen of Eastern
Baltica. Memoir, 30. Geological Society, 191–206.
Gehrels, G.E. (2000) Introduction to detrital zircon studies of
Paleozoic and Triassic strata in western Nevada and
northern California, in Soreghan, M.J., and Gehrels,
G.E., eds., Paleozoic and Triassic paleogeography and
tectonics of western Nevada and northern California.
Special Paper, 347. Boulder, CO, Geological Society of
America, 1–17.
Gehrels, G.E., Dickinson, W.R., Ross, G.M., Stewart, J.H.,
and Howell, D.G. (1995) Detrital zircon reference for
Cambrian to Triassic strata of western North America.
Geology, 23, 831–834.
Gower, C.F., Kamo, S., and Krogh, T.E. (2008) Indentor
tectonism in the eastern Grenville Province. Precam-
brian Research, 167, 201–212.
Gower, C.F., and Krogh, T.E. (2003) A U-Pb geochronolog-
ical review of Pre-Labordian and Labradorian geological
history of the eastern Grenville Province, in Brisebois, D.,
and Clark, T., eds., G
eologie et ressources min
erales de la
partie est de la Province de Grenville. Qu
ebec, Minist
ere
des Ressources naturelles, DV 2002–03, 142–172.
Gray, M.B., and Zeitler, P.K. (1997) Comparison of
clastic wedge provenances in the Appalachian foreland
using U-Pb ages of detrital zircons. Tectonics, 16,
151–160.
Grove, M., and Kimbrough, D.L. (2008) Evolution of the
Colorado River: culmination of the major Cenozoic
Transformation of SW North America Drainage Patterns,
American Geophysical Union, Fall Meeting 2008,
abstract #T33D-2101.
Hadlari, T., Rainbird, R.H., and Pehrsson, S.J. (2004) Geol-
ogy, Schultz Lake, Nunavut, Geological Survey of
Canada Open File 1839, 1:250,000.
Harland, W.B. (1997) The geology of Svalbard. Memoir, 17.
London, Geological Society of London, 521 p.
Hauser, E.C. (1993) Grenville foreland thrust belt hidden
beneath the U.S. midcontinent. Geology, 21, 61–64.
Hauser, E.C. (1996) Midcontinent rifting in a Grenville
embrace. Geological Society of America Special Paper
308, 67–75.
Hawkesworth, C., Cawood, P., Kemp, T., Storey, C., and
Dhuime, B. (2009) Geochemistry: a matter of preserva-
tion. Science, 323, 49–50.
Hedgman, C.A. (1992) Petrology and provenance of a con-
glomerate facies of the Jacobsville sandstone: Ironwood
to Bergland Michigan. Geological Society of America,
Abstracts with Programs, 24, A329.
Hoffman, P.F. (1989) Precambrian geology and tectonic
history of North America, in Bally, A.W., and Palmer,
A.R., eds., The geology of North America: an overview,
vol. A. Boulder, CO, Geological Society of America,
447–512.
Hoffman, P.F. (1991) Did the breakout of Laurentia
turn Gondwanaland inside-out? Science, 252, 1409–
1412.
Hollings, P., Fralick, P., and Cousens, B. (2007) Early history
of the Midcontinent Rift inferred from geochemistry and
sedimentology of the Mesoproterozoic Osler Group,
northwestern Ontario. Canadian Journal of Earth
Sciences, 44, 389–412.
Iizuka, T., Hirata, T., Komiya, T., Rino, S., Katayama, I.,
Motoki, A., and Maruyama, S. (2005) U-Pb and Lu-Hf
isotope systematics of zircons from the Mississippi River
sand: implications for reworking and growth of conti-
nental crust. Geology, 33, 485–488.
Jamieson, R.A., and Beaumont, C. (1989) Deformation and
metamorphism in convergent orogens: a model for uplift
and exhumation of metamorphic terrains. Evolution of
metamorphic belts, 117–129.
Jamieson, R.A., Beaumont, C., Nguyen, M.H., and Culshaw,
N.G. (2007) Synconvergent ductile flow in variable-
strength continental crust: numerical models with appli-
cation to the western Grenville orogen. Tectonics, 26
TC5005.
Kalliokoski, J. (1982) Jacobsville Sandstone, in Wold, R.J.,
and Hinze, W.J., eds., Geology and Tectonics of the Lake
superior Basin, Geological Society of America, Memoir
156, 147–155.
Kalsbeek, F., Thrane, K., Nutman, A.P., and Jepsen, H.F.
(2000) Late Mesoproterozoic to early Neoproterozoic
history of the East Greenland Caledonides: evidence
for Grenvillian orogenesis. Journal of the Geological
Society, London, 157, 1215–1225.
