You are on page 1of 26

Precambrian Research 139 (2005) 121–146

U–Pb age data from the Sunsas region of Eastern Bolivia, evidence
for the allochthonous origin of the Paragua Block
S.D. Boger a,∗ , M. Raetz a , D. Giles a,∗ , E. Etchart b , C.M. Fanning c
a Australian Crustal Research Centre, School of Geosciences, Monash University, Clayton, Vic. 3400, Australia
b BHP Billiton Ltd., 6 Holland Street, Johannesburg 2001, Republic of South Africa
c Research School of Earth Sciences, The Australian National University, Canberra, ACT 0200, Australia

Received 27 October 2004; received in revised form 2 May 2005; accepted 9 May 2005

Abstract

Ion microprobe U–Pb dating of detrital zircons from the Chiquitania and San Ignacio groups of Eastern Bolivia show that these
rocks were derived mostly from a 1765 Ma source and were deposited after 1690 Ma. The emplacement of the Lomas Maneches
granitoid suit, dated at 1690–1660 Ma, accompanied the deposition of these metasediments. These rocks, which collectively define
the basement of Eastern Bolivia, were subsequently deformed and metamorphosed during two discrete periods of orogenesis.
The oldest of these events, the San Ignacio Orogeny, deformed the terrane through three successive phases of deformation and
was accompanied by voluminous syn-tectonic granite intrusion and migmatisation of the 1690 Ma sediments and intrusions.
U–Pb age data from igneous and metamorphic zircon constrain the San Ignacio Orogeny to between 1340 and 1320 Ma, while the
orientation of structures suggests deformation was the result of W–NW directed shortening. A 300 million years hiatus, during
which the overlying Sunsas Group was deposited, separated the San Ignacio Orogeny from the Sunsas Orogeny. Shortening
during Sunsas orogenesis resulted in an intense NW trending structural overprint in the southern and western extremes of the
study area. The intensity of deformation decreased toward the northeast such that in the San Ignacio region, Sunsas tectonism
resulted in distinct dome and basin fold interference patterns. Late syn- to post-tectonic granites were emplaced in the region
of highest Sunsas strain. One of these intrusions, the Taperas granite, yielded an age of ca. 1075 Ma. When taken collectively,
these new data highlight a number of important points: (1) The deposition of the Chiquitania and San Ignacio groups, and the
synchronous emplacement of the Lomas Maneches granitoid suite (1690–1660 Ma) have no temporal equivalents in the either
the Rondônia and Mato Grosso regions of Brazil to the north or east, or the Proterozoic pre-Andean inliers to the west; (2) the
trend of the structures attributed to the San Ignacio Orogeny are oriented at a high-angle to the NW trending pre-Sunsas margin
of Amazonia and; (3) the Sunsas Orogeny represents the only geological event with a common age and direction of crustal
shortening recorded in both the Amazon Craton and the rocks of Eastern Bolivia. These data are interpreted to suggest that the
rocks of Eastern Bolivia represent a geologically distinct microcontinent (the Paragua Craton) that was accreted to the southern
margin of Amazon Craton during the Sunsas Orogeny.
© 2005 Elsevier B.V. All rights reserved.

Keywords: Bolivia; South America; Amazon Craton; Sunsas Orogeny; San Ignacio Orogeny; Grenville; SHRIMP geochronology; Rodinia

∗ Corresponding authors.
E-mail addresses: sdboger@unimelb.edu.au, stevenboger@hotmail.com (S.D. Boger), david.giles@sci.monash.edu.au (D. Giles).

0301-9268/$ – see front matter © 2005 Elsevier B.V. All rights reserved.
doi:10.1016/j.precamres.2005.05.010
122 S.D. Boger et al. / Precambrian Research 139 (2005) 121–146

1. Introduction Rodinia palaeogeography. Nevertheless, this region has


seen no significant geological work since the “Proyecto
The thick Neoproterozoic rift related sediments Precámbrico”, an Anglo-Bolivian technical coopera-
found along the eastern and western margins of Lauren- tion program that first mapped the Proterozoic rocks of
tia (Bond et al., 1984) have been interpreted as evidence Bolivia in the late 1970s and early 1980s (Litherland
for the existence of the Mesoproterozoic Rodinia super- and Bloomfield, 1981; Berrrangé and Litherland, 1982;
continent (Dalziel, 1991; Hoffman, 1991; Moores, Litherland et al., 1996, 1989). These studies published
1991; Brookfield, 1993; Borg and DePaolo, 1994). a number of Rb–Sr and K–Ar ages that broadly overlap
Although the exact configuration of Rodinia remains with the youngest recognised events in Laurentia. How-
a point of considerable conjecture, its gross geome- ever, more recent studies from neighbouring regions
try centred on the north American cratonhas remained of Brazil have shown that these isotopic systems have
largely unchanged. East Antarctica and Australia, con- consistently post-dated the U–Pb ages from the same
tiguous prior to the break-up of Gondwana, remains the rocks by as much as 200 million years (Bettencourt et
preferred choice as the conjugate to Laurentia’s west- al., 1999). Thus, the significance of the existing age
ern margin (Hoffman, 1991; Moores, 1991; Brookfield, data from Eastern Bolivia is equivocal. The present
1993; Burrett and Berry, 2000; Berry et al., 2001; study provides new structural data, coupled with U–Pb
Karlstrom et al., 2001; Wingate et al., 2002). For Lau- SHRIMP ages obtained from some of the key intru-
rentia’s southern and eastern margins, reconstructions sions from the region. With these data we are able to
have emphasised a long-lived connection between Lau- better define the formation age of the basement rocks
rentia and Amazonia (Bond et al., 1984; Dalziel, 1991, from this region, as well as the timing and kinematics
1992; Hoffman, 1991). of both the San Ignacio and Sunsas orogenies. These
Although the Laurentia-Amazonia connection results place new constraints on the geologic evolution
within Rodinia is generally accepted, conflicting of the south-western margin of the Amazon Craton, a
interpretations have correlated the Amazon Craton key region for testing the commonly inferred Protero-
with various parts of the +3000 km long southern and zoic connection between Laurentia and Amazonia.
eastern margin of Laurentia. For example, Tohver et
al. (2002) paired the Amazon Craton with the south-
western Llano segment of Laurentia based on the low 2. Geologic setting
latitude ca. 1200 Ma palaeomagnetic pole obtained for
Amazonia. In contrast, broad geological similarities The Amazon Craton is divided into four broad
lead Hoffman (1991) and Sadowski and Bettencourt tectonic provinces that form NW–SE trending
(1996) to correlate Amazonia with the central Grenville belts that young outwards from an Archaean core
Province of Ontario and New York, while Loewy et (Teixeira et al., 1989; Tassinari and Macambira,
al. (2003) proposed an intermediate fit based on the 1999; Santos et al., 2000; Tassinari et al., 2000).
Pb isotopic affinities between Amazonian crust and From northeast to southwest these provinces are
that which defines the basement to the southern and defined as the Ventuari-Tapajós (1950–1800 Ma), Rio
central Appalachians (see also Tohver et al., 2004a). Negro-Juruena (1800–1550 Ma), Rondônia-San Igna-
Most of the uncertainty regarding the Laurentia- cio (1450–1300 Ma) and Sunsas (1250–1000 Ma)
Amazonia fit is a function of an imbalance in reliable provinces (Fig. 1). Although many aspects of the defor-
data from the two margins. While extensive geologic mation and intrusion histories within these provinces
data has been collected from the eastern margin of remains unresolved, it is generally believed that each
Laurentia, very little is actually known about the con- belt represents a successive continental margin arc built
jugate margin of South America. Critical to address- upon the basement of the preceding terrane (Geraldes
ing this imbalance is the Sunsas region of Eastern et al., 2001; Tassinari and Macambira, 1999). Of
Bolivia. Located toward the southwest margin of Ama- these provinces, part of the Rondônia-San Ignacio
zonia (Fig. 1), the Sunsas region defines the youngest Province and most of the Sunsas Province are inferred
tectonic province in the Amazon Craton and, being to define the basement rocks of Eastern Bolivia. How-
broadly of Grenville age, is critical to understanding ever, the direct continuation of these provinces from
S.D. Boger et al. / Precambrian Research 139 (2005) 121–146 123

Fig. 1. Geology of South America. (a) Distribution of gross tectonic units. Area of interest and the borders of Bolivia are highlighted. AC,
Amazon Craton and AAB, Arequipa-Antofalla Basement. (b) Enlargement of the Amazon Craton illustrating the location and approximate
boundaries of the constituent tectonic provinces (after Teixeira et al., 1989 and Tassinari and Macambira, 1999). As for part (a), the area of
interest is highlighted.

the Rondônia and Mato Grosso regions of Brazil into were reworked and intruded by granites during the
Eastern Bolivia has been challenged by a number of Mesoproterozoic and Ordovician (Shackleton et al.,
recent studies. For example, Tohver et al. (2004b) have 1979; Wasteneys et al., 1995; Wörner et al., 2000;
proposed that the rocks of Eastern Bolivia, defined Loewy et al., 2004). The distinct Proterozoic history
as the Paragua Craton (Litherland et al., 1989), rep- of the Arequipa-Antofalla Basement has led a num-
resent an accreted terrain of sub-continental size. ber of studies to infer an exotic origin for these rocks.
Tohver et al. (2004b) have argued that the deep-water Similar to the Paragua Craton, the Arequipa-Antofalla
metasediments of the Nova Brasilândia belt (Fig. 2), Basement is considered to have accreted to the mar-
together with an underlying east–west trending mag- gin of the Amazon Craton during the Sunsas Orogeny
netic anomaly, defined the southern margin of Amazo- (Loewy et al., 2004).
nia prior to the Sunsas aged collision and suturing of Although the outward accretionary growth of the
the Paragua Craton. Amazon Craton is generally accepted, the rocks that
Croping out in the Andes west of the Ama- define the Paragua Craton remain poorly understood.
zon Craton are the isolated east–west trending base- These rocks crop out almost entirely within the eastern
ment inliers that collectively define the Arequipa- third of Bolivia (Fig. 1), a flat to undulating region that
Antofalla Basement Province (AAB—Fig. 1). The is mostly covered by a Late Tertiary lateritic peneplain
Arequipa-Antofalla Basement consists of Palaeopro- where rock exposures are limited to isolated elevated
terozoic granulite facies rocks (1900–1800 Ma) that outcrops or along the incised margin of the peneplain in
124 S.D. Boger et al. / Precambrian Research 139 (2005) 121–146

Fig. 2. Geologic Map of the south-western portion of the Amazonian Craton. Thick dashed lines mark the border between Bolivia and Brazil
and the province boundaries within Brazil. Solid line outlines the region of geophysical coverage. Rio Negro Front (fine dashed line) marks the
inferred boundary between the Paragua Craton and rocks reworked by the Sunsas Orogeny (Litherland et al., 1989). C: Concepcı́on; SI, San
Ignacio; SJ, San Javier; SM, San Miguel; SR, San Ramon and SRa, San Rafael—major towns in Eastern Bolivia.

modern creek and river systems. The British Geologic the San Ignacio Group. These rocks pre-date the San
Survey and the Servicio Geológio de Bolivia under- Ignacio Orogeny (>1400 Ma) and are considered to
took the original geological work in Eastern Bolivia be Palaeo- or Mesoproterozioc in age (Berrrangé and
(e.g. Berrrangé and Litherland, 1982; Litherland et Litherland, 1982). Two cover-sequences, the Sunsas
al., 1996). These studies subdivided the exposed and the Tucavaca groups (Fig. 2), post-date the San
rocks into a metamorphic basement, consisting of Ignacio Orogeny. The Sunsas Group unconformably
the Lomas Maneches and Chiquitania complexes and overlies the Lomas Maneches Complex, Chiquitania
S.D. Boger et al. / Precambrian Research 139 (2005) 121–146 125

