You are on page 1of 16

TERTIARY PROVENANCE HISTORY OF THE NORTHERN AND CENTRAL ALTIPLANO

(CENTRAL ANDES, BOLIVIA): A DETRITAL RECORD OF PLATEAU-MARGIN TECTONICS

B.K. HORTON1, B.A. HAMPTON 2, B.N. LAREAU 2, AND E. BALDELLÓN3


1 Department of Earth and Space Sciences, University of California, Los Angeles, California 90095-1567, U.S.A.
2 Department of Geology and Geophysics, Louisiana State University, Baton Rouge, Louisiana 70803, U.S.A.
3 Servicio Nacional de Geologı́a y Minerı́a (SERGEOMIN), Casilla 2729, 1673 Calle Federico Zuazo, La Paz, Bolivia

e-mail: horton@ess.ucla.edu

ABSTRACT: The internally drained Altiplano plateau in the high-ele- complex regions of mountain belts. For continental hinterland plateaus,
vation hinterland of the central Andes contains an extremely thick (up such as the Tibetan plateau of central Asia and Altiplano–Puna plateau of
to 12 km) succession of Tertiary nonmarine strata that recorded the the central Andes (the two largest elevated regions on Earth), provenance
topographic evolution of eastern and western plateau margins. In this studies may help resolve questions on the timing and magnitude of defor-
study, temporal and spatial variations in detrital composition (based mation, magmatism, and erosion for both the margins and interiors of pla-
on 113 sandstone thin sections and 31 conglomerate clast counts), com- teaus. Such information may ultimately prove beneficial in constraining the
bined with sediment dispersal patterns and regional stratigraphic cor- mechanisms of surface uplift in orogenic plateaus, an issue fundamental to
relations, help delineate the Tertiary history of Altiplano basin devel- understanding mountain building in Tibet (Yin and Harrison 2000), the
opment in relationship to thrust deformation and arc magmatism along Andes (Isacks 1988), and possible ancient plateau systems (Russo and Sil-
plateau margins. ver 1996; Chase et al. 1998).
Stratigraphic and provenance data suggest limited subsidence and Although compositional analyses of hinterland plateaus are limited, the
west-directed sediment dispersal during the early Tertiary history of utility of provenance studies for the interiors of other orogenic systems has
the Altiplano region. By late Eocene time, development of a rapidly been well established. For example, Mesozoic–early Tertiary hinterland
subsiding basin (probable foredeep) was underway, as recorded by ex- basins in the Canadian and Alaskan Cordillera provide accurate records of
tensive east-directed fluvial systems of the 3–6.5-km-thick Potoco For- Cordilleran thrust-belt development and episodes of terrane accretion along
mation. Potoco compositional data are compatible with sediment sourc- the western margin of North America (Eisbacher et al. 1974; Ricketts et
es in the incipient Western Cordillera and regions farther west, in- al. 1992; Trop and Ridgway 1997; Johnsson 2000). Similarly, analyses of
cluding (1) a probable thrust-belt source terrane composed of Paleo- large continental basins in central Asia north of the Tibetan plateau have
zoic–Mesozoic strata and Precambrian granitic/gneissic basement and defined compositional records of Mesozoic through Cenozoic collisional
(2) a subordinate magmatic-arc source terrane composed of Tertiary mountain building and associated unroofing of diverse accreted terranes
volcanic rocks. Increased abundances of plagioclase grains and volca- (Graham et al. 1993; Hendrix 2000; Shao et al. 2001). These studies sug-
nic lithic fragments upsection indicate a progressively greater arc in- gest, by analogy, that provenance investigations aimed at orogenic plateaus