Kirkland, C.L., Daly, J.S., and Whitehouse, M.J. (2007)
Provenance and terrane evolution of the Kalak Nappe
Complex, Norwegian Caledonides: implications for Neo-
proterozoic palaeogeography and tectonics. Journal of
Geology, 115, 21–41.
Kirkland, C.L., Strachan, R.A., and Prave, A.R. (2008) Detri-
tal zircon signature of the Moine Supergroup, Scotland:
contrasts and comparisons with other Neoproterozoic
successions within the circum-North Atlantic region.
Precambrian Research, 163, 332–350.
598 Part 5: Plate Interior Basins and Basin Types
Kokonyangi, J.W., Kampunzu, A.B., Armstrong, R.,
Yoshida, M., Okudaira, T., Arima, M., and Ngulube, D.
A. (2006) The Mesoproterozoic Kibaride belt (Katanga,
SE D.R. Congo). Journal of African Earth Sciences, 46,
1–35.
Krabbendam, M., Prave, T., and Cheer, D. (2008) A fluvial
origin for the Neoproterozoic Morar Group, NW Scot-
land; implications for Torridon-Morar Group correlation
and the Grenville Orogen foreland basin. Journal of the
Geological Society, 165, 379–394.
Li, Z.X., Bogdanova, S.V., Collins, A.S., Davidson, A., De
Waele, B., Ernst, R.E., Fitzsimons, I.C.W., Fuck, R.A.,
Gladkochub, D.P., Jacobs, J., Karlstrom, K.E., Lu, S.,
Natapov, L.M., Pease, V., Pisarevsky, S.A., Thrane, K.,
and Vernikovsky, V. (2008) Assembly, configuration,
and break-up history of Rodinia: a synthesis. Precam-
brian Research, 160, 179–210.
Link, P.K., Christie-Blick, N., Devlin, W.J., Elston, D.P.,
Horodyski, R.J., Levy, M., Miller, J.M.G., Pearson, R.C.,
Prave, A., Stewart, J.H., Winston, D., Wright, L.A., and
Wrucke, C.T. (1993) Middle and late Proterozoic strati-
fied rocks of the western U.S. Cordillera, Colorado Pla-
teau, and Basin and Range province, in Reed, J.C.J.,
Bickford, M.E., Houston, R.S., Link, P.K., Rankin, D.
W., Sims, P.K., and Van Schmus, W.R., eds., Precam-
brian: coterminous U. S. Boulder, Colorado, Geological
Society of America, The Geology of North America, C-2,
463–595.
Long, D.G.F., Rainbird, R.H., Turner, E.C., and MacNaugh-
ton, R.B. (2008) Early Neoproterozoic strata (Sequence B)
of mainland northern Canada and Victoria and Banks
islands: a contribution to the Geological Atlas of the
Northern Canadian Mainland Sedimentary Basin, Geo-
logical Survey of Canada Open File 5700, 22.
Mapes, R.W., Coleman, D.S., Nogueira, A.C.R., and Housh,
T.B. (2004) How far do zircons travel? Evaluating
the significance of detrital zircon provenance using
the modern Amazon River fluvial system, Geological
Society of America, Abstracts with Programs, Volume
36, 78.
Master, S., Rainaud, C., Armstrong, R.A., Phillips, D., and
Robb, L.J. (2005) Provenance ages of the Neoproterozoic
Katanga Supergroup (Central African Copperbelt) with
implications for basin evolution. Journal of African Earth
Sciences, 42, 41–60.
McLennan, S.M., Bock, B., Compston, W., Hemming, S.R.,
and McDaniel, D.K. (2001) Detrital zircon geochronology
of Taconian and Acadian foreland sedimentary rocks in
New England. Journal of Sedimentary Research, 71,
305–317.
Medaris, L.G., Singer, B.S., Dott, R.H., Naymark, A., John-
son, C.M., and Schott, R.C. (2003) Late Paleoproterozoic
climate, tectonics, and metamorphism in the Southern
Lake Superior region and Proto-North America: evidence
from Baraboo interval quartzites. Journal of Geology, 111,
243–257.
Medaris, L.G., Van Schmus, W.R., Loofboro, J., Stonier, P.J.,
Zhang, X., Holm, D.K., Singer, B.S., and Dott, R.H. (2007)
Two Paleoproterozoic (Statherian) siliciclastic metase-
dimentary sequences in central Wisconsin:the end of the
Penokean Orogeny and cratonic stabilization of the
southern Lake Superior region. Precambrian Research,
157, 188–202.
Mezger, K., and Cosca, M.A. (1999) The thermal history of
the Eastern Ghats Belt (India) as revealed by U-Pb and
40Ar/39Ar dating of metamorphic and magmatic miner-
als: implications for the SWEAT correlation. Precam-
brian Research, 94, 251–271.
Moecher, D.P., and Samson, S.D. (2006) Differential zircon
fertility of source terranes and natural bias in the detrital
zircon record: implications for sedimentary provenance
analysis. Earth and Planetary Science Letters, 247,
252–266.