Complex and the San Ignacio Group and was deformed and has been used to argue that these rocks are probably
together with the basement during the subsequent Sun- of Palaeoproterozoic age (Litherland et al., 1989).
sas Orogeny (∼1000 Ma). The Tucavaca Group was
deposited after the Sunsas Orogeny and forms a mostly 2.2. San Ignacio Orogeny
undeformed cover sequence of Late Neoproterozoic
or Cambrian age. The majority of deformation and metamorphism
recorded in the basement rocks occurred as a result
2.1. Basement rocks of Eastern Bolivia of the San Ignacio Orogeny. Rb–Sr and K–Ar ages
obtained from both the basement rocks and from the
The metamorphic basement consists of an inter- syn-tectonic San Ignacio granite suite constrain this
calated package of layer-parallel granitic intrusions event to between 1400 and 1250 Ma (Litherland et al.,
and metasediments that vary in grade from granulite 1989). This event is interpreted to have affected all of
to lower greenschist. The Lomas Maneches Complex Eastern Bolivia and to have resulted in up to three over-
defines the structurally deepest and consequently printing fold generations. Metamorphism accompanied
highest-grade rocks. They are exposed in the cores the earliest stages of deformation and established the
of regional antiforms (Berrrangé and Litherland, observed metamorphic sequence. Subsequent defor-
1982; Litherland and Bloomfield, 1981). The Lomas mation resulted in the formation of upright NE-trending
Maneches Complex consists predominantly of granitic (F2 ) followed by upright NW-trending (F3 ) folds. The
sills, often up to 1000 m thick, that are interleaved latter define the gross structural grain (NW–SE) and
with lesser metasedimentary rocks of calc-silicate or control the differential exposure of the basement rocks
psammitic composition (Berrrangé and Litherland, (Litherland et al., 1989).
1982). The Chiquitania Complex, lying structurally
above the Lomas Maneches Complex, is comprised 2.3. Sunsas Orogeny
of a monotonous sequence of variably migmatitic
semipelitic gneiss interleaved locally with granite After the San Ignacio Orogeny, the basement rocks
gneisses similar to those seen in the Lomas Maneches were uplifted and eroded before being overlain along
Complex (Berrrangé and Litherland, 1982; Litherland a marked unconformity by the sediments of the Sun-
and Bloomfield, 1981). Metamorphic grade falls sas Group. These units were deposited between 1300
from muscovite-absent upper-amphibolite facies and 1000 Ma and crop out sporadically over the region.
conditions near the base of the Chiquitania Complex, The Sunsas Group locally attained a thickness of up to
to muscovite-present medium-grade conditions near 6 km and is generally comprised of a basal conglomer-
the contact with the overlying San Ignacio Group. ate overlain successively by two psammitic sequences
Litherland and Bloomfield (1981) suggested that the and an intervening pelitic unit (Berrrangé and Lither-
San Ignacio Group and some of the gneisses of the land, 1982). Cross-beds and ripple marks are ubiqui-
Chiquitania Complex might have constituted a cover tous in the psammitic units, whilst clay gall breccias
sequence to the Lomas Maneches Complex. However, indicative of subaerial desiccation were observed in
the correlation of lithostratigraphic units across the the pelites (Berrrangé and Litherland, 1982). Berrrangé
contact led Litherland et al. (1989) to revise this and Litherland (1982) inferred that the Sunsas Group
interpretation and suggest that the two sequences were was deposited in an alluvial to deltaic environment that
continuous. The low- to medium-grade rocks ascribed Saes et al. (1992) subsequently inferred was the result
to the San Ignacio Group are estimated to have attained of a developing intra-continental rift. In the south–east,
a total thickness of over 10 km and consist of quartzites, the Sunsas Group was intruded by the Rincón del
feldspathic psammites, and pelitic schists. The latter Tigre Complex, a 4.5 km thick layered igneous com-
commonly contain most of the typical Barrovian index plex (Prendergast, 2000), that was emplaced into the
minerals (Litherland et al., 1989). The depositional and upper psammitic units (Vibosi Group) of the Sunsas
emplacement age of these rocks is not well constrained. Group. The Rincón del Tigre Complex has yielded an
A Rb–Sr model age of 1960 Ma was obtained from Rb–Sr age of ca. 990 Ma, an age that is identical to that
samples collected from the Lomas Maneches Complex obtained for the late syn- to post-tectonic Sunsas gran-
126 S.D. Boger et al. / Precambrian Research 139 (2005) 121–146

ites (Litherland et al., 1996; Litherland et al., 1989). (Fig. 2). The region is interpreted to have cratonised
The rocks of the Aguapeı́ Group, the successions at Ser- after the San Ignacio Orogeny based on the preservation
rania Huanchaca and the rocks of the Nova Brasilândia of 1400–1200 Ma K–Ar ages and the flat lying nature of
Group are all thought to temporally correlate with the the (Sunsas equivalent?) Huanchaca Group (Berrrangé
Sunsas Group (Litherland et al., 1989; Santos et al., and Litherland, 1982; Litherland et al., 1989).
2000). Although rocks from both the Aguapeı́ and Nova The geologic evolution post-dating the Sunsas
Brasilândia belts yield similar aged detrital zircons, Orogeny is limited to the distal effects of the Braziliano
the continental sediments of the Aguapeı́ and Sunsas Orogeny along the southern margin of the Tucavaca
groups (Saes et al., 1992) are lithologically distinct Group, the emplacement of the Huanchaca and Mari-
from the deep-water sediments of the Nova Brasilândia mai dolerite dyke suites, and the intrusion of the
Group (Tohver et al., 2004b). Velasco Alkaline Complex (Berrrangé and Litherland,
Sunsas deformation was confined to three orogenic 1982). Both the Huanchaca and Marimai dykes and
belts: the WNW trending Sunsas belt, the NNW trend- the Velasco Alkaline Complex cross-cut structures
ing Aguapeı́ belt and the W trending Nova Brasilândia attributed to both the San Ignacio and Sunsas orogenies
belt (Geraldes et al., 1997; Litherland et al., 1989; and intrude the Chiquitania Complex and San Ingacio
Tohver et al., 2004b). Sunsas deformation is most Group in the middle of the study area (Fig. 3). The
intensely developed in the southwest of the study Huanchaca and Marimai dolerites are regarded as sep-
area (Sunsas belt—Fig. 2) where upright NW trending arate intracontinental phases of basic igneous activity.
folds formed contemporaneous with low to medium The Marimai have a distinct east–west trend and can
grade metamorphism. Late syn- to post-tectonic gran- be clearly identified geophysically. The Velasco Alka-
ites were also emplaced along this belt (Litherland et line Complex forms a linear NE–SW trend and was
al., 1989). Deformation at this time reset the Rb–Sr emplaced during the Late Jurassic or Earliest Creta-
and K–Ar isotope systems south of the Rio Negro Front ceous (∼140 Ma). It is thought to be related to the initial
(Fig. 2). Further to the northeast, non-penetrative defor- rifting between Africa and South America (Darbyshire
mation and little or no metamorphism is described. and Fletcher, 1979).
Along the Aguapeı́ belt, Geraldes et al. (1997) describe
northwest trending (upright to northeast dipping) folds
and southwest directed thrusting of the basement over 3. Structural analysis
the Aguapeı́ Group (Fig. 2). K–Ar ages from this region
yield ages between 960 and 920 Ma (Geraldes et al., Geological analysis of the study area was conducted
1997) that are similar to the Rb–Sr and K–Ar ages from using structural observations made via road and jungle
the Sunsas region (Berrrangé and Litherland, 1982; traverses together with the interpretation of airborne
Litherland et al., 1989). magnetic and radiometric data collected by the Boli-
A similar direction of shortening is implied along the vian Government between 1996 and 1997. Field struc-
Nova Brasilândia belt (Luft et al., 2000; Tohver et al., tural observations were concentrated in the two regions
2004b), although metamorphic grade is higher—upper- south of the Rio Negro Front. One was in the vicinity
amphibolite to granulite facies (Tohver et al., 2004b). of San Javier and Concepción. The second was in the
Within the Nova Brasilândia belt orogenesis is con- vicinity of San Ignacio and San Rafael (Fig. 2). The
strained by U–Pb zircon ages of approximately geophysical images together with the field data have
1110 Ma from syn-tectonic orthogneiss (1113 ± 56) been used re-map the region (Fig. 3). Our map does not
and leucogranite (1110 ± 8 Ma) and by U–Pb monazite differ significantly from that presented by the British
and titanite ages of 1100–1080 and 1085–1020 Ma Geologic Survey (Litherland, 1984). However, we were
(Rizzotto, 1999; Luft et al., 2000; Tohver et al., 2004b). not able to define a boundary between the Chiquitania
Cooling from peak metamorphism is recorded by horn- Complex and the Lomas Maneches Complex and have
blende and biotite 40 Ar/39 Ar ages of ca. 970 and ca. hence grouped these rocks together (Fig. 3). Neverthe-
940 Ma, respectively (Tohver et al., 2004b). The region less, we are better able to define the structural trends
in between these belts is referred to as the Pensamiento preserved in the basement as well as the distribution
Complex, a vast region of San Ignacio-aged granite of the different generations of intrusions. We are also
S.D. Boger et al. / Precambrian Research 139 (2005) 121–146
Fig. 3. Regional geology map with basement trends for the study area. Boundaries of recent alluvium taken from the 1:1,000,000 scale British Geological Survey map of the region
(Litherland, 1984). Numbered stars refer to sample numbers and locations described in the text.

127
128 S.D. Boger et al. / Precambrian Research 139 (2005) 121–146

better able to delineate the extent of the Sunsas Group


southwest of San Rafael and the extent of the Mari-
mai dyke suite between San Miguel and San Ignacio
(Fig. 3).

3.1. San Ignacio region

The basement rocks from the San Ignacio region


preserve evidence of four phases of deformation. The
earliest recognised foliation (S1 ) is defined by a pre-
ferred mineral orientation that is always concordant
with original sedimentary layering (S0 ). In both the
metasediments and the granites of the Lomas Maneches
and Chiquitania complexes, this fabric was defined by a
preferred orientation of biotite or by sub-centimetre to
centimetre scale leucosomes. In the San Ignacio Group,
S1 was invariably defined by muscovite. This fabric was
folded into layer-parallel isoclinal metre to sub-metre
scale F2 folds. Although S0 /S1 was tightly folded dur-
ing D2 , there is little development of an axial planar
fabric in the gneissic rocks of the Lomas Maneches or Fig. 4. Structural data from the San Ignacio region. (a) Plot of poles
to S0 /S1 . (b) Kamber contour plot of (a). Counting circle = 5.9%,
Chiquitania complexes. In contrast, the schists of the contour interval = 1σ and significance level = 2σ. (c) Fold plunge data
San Ignacio Group contain a well-developed muscovite for D3 N–NE trending folds. (d) Fold plunge data for D4 NW trending
crenulation cleavage (S2 ). The San Ignacio granite suite folds.
also contains a foliation that is parallel to S0 /S1 . How-
ever, in these granites this fabric is nowhere folded by (S0 /S1 ) as well as form along the axial plane of the
F2 , a characteristic that differentiates this granite suite north–north-east trending fold set. We imply from this
from the isoclinally folded intrusive rocks within the that both the development of the layer-parallel folia-
Lomas Maneches and Chiquitania complexes. The San tion and the north–north-east trending folds occurred
Ignacio granites are consequently considered to post- at similar metamorphic grades and were consequently
date the onset of orogenesis, but to have crystallised also similar in age. In contrast, the north–west trend-
prior to the end of D1–2 . ing folds reorient the strata, but do not appear to be
Subsequent deformation resulted in the formation associated with migmatite formation. They are there-
of two nearly orthogonal meso- to macroscale upright fore considered to be younger and to have developed
fold generations. This can be seen in the distribution of at lower metamorphic grade. Thus, we consider the
poles to S0 /S1 and in the orientation of fold axes and north–north-east trending folds (F3 ) to pre-date the
intersection lineation data (Fig. 4). These data show northwest trending folds (F4 ).
fold axes trends to the north–north-east, and to the
north–west. Both fold sets doubly plunge at shallow 3.2. San Javier region
to moderate angles and have vertically oriented axial
surfaces. These two generations of folds are responsi- The San Javier region preserves a similar sequence
ble for the distinct dome and basin interference patterns of deformation events. The earliest fabric observed is
that dominate the outcrop pattern of San Ignacio–San a composite S0 /S1 foliation that defines sub-metre to
Rafael region (Fig. 3). metre-scale isoclinal folds (F2 ) in the metasediments.
We were unable to determine the relative timing of Similar to the San Ignacio region, a layer-parallel folia-
these two fold generations via overprinting relation- tion also occurs in the San Ignacio granitoids, but again
ships. However, in the Chiquitania Complex, leuco- this foliation is not folded by the earliest generation of
somes were observed to lie parallel with the layering folds (F2 ) recognised in the adjacent metasediments. In
S.D. Boger et al. / Precambrian Research 139 (2005) 121–146 129

Fig. 5. Structural data from the San Javier region. (a) Plot of poles to S0 /S1 . (b) Kamber contour plot of (a). Counting circle = 16.1%, contour
interval = 1σ and significance level = 2σ. For (a) and (b) note the more linear distribution around a single great circle when compared to the
scattered results from the San Ignacio region (Fig. 4). (c) Measured fold axes for upright NW trending folds (D4 ). The great circle defines the
best-fit fold axis. Note that this calculated fold axis dissects the D4 fold plunge data from the San Ignacio area (Fig. 4d).