fluence, due to either enhanced volcanism or eastward arc migration. may offer insight into the tectonic and sedimentary processes related to
By late Oligocene time, west-directed fluvial and alluvial-fan systems plateau development.
of the upper Potoco Formation and the ;3-km-thick Coniri Formation In the Andes, numerous sedimentary basins occur within the high-ele-
reached the Altiplano basin, providing a minimum uplift age for the vation interior of the orogenic belt. These basins include zones of modern
backthrust belt along the western flank of the Eastern Cordillera. Pa- sediment accumulation as well as exposed fill of older Tertiary basins. The
leozoic-clast conglomerate and relatively quartz-rich sandstone con- Altiplano basin of the Bolivian central Andes is the most extensive (roughly
taining sedimentary and metasedimentary lithic fragments show un- 150,000 km 2), contains the thickest fill (up to 12 km), and is among the
roofing of predominantly Paleozoic strata in the backthrust belt. Con- highest (3.7 km average elevation) of the Andean hinterland basins (Ahlfeld
tinued late Oligocene to Quaternary accumulation in the Altiplano re- and Branisa 1960; Martinez 1980; Sempere et al. 1990a; Sempere et al.
veals that the basin persisted as a depositional region during a 1990b; Allmendinger et al. 1997). Previous studies have documented the
well-documented period of Neogene shortening and uplift in the An- sedimentology, stratigraphy, and structure of several Andean hinterland ba-
dean thrust belt farther east. This intra-orogenic (hinterland) deposi- sins, including basins in Ecuador (Hungerbühler et al. 1995; Steinmann et
tion between eastern and western highs during ongoing shortening sug- al. 1999), Peru (Mégard et al. 1984; Marocco et al. 1995), and Argentina
gests that the Altiplano basin existed as a crustal-scale piggyback basin, (Vandervoort 1993; Vandervoort et al. 1995; Kraemer et al. 1999). How-
possibly in an internally drained plateau-like geomorphic setting, since ever, the detailed depositional history, sediment source areas, structural
;25 Ma. controls, and timing of uplift for the Altiplano remain poorly understood.
Provenance data and independent structural and thermochrono- The goal of this investigation is to shed light on the structural and mag-
logical evidence for spatially and temporally varying tectonic histories matic history of the Altiplano plateau margins on the basis of the Tertiary
along the eastern and western plateau margins imply that several provenance record, and in doing so, identify possible factors contributing
modes of basin development were active during growth of the Altipla- to the growth of the plateau and accumulation of an extremely thick Ter-
no. Although flexural subsidence due to thrust loading along both pla- tiary succession. We present here an analysis of the composition, stratig-
teau margins, initially in the west and subsequently in the east, may raphy, and paleocurrent patterns of Paleocene to lower Miocene strata in
explain most Tertiary subsidence, possible closed-basin conditions due the northern and central Altiplano. Our study illustrates the temporal and
to structural damming could have promoted additional sediment ac- compositional variability associated with basin evolution in a high-eleva-
cumulation from late Oligocene to Quaternary time. tion, orogenic plateau setting and may have implications for the tectono-
geomorphic conditions conducive to plateau development.