Moore, J.M., and Thompson, P.H. (1980) The Flinton Group:
a late Precambrian sedimentary sequence in the Gren-
ville Province of eastern Ontario. Canadian Journal of
Earth Sciences, 17, 1685–1707.
Moores, E.M. (1991) Southwest U.S.-East Antarctic
(SWEAT) connection: a hypothesis. Geology, 19,
425–428.
Morey, G.B., and Ojakangas, R.W. (1992) Keweenawan
rocks of eastern Minnesota and northwestern Wisconsin,
in Wold, R.J., and Hinze, W.J., eds., Geology and Tec-
tonics of the Lake Superior Basin, Geological Society of
America Memoir 156, 135–146.
Morley, C.K., Woganan, N., Sankumarn, N., Hoon, T.B.,
Alief, A., and Simmons, M. (2001) Late Oligocene-
Recent stress evolution in rift basins of northern and
central Thailand; implications for escape tectonics.
Tectonophysics, 334, 115–150.
Mueller, P.A., Foster, D.A., Mogk, D.W., Wooden, J.L.,
Kamenov, G.D., and Vogl, J.J. (2007) Detrital mineral
chronology of the Uinta Mountain Group: implications
for the Grenville flood in southwestern Laurentia.
Geology, 35, 431–434.
Murphy, J.B., Nance, R.D., and Cawood, P.A. (2009) Con-
trasting Modes of Supercontinent Formation and the
Conundrum of Pangea. Gondwana Research, 15,
408–420.
Nance, R.D., Worsley, T.R., and Moody, J.B. (1988) Der
Superkontinent-Zyklus. The supercontinent cycle.
Spektrum der Wissenschaft, 1988, 80–87.
Nesbitt, H.W., Young, G.M., McLennan, S.M., and Keays, R.
R. (1996) Effects of chemical weathering and sorting on
the petrogenesis of siliciclastic sediments, with implica-
tions for provenance studies. Journal of Geology, 104,
525–542.
Nicholson, P.G. (1993) A basin reappraisal of the Protero-
zoic Torridon Group, northwest Scotland, in Frostick, L.
E., and Steel, R.J., eds., Tectonic controls and signatures
in sedimentary successions. Special Publication, 20.
Ghent, Belgium, International Association of Sedimen-
tologists, 183–202.
Ojakangas, R.W., Morey, G.B., Southwick, D.L., Eriksson,
P.G., Catuneanu, O., Aspler, L.B., Chiarenzelli, J.R.,
and Martins-Neto, M.A. (2001) Paleoproterozoic
basin development and sedimentation in the
Lake Superior region, North America. The influence
of magmatism, tectonics, sea level change and
paleo-climate on Precambrian basin evolution;
change over time. Sedimentary Geology, 141–142,
319–341.
Pisarevsky, S.A., Wingate, M.T.D., Powell, C.M., Johnson,
S., and Evans, D.A.D. (2003) Models of Rodinia assembly
and fragmentation, Geological Society Special Publica-
tion, Volume 206, 35–55.
The Great Grenvillian Sedimentation Event 599
Potter, P.E. (1978) Petrology and chemistry of modern big
river sands. Journal of Geology, 86, 423–449.
Rainbird, R.H., Hadlari, T., Aspler, L.B., Donaldson, J.A.,
LeCheminant, A.N., and Peterson, T.D. (2003) Sequence
stratigraphy and evolution of the Paleoproterozoic intra-
continental Baker Lake and Thelon basins, western
Churchill Province, Nunavut, Canada. Precambrian
Research, 125, 21–53.
Rainbird, R.H., Hamiltom, M.A., and Young, G.M. (2001)
Detrital zircon geochronology and provenance of the
Torridonian, NW Scotland. Journal of the Geological
Society [London], 158, 15–27.
Rainbird, R.H., Heaman, L.M., and Young, G.M. (1992)
Sampling Laurentia: detrital zircon geochronology offers
evidence for an extensive Neoproterozoic river system
originating from Grenville orogen. Geology, 20, 351–354.
Rainbird, R.H., Jefferson, C.W., and Young, G.M. (1996) The
early Neoproterozoic sedimentary Succession B of north-
west Laurentia: correlations and paleogeographic signif-
icance. Geological Society of America Bulletin, 108,
454–470.
Rainbird, R.H., McNicoll, V.J., Th
eriault, R.J., Heaman, L.
M., Abbott, J.G., Long, D.G.F., and Thorkelson, D.J.
(1997) Pan-continental River System Draining Grenville
Orogen Recorded by U-Pb and Sm-Nd Geochronology of
Neoproterozoic Quartzarenites and Mudrocks, North-
western Canada. Journal of Geology, 105, 1–18.
Rainbird, R.H., Stern, R.A., Khudoley, A.K., Kropachev, A.