contrast to the San Ignacio region, the San Javier region Cathodoluminescent (CL) imaging was conducted to
is dominated by upright northwest trending folds and assess the internal structure of the unknown zircons. A
a similarly oriented structural grain (Fig. 3). A plot of primary beam of O− ions was used to sputter positive
poles to S0 /S1 from San Javier defines a single north- secondary ions from areas ∼25 ␮m in diameter from
east trend and a shallow northwest or southeast plunge individual sectioned zircons. Zr2 O+ , 206 Pb+ , 207 Pb+ ,
(Fig. 5). Measured fold axes are sub-vertical and trend 208 Pb+ , 238 U+ , 232 ThO+ , and 238 UO+ , were measured

northwest, parallel to the youngest folds (F4 ) observed in cycles by magnetic field switching, seven cycles per
in the San Ignacio region. We observed no direct evi- data set. The exception to this is sample 221D where
dence for the earlier generation of upright folds (F3 ) three cycles of magnetic field switching per data set
observed in the San Ignacio region. However, refold- was applied. Analysis of unknowns was interspersed
ing has been reported by previous workers (Berrrangé with analyses of the standard AS3 in order to monitor
and Litherland, 1982) and thus we infer that the San the differential fractionation between U and Pb. Data
Javier region was also deformed during D3 . However, were examined, processed and reduced using SQUID
D4 strain in this area was considerably higher than in software of Ludwig (2001). Ages presented in the text
the San Ignacio region and resulted in a more pervasive are stated with 2σ confidence limits.
overprint. This region of more intense D4 strain corre-
sponds with the west to northwest trending Sunsas belt 4.1. Sunsas suite
along which the Sunsas granite suite intrudes (Fig. 3).
These granites mostly post-date the northwest trending The Sunsas suite intrudes as circular or slightly
F4 folds. elliptical plutons that are essentially undeformed (post-
D4 ). We obtained U–Pb data from the Taperas granite
(sample 301), a relatively small nearly circular plu-
4. Geochronology ton that crops out to the south of San Javier (Fig. 3).
This granite contains biotite, quartz, plagioclase and
Zircons for U–Pb SHRIMP analysis were sepa- alkali feldspar. Muscovite is also present, but forms
rated by standard heavy liquid and magnetic separation a minor constituent, as does sphene. Centimetre-scale
procedures, then by hand picking. They were then feldspar phenocrysts are common and typically show
mounted in epoxy resin discs along with fragments rapakivi textures—plagioclase mantled by K-feldspar.
of zircon standard AS3. The discs were polished and Zircons obtained from this granite are large, commonly
Au-coated before being analysed on the SHRIMP RG >200 ␮m and mostly euhedral with aspect ratios in the
ion-microprobe at the Australian National University. order of 3–4:1. The U content of zircon is variable,
130 S.D. Boger et al. / Precambrian Research 139 (2005) 121–146

Fig. 7. U–Pb concordia diagram for the Taperas granite (Sunsas


suite). 206 Pb/207 Pb ages are stated to 2σ (95%) confidence limits
while the illustrated error ellipses reflect 1σ confidence limits (68%).

simple concentrically zoned structure of the zircons,


together with the single age population obtained, we
interpret the age of 1076 ± 18 Ma to date the emplace-
ment of the Taperas granite (Fig. 7). Given the late-syn-
to post-tectonic timing implied for the Taperas granite
(Fig. 3), we interpret all of the observed stages of duc-
Fig. 6. Cathodoluminescence images of zircon morphologies from tile deformation (D1–4 ) to have occurred prior to this
the Sunsas and San Ignacio granite suites. (a) Taperas granite (sample time.
301) and (b) San Rafael Granite (sample 228). Analysis spots and
obtained ages are shown.
4.2. San Ignacio suite

from 20 to 700 ppm while Th/U ratios vary between The San Ignacio suite is a group of porphyritic
0.3 and 1.6. The cathodoluminescent (CL) response of granites that intrude at, or structurally above, the con-
these zircons is also variable. The most common zircon tact between the San Ignacio Group and the underlying
morphology shows subtle concentric growth zoning Chiquitania Gneiss Complex. Centimetrescale phe-
(Fig. 6a). However, in some grains there is a concordant nocrysts of K-feldspar and the absence of both leuco-
but very distinct high-CL core. These cores are com- somes and isoclinal folds differentiate these rocks from
monly structureless, although in some instances there the granitoids that intrude the Lomas Maneches and
is oscillatory zoning parallel with that observed in the Chiquitania complexes. These characteristics suggest
low-CL rims (Fig. 6a). Other grains are dominated by that the San Ignacio suite is younger than the granitoids
high-CL zircon and show no evidence for a low-CL in the Lomas Maneches and Chiquitania complexes,
rim. and were not melted subsequent to its emplacement.
Eighteen analyses from various positions within dif- We consider the San Ignacio suite to have been
ferent zircon grains were obtained from this sample emplaced synchronous with the early stages of the
(Table 1). Both the low- and high-CL response regions San Ignacio Orogeny (prior to both periods of upright
give similar ages. Most analyses sit on or near concor- folding).
dia (Fig. 7) and using all but two the two most discor- We obtained U–Pb data from the San Rafael Gran-
dant data (7.1 and 8.1) we obtain a 207 Pb/206 Pb age ite (sample 228) that crops out south of the main road
of 1076 ± 18 Ma (MSWD = 1.4). Given the relatively between San Rafael and San Miguel (Fig. 3). This
Table 1
U–Th–Pb compositions of zircon from the Sunas and San Ignacio suites
204
Grain U (ppm) Th (ppm) Th/U Pb* (ppm) Pb/206 Pb f206 (%) Radiogenic ratios Age (Ma) % Disc
spot 206
Pb/238 U ± 207
Pb/235 U ± 207
Pb/206 Pb ± 206
Pb/238 U ± 207
Pb/206 Pb ±
Sample 301—Taperas granite (Sunsas suite)
1.1 292 164 0.56 47 0.00010 0.17 0.189 0.003 1.977 0.042 0.076 0.001 1116 15 1091 32 −2
2.1 683 326 0.48 106 0.00017 0.28 0.180 0.002 1.870 0.033 0.075 0.001 1069 11 1073 27 0
3.1 19 25 1.34 3 0.00165 2.81 0.167 0.009 1.697 0.306 0.074 0.013 994 50 1037 347 4
4.1 205 134 0.65 31 0.00111 1.90 0.172 0.005 1.723 0.113 0.073 0.004 1023 27 1004 119 −2
5.1 27 38 1.39 4 – <0.01 0.184 0.004 2.013 0.069 0.079 0.002 1089 21 1180 53 8
6.1 192 117 0.61 29 0.00018 0.31 0.176 0.004 1.789 0.061 0.074 0.002 1046 23 1032 51 −1
7.1 27 38 1.40 4 – <0.01 0.195 0.006 2.430 0.193 0.090 0.007 1149 31 1432 140 25

S.D. Boger et al. / Precambrian Research 139 (2005) 121–146


8.1 413 186 0.45 45 0.00171 2.93 0.123 0.002 1.195 0.077 0.071 0.004 745 9 949 129 27
11.1 92 121 1.32 14 0.00039 0.67 0.178 0.003 1.692 0.085 0.069 0.003 1053 16 903 99 −14
12.1 47 71 1.51 7 0.00068 1.17 0.172 0.004 1.677 0.102 0.071 0.004 1024 21 947 115 −8
13.1 348 223 0.64 50 0.00014 0.23 0.167 0.003 1.701 0.039 0.074 0.001 994 16 1040 31 5
14.1 515 139 0.27 65 0.00089 1.52 0.144 0.002 1.397 0.083 0.070 0.004 866 13 942 118 9
15.1 60 95 1.59 9 0.00076 1.28 0.180 0.003 1.872 0.094 0.075 0.004 1069 16 1076 96 1
16.1 29 23 0.80 4 0.00067 1.16 0.176 0.005 1.802 0.068 0.074 0.002 1046 31 1048 51 0
17.1 257 176 0.68 39 0.00015 0.25 0.178 0.002 1.865 0.036 0.076 0.001 1058 11 1091 31 3
18.1 565 282 0.50 87 0.00004 0.07 0.178 0.002 1.837 0.023 0.075 0.000 1058 10 1060 13 0
19.1 422 638 1.51 67 0.00005 0.08 0.184 0.002 1.916 0.029 0.075 0.001 1090 13 1080 16 −1
20.1 213 344 1.62 32 – <0.01 0.176 0.002 1.865 0.031 0.077 0.001 1044 14 1118 18 7
Sample 228—San Rafael Granite (San Ignacio suite)
1.1 1793 1871 1.04 326 0.00067 1.11 0.210 0.003 2.485 0.060 0.086 0.002 1226 17 1338 36 9
2.1 578 471 0.82 88 0.00287 4.55 0.169 0.002 2.370 0.173 0.102 0.007 1006 12 1657 133 65
3.1 1446 68 0.05 290 0.00002 0.03 0.234 0.002 2.762 0.030 0.086 0.000 1353 13 1333 6 −1
4.1 5890 755 0.13 681 0.00141 2.35 0.131 0.001 1.443 0.070 0.080 0.004 796 8 1189 94 49
4.2 155 59 0.38 43 0.00099 1.56 0.316 0.004 4.431 0.133 0.102 0.003 1769 20 1657 50 −6
5.1 1692 2555 1.51 333 0.00006 0.09 0.229 0.003 2.706 0.035 0.086 0.000 1327 15 1335 7 1
6.1 1711 906 0.53 334 0.00001 0.01 0.227 0.002 2.709 0.030 0.086 0.000 1320 13 1348 6 2
7.1 5363 2841 0.53 1186 0.00002 0.04 0.257 0.003 3.073 0.034 0.087 0.000 1477 14 1351 7 −8
8.1 1938 242 0.12 403 0.00016 0.27 0.241 0.004 2.872 0.051 0.086 0.000 1393 21 1346 11 −3
9.1 336 359 1.07 71 0.00049 0.77 0.244 0.003 3.489 0.064 0.104 0.001 1409 14 1689 26 20
10.1 2779 753 0.27 541 0.00007 0.11 0.226 0.003 2.679 0.037 0.086 0.000 1315 15 1336 9 2
11.1 4058 1297 0.32 485 0.00065 1.08 0.138 0.001 1.509 0.032 0.079 0.001 831 8 1184 37 42
12.1 5527 2000 0.36 950 0.00013 0.21 0.200 0.002 2.277 0.024 0.083 0.000 1174 11 1262 7 7
13.1 338 227 0.67 89 0.00002 0.03 0.307 0.004 4.380 0.059 0.103 0.000 1726 19 1687 9 −2
14.1 230 111 0.48 47 0.00091 1.43 0.236 0.003 3.365 0.081 0.104 0.002 1363 14 1690 39 24
15.1 4467 1299 0.29 676 0.00078 1.29 0.174 0.002 2.012 0.042 0.084 0.002 1033 10 1292 36 25
16.1 1295 82 0.06 262 0.00004 0.06 0.235 0.002 2.789 0.031 0.086 0.000 1363 13 1336 7 −2
17.1 4001 155 0.04 621 0.00122 2.03 0.177 0.002 2.019 0.067 0.083 0.003 1050 10 1264 61 20
18.1 3511 1641 0.47 653 0.00028 0.46 0.216 0.002 2.507 0.032 0.084 0.001 1258 12 1301 13 3
19.1 3815 520 0.14 338 0.00059 1.01 0.102 0.001 1.038 0.021 0.074 0.001 626 6 1035 35 65
Uncertainties given at the 1σ level. f206 (%) denotes the percentage of 206 Pb that is common Pb. % Disc denotes the degree of concordance (0% denotes a concordant analysis).