INTRODUCTION GEOLOGIC SETTING

Sedimentary basins exposed in orogenic interiors (hinterlands) can pro- The Altiplano occupies the hinterland of the Bolivian central Andes. It
vide records of erosional denudation in the highest and most structurally consists of a 200-km-wide, internally drained, low-relief region at 3.7 km

JOURNAL OF SEDIMENTARY RESEARCH, VOL. 72, NO. 5, SEPTEMBER, 2002, P. 711–726


Copyright q 2002, SEPM (Society for Sedimentary Geology) 1527-1404/02/072-711/$03.00
712 B.K. HORTON ET AL.

elevation between two mountain ranges bearing peaks over 6 km high during post-early Miocene growth of a crustal-scale monocline (Mortimer
(Isacks 1988). To the west is the narrow Western Cordillera, a modern 1981; Isacks 1988; Lamb et al. 1997; Wörner et al. 2000a). However, the
volcanic arc, and to the east is the extensive Eastern Cordillera, the uplifted greatest magnitude of deformation in the Precordillera and western slope
interior of the Andean thrust belt (Fig. 1). is primarily of Late Cretaceous–Paleogene age. During this time, thrusting
The Altiplano exhibits an exceptionally thick (up to 12 km) succession and related strike-slip faulting accommodated several tens of kilometers of
of Tertiary nonmarine strata (Ahlfeld and Branisa 1960; Martinez 1980; east–west shortening (Chong 1977; Chong and Reutter 1985; Hammer-
Sempere et al. 1990a; Sempere et al. 1990b; Allmendinger et al. 1997). schmidt et al. 1992; Hartley et al. 1992; Scheuber and Reutter 1992; Scheu-
Although numerous stratigraphic units have been identified (Evernden et ber et al. 1994; Kuhn and Reuther 1999; Arriagada et al. 2000).
al. 1977; Suárez and Diaz 1996; Lamb and Hoke 1997), the basic Upper On the opposite side of the Altiplano, the Eastern Cordillera forms a
Cretaceous through Cenozoic stratigraphy (Fig. 2) can be summarized brief- 200-km-wide zone of rugged topography with peak elevations up to 6.4
ly as follows (Horton et al. 2001): (1) Maastrichtian to mid-Paleocene shal- km (Fig. 1). A topographic divide near the western edge of the Eastern
low marine and lacustrine carbonate, and distal fluvial siltstone and sand- Cordillera separates the Altiplano drainage system from river networks that
stone (El Molino and Santa Lucia formations, 250–900 m thick); (2) mostly flow to the South American craton to the east. Paleozoic sedimentary and
upper Eocene through Oligocene fluvial sandstone (Potoco Formation, low-grade metasedimentary rocks, primarily of clastic origin, dominate the
3000–6500 m thick); and (3) upper Oligocene to Quaternary fluvial and Eastern Cordillera (Pareja et al. 1978; Roeder 1988; Reutter et al. 1994;
alluvial-fan conglomerate (upper Oligocene–lower Miocene Coniri For- Kley et al. 1997; McQuarrie and DeCelles 2001). Carbonates are rare and
mation) and overlying volcaniclastic section (numerous lower Miocene– no Precambrian basement is exposed. In places, upper Miocene ignimbrites,
Quaternary stratigraphic units) (1000–4000 m total thickness). Age assign- Neogene intrusions, and narrow synclinal belts of Cretaceous–Tertiary stra-
ments are based on magnetostratigraphy and mammal fossils from Upper ta overlap or intrude the Paleozoic section (e.g., Gubbels 1993; Lamb et
Cretaceous to mid-Paleocene El Molino and Santa Lucia strata (Marshall al. 1997; Horton and DeCelles 2001). The Eastern Cordillera is a distinctive
et al. 1997; Sempere et al. 1997), late Eocene–Oligocene palynological ages belt of oppositely verging fold–thrust structures (Roeder 1988; Roeder and
from the Potoco Formation (Horton et al. 2001), late Oligocene–Pliocene Chamberlain 1995; Kley et al. 1997; McQuarrie and DeCelles 2001). Along
40Ar/39Ar and K–Ar ages from interbedded volcanic rocks (Evernden et al. the western flank of the Eastern Cordillera, a west-vergent backthrust belt
1977; Lavenu et al. 1989; MacFadden et al. 1990; Marshall et al. 1992; accommodated 100–150 km of roughly mid-Tertiary shortening (Roeder
Kennan et al. 1995; MacFadden et al. 1995; Lamb and Hoke 1997; Kay and Chamberlain 1995; McQuarrie and DeCelles 2001). To the east, the
et al. 1998), and magnetostratigraphy of Oligocene–Miocene strata east-vergent Andean thrust belt, which includes the central to eastern sector
(MacFadden et al. 1990; MacFadden et al. 1995; Kay et al. 1998; Roperch of the Eastern Cordillera as well as the Subandean Zone farther east, ac-
et al. 1999). Available age constraints suggest that rocks of the upper Po- commodated 100–150 km of primarily Neogene shortening (Roeder 1988;