P., Heaman, L.M., and Sukhorukov, V.I. (1998) U-Pb
geochronology of Riphean sandstone and gabbro from
southeast Siberia and its bearing on the Laurentia-Siberia
connection. Earth and Planetary Science Letters,
164 (3–4), 409–420.
Rainbird, R.H., Stern, R.A., Rayner, N.M., and Jefferson, C.
W. (2007) Age, provenance, and regional correlation of
the Athabasca Group, Alberta and Saskatchewan, con-
strained by igneous and detrital zircon geochronology, in
Jefferson, C.W., and Delaney, G.D., eds., EXTECH IV:
Geology and Uranium EXploration TECHnology of the
Proterozoic Athabasca Basin, Saskatchewan and
Alberta. Geological Survey of Canada Bulletin, 588,
193–209.
Rainbird, R.H., and Young, G.M. (2009) Colossal Rivers,
Massive Mountains and Supercontinents. Earth, 54,
52–61.
Ramaekers, P., and Catuneanu, O. (2004) Development and
Sequences of the Athabasca Basin, Early Proterozoic,
Saskatchewan and Alberta, Canada, in Eriksson, P.G.,
Altermann, W., Nelson, D.R., Mueller, W.U., and Catu-
neanu, O., eds., Tempos and events in Precambrian time,
Amsterdam, Elsevier, chap. 8.3.
Ramaekers, P., Jefferson, C.W., Yeo, G.M., Collier, B., Long,
D.G.F., Drever, G., McHardy, S., Jiricka, D., Cutts, C.,
Wheatley, K., Catuneanu, O., Bernier, S., Kupsch, B., and
Post, R. (2007) Revised geological map and stratigraphy
of the Athabasca Group, Saskatchewan and Alberta; in
EXTECH IV: Geology and Uranium EXploration TECH-
nology of the Proterozoic Athabasca Basin, Saskatche-
wan and Alberta, in Jefferson, C.W., and Delaney, G.,
eds., Geological Survey of Canada, Bulletin 588,
155–192.
Rino, S., Kon, Y., Sato, W., Maruyama, S., Santosh, M., and
Zhao, D. (2008) The Grenvillian and Pan-African
orogens: world’s largest orogenies through geologic
time, and their implications on the origin of superplume.
Gondwana Research, 14, 51–72.
Rivers, T. (1997) Lithotectonic elements of the Grenville
Province: review and tectonic implications. Precam-
brian Research, 86, 117–154.
Ross, G.M. (1983) Proterozoic aeolian quartz arenites from
the Hornby Bay Group, Northwest Territories, Canada:
implications for Precambrian aeolian processes. Precam-
brian Research, 20, 149–160.
Santos, J.O.S., Hartmann, L.A., McNaughton, N.J., Easton,
R.M., Rea, R.G., and Potter, P.E. (2002) Sensitive high
resolution ion microprobe (SHRIMP) detrital zircon geo-
chronology provides new evidence for a hidden Neopro-
terozoic foreland basin to the Grenville Orogen in the
eastern Midwest U.S. Canadian Journal of Earth
Sciences, 39, 1505–1515.
Santos, J.O.S., Potter, P.E., Reis, N.J., Hartmann, L.A.,
Fletcher, I.R., and McNaughton, N.J. (2003) Age, source,
and regional stratigraphy of the Roraima Supergroup and
Roraima-like outliers in northern South America based
on U-Pb geochronology. Geological Society of America
Bulletin, 115, 331–348.
Santos, J.O.S., Rizzotto, G.J., Potter, P.E., McNaughton, N.J.,
Matos, R.S., Hartmann, L.A., Chemale Jr, F., and Quad-
ros, M.E.S. (2008) Age and autochthonous evolution of
the Suns~
as Orogen in West Amazon Craton based on
mapping and U-Pb geochronology. Precambrian
Research, 165, 120–152.
Sch
arer, U., and Gower, C.F. (1988) Crustal evolution in
eastern Labrador; constraints from precise U-Pb ages.
Precambrian Research, 38, 405–421.
Schumm, S.A. (1968) Speculations concerning paleohydro-
logic controls on terrestrial sedimentation. Geological
Society of America Bulletin, 79, 1573–1588.
Seng
or, A.M.C., Burke, K., and Dewey, J.F. (1978) Rifts at
high angles to orogenic belts; tests for their origin and the
Upper Rhine Graben as an example. American Journal of
Science, 278, 24–40.
Shrake, D.L., Carlton, R.W., Wickstrom, L.H., Potter, P.E.,
Benjamin, R.H., Wolfe, P.J., and Sitler, G.W. (1991) Pre-
Mount Simon basin under the Cincinatti Arch. Geology,
19, 139–142.