131
132 S.D. Boger et al. / Precambrian Research 139 (2005) 121–146

granite is pale grey in colour and contains phenocrysts with an MSWD of 2.8. Excluding one of the datum
of K-feldspar that occur in a matrix of finer-grained points (1.1), the MSWD can be reduced to 0.99.
K-feldspar, plagioclase, quartz, biotite and muscovite. The resultant upper intercept is unchanged with a
Zircons from this sample are large (>200 ␮m) and mean 207 Pb/206 Pb age of 1337 ± 7 Ma. From the five
commonly show well-formed euhedral terminations. core analyses, an upper intercept 207 Pb/206 Pb age
Cathodoluminescent images of these zircons show of 1686 ± 16 Ma (MSWD = 0.14) was obtained. The
that they contain distinct highly luminescent cores lower intercepts for the rim and core analyses are
surrounded by large mantles of new low-CL zircon 317 ± 52 and −9 ± 290 Ma, respectively. Neither age
(Fig. 6b). The cores commonly show oscillatory zon- is considered to have geological significance. We inter-
ing that terminates abruptly against broken or partly pret the upper intercept age obtained from the rim
resorbed boundaries of the zircon core (Fig. 6b). The analyses to constrain the age of emplacement of the
zircon cores are characterised by U concentrations San Rafael Granite (1334 ± 12 Ma) and infer from the
of <600 ppm, and contrast with the surrounding low- core ages (1686 ± 16 Ma) that this granite melted from
CL mantles where U concentrations are commonly a Palaeoproterozoic protolith.
between 1500 and 6000 ppm. The low-CL rims show
subtle concentric zoning, although this tends not to be 4.3. Lomas Maneches suite
as well developed as in the cores. We interpret the cores
of the zircons to have been inherited, whilst the euhe- We define the Lomas Maneches suite as consist-
dral low-CL rims we infer to have grown at the time of ing of the granitic sills that form the dominant rock
granite emplacement and crystallisation. type in the Lomas Maneches Complex as defined by
Twenty analyses of 19 zircon grains were col- Berrrangé and Litherland (1982). Granites of this suite
lected from this sample. Five were from the zir- also intrude the Chiquitania Complex, although are
con cores; the remaining 15 were of the zircon rims unknown in the San Ignacio Group. In contrast to
(Table 1). Analyses from both core and rim loca- the granitoids of the San Ignacio suite, these intru-
tions show a distribution of concordant and discor- sions invariably contain all of the structures preserved
dant data (Fig. 8). The upper intercept from the rim in their host rocks and commonly show evidence for
analyses yielded a 207 Pb/206 Pb age of 1334 ± 12 Ma partial melting subsequent to their intrusion. They are
consequently interpreted to pre-date the San Ignacio
Orogeny.
Sample 240 was collected from the Lomas
Maneches region east of San Rafael (Fig. 3). The
rock, an orthopyroxene bearing granitoid, is bluish
grey in colour and mostly medium grained without
obvious feldspar phenocrysts. Its mineral assemblage
consists of orthopyroxene, minor biotite and/or horn-
blende, K-feldspar, quartz and magnetite. The rock
contains a well-developed gneissic foliation (S1 ) that
is locally cross-cut by coarse-grained leucosomes that
commonly contain more abundant and coarser-grained
orthopyroxene.
Zircons obtained from this granite are between
100 and 300 ␮m long and have aspect ratios close to
3:1. The U content of zircon is variable, from 40 to
1100 ppm, while Th/U ratios vary between 0.3 and
Fig. 8. U–Pb concordia diagram for the San Rafael Granite (San
1.8. Cathodoluminescent images of these zircons show
Ignacio suite). 206 Pb/207 Pb ages are stated to 2σ (95%) confidence
limits while the illustrated error ellipses reflect 1σ confidence limits that most grains have distinct concentric growth zoning
(68%). Light grey error ellipses are from core analyses, dark grey (Fig. 9a) which is often truncated by the margins of the
are from rim analyses. zircon grains. In some instances, concentric zoned zir-
Table 2
U-Th-Pb compositions of zircon from the Lomas Maneches granite suite
204
Grain U (ppm) Th (ppm) Th/U Pb* (ppm) Pb/206 Pb f206 (%) Radiogenic ratios Age (Ma) % Disc
spot 206
Pb/238 U ± 207
Pb/235 U ± 207
Pb/206 Pb ± 206
Pb/238 U ± 207
Pb/206 Pb ±
Sample 240—opx bearing gneissic granite (Lomas Maneches suite)
1.1 213 192 0.90 39 0.00018 0.30 0.213 0.003 2.478 0.048 0.084 0.001 1245 14 1302 28 5
1.2 277 306 1.11 69 0.00016 0.25 0.291 0.003 4.021 0.057 0.100 0.001 1646 17 1629 16 −1
2.1 182 267 1.47 36 0.00008 0.12 0.228 0.003 2.671 0.061 0.085 0.002 1323 15 1317 38 0
2.2 281 185 0.66 70 0.00004 0.06 0.289 0.003 4.055 0.056 0.102 0.001 1634 17 1660 14 2
2.3 144 195 1.35 28 0.00008 0.13 0.222 0.003 2.587 0.044 0.084 0.001 1293 15 1303 22 1
3.1 58 57 0.97 15 0.00034 0.53 0.299 0.006 4.162 0.131 0.101 0.003 1688 28 1640 47 −3
−1

S.D. Boger et al. / Precambrian Research 139 (2005) 121–146


4.1 174 236 1.36 44 0.00003 0.05 0.295 0.004 4.145 0.064 0.102 0.001 1667 18 1658 17
5.1 48 43 0.91 12 0.00003 0.05 0.283 0.015 4.193 0.278 0.107 0.004 1608 77 1755 70 9
6.1 69 123 1.79 14 0.00008 0.12 0.229 0.003 2.720 0.082 0.086 0.002 1330 17 1340 51 1
6.2 272 299 1.10 61 0.00002 0.03 0.260 0.008 3.628 0.128 0.101 0.001 1489 43 1648 25 11
7.1 70 56 0.81 19 0.00011 0.18 0.316 0.004 4.463 0.094 0.102 0.002 1769 22 1670 29 −6
8.1 101 108 1.06 26 0.00010 0.16 0.300 0.004 4.280 0.072 0.103 0.001 1692 19 1686 20 0
9.1 145 144 0.99 32 0.00011 0.18 0.256 0.003 3.272 0.056 0.093 0.001 1470 17 1480 22 1
10.1 1101 320 0.29 218 0.00002 0.04 0.230 0.003 2.690 0.034 0.085 0.000 1336 14 1309 9 −2
10.2 242 201 0.83 61 0.00004 0.06 0.296 0.003 4.217 0.057 0.103 0.001 1670 17 1686 13 1
11.1 260 224 0.86 66 0.00009 0.15 0.294 0.003 4.130 0.054 0.102 0.001 1663 17 1657 12 0
11.2 193 154 0.80 52 0.00010 0.16 0.311 0.004 4.509 0.070 0.105 0.001 1748 18 1714 18 −2
12.1 395 415 1.05 102 0.00013 0.20 0.300 0.004 4.268 0.060 0.103 0.001 1690 18 1683 13 0
13.1 132 172 1.31 34 0.00015 0.23 0.304 0.004 4.329 0.080 0.103 0.001 1711 19 1684 25 −2
14.1 173 195 1.12 43 0.00009 0.15 0.285 0.003 3.946 0.060 0.100 0.001 1618 17 1630 17 1
15.1 1029 311 0.30 201 0.00005 0.07 0.227 0.003 2.686 0.038 0.086 0.000 1320 16 1332 8 1
Sample 261—hb bearing gneissic granite (Lomas Maneches suite)
1.1 313 204 0.65 85 0.00004 0.07 0.317 0.004 4.528 0.061 0.104 0.001 1776 18 1688 12 −5
2.1 372 229 0.62 89 0.00005 0.08 0.279 0.005 3.990 0.079 0.104 0.001 1588 27 1689 11 6
3.1 94 60 0.64 25 0.00017 0.27 0.311 0.004 4.435 0.097 0.103 0.002 1747 22 1685 30 −4
4.1 594 534 0.90 154 0.00003 0.05 0.302 0.003 4.317 0.054 0.104 0.001 1702 17 1690 11 −1
5.1 1053 879 0.84 245 0.00005 0.08 0.270 0.003 4.039 0.119 0.108 0.003 1542 15 1773 50 15
5.2 238 133 0.56 65 0.00003 0.05 0.317 0.010 4.516 0.159 0.103 0.002 1776 47 1683 32 −5
6.1 95 58 0.61 25 0.00008 0.13 0.311 0.004 4.384 0.085 0.102 0.001 1743 22 1668 24 −4
7.1 499 320 0.64 124 – <0.01 0.289 0.004 4.143 0.055 0.104 0.000 1637 18 1696 8 4
8.1 433 210 0.48 107 0.00003 0.05 0.287 0.003 4.115 0.051 0.104 0.001 1625 16 1699 10 5
9.1 355 224 0.63 88 0.00005 0.08 0.287 0.004 4.090 0.063 0.103 0.001 1629 20 1682 11 3
10.1 457 290 0.63 118 0.00003 0.04 0.302 0.003 4.321 0.051 0.104 0.000 1700 16 1694 9 0
11.1 635 354 0.56 156 0.00002 0.03 0.286 0.003 4.056 0.047 0.103 0.000 1624 15 1673 8 3
12.1 655 327 0.50 168 0.00002 0.04 0.299 0.003 4.273 0.049 0.104 0.000 1688 16 1688 7 0
13.1 523 276 0.53 131 0.00006 0.09 0.292 0.003 4.212 0.049 0.105 0.000 1651 16 1709 9 3
14.1 113 52 0.46 28 0.00004 0.07 0.286 0.004 4.036 0.064 0.102 0.001 1620 18 1669 17 3
15.1 122 73 0.60 30 0.00005 0.09 0.291 0.005 4.123 0.074 0.103 0.001 1645 23 1676 15 2
Uncertainties given at the 1σ level. f206 (%) denotes the percentage of 206 Pb that is common Pb. % Disc denotes the degree of concordance (0% denotes a concordant analysis).

133
134 S.D. Boger et al. / Precambrian Research 139 (2005) 121–146

MSWD of less than one (0.78). The remaining core


analyses (5.1 and 11.2) yield older ages of 1755 ± 70
and 1714 ± 18 Ma, both of which are interpreted to be
of inherited origin. The six analyses of zircon rims yield
a weighted mean 207 Pb/206 Pb age of 1320 ± 11 Ma
(MSWD = 1.0). This age is within error of the age
obtained from sample 228 described above. An addi-
tional grain (9.1) yielded an age of 1480 ± 22 Ma.
This result is interpreted to represent a mixing age
between the zircon core and the rim (poorly positioned
analysis spot) and is not considered to have any geo-
logical meaning. Based on these results, we infer that
the granite from which sample 240 was collected was
emplaced at 1663 ± 13 Ma, then partly re-melted at
1320 ± 11 Ma.
Sample 261 was collected from a prominent
outcrop beside the main road south from San Rafael
(Fig. 3). The rock is a medium-grained, pale-grey or
pinkish granitoid that contains hornblende, biotite,
K-feldspar, quartz and magnetite. It was emplaced
close to the contact between the Chiquitania Complex
and the overlying San Ignacio Group (Fig. 3). The
rock contains a strong gneissic foliation that is folded
by sub-metre to metrescale F2 and F3 folds. Similar
Fig. 9. Cathodoluminescence images of zircon morphologies from to sample 240, this rock contains quartzo-feldspathic
granites from the Lomas Maneches suite: (a) sample 240 and (b) leucosomes that form both layer-parallel and axial
sample 261. Analysis spots and obtained ages are given. planar segregations of coarser-grained quartz and
K-feldspar.
Zircons from sample 261 are generally less than
cons form cores surrounded by a two-part overgrowth 200 ␮m long and have aspect ratios between 2:1 and
of structureless zircon. These overgrowths consist of 3:1. The grains are mostly euhedral in shape but consis-
an inner layer of low-CL zircon, which is mantled tently show rounded terminations. In CL, zircons from
by an outer layer that shows a similar CL intensity this sample generally show a simple concentric zon-
as the zircon core. The U–Th concentrations for the ing pattern (Fig. 9b). Distinct cores can be observed in
cores and the rims are indistinguishable (Table 2). The transmitted light in some grains, although these are rel-
cores are interpreted to represent zircon grown at the atively uncommon and difficult to distinguish from the
time of emplacement, whilst the rims are thought to rims in CL. Based on this straightforward morphol-
reflect a subsequent phase of melting, the existence of ogy, we interpret the zircons to represent magmatic
which is inferred from the presence of the observed grains that grew at the time of granite emplacement.
leucosomes. Sixteen analyses of 15 zircon grains were obtained
Twenty-one analyses of 15 zircon grains were from this sample (Table 2). Fifteen of these analyses
obtained from this sample (Table 2). Of these, fourteen yield a concordant 207 Pb/206 Pb age of 1689 ± 5 Ma
analyses were of zircon cores and the remainder were with an MSWD of 1.06 (Fig. 10b). The remaining
of zircon rims. Twelve of the core analyses lie on or datum was obtained from the core of a zircon and yields
near concordia and yield a weighted mean 207 Pb/206 Pb an age of 1773 ± 50 Ma. This age is similar to the older
age of 1663 ± 13 Ma (MSWD = 1.6) (Fig. 10a). A sub- core ages obtained from sample 240 (1755 ± 70 and
set of these, ignoring the two youngest analyses, yields 1714 ± 18 Ma) and is taken to be inherited from the
a statistically identical age of 1670 ± 11 Ma with an source of the granite.
S.D. Boger et al. / Precambrian Research 139 (2005) 121–146 135

Fig. 10. U–Pb concordia diagrams for the Lomas Maneches suite: (a) Sample 240 and (b) sample 261. 206 Pb/207 Pb ages are stated to 2σ (95%)
confidence limits while the illustrated error ellipses reflect 1σ confidence limits (68%). Light grey error ellipses are from core analyses, dark
grey are from rim analyses. White error ellipses are discordant or suspect analyses ignored in the age determination given.