toco Formation and lower Coniri Formation are age-equivalent strata. This Sheffels 1990; Roeder and Chamberlain 1995; Kley 1996; Baby et al. 1997;
study focuses on the most complete exposures of the thick Paleocene to Kley et al. 1997).
lower Miocene succession represented by the Santa Lucia, Potoco, and On the basis of the above descriptions, possible sediment sources for the
Coniri formations. Altiplano basin include a western region composed of Precambrian base-
ment, Paleozoic–Mesozoic strata, and Tertiary igneous rocks, and an east-
POTENTIAL SEDIMENT SOURCE AREAS ern region composed of primarily Paleozoic strata with limited amounts of
Mesozoic strata and Neogene igneous rocks. It should be emphasized that
Possible sediment sources for the Altiplano basin include the narrow the structural and topographic expression of the modern plateau margins
Western Cordillera and Precordillera to the west and the extensive Eastern may be substantially different from their early to mid-Tertiary appearance.
Cordillera to the east (Fig. 1). The Western Cordillera is an 80-km-wide In addition, given the large extent of Neogene volcanic cover, the modern
mountain range forming a drainage divide along the Chile–Bolivia border. bulk-rock compositions at the surface of the Precordillera, the Western
The range stands 1–2 km above the surface of the Altiplano. Although Cordillera, and the Eastern Cordillera may not be accurate representations
Neogene–Quaternary volcanic rocks and stratovolcanos cover most of the of their early to mid-Tertiary composition.
Western Cordillera (Pareja et al. 1978; Isacks 1988; Marsh et al. 1992;
Reutter et al. 1994; Wörner et al. 2000a), several lines of evidence suggest METHODS
that Precambrian granitic and gneissic basement compose the bulk of the
range. Evidence includes xenoliths of granite and gneiss in Neogene ig- Our investigation centers on sandstone compositional data from petro-
neous rocks (Jiménez de Rı́os 1992), dated Precambrian granitic and graphic analyses (point counts), conglomerate compositional data from
gneissic clasts in middle to upper Tertiary strata (Evernden et al. 1977; clast counts, sedimentologic data derived from paleocurrent measurements
Jiménez de Rı́os 1992), dated Precambrian gneiss and granite from drill- and facies analyses, and regional stratigraphic correlations.
hole samples (Fig. 1) (Lehmann 1978; Tosdal 1996), and limited surface Compositional data were collected from ten stratigraphic sections mea-
exposures of dated Precambrian rocks beneath the volcanic cover (Fig. 1) sured within a nearly continuous belt of Paleocene–Miocene exposures in
(Troëng et al. 1994; Wörner et al. 2000b). Although these data provide the northern and central Altiplano (Fig. 3). For sandstones, a standard pet-
insight into the crustal composition of the Western Cordillera, the structural rographic thin section from each sample, primarily medium-grained sand-
relationships and pre-Neogene tectonic history of the region remain largely stone, was cut and stained for plagioclase and potassium feldspar. 113 thin
inaccessible beneath the Neogene–Quaternary volcanic cover. sections were analyzed using a modified Gazzi-Dickinson method of point
To the west of the Western Cordillera lies a 50-km-wide region at 2–4 counting (Gazzi 1966; Dickinson 1970; Ingersoll et al. 1984). At least 400–
km elevation called the Precordillera (Fig. 1). The Precordillera is com- 500 framework grains were identified per thin section. In cases of diage-
posed of Mesozoic arc-related sedimentary and igneous rocks, Paleozoic netic alteration (including compaction, partial alteration of feldspar to clay
strata, and Precambrian basement (Chong 1977; Chong and Reutter 1985; minerals, and cementation), the original framework grains were recogniz-
Hartley et al. 1992; Scheuber and Reutter 1992; Reutter et al. 1994). Farther able and identified as the original grain constituent. For these samples, lithic
west, Andean topography consists of a west-facing slope that, over a 50 fragments are represented primarily by volcanic grains (Fig. 4A), meta-
km distance, drops uniformly in elevation down to a 1-km-high forearc morphic and metasedimentary grains (Fig. 4B), mudrock and sandstone
basin known as the Longitudinal Valley, Central Depression, or Pampa del grains (Fig. 4C), and minor amounts of carbonate grains (Fig. 4D). Pet-
Tamuragal. The western Andean slope has been attributed to surface tilting rographic counting parameters (Table 1), raw point-count data (Appendix
TERTIARY PROVENANCE HISTORY OF THE ALTIPLANO, CENTRAL ANDES 713