Sloss, L.L. (1988) Tectonic evolution of the craton in Phan-
erozoic time, in Sloss, L.L., ed., Sedimentary cover:
North American craton, vol. D-2. Boulder, CO, Geolog-
ical Society of America.
Soegaard, K., and Eriksson, K.A. (1989) Origin of thick, first-
cycle quartz arenite successions: evidence from the
1.7 Ga Ortega Group, northern New Mexico. Precambrian
Research, 43, 129–141.
Squire, R.J., Campbell, I.H., Allen, C.M., and Wilson, C.J.L.
(2006) Did the Transgondwanan Supermountain trigger
the explosive radiation of animals on Earth? Earth and
Planetary Science Letters, 250, 116–133.
Stewart, A.D. (2002) The Later Proterozoic Torridonian
rocks of Scotland; their sedimentology, geochemistry
and origin, Geological Society Memoir No. 24, 136 p.
Stewart, J.H., Gehrels, G.E., Barth, A.P., Link, P.K., Christie,
B.N., and Wrucke, C.T. (2001) Detrital zircon provenance
of Mesoproterozoic to Cambrian arenites in the Western
United States and northwestern Mexico. Geological
Society of America Bulletin, 113, 1343–1356.
600 Part 5: Plate Interior Basins and Basin Types
Thomas, W.A., Becker, T.P., Samson, S.D., and Hamilton,
M.A. (2004) Detrital zircon evidence of a recycled oro-
genic foreland provenance for Alleghanian clasticwedge
sandstones. Journal of Geology, 112, 23–37.
Timmons, J.M., Karlstrom, K.E., Heizler, M.T., Bowring, S.
A., Gehrels, G.E., and Crossey, L.J. (2005) Tectonic infer-
ences from the ca. 1255–1100 Ma Unkar Group and
Nankoweap Formation, Grand Canyon: intracratonic
deformation and basin formation during protracted
Grenville orogenesis. Geological Society of America Bul-
letin, 117, 1573–1595.
Tucker, R.D., and Gower, C.F. (1994) A U-Pb geochronolog-
ical framework for the Pinware terrane, Grenville Prov-
ince, Labrador, Canada. Journal of Geology, 102, 67–78.
Veevers, J.J. (2004) Gondwanaland from 650–500 Ma assembly
through 320 Ma merger in Pangea to 185–100 Ma breakup:
supercontinental tectonics via stratigraphy and radiomet-
ric dating. Earth-Science Reviews, 68, 1–132.
Watt, G.R., and Thrane, K. (2001) Early Neoproterozoic
events in East Greenland. Precambrian Research, 110,
165–184.
Wheeler, J.O., Hoffman, P.F., Card, K.D., Davidson, A.,
Sanford, B.V., Okulitch, A.V., and Roest, W.R. (1997)
Geological Map of Canada, Geological Survey of Canada,
Map D1860A.
Whitmeyer, S.J., and Karlstrom, K.E. (2007) Tectonic model
for the Proterozoic growth of North America. Geosphere,
3, 220–259.
Williams, G.E. (2001) Neoproterozoic (Torridonian) alluvial
fan succession, Northwest Scotland, and its tectonic
setting and provenance. Geological Magazine, 138,
471–494.
Wirth, K.R., Vervoort, J., Craddock, J., Davidson, C., Finley-
Blasi, L., Kerber, L., Lundquist, B., Vorhies, S., Walker, E.
(2006) Source rock ages and patterns of sedimentation in
the Lake Superior region: results of preliminary U-Pb
detrital zircon studies. Abstracts with Programs, 38.
Boulder, CO, Geological Society of America.
Young, G.M. (1978) Proterozoic (<1.7 b.y.) stratigraphy,
paleocurrents and orogeny in North America. Egyptian
Journal of Geology, 22, 45–64.
Young, G.M. (1979) Correlation of middle and upper Pro-
terozoic strata of the northern rim of the North American
craton. Transactions of the Royal Society of Edinburgh,
70, 323–336.
Young, G.M., Long, D.G.F., Fedo, C.M., and Nesbitt, H.W.
(2001) Paleoproterozoic Huronian basin: product of
a Wilson cycle punctuated by glaciations and a
meteorite impact. Sedimentary Geology, 141–142,
233–254.
The Great Grenvillian Sedimentation Event 601
... A discordance filter of 20% was used for the U-Pb ages following ref. 75 . The detrital zircon U-Pb ages are plotted as histograms and KDEs using DensityPlotter 76 . ...