4.4. Chiquitania Complex entirely from the leucocratic portion of the rock. In
this sample, zircons are characterised by rounded cores
Two samples (235A and 235A-1) were obtained with concentrically zoned high-CL morphology, simi-
from an outcrop of biotite-bearing felsic gneiss from lar to most grains in sample 235A. However, the high-U
the Chiquitania Complex approximately 20 km east of rims in sample 235A-1 are well developed and consti-
San Rafael (Fig. 3). This rock is comprised of pla- tute at least a third of most zircon grains (Fig. 11b).
gioclase, alkali feldspar, quartz, minor biotite (<5%) In both samples, we infer the rims to be of metamor-
and magnetite. The rock did not contain feldspar phe- phic origin—having grown during the melting of these
nocrysts common to the intrusive rocks of both the rocks (Williams, 2001). The extent to which the rims
Lomas Maneches and San Ignacio suites. The rock have developed is interpreted to reflect the location of
was therefore interpreted to be of sedimentary or vol- the sample, wider and better-developed rims formed
canic origin. Partial melting of the rock is inferred from in the leucosome (235A-1), and narrower and more
the extensive development of leucosomes. These leuco- poorly developed rims in the country rock (235A). In
somes are more coarse-grained than the rest of the rock both samples, the zircon cores are taken has having
and consist predominantly of K-feldspar and quartz. been inherited from the protolith of the paragneiss.
They form bands parallel with the dominant foliation For sample 235A-1, 19 analyses from 15 zircon
as well as segregations that cross-cut this foliation at grains were obtained (Table 3). Fourteen of these were
high angles, commonly forming along the axial plane of zircon rims, which produce a scattered distribution
of mesoscale northeast trending D3 folds. on or near the concordia at ∼1350 Ma. The ten least
Sample 235A was obtained from a relatively leu- discordant of these analyses yield a mean 207 Pb/206 Pb
cosome poor portion of the outcrop. Zircons from this age of 1333 ± 6 Ma (MSWD = 2.6; Fig. 12a). We inter-
sample are of variable size with partly rounded or bro- pret this age to reflect the time of partial melting (peak
ken terminations (Fig. 11a). The zircons are mostly metamorphism) in the Chiquitania Complex. This age
luminescent and show variably developed oscillatory is identical to the emplacement age obtained from the
zoning. The U content of the grains is mostly less San Rafael Granite (1337 ± 7 Ma) and the rim age
than 200 ppm. On many of the grains, there is a nar- (1320 ± 11 Ma) obtained from sample 240 from the
row and poorly developed high-U (>1000 ppm) and Lomas Maneches suite. The remaining 5 analyses were
low-CL rim (Fig. 11a). Sample 235A-1 was collected of zircon cores. Four of these analyses lie on concordia
136
Table 3
U–Th–Pb compositions of zircon from paragneisses
204
Grain U (ppm) Th (ppm) Th/U Pb* (ppm) Pb/206 Pb f206 (%) Radiogenic ratios Age (Ma) % Disc
spot
206
Pb/238 U ± 207
Pb/235 U ± 207
Pb/206 Pb ± 206
Pb/238 U ± 207
Pb/206 Pb ±
Sample 235A—Chiquitania Paragneiss (country rock)
1.1 194 199 1.02 30 0.00068 1.07 0.176 0.002 2.553 0.062 0.105 0.002 1047 11 1713 39 64
2.1 107 90 0.84 30 0.00010 0.15 0.321 0.004 4.766 0.071 0.108 0.001 1797 19 1758 15 −2
3.1 118 116 0.98 27 0.00013 0.20 0.262 0.004 3.875 0.064 0.107 0.001 1501 19 1753 17 17
4.1 1089 45 0.04 222 0.00001 0.02 0.237 0.003 2.882 0.034 0.088 0.000 1371 14 1387 9 1
5.1 80 83 1.04 22 0.00004 0.06 0.315 0.005 4.762 0.093 0.110 0.001 1767 26 1792 19 1
6.1 486 46 0.09 102 0.00004 0.06 0.244 0.003 3.174 0.041 0.094 0.000 1410 15 1511 10 7
7.1 106 142 1.33 30 0.00000 0.00 0.323 0.004 4.848 0.070 0.109 0.001 1804 19 1781 13 −1

S.D. Boger et al. / Precambrian Research 139 (2005) 121–146


8.1 1120 105 0.09 244 0.00004 0.06 0.253 0.003 3.268 0.040 0.094 0.001 1456 13 1499 12 3
9.1 194 158 0.82 52 0.00005 0.08 0.313 0.004 4.657 0.060 0.108 0.001 1754 18 1766 11 1
10.1 150 144 0.96 42 0.00004 0.07 0.323 0.004 4.802 0.064 0.108 0.001 1807 18 1760 11 −3

Sample 235A-1—Chiquitania Paragneiss (leucosomes)


1.1 2570 273 0.11 530 0.00000 0.01 0.240 0.003 2.827 0.032 0.085 0.973 1387 14 1324 5 −5
2.1 115 84 0.73 32 0.00007 0.11 0.324 0.005 4.942 0.104 0.111 0.699 1807 23 1812 27 0
2.2 1888 131 0.07 375 0.00003 0.06 0.231 0.003 2.817 0.036 0.088 0.959 1341 15 1390 7 4
3.1 674 51 0.08 127 0.00059 0.96 0.217 0.002 2.664 0.052 0.089 0.549 1265 12 1408 31 11
4.1 2194 211 0.10 456 0.00002 0.03 0.242 0.003 2.872 0.031 0.086 0.965 1396 13 1342 5 −4
5.1 2506 260 0.10 512 0.00000 0.00 0.238 0.002 2.804 0.030 0.085 0.970 1376 13 1326 5 −4
6.1 2791 297 0.11 542 0.00001 0.01 0.226 0.002 2.682 0.030 0.086 0.975 1314 13 1338 5 2
7.1 153 199 1.30 41 0.00007 0.11 0.312 0.004 4.710 0.075 0.109 0.823 1753 20 1788 16 2
8.1 260 174 0.67 36 0.00012 0.18 0.161 0.002 2.382 0.038 0.108 0.776 960 11 1760 18 83
9.1 2409 254 0.11 474 0.00001 0.01 0.229 0.002 2.689 0.029 0.085 0.966 1329 12 1319 5 −1
10.1 1919 170 0.09 372 0.00001 0.02 0.225 0.002 2.670 0.030 0.086 0.942 1310 13 1337 7 2
11.1 210 209 1.00 54 0.00005 0.08 0.302 0.004 4.473 0.066 0.107 0.843 1701 18 1757 14 3
11.2 2133 80 0.04 426 0.00001 0.02 0.232 0.002 2.757 0.031 0.086 0.968 1347 13 1340 5 −1
12.1 2099 182 0.09 412 0.00002 0.03 0.228 0.002 2.694 0.029 0.086 0.963 1325 12 1330 6 0
13.1 2224 635 0.29 585 0.00001 0.01 0.306 0.003 4.374 0.047 0.104 0.959 1721 16 1690 6 −2
13.2 4917 45 0.01 951 0.00001 0.01 0.225 0.002 2.676 0.028 0.086 0.981 1309 12 1342 4 3
14.1 2208 218 0.10 402 0.00072 1.19 0.209 0.002 2.511 0.050 0.087 0.520 1225 12 1361 33 11
14.2 481 212 0.44 94 0.00100 1.61 0.224 0.003 2.939 0.089 0.095 0.377 1302 13 1533 53 18
15.1 2795 310 0.11 528 0.00002 0.02 0.220 0.002 2.593 0.030 0.086 0.970 1281 13 1329 6 4
Sample 221D—San Ignacio Group psammite
1.1 226 73 0.32 58 0.000062 0.10 0.301 0.005 4.536 0.080 0.109 0.001 1698 23 1786 14 5
2.1 361 76 0.21 87 0.000032 0.05 0.280 0.003 4.152 0.056 0.107 0.001 1593 17 1756 12 10
3.1 252 56 0.22 68 – <0.01 0.316 0.004 4.687 0.075 0.108 0.001 1770 19 1759 19 −1
4.1 294 56 0.19 79 0.000041 0.06 0.312 0.004 4.651 0.062 0.108 0.001 1749 18 1770 11 1
5.1 119 52 0.43 32 0.000065 0.10 0.313 0.006 4.651 0.099 0.108 0.001 1755 27 1763 22 0
6.1 98 78 0.80 26 0.000129 0.20 0.309 0.005 4.488 0.087 0.105 0.001 1735 23 1721 23 −1
7.1 159 120 0.76 43 – <0.01 0.317 0.004 4.762 0.088 0.109 0.001 1776 21 1781 23 0
8.1 416 194 0.47 108 0.000024 0.04 0.301 0.004 4.285 0.062 0.103 0.001 1698 18 1681 13 −1
9.1 295 45 0.15 76 0.000041 0.06 0.299 0.004 4.444 0.069 0.108 0.001 1689 19 1759 17 4
10.1 535 297 0.55 129 0.000169 0.26 0.281 0.003 4.157 0.066 0.107 0.001 1595 17 1755 19 10
11.1 430 148 0.34 118 – <0.01 0.321 0.004 4.800 0.065 0.109 0.001 1794 18 1774 12 −1
12.1 86 46 0.54 22 – <0.01 0.300 0.005 4.290 0.095 0.104 0.001 1692 25 1690 26 0
13.1 73 43 0.59 20 0.000141 0.22 0.315 0.005 4.598 0.132 0.106 0.002 1764 26 1731 43 −2
14.1 652 243 0.37 119 0.000341 0.53 0.212 0.002 3.114 0.050 0.107 0.001 1239 13 1742 20 41
15.1 128 70 0.55 28 – <0.01 0.254 0.004 3.975 0.097 0.114 0.002 1459 20 1857 34 27
16.1 260 133 0.51 66 0.000048 0.08 0.297 0.004 4.243 0.077 0.104 0.001 1676 19 1690 24 1
17.1 473 12 0.03 132 0.000006 0.01 0.325 0.004 4.870 0.064 0.109 0.001 1816 19 1775 11 −2
18.1 125 77 0.61 31 0.000076 0.12 0.290 0.004 4.188 0.086 0.105 0.002 1639 21 1713 27 4
19.1 256 151 0.59 53 0.000071 0.11 0.241 0.003 3.510 0.062 0.106 0.001 1393 17 1724 21 24
20.1 114 79 0.70 31 0.000218 0.34 0.312 0.005 4.595 0.110 0.107 0.002 1753 23 1743 35 −1
21.1 202 38 0.19 58 0.000110 0.17 0.331 0.004 4.914 0.098 0.108 0.002 1845 22 1758 27 −5
22.1 372 58 0.16 89 0.000087 0.14 0.278 0.003 3.963 0.059 0.103 0.001 1581 17 1686 16 7
23.1 144 137 0.95 40 – <0.01 0.324 0.005 4.936 0.096 0.111 0.001 1808 22 1809 24 0
24.1 119 84 0.71 32 – <0.01 0.315 0.005 4.752 0.095 0.109 0.002 1765 22 1790 25 1