FIG. 1.—Simplified geologic map of the


northern and central Altiplano (after Pareja et al.
1978) showing location of measured sections
within the northern (N), central (C), and
southern (S) study regions. Topographic margins
of the plateau are defined by the Western
Cordillera (along the Chile–Bolivia border) and
the Eastern Cordillera. Outcrop and drill-hole
locations of Precambrian basement after Wörner
et al. (2000b) and Lehmann (1978), respectively.
Large areas of Quaternary alluvium denote the
depositional surface of the modern Altiplano
basin. Stippled and shaded patterns represent
exposed Paleocene to lower Miocene strata of
the Altiplano basin. Inset map shows regional
setting within the central Andes, including
international borders (dashed lines), geological
zonations (labeled), elevated region over 3 km
high (shaded), and eastern mountain front of
Andean thrust belt (barbed line).

1, archived; see acknowledgments section), and recalculated detrital modes DeCelles et al. 1983) and clast imbrication in conglomerate, document
(Table 2) are based on methods previously defined by Ingersoll et al. (1984) changes in paleoflow within the measured stratigraphic sections.
and Dickinson (1985). Recalculated parameters are reported for Q-F-L, Stratigraphic correlations among the measured sections (Fig. 3) were
Qm-P-K, and Lv-Lm-Ls ternary diagrams. based on similar depositional ages, and, in the absence of age control,
Conglomerates are typically clast-supported, moderately organized beds diagnostic lithostratigraphic elements. The mid-Paleocene Santa Lucia For-
containing subangular to subrounded pebbles to boulders. Clast counts were mation (Marshall et al. 1997; Sempere et al. 1997) provides a lower marker
performed on the outcrop at a stratigraphic spacing of roughly 100–150 m. unit that can be traced over large parts of the Altiplano. Red siltstone and
Each of the 31 conglomerate clast counts involved lithologic identification minor sandstone of the Santa Lucia Formation overlie the carbonate-rich
of 100–125 clasts. Clast types include the following: granite; orthogneiss; El Molino Formation and underlie a distinctive paleosol zone reported by
basalt; welded tuff; several varieties of quartz-rich, fine- to medium-grained Horton et al. (2001) in the basal Potoco Formation (Figs. 2, 3). In areas
sandstone and quartzite; fossiliferous carbonate; laminated carbonate; mas- where the Santa Lucia is covered, lower correlations are based on either a
sive vein quartz; chert; massive mudrock; and pebble–cobble conglomerate. basal gypsum (Chuquichambi Member) of the Potoco Formation that yields
Sedimentological analyses, discussed in part by Horton et al. (2001), late Eocene palynomorph assemblages (Horton et al. 2001; Hampton 2002)
were conducted to determine sediment dispersal patterns and ancient de- or the lowest east-derived conglomeratic deposits of the Coniri Formation
positional systems for the Altiplano basin. Paleocurrent analyses, primarily (Sempere et al. 1990b; LaReau and Horton 2000). Regional correlations of
from measurements of trough-cross limbs in sandstone (method I of upper stratigraphic intervals are based on the stratigraphically lowest vol-
714 B.K. HORTON ET AL.

nian sandstone and quartzite are generally yellow to red (e.g., the Sica Sica
and Vila Vila formations), Silurian sandstone and quartzite are primarily
gray to green (e.g., Catavi and Llallagua formations) (Suárez and Diaz
1996).
On the basis of over 2500 paleocurrent measurements, it is clear that
major sediment sources existed to both the east and west. Lithofacies anal-
yses indicate depositional processes associated with fluvial and, to a lesser
degree, alluvial-fan and lacustrine sedimentation (LaReau and Horton 2000;
Horton et al. 2001; Hampton 2002).
Within the exposed belt of Paleocene–Miocene rocks, three geographic
regions of differing lithostratigraphy and detrital composition have been
identified and are herein referred to as the southern, central, and northern
study regions (Figs. 1, 3). The southern region contains a diagnostic pa-
leosol interval in the basal Potoco Formation; the central region is distin-
guished by the greatest exposed thickness of the Potoco Formation, which
occurs in the Corque syncline; and the northern region contains exposures
of both the Potoco Formation and conglomeratic Coniri Formation (Fig.
3).

Southern Region
Description.—Measured sections 1–3 (Julchi, Maicoma, and Andamarca
sites) all exhibit striking upsection variations in sandstone composition
from the lower Santa Lucia Formation to the upper Potoco Formation (Fig.
5). At the base, 11 Santa Lucia samples define a generally quartzofeld-
spathic composition (Fig. 5A). Monocrystalline quartz (Fig. 5B) is the dom-
FIG. 2.—Simplified Cenozoic stratigraphic framework for the northern and central
Altiplano. Ages are based on palynological assemblages, magnetostratigraphy, and inant grain type and lithic fragments are mainly of sedimentary or meta-
isotopic ages of tuffs (see text). Ruled area indicates zone of limited sediment ac- sedimentary origin (Fig. 5C). The 17 Potoco sandstones sampled upsection
cumulation and