Article
Full-text available
The interplay between continental motions during the assembly of Pangaea and late Palaeozoic climate change, including severe glaciation and global aridification, remains enigmatic. Here we identify the provenance of Permian–Early Triassic sediments that recorded climate change of North China and estimate palaeoelevation to constrain tectonic-climate interaction during the assembly of northeast Pangaea. Detrital zircon U-Pb-Hf analysis indicates the sediments were locally sourced from the ancient basement and associated with a late Palaeozoic (410–260 million years ago) continental arc, devoid of input from juvenile arcs of the Altaids. These sediments were interpreted as deposited in a retroarc foreland basin ascribed to subduction of the Palaeo-Asian Ocean. Crustal thickness estimated from whole-rock La/Yb yields an average value of 58 ± 11 km, which corresponds to a palaeoelevation of 3.8 ± 0.7 km. The results reveal the existence in North China of an orogenic plateau comparable to the Altiplano of the Andes that blocked moisture transport from the ocean and served as an important orographic barrier to trigger Permian aridification.
... Synorogenic sediments (depositional ages of ca. 1250-1000 Ma) include distal fluvial deposits of a Grenville clastic wedge exposed in northwest Canada and the Grand Canyon (Timmons et al., 2005;Rainbird et al., 2012Rainbird et al., , 2017Dehler et al., 2017;Mulder et al., 2017), proximal sediments deposited before ca. 1200 Ma in restricted basins within the Grenville orogen of Ontario, Quebec, and the Adirondacks (e.g., the "Grenville Supergroup"; Rivers et al., 1989Rivers et al., , 2012Peck et al., 2019), and proximal sediments in Appalachian massifs that were all overprinted by Shawinigan or Ottawan metamorphism (Tollo et al., 2017;Johnson et al., 2020;Moecher et al., 2020). ...
Chapter
Full-text available
The North American continent has a rich record of the tectonic environments and processes that occur throughout much of Earth history. This Memoir focuses on seven “turning points” that had specific and lasting impacts on the evolution of Laurentia: (1) The Neoarchean, characterized by cratonization; (2) the Paleoproterozoic and the initial assembly of Laurentia; (3) the Mesoproterozoic southern margin of Laurentia; (4) the Midcontinent rift and the Grenville orogeny; (5) the Neoproterozoic breakup of Rodinia; (6) the mid-Paleozoic phases of the Appalachian-Caledonian orogen; and (7) the Jurassic–Paleogene assembly of the North American Cordillera. The chapters in this Memoir provide syntheses of current understanding of the geologic evolution of Laurentia and North America, as well as new hypotheses for testing.
Article
Full-text available
Detrital zircon geochronology plays a crucial role in provenance analysis, serving as one of the fundamental strategies. The age spectrum of detrital zircons collected from the sedimentary unit of interest is often compared or correlated with that of potential source terranes. However, biases in the age data can arise due to factors related to detrital sampling, analysis techniques, and nonlinear geological mechanisms. The current study reviewed two sets of detrital zircon datasets established in 2011 and 2021 to discuss the origins of the Tibetan Plateau. These datasets collected from different media effectively demonstrate a progressive understanding of provenance affinity among the main terranes on the Tibetan Plateau. This highlights issues regarding weak and unclear temporal connections identified through analyzing the age spectrum for provenance analysis. Within this context, a local singularity analysis approach is currently employed to address issues associated with unclear and weak provenance information by characterizing local variations in nonlinear behaviors and enhancing detection sensitivity towards subtle anomalies. This new graphical approach effectively quantifies temporal variations in detrital zircon age populations and enhances identification of weak provenance information that may not be readily apparent on conventional age spectra.
Chapter
The eastern Great Smoky Mountains basement complex consists of the following components: (1) ca. 1350–1325 Ma orthogneiss and mafic xenoliths that represent some of the oldest crust in Appalachian Grenville massifs (similar to “pre-Grenville” basement components in the Adirondack, Green Mountain, Hudson Highland, and Shenandoah massifs); (2) ca. 1150 Ma augen orthogneisses and granitic orthogneisses correlating with the Shawinigan phase of Grenville magmatism; and (3) paragneisses (cover rocks) that have either pre- or syn-Grenville (i.e., Mesoproterozoic) versus post-Grenville (Neoproterozoic) depositional ages, and that experienced Taconian metamorphism and migmatization. Mesoproterozoic paragneisses contain major zircon age modes that require a component of Proterozoic crust in the source region. The Neoproterozoic paragneisses exhibit the archetypical “Grenville doublet” in detrital zircon age distributions that matches the age distribution of Ottawan and Shawinigan magmatic/metamorphic events in eastern Laurentia. Most zircon U-Pb age systematics exhibit variable lead loss interpreted to result from high-grade Taconian (ca. 450 Ma) regional metamorphism and migmatization. Neodymium mantle model ages (TDM) for ortho- and paragneisses range from 1.8 to 1.6 Ga, indicating that all rocks were derived from recycling of Proterozoic crust (i.e., they are not juvenile), which is consistent with Proterozoic detrital zircon ages in pre- to syn-Grenville paragneisses. Lead isotope compositions confirm the presence of an exotic (Amazonian) crustal component in the source region for the protoliths of the pre-Grenville orthogneisses and xenoliths, and that this exotic component was incorporated to varying degrees in the evolution of the basement complex. The oldest age component may represent an Amazonian pre-Grenville analog to the ca. 1.35 Ga native Laurentian crust present in Adirondack and northern Appalachian basement massifs.