S.D. Boger et al. / Precambrian Research 139 (2005) 121–146


25.1 73 41 0.56 19 – <0.01 0.306 0.005 4.728 0.122 0.112 0.002 1719 25 1836 36 7
26.1 674 320 0.48 141 0.000166 0.26 0.243 0.003 3.503 0.051 0.104 0.001 1404 15 1704 16 21
27.1 235 162 0.69 59 0.000013 0.02 0.292 0.004 4.140 0.066 0.103 0.001 1651 19 1677 18 2
28.1 776 104 0.13 141 0.000534 0.84 0.210 0.003 3.025 0.056 0.105 0.001 1228 13 1707 26 39
30.1 635 470 0.74 129 – <0.01 0.236 0.003 3.346 0.047 0.103 0.001 1364 14 1678 14 23
31.1 415 211 0.51 100 0.000007 0.01 0.281 0.003 4.153 0.057 0.107 0.001 1596 17 1753 13 10
32.1 313 177 0.57 84 0.000032 0.05 0.313 0.004 4.810 0.075 0.111 0.001 1756 19 1822 17 4
33.1 227 142 0.62 58 0.000082 0.13 0.299 0.004 4.431 0.083 0.108 0.001 1685 21 1759 23 4
34.1 224 116 0.52 55 – <0.01 0.287 0.004 4.244 0.074 0.107 0.001 1627 21 1753 17 8
35.1 363 231 0.64 88 0.000477 0.73 0.281 0.003 4.524 0.087 0.117 0.002 1596 18 1908 26 20
36.1 499 157 0.32 96 0.000133 0.21 0.223 0.003 3.295 0.049 0.107 0.001 1296 14 1754 15 35
37.1 294 185 0.63 76 0.000111 0.18 0.302 0.004 4.276 0.063 0.103 0.001 1702 19 1672 15 −2
38.1 196 109 0.55 51 0.000051 0.08 0.302 0.004 4.501 0.085 0.108 0.001 1702 20 1766 25 4
39.1 257 127 0.49 71 0.000108 0.17 0.324 0.004 4.831 0.073 0.108 0.001 1807 21 1770 15 −2
40.1 200 106 0.53 52 – <0.01 0.304 0.004 4.405 0.077 0.105 0.001 1711 20 1717 21 0
41.1 269 137 0.51 73 0.000001 <0.01 0.317 0.004 4.730 0.071 0.108 0.001 1776 19 1769 15 0
42.1 330 154 0.47 86 0.000051 0.08 0.304 0.004 4.554 0.065 0.109 0.001 1713 18 1775 13 4
43.1 217 78 0.36 59 – <0.01 0.317 0.004 4.780 0.073 0.109 0.001 1777 20 1787 15 1
44.1 172 89 0.51 46 – <0.01 0.308 0.004 4.506 0.082 0.106 0.001 1733 21 1731 22 0
45.1 247 126 0.51 66 – <0.01 0.310 0.012 4.476 0.177 0.105 0.001 1741 58 1709 20 −2
46.1 295 59 0.20 82 – <0.01 0.323 0.004 4.812 0.073 0.108 0.001 1806 19 1765 16 −2
47.1 149 63 0.42 41 – <0.01 0.323 0.005 4.817 0.089 0.108 0.001 1803 22 1770 22 −2
48.1 192 58 0.30 53 0.000149 0.23 0.317 0.004 4.737 0.098 0.108 0.002 1776 21 1771 29 0
49.1 360 250 0.70 91 – <0.01 0.294 0.004 4.235 0.061 0.105 0.001 1661 19 1706 13 3
50.1 116 63 0.54 31 – <0.01 0.311 0.005 4.565 0.100 0.106 0.002 1748 23 1737 30 −1
Uncertainties given at the 1σ level. f206 (%) denotes the percentage of 206 Pb that is common Pb. % Disc denotes the degree of concordance (0% denotes a concordant analysis).

137
138 S.D. Boger et al. / Precambrian Research 139 (2005) 121–146

1760 Ma and the origin suggesting some recent Pb


loss.
For sample 235A, seven of the ten analyses were
of zircon cores. Five of these plots on concordia
with yield a mean 207 Pb/206 Pb age of 1764 ± 12 Ma
(MSWD = 0.76). The remaining two core analyses are
discordant and when combined with the other core data
produce a line whose lower intercept lies near the origin
(Fig. 12b). The remaining three analyses were of zir-
con rims. These yield ages between 1380 and 1510 Ma,
the youngest of which (1387 ± 9) probably has some
geological significance given the similarity to rim ages
obtained from sample 235A-1. The other two are inter-
preted to represent partial mixing between the core
(∼1760 Ma) and the rim (∼1330 Ma) portions of the
analysed zircon grains.
Combined, these data suggest that the paragneiss
protolith was derived from a predominantly Palaeopro-
terozoic source that formed at about 1765 Ma. How-
ever, the presence of detrital zircons as young as ca.
1690 Ma suggest that the paragneiss was not deposited
prior to this time. The variably developed mantles on
the zircons suggest that the sample was extensively
melted at 1333 ± 6 Ma.

4.5. San Ignacio Paragneiss

Sample (221D) was collected from within the San


Ignacio Group between San Rafael and San Ignacio
(Fig. 3). The sample was collected from a more psam-
mitic layer within a sequence that is otherwise dom-
inated by pelitic schists. The rock consists of quartz,
muscovite and minor disseminated staurolite. It con-
tains less muscovite and more quartz relative to its
more pelitic host rocks and lacks garnet and biotite,
both of which are common in the metapelites. Zir-
Fig. 11. Cathodoluminescence images of zircon morphologies from cons from sample 221D preserve reasonably euhedral
paragneisses from the Chiquitania and San Ignacio groups. (a) Chiq- grain-shapes and are uniformly clear in transmitted
uitania Paragneiss (country rock—sample 235A), (b) Chiquitania light. In CL, the zircons are mostly luminescent and
Paragneiss (leucosome—sample 235A-1) and (c) San Ignacio Parag- show distinct oscillatory zoning. However, they lack
neiss (sample 221D). Analysis spots and obtained ages are given.
evidence for the high-U (low-luminosity) phase of
new zircon growth seen in samples 235A and 235A-1.
between 1690 and 1830 Ma. Two of these yield ages of Fifty grains from this sample were analysed (Table 3).
1788 ± 16 and 1757 ± 14 Ma and are similar to the age The largest population (29 grains) yield a concordant
of the zircon cores from sample 235A (Fig. 12b). The 207 Pb/206 Pb age of 1764 ± 6 Ma (MSWD = 0.94), iden-

other two analyses (1812 ± 27 and1690 ± 6 Ma) point tical to the core ages obtained from samples 235A
to a more varied source of detritus. The remaining and 235A-1. Three grains yield older ages between
datum is discordant and falls on a line between 1840 and 1910 Ma. A younger population with an
S.D. Boger et al. / Precambrian Research 139 (2005) 121–146 139

Fig. 12. U–Pb concordia diagrams for the Chiquitqnia Paragneiss. 206 Pb/207 Pb ages are stated to 2σ (95%) confidence limits while the illustrated
error ellipses reflect 1σ confidence limits (68%). Light grey error ellipses are from core analyses, dark grey are from rim analyses. White error
ellipses are discordant or suspect analyses ignored in the age determination given.

age of ∼1690 Ma was also obtained (14 grains). The constrained by two samples dated at 1663 ± 13 and
remaining data yield a spread of ages between 1690 1689 ± 5 Ma. These ages overlap with the youngest
and 1760 Ma. These results suggest that, similar to detrital zircons obtained from the Chiquitania Com-
the Chiquitania Paragneiss described above, the pro- plex and the San Ignacio Group, a result that implies
tolith of the San Ignacio Paragneiss was derived from that the Lomas Maneches suite was probably emplaced
a Palaeoproterozoic source. These data further suggest just after the deposition of the Chiquitania Complex,
that the San Ignacio Group was not deposited prior to into which the granites intrude, but before the depo-
1690 Ma, the youngest age obtained from the popula- sition of the San Ignacio Group, where these granites
tion of detrital zircons. are unknown. The relatively rare inherited zircons from
the Lomas Maneches suite gave ages between 1715 and
1775 Ma, an interval that overlaps with the main age of
5. Synopsis of new data detrital zircons obtained from their host sediments.
Following the deposition of the protoliths of the
Zircon grains recovered from widely spaced out- Chiquitania Complex and San Ignacio Group and the
crops of the Chiquitania Complex (middle of the synchronous emplacement of the Lomas Manaches
sequence) and the San Ignacio Group (top) both con- suite, there was a ca. 300 million years period of
tain a dominant population of detrital zircons with an geologic quiescence until the San Ignacio Orogeny.
age of ca. 1765 Ma (1765 ± 12—Chiquitania Complex, This event was recorded regionally in the study area
1764 ± 6 Ma—San Ignacio Group). Zircons from both and resulted in the partial melting of the Chiquitania
sedimentary packages also yield detrital zircon ages no and Lomas Maneches complexes and, at higher struc-
older than ca. 1910 Ma, or younger than ca. 1690 Ma. tural levels, the widespread emplacement of layer-
The latter age is taken to provide a maximum age parallel San Ignacio granitoids near the base of the
of deposition for both sedimentary sequences. Col- San Ignacio Group. This is inferred from the con-
lectively, these data suggest that both the Chiquitania temporaneous emplacement of the San Rafael Granite
Complex and the San Ignacio Group were shed from (1337 ± 12 Ma) and growth of metamorphic zircon in
the same Palaeoproterozoic source region during what the Chiquitania Complex (1333 ± 6 Ma) and the Lomas
was probably a single continuous period of sedimen- Maneches suite (1320 ± 11). Structures attributable to
tation. The intrusion of the Lomas Maneches suite is D1 –D2 deformation developed during this interval,
140 S.D. Boger et al. / Precambrian Research 139 (2005) 121–146

while the similarity in metamorphic grade between D2 located to the east of the study area. It is unclear if these
and D3 (both were associated with leucosome forma- reflect the time of deformation in this region or reflect
tion), suggest D3 upright folds also occurred at about cooling from the Sunsas Orogeny. The latter inference
this time. We imply that these structures, which we col- is supported by the observation that the Sunsas, Nova
lectively attribute to the San Ignacio Orogeny, were the Brasilândia and Aguapéi regions all record similarly
result W–NW directed shortening based on the N–NE directed NE–SW tectonics, while Rb–Sr, K–Ar, and
trend of the D3 folds. hornblende and biotite 40 Ar/39 Ar ages from both the
In contrast to previous studies (e.g. Berrrangé and Sunsas and Nova Brasilândia regions (Litherland et
Litherland, 1982; Litherland et al., 1989) we attribute al., 1989; Tohver et al., 2004b) overlap in age with
the subsequent development of NW to W trending folds the K–Ar ages reported from the Aguapéi belt. In both
(F4 ) entirely to the Sunsas Orogeny. This is consistent of these regions the Rb–Sr and K–Ar systems close
with the observation that a single generation of NW approximately 100 million years after orogenesis as
trending folds are found within the Sunsas Group and, constrained by U–Pb data from a variety of minerals.
within both the Sunsas Group and the basement these
folds developed at relatively low metamorphic grade.
Sunsas deformation was most intense along the south- 6. The Paragua Block—an allochthonous
ern and western margins of the study area (Sunsas belt) microcontinent?
where it resulted in the almost complete overprinting
of the San Ignacio structures (e.g. San Javier region). The U–Pb zircon age data presented above suggests
This region also defines the locus of Sunsas plutonism that the Proterozoic rocks of Eastern Bolivia evolved
(Fig. 3). To the north and east of this region, there is in four distinct geologic stages. These are: (1) the
a marked decrease in the thermal and structural over- Palaeoproterozoic deposition of the protoloiths to the
print related to the Sunsas Orogeny. In the San Ignacio Chiquitania Complex and San Ignacio Group, and the
region, we attribute the distinct type-2 fold interfer- emplacement of Lomas Maneches suite (≤1690 Ma)
ence patterns to a NW-trending Sunsas overprint of the that collectively define the basement rocks of Eastern
N–NE trending San Ignacio structures. Slightly to the Bolivia; (2) the deformation and mid- to high-grade
north of here, the Rio-Negro front delineates the tran- metamorphism of these rocks during the San Ignacio
sition from K–Ar ages of ca. 1000 Ma to those of ca. Orogeny (1340–1320 Ma); (3) the deposition of the
1300 Ma (Litherland et al., 1989) suggesting the ther- Sunsas group unconformably above the basement rocks
mal effects of Sunsas orogenesis wane away from the and; (4) the relatively low-grade folding of the entire
main zone of Sunsas deformation in the southwest of sequence during the Sunsas Orogeny (1110–1070 Ma).
the study area. This four-stage evolution differentiates this area from
The age of Sunsas deformation is constrained to pre- the adjacent regions of Rondônia and Mato Grosso
date 1076 ± 18 Ma, the age obtained from the late-syn located to the north and east (Fig. 13).
to post-orogenic Taperas granite. This age is strikingly To the north of the study area, the Rondônia region
similar the ages obtained from the Nova Brasilândia (Fig. 2) preserves a complicated polycyclic history
region the immediate north (Fig. 2) where Rizzotto (Fig. 13). The oldest rocks crop out in the Rio Negro-
(1999) and Luft et al. (2000) present ages from syn- Juruena Province a belt of Palaeoproterozoic rocks
tectonic intrusions dated at ca. 1110 Ma and Tohver dated between 1750 and 1700 Ma (Tassinari et al.,
et al. (2004b) obtained ages between 1095–1080 and 1996; Payolla et al., 2002). The subsequent evolution
1085–1020 Ma from metamorphic monazite and titan- of this area consists of multiple phases of orogene-
ite. Together with the emplacement age of the Taperas sis and synchronous intrusion of granitoids into both
granite (1076 ± 18 Ma), these data suggest the Sun- deforming and cratonised crust. Major events occurred
sas Orogeny occurred over an approximate 40 million between 1600 and 1530 Ma, 1440 and1420 Ma and
years interval between 1110 and 1070 Ma and was again between 1330 and 1300 Ma (Tassinari and
widely manifested in both Eastern Bolivia and West- Macambira, 1999; Payolla et al., 2002; Tohver et al.,
ern Brazil. Post-Sunsas K–Ar ages of 960–920 Ma are 2005a). An almost continuous emplacement of gran-
recorded from the Aguapéi belt (Geraldes et al., 1997) itoids between 1420 and 1310 Ma is also described
S.D. Boger et al. / Precambrian Research 139 (2005) 121–146 141

Fig. 13. Time–space diagram of the Proterozoic intrusive, sedimentary and orogenic events from the south-western margin of South America.
Data is from this study and references cited in the text.

(Bettencourt et al., 1999; Payolla et al., 2002). The Eastern Bolivia. This belt cross-cuts the structural trend
youngest ductile period of deformation recognised in of pre-Sunsas structures in Rondônia (Tohver et al.,
Rondônia resulted in a 200 km wide NNW trending 2004b) and marks the boundary between the Amazon
network of sinistral shear zones collectively called the (exposed in Rondônia) and Paragua (exposed in Eastern
Ji-Paraná shear zone (Scandolâra et al., 1999; Tohver et Bolivia) cratons. The Nova Brasilândia belt is inter-
al., 2005a). This event, dated at 1200–1150 Ma, over- preted to define a Sunsas aged suture by Tohver et al.
prints ca. 1350 Ma crust (Tohver et al., 2005a, 2005b). (2004b).
When one compares the history from Rondônia to An equally significant mismatch in geologic his-
that from Eastern Bolivia, it is clear that the base- tories is observed between the Mato Grosso region
ment rocks from Rondônia (1750–1700 Ma) predate of Brazil and the rocks described here from East-
and evolved in a different tectonic setting (mag- ern Bolivia. Lying to the immediate east of the
matic arc) when compared to the sedimentary environ- study area (Fig. 2), the rocks in the Mato Grosso
ment in which the basement rocks of Eastern Bolivia region records evidence of 1790–1740, 1570–1520 and
were deposited/emplaced between 1690 and 1660 Ma. 1480–1420 Ma arc magmatism (Tassinari et al., 2000;
Although the 1340–1320 Ma San Ignacio potentially Geraldes et al., 2001; Pinho et al., 2003). These events
correlates with equivalent aged deformation and meta- young westward and, respectively, define the Alto Juaru
morphism in Rondônia, the polycyclic intrusion of terrane and the Cachoeirinha and St Helena Orogenies
granitoids and high-grade events that span the Late (Fig. 13). Similar to Rondônia, these events have no
Palaeoproterozoic and Early to Middle Mesoprotero- equivalents in Eastern Bolivia. Thus, similar to the
zoic in Rondônia (1600–1150 Ma), is not replicated Nova Brasilândia belt, the Aguapéi Thrust belt and
in Eastern Bolivia. Similarly the Ji-Paraná shear zone the Guaporé River valley (Fig. 2) define an abrupt
(1200–1150 Ma) is absent in Eastern Bolivia. These chronologic boundary between the outward younging
structures are not reported from outcrop geology, arc terranes of Mato Grosso and an older package of
nor were such distinct NNW trending linear features metasedimentary and igneous rocks that crop out in
observed geophysically. The east–west trending Nova Eastern Bolivia.
Brasilândia belt separates the rocks that record the With this in mind, we conclude that the rocks
above-described evolution of Rondônia, from those of of Eastern Bolivia (the Paragua Craton) were
142 S.D. Boger et al. / Precambrian Research 139 (2005) 121–146

allochthonous with respect to both the Rondônia and 6.2. Sunsas accretion of the Paragua Craton (1)
Mato Grosso regions of western Brazil. If this infer-
ence is correct, the important question then becomes In a recent paper Tohver et al. (2004b) argued that
how and when did the Paragua Craton accrete to the Nova Brasilândia belt represents the Sunsas aged
the southwest margin of Amazonia? The two possi- suture between the Paragua and Amazon cratons.
bilities are during the San Ignacio (1340–1320 Ma) Tohver et al. (2004b) correlated the Sunsas aged
or Sunsas (1100–1070 Ma) orogenies. Within a Nova Brasilândia belt with a continuous magnetic
Sunsas model, two configurations can be envisaged anomaly that divides the Amazon Craton into northern
depending on the significance attributed to the and southern halves, in effect implying that both the
Aguapéi Thrust belt. These possibilities are discussed Paragua Craton and the Mato Grosso region were
below. separate from the Rondônia region prior to Sunsas oro-
genesis (Fig. 14b). If correct, these two regions, with
6.1. San Ignacio accretion of the Paragua Craton their distinct geologic histories, need to have formed a
single microcontinent prior to Sunsas collision. Similar
If one considers the San Ignacio option first to the scenario presented above, this could arguably
(Fig. 14a), the age of this event (1340–1320 Ma) is have been achieved via San Ignacio aged collision
consistent with the progressive westward younging of between the Paragua Craton and the Mato Grosso
events observed in the Amazon Craton, while the accre- region. The Sunsas and Aguapeı́ groups would then
tion of an exotic terrane at this time would explain the represent a within-plate rift (Saes et al., 1992) formed
termination of slightly older (ca. 1420 Ma) arc magma- along the axis of the San Ignacio aged suture and be
tism along the margin of Amazonia. Although no defor- unrelated to the similar aged but lithologically distinct
mation of definitive San Ignacio age is described in deposits of the Nova Brasilândia Group (Tohver et
Mato Grosso, a possible continuation of the San Ignacio al., 2004b). In this scenario, Sunsas aged orogene-
belt is observed in Rondônia where 1330–1300 Ma oro- sis recognised along Nova Brasilândia and Sunsas
genesis has been reported (Payolla et al., 2002; Tohver belts would represent the suture between combined
et al., 2005a). Arguably, the 1400–1300 Ma anorogenic Paragua-Mato Grosso craton and the Amazon Craton
plutons recognised in Rondônia (Bettencourt et al., on the one hand, and the suture between Paragua-Mato
1999) could also represent the inboard manifestation of Grosso craton and the Arequipa-Antofalla Basement
San Ignacio collision of the Paragua Craton. This sce- (Loewy et al., 2004) on the other. The Aguapeı́ belt
nario would imply a within-plate origin for the Sunsas, would then represent a zone of within plate reactivation
Nova Brazilandia and Aguapeı́ groups. The Sunsas and associated with Sunsas accretion along the northern
Aguapeı́ groups consist mostly of terrestrial or shallow and southwestern margins of the Paragua Craton
marine deposits and are implied by Saes et al. (1992) (Fig. 14b).
to have been deposited in an intra-plate rift. In this set-
ting the turbiditic sediments of the Nova Brasilândia 6.3. Sunsas accretion of the Paragua Craton (2)
(Rizzotto, 1999; Luft et al., 2000) are perhaps more
difficult to explain, although could simply reflect depo- The alternate scenario involving Sunsas aged accre-
sition of these rocks in the rift depocenter. Assuming tion of the Paragua Craton involves Sunsas aged col-
the suturing of the Paragua and Amazon cratons was lision along a continuous Nova Brasilândia-Aguapeı́
achieved during the San Ignacio Orogeny, Sunsas oro- belt. This alternative implies that the Rondônia and
genesis in the Aguapeı́ and the Nova Brasilândia belts Mato Grosso Provinces were continuous within the
would then reflect zones of intense within-plate short- Amazon Craton (Fig. 14c), an inference that is sup-
ening localised along the axis of an intraplate Aguapeı́- ported by the good correlation in the geologic histories
Nova Brasilândia rift (Fig. 14a). Deformation at this preserved in both of these areas (Fig. 13). A model
time would presumably have been driven by the colli- implying that the Aguapeı́ and Nova Brasilândia belts
sion and accretion of the Arequipa-Antofalla Basement define part of a continuous suture would imply a com-
(AAB) to the western margin of the Paragua Craton mon link between the sediments that define these belts.
(Loewy et al., 2004). Although, Tohver et al. (2004b) argued against this cor-
S.D. Boger et al. / Precambrian Research 139 (2005) 121–146 143

Fig. 14. Cartoon of the three possible scenarios for the accretion of the Paragua Block and the Arequipa-Antofalla Basement to the western
margin of the Amazon Craton.

relation on the basis of differing depositional settings, sistent with a model implying Sunsas suturing along
the youngest detrital zircons from both the Aguapeı́ a continuous Nova Brasilândia-Aguapeı́ belt is the
and Nova Brasilândia sequences yield within error observation that the San Ignacio structures observed
ages of ca. 1220 Ma (Santos et al., 2000) suggesting within the Paragua Craton show a markedly differ-
both sequences were deposited at much the same time ent trend to that in the basement rocks of Rondônia
and eroded from potentially the same source. Con- or Mato Grosso (Fig. 14c). This suggests that the
144 S.D. Boger et al. / Precambrian Research 139 (2005) 121–146

Paragua Block was accreted with an established San and Macambira, 1999; Tassinari et al., 2000; Santos
Igancio structural grain. This argues against the first et al., 2000). An approximate 100 million years hiatus
two scenarios described above which imply San Igna- then appears to have separated the youngest of these
cio orogenesis should form a continuous belt with the events from the development of the Ji-Paraná network
similarly aged events recognised in Rondônia, or repre- of shear zones (Tohver et al., 2005a, 2005b). The onset
sent a westward continuation of orogenesis recognised of the Sunsas Orogeny at ca. 1110 Ma marked a new,
in the Mato Grossso. The absence of the Ji-Paraná but relatively short-lived period of accretion along the
shear zone (Scandolâra et al., 1999; Tohver et al., western margin of Amazonia. In contrast to the preced-
2005b) within the Paragua Craton further questions the ing periods of orogenesis, the Sunsas Orogeny resulted
continuity of structures across the Nova Brasilândia in the growth of the Amazon Craton by the accre-
belt. The absence of San Ignacio aged deformation in tion of two allochthonous lithospheric blocks—the
the Mato Grosso region (Geraldes et al., 2001) simi- Paragua Craton and the Arequipa-Antofalla Basement
larly questions the continuity of structures across the (Litherland et al., 1989; Loewy et al., 2004; Tohver et
Aguapeı́ belt. Both the Nova Brasilândia and Aguapeı́ al., 2004b). Both of these regions preserve an alternate
belts define zones of Sunsas high-strain and form dis- geologic history than the margin onto which they
tinct isotopic/structural boundaries that in this sce- accreted (Fig. 13) and show structural trends that are
nario define the northern and eastern margins of the at an angle to the NW trending palaeo-continental
Paragua Craton. Similar to the previous two models, margin of Amazonia (Fig. 14c). We suggest that the
the Sunsas belt would again represent the Sunsas-aged majority of Sunsas deformation, metamorphism, and
suture between the Arequipa-Antofalla Basement and plutonism was localised along the Paragua-Amazonia
the Paragua Craton. and Paragua-AAB sutures located along the northern
Nevertheless, this third model does not provide and eastern, and southern and western, margins
a panacea for interpreting the Mesoproterozoic evo- of the Paragua Craton. Only mild structural and
lution of the sothwestern margin of Amazonia as it isotopic reworking was recorded between these
does not account for the linear magnetic anomaly that belts.
underlies the Nova Brasilândia belt and which appears Post-Sunsas intrusive activity occurred between
to divide the Mato Grosso and Paragua regions from 1000 and 970 Ma and is recorded in the Negro-
the remainder of the Amazon Craton (Tohver et al., Juruena and the Novas Brasilândia regions of Rondônia
2004b). Arguably this feature could be unrelated to (Bettencourt et al., 1999; Luft et al., 2000) and in the
Sunsas orogenesis (i.e. Neoproterozoic or younger). Arequipa-Antofalla Basement (Wasteneys et al., 1995;
However, without definitive age data from this feature, Wörner et al., 2000; Loewy et al., 2004). In Rondônia,
this inference is no more or less valid that that proposed the Younger Granites of Rondônia (Bettencourt et
by Tohver et al. (2004b). Thus, we suggest that without al., 1999) cross-cut elements of the Sunsas Orogeny
being able to demonstrate a definitive separation or and form a north–north-east trending belt of intru-
connection between the Paragua Craton and the sions that lie at a high-angle to the northwest trending
Mato Grosso and/or Rondônia regions in the interval palaeo-margin of Amazonia (Bettencourt et al., 1999;
between the San Ignacio and Sunsas orogenies, all Luft et al., 2000). Although interpreted to represent
of the models described above remain plausible. the distal magmatic effects of the Sunsas collision
Nevertheless, we prefer the interpretation that the (Bettencourt et al., 1999), the Younger Granites of
Paragua Block was accreted to the margin of a coherent Rondônia (<1000 Ma) appreciably post-date the end
Amazon Craton during the Sunsas Orogeny (Fig. 14c) of the Sunsas Orogen (ca. 1075 Ma) as defined by this
as this scenario seems most consistent with the existing study. This period of plutonism does coincide with the
data. closure of the Rb–Sr and K–Ar systems in the Sunsas,
We thus envisage the evolution of the western mar- Aguapeı́ and Nova Brasilândia regions, although intru-
gin of the Amazon Craton to have evolved by the con- sions of this age appear to be absent from the Paragua
tinuous or semi-continuous addition of new continental Craton. We consider the emplacement of these rocks to
crust via successive periods of arc magmatism from be unrelated to the accretionary growth of the Amazon
ca. 2000 to 1300 Ma (Teixeira et al., 1989; Tassinari Craton.
S.D. Boger et al. / Precambrian Research 139 (2005) 121–146 145

Acknowledgements Dalziel, I.W.D., 1992. Antarctica: a tale of two supercontinents.