Article
Fluvial point bar deposits are described from the ca. 1 Ga Diabaig Formation, representing some of the most compelling architectural evidence of planform sinuosity recognized in pre-vegetation sedimentary strata to date. The stratigraphic architecture of the deposits indicates different planform transformations in response to meander-bend expansion and downstream migration. Point bar architectural motifs are locally abundant in the unit and this is attributed to elevated amounts of mud-calibre cohesive sediment. Mud afforded landscape stability that, together with limited stream power, was sufficient to corral sediment into single-thread, sinuous channels. Significant volumes of fine-grained sediment were encouraged to accumulate by the local basin setting of several topographically confined endorheic depressions. These conditions ceased once the hydrologically closed valleys were filled, with the overlying sandstone-dominated Applecross Formation subsequently deposited by low-sinuosity, braided rivers that bypassed fine-grained sediment oceanward. Other endorheic basins conducive to the formation of depositional meanders should be expected to have been relatively widespread on pre-vegetation Earth, though many individually limited in size, confined by their inherited topographies. Depositional architectures representative of sinuous channel planforms remain scarce within more extensive, externally drained basins.
Article
Full-text available
Abstract: Multidisciplinary geochronological, isotopic, chemical, and facial studies in the Malyi Karatau Range (MK) of South Kazakhstan elucidate the Precambrian stratigraphic framework and evolution of the Ishim Middle Tianshan microcontinent (IMT) in the western Central Asian Orogenic Belt. Detrital zircon and apatite U-Pb ages for siliciclastic rocks and chemostratigraphic Sr-C isotopic data for carbonates indicate that they deposited during 800–730 Ma. The sediments are dominated by deeper marine facies in the southwest and shallow marine facies in the northeast. According to paleocurrent indicators, the main provenance located to the west of the basin. Based on the detrital zircon age spectra, the source terrane represented a Paleoproterozoic to Archean crustal block, reworked by an early Neoproterozoic c. 850–720 Ma continental arc. Erosion of Archean and Paleoproterozoic crust is also evidenced by negative eNd values for the sandstones. The petrographic and chemical compositions of the sandstones are consistent with a continental arc source; however anomalously high concentrations of chromium in some layers point to the presence of ultramafic rocks in the source terrane. The ages of metamorphic zircons indicate a high-grade metamorphic event at provenance at c. 2.0 Ga and the ages of detrital apatites suggest a reset of U-Pb isotope system in apatite at c. 1.8–1.9 Ga. In the early Neoproterozoic, the MK located between the continental arc in the west and the oceanic basin in the east and represented a fore-arc basin. The similarity of the Precambrian magmatic and metamorphic histories and sedimentary facies indicates that the IMT, the Tarim and Yangtze cratons constituted a single Precambrian Ulutau-Tarim-Yangtze Continent (UTY). Judging from the ages of continental arcs that evolved on the northern and southern sides of the UTY, it represented an independent continent at c. 950–840 Ma and was incorporated in Rodinia at c. 830–820 Ma.
Article
Full-text available
Lithic and quartz arenites of the Central Appalachian Basin deposited by late Paleozoic Laurentian fluvial systems are widely interpreted to be sourced by recycling of late Precambrian and early Paleozoic clastic sequences in the Appalachian Orogen. U-Pb and (U-Th)/He age distributions for detrital-zircon and Th-Pb age distributions for detrital-monazite, detrital-zircon and monazite textures (including detrital diagenetic monazite, which prove recycling), sandstone petrology, heavy-mineral abundances, and other detrital proxies are all accounted for by the following: 1) lithic arenite is directly sourced from late Neoproterozoic metasediments in the proximal Appalachian fold and thrust belt via transverse drainages, 2) the late Neoproterozoic sediments were recycled from early Neoproterozoic, post-Grenvillian clastic sequences, 3) Cambrian quartz arenites along the Laurentian margin were recycled from Neoproterozoic sequences with local cratonic input, 4) although dominated by sediment of ultimate Grenvillian age, quartz arenites require ∼ 40% of Paleoproterozoic and Archean input, interpreted to be recycled from late Neoproterozoic to Devonian clastic sequences of the northern Appalachians and/or southern (Scottish–Irish) Caledonides in the distal reaches of a longitudinal drainage system. Ordovician to Mississippian clastic sequences and their metamorphosed equivalents in the Appalachian crystalline core were also likely sediment sources. Quartz arenite does not result from mixing of lithic arenite with other sources because of differences in textural and compositional maturity, and in heavy-mineral characteristics. Input from the Laurentian craton, commonly cited as the source for Paleoproterozoic and Archean detrital zircon in the eastern Laurentian clastic systems, is untenable here because of: 1) the presence of Paleozoic monazite derived from Appalachian Neoproterozoic and early Paleozoic metasediments, 2) abundant detrital chromite, and 3) abundant Paleozoic detrital muscovite. Multiple recycling explains all observed sedimentologic and mineralogic characteristics of the two lithic types. Incorporation of published detrital-zircon data for Paleozoic to modern clastic sequences in eastern Laurentia reveals that Grenville-age zircons experienced at least five and potentially ten recycling events since entering the clastic system in the Neoproterozoic. Recycling also explains the abundance of quartz pebbles in conglomerates of the quartz-arenite lithofacies, and the range of detrital-muscovite 40Ar/39Ar ages in quartz arenites of the Appalachian Basin.