Annu. Rev. Earth Planet. Sci. 20, 501–526.
Darbyshire, D.P.F., Fletcher, C.J.N., 1979. A Mesozoic alkaline
The fieldwork for this project was completed with
province in Eastern Bolivia. Geology 7, 545–548.
the assistance of BHP Billiton, who provided financial Geraldes, M.C., Figueiredo, B.R., Tassinari, C.C.G., Erert, H.D.,
support and access to proprietary geophysical datasets. 1997. Middle Proterozoic vein-hosted gold deposits in the Pontes
BHP Billiton staff also gave invaluable assistance from e Lacerda region, southwestern Amazonian Craton, Brazil. Int.
both the Santiago (Chile) and Santa Cruz (Bolivia) Geol. Rev. 39, 438–448.
Geraldes, M.C., van Schmus, W.R., Condie, K.C., Bell, S., Teixeira,
offices. To these people the authors would like to
W., Babinski, M., 2001. Proterozoic geologic evolution of the SW
extend their sincere thanks. Data collection and anal- part of the Amazonian Craton in the Mato Grosso state, Brazil.
ysis, and the development of the concepts presented Precambrian Res. 111, 91–128.
here, occurred between 2002 and 2003 during which Hoffman, P.F., 1991. Did the breakout of Laurentia turn Gondwana-
time Boger and Raetz were supported by an Australian land inside out? Science 252, 1409–1412.
Karlstrom, K.E., Åhäll, K.-I., Harlan, S.S., Williams, M.L., McLel-
Research Council (ARC) Linkage grant awarded to
land, J., Geissman, J.W., 2001. Long-lived (1.8–1.0 Ga) conver-
Martin Krabbendam. Boger is currently supported by gent orogen in southern Laurentia, its extensions to Australia
an Australian Postdoctoral Fellowship awarded as part and Baltica, and implications for Rodinia. Precambrian Res. 111,
of an ARC Discovery Grant (DG0343406) to Dr’s 5–30.
Christopher Wilson and Ian Fitzsimons. Finally, the Litherland, M., 1984. Mappa Geologico Del Area Del Projecto Pre-
cambrio (1976–83), 1:1,000,000 map sheet. British Geological
authors would like to thank Dr’s Eric Tohver and
Survey, London.
Staci Loewy who provided thoughtful and construc- Litherland, M., Annells, R.N., Appleton, J.D., Berrrangé, J.P.,
tive reviews. These comments were of considerable Bloomfield, K., Burton, C.C.J., Darbyshire, D.P.F., Fletcher,
assistance in shaping the published version of this C.J.N., Hawkins, M.P., Klinck, B.A., Llanos, A., Mitchell, W.I.,
manuscript. O’Conner, E.A., Pitfield, P.E.J., Power, G., Webb, B.C., 1986.
The Geology and Mineral Resources of the Bolivian Precam-
brian Shield, 9. British Geological Survey, Overseas Memoir, p.
153.
References Litherland, M., Annells, R.N., Darbyshire, D.P.F., Fletcher, C.J.N.,
Hawkins, M.P., Klinck, B.A., Mitchell, W.I., O’Conner, E.A.,
Pitfield, P.E.J., Power, G., Webb, B.C., 1989. The Proterozoic of
Berrrangé, J.P., Litherland, M., 1982. Synopsis of the Geology and
Eastern Bolivia and its relationship to the Andean mobile belt.
Mineral Potential of the Proyecto Precámbrico Area. Servicio
Precambrian Res. 43, 157–174.
Geologico de Bolivia, Santa Cruz.
Litherland, M., Bloomfield, K., 1981. The Proterozoic history of
Berry, R.F., Jenner, G.A., Meffre, S., Tubrett, M.N., 2001. A North
Eastern Bolivia. Precambrian Res. 15, 157–179.
American provenance for Neoproterozoic to Cambrian sand-
Loewy, S.L., Connelly, J.N., Dalziel, I.W.D., 2004. An orphaned
stones in Tasmania. Earth Planet. Sci. Lett. 192, 207–222.
basement block: the Arequipa-Antofalla Basement of the central
Bettencourt, J.S., Tosdal, R.M., Leite, W.B., Payolla, B.L., 1999.
Andean margin of South America. Geol. Soc. Am. Bull. 116,
Mesoproterozoic rapakivi granites of the Rondônia Tin Province,
171–187.
southwestern border of the Amazonian craton, Brazil—I. Recon-
Loewy, S.L., Connelly, J.N., Dalziel, I.W.D., Gower, C.F.,
naissance U–Pb geochronology and regional implications. Pre-
2003. Eastern Laurentia in Rodinia: constraints from whole-
cambrian Res. 95, 41–67.
rock Pb and U/Pb geochronology. Tectonophysics 375, 169–
Bond, G.C., Nickeson, P.A., Kominz, M.A., 1984. Breakup of a
197.
supercontinent between 625–555 Ma: new evidence and impli-
Ludwig, K., 2001. Squid 1.02: A Users Manual. Berkeley
cations for continental histories. Earth Planet. Sci. Lett. 70,
Geochronology Centre, Special Publication No. 2.
325–345.
Luft, J.L., Rizzotto, G., Chemale, F., de Lima, E.F., 2000.
Borg, S.G., DePaolo, D.J., 1994. Laurentia, Australia, and Antarctica
Análise Geológico-Estrutural do Cinturão Sunsás na Região de
as a Late Proterozoic supercontinent: constraints from isotope
Brasilándia—Sudeste de Rondônia. Pesquisas em Geociências
mapping. Geology 22, 307–310.
27, 65–78.
Brookfield, M.E., 1993. Neoproterozoic Laurentia-Australia fit.
Moores, E.M., 1991. Southwest U.S.–East Antarctica (SWEAT) con-
Geology 21, 683–686.
nection: a hypothesis. Geology 19, 425–428.
Burrett, C., Berry, R., 2000. Proterozoic Australia–western United
Payolla, B.L., Bettencourt, J.S., Kozuch, M., Leite, W.B., Fetter,
States (AUSWUS) fit between Laurentia and Australia. Geology
A.H., van Schmus, W.R., 2002. Geological evolution of the base-
28, 103–106.
ment rocks in the east-central part of Rondônia Tin Province,
Dalziel, I.W.D., 1991. Pacific margins of Laurentia and East
SW Amazonian craton, Brazil: U–Pb and Sm–Nd isotopic con-
Antarctica–Australia as a conjugate rift pair: evidence and impli-
straints. Precambrian Res. 119, 141–169.
cations for an Eocambrian supercontinent. Geology 19, 598–601.
146 S.D. Boger et al. / Precambrian Research 139 (2005) 121–146

Pinho, M.A.S.B., Chemale, F., van Schmus, W.R., Pinho, F.E.C., Teixeira, W., Tassinari, C.C.G., Cordani, U.G., Kawashita, K., 1989.
2003. U–Pb and Sm–Nd evidence for 1.76–1.77 Ga magma- A review of the geochronology of the Amazonian craton: tectonic
tism in the Moriru region, Mato Grosso, Brazil: implications for implications. Precambrian Res. 42, 213–227.
province boundaries in the SW Amazon Craton. Precambrian Tohver, E., van der Pluijm, B.A., Mezger, K., Scandolara, J.E.,
Res. 126, 1–25. Essene, E., 2005a. Two stage history of the SW Amazon Craton
Prendergast, M.D., 2000. Layering and precious metals mineraliza- in the late Mesoproterozoic: identifying a cryptic suture zone.
tion in the Rincón del Tigre Complex, Eastern Bolivia. Econ. Precambrian Res. 137, 35–59.
Geol. 95, 113–130. Tohver, E., van der Pluijm, B.A., Scandolara, J.E., Essene, E., 2005b.
Rizzotto, G.J., 1999. Petrologia e Geotectônia do Grupo Novo Late Mesoproterozoic deformation of SW Amazonia (Rondônia,
Brazilándia, Rondônia. M.Sc. thesis, Univ. Fed. Rio Grande do Brazil): geochronological and structural evidence for collision
Sul, Porto Allegre, Brazil, 131 pp. with southern Laurentia. J. Geol. 113, 309–324.
Sadowski, G.R., Bettencourt, J.S., 1996. Mesoproterozoic tectonic Tohver, E., Bettencourt, J.S., Tosdal, R., Mezger, K., Leite,
correlations between eastern Laurentia and the western border of W.B., Payolla, B.L., 2004a. Terrane transfer during the
the Amazon Craton. Precambrian Res. 76, 213–227. Grenville orogeny: tracing the Amazonian ancestry of southern
Saes, G.S., Leite, J.A.D., Alvarenga, C.J.S., 1992. Evolução tectono- Appalachian basement through Pb and Nd isotopes. Earth Planet.
sedimentar do Grupo Aguapeı́, Proterozóico Médio na porçã Sci. Lett. 228, 161–176.
meridional do Cráton Amazônico: Mato Grosso e oriente Boli- Tohver, E., van der Pluijm, B.A., Mezger, K., Essene, E., Scandolara,
viano. Revista Brasileira de Geociências 23, 31–37. J.E., 2004a. Significance of the Nova Brasilândia metased-
Santos, J.O.S., Hartmann, L.A., Gaudette, H.E., Groves, D.I., imentary belt in western Brazil: Refining the Mesoprotero-
McNaughton, N.J., Flecher, I.R., 2000. A new understanding of zoic boundary of the Amazon Craton. Tectonics, TC6004,
the provinces of the Amazon Craton based on integration of field doi:10.1029/2003TC001563.
mapping and U–Pb and Sm–Nd geochronology. Gondwana Res. Tohver, E., van der Pluijm, B.A., van der Voo, R., Rizzotto, G.,
3, 489–506. Scandolara, J.E., 2002. Paleogeography of the Amazon Craton
Scandolâra, J.E., Rizzotto, G.J., Amorim, J.L., Bahia, R., Quadros, at 1.2 Ga: early Grenvillian collision with the Llano segment of
M.L., da Silva, C.R., 1999. Geological Map of Rondônia, Scale Laurentia. Earth Planet. Sci. Lett. 199, 185–200.
1:1,000,000. Companhia de Pesquisa de Recursos Minerais, Rio Wasteneys, H.A., Clark, A.H., Farrar, E., Langridge, R.J., 1995.
de Janeiro. Grenvillian granulite facies metamorphism in the Arequipa Mas-
Shackleton, R.M., Ries, A.C., Coward, M.P., Cobbold, P.R., 1979. sif, Peru: a Laurentia-Gondwana link. Earth Planet. Sci. Lett. 132,
Structure, metamorphism and geochronology of the Araquipa 63–73.
Massif of coastal Peru. J. Geol. Soc. (Lond.) 136, 195–214. Williams, I.S., 2001. Response of detrital zircon and monazite, and
Tassinari, C.C.G., Bettencourt, J.S., Geraldes, M.C., Macambira, their U–Pb isotopic systems, to regional metamorphism and host-
M.J.B., Lafon, J.M., 2000. The Amazon Craton. In: Cordani, rock partial melting, Cooma Complex, southeastern Australia.
U.G., Milani, E.J., Thomas Filho, A., Campos, D.A. (Eds.), Tec- Aust. J. Earth Sci. 48, 557–580.
tonic Evolution of South America. 31st International Geological Wingate, M.T.D., Pisarevsky, S.A., Evans, D.A.D., 2002. Rodinia
Congress, Rio de Janeiro, pp. 41–96. connections between Australia and Laurentia: no SWEAT, no
Tassinari, C.C.G., Macambira, M.J.B., 1999. Geochronological AUSWUS? Terra Nova 14, 121–128.
provinces of the Amazonian Craton. Episodes 22, 174–182. Wörner, G., Lezaun, J., Beck, A., Heber, V., Lucassen, F., Zin-
Tassinari, C.C.G., Cordani, U.G., Nutman, A.P., Van Schmus, W.R., ngrebe, E., Rössling, R., Wilke, H.G., 2000. Precambrian and
Bettencourt, J.S., Taylor, P.N., 1996. Geochronological system- Early Palaeozoic evolution of the Andean basement at Belen
atics on basement rocks from the Rı́o Negro-Juruena (Amazonian (northern Chile) and Cerro Uyarani (western Bolivia Altiplano).
Craton) and tectonic implications. Int. Geol. Rev. 38, 161–175. J. South Am. Earth Sci. 13, 717–737.

You might also like