Article
Full-text available
Geochemical and geochronological data from the Pinguicula Group and unit PR1 of the lower Fifteenmile Group (Yukon, Canada) provide information on sediment provenance and timing of break-up of supercontinent Columbia and seaway development on Laurentia’s northwestern margin. The older unit PR1, in the Coal Creek inlier, has a near-unimodal detrital zircon population with an age of 1499 ± 3 Ma. The Pinguicula Group detrital zircon data, in the Wernecke and Hart River inliers, display a polymodal detrital zircon population with a maximum age of <1322 ± 23 Ma. Using detrital zircon signatures, Sm–Nd isotopic data, and C-isotopic signatures, lithostratigraphic correlations between the Pinguicula Group in the Wernecke and Hart River inliers are confirmed, whereas the Pinguicula Group and unit PR1 are no longer considered correlative. The zircon population in unit PR1 requires a proximal source, but sources of this age are not known in western Laurentia. Based on detrital zircon and Sm–Nd data, sediment in unit PR1 was derived from the Mt. Isa inlier in northeastern Australia sometime after 1460 Ma. Unit PR1 correlates with older Mesoproterozoic successions, including the Belt-Purcell, that were deposited during break-up of supercontinent Columbia, and contain sediment from Australia and the Mawson continent. Mesoproterozoic successions deposited after 1.45 Ga, including the Missoula Group, lack North American Magmatic Gap (NAMG)-aged zircon and instead record sediment provenance from southern Laurentia, as north Australia and the Mawson continent rifted from Laurentia’s western margin. The Pinguicula Group has few NAMG-aged grains that were probably recycled from older Mesoproterozoic basins.
Article
Full-text available
The Mesoproterozoic Husky Creek Formation is an interlayered redbed and basalt package that overlies a ca. 2.5 km thick, regionally extensive basaltic plateau linked to the 1.27 Ga Mackenzie Large Igneous Province. This succession provides an opportunity to study the interaction between contemporaneous fluvial sedimentation and flood basalt volcanism. The Husky Creek Formation is approximately 1900 m thick, consisting predominantly of fluvial channel and subordinate floodplain and eolian strata dominated by lithic detritus intercalated with basalt flows. U–Pb dating of detrital zircon grains collected from four stratigraphic levels in the Husky Creek Formation reveals two main age groupings: (1) a 1270 Ma peak attributed to the Mackenzie Large Igneous Province; and (2) Pre-1.6 Ga detrital zircon ages, which reflect sources from a broad region of northwestern Laurentia and are interpreted to be recycled from underlying sedimentary rocks of the Hornby Bay Basin. An upsection decrease in the proportion of ca. 1270 Ma zircon grains, coupled with increasing sandstone compositional maturity, reflects the expansion of the drainage basin during burial of a volcanic pile, with recycling becoming more dominant as the Husky Creek Formation accumulated. The Husky Creek Formation was deposited in the waning stages of the Mackenzie Igneous Event by west-flowing rivers into a geographically restricted basin above an extensive mafic volcanic plateau. This paper provides insight into fluvial basin development during the aftermath of one of Earth's largest igneous events.
Article
Existing models of Torridon Group (Applecross Formation) genesis suggest deposition by at least two large alluvial fans within a 75-100-km-wide rift graben cutting Archaean crust. Results presented here reveal the presence of large-scale bar structures (up to 9 m thick) within the alluvium, indicating the occurrence of major rivers within the fluvial system. An alluvial braidplain origin for the Applecross Formation is preferred to the existing alluvial fan model, based on the evidence for rivers over 500 km long, together with the consistently E-SE regional palaeoflow. These scale considerations, combined with the lack of evidence for both syndepositional tectonism within the basin and for active basin-bounding fault margins, collectively favour deposition within a significantly larger basin than that previously proposed. A larger scale, later stage, extensional basin formed by post-rifting thermal relaxation processes, accompanied by a component of passive subsidence-driven normal faulting along the Outer Isles Fault, is proposed for deposition of the Torridon Group. -from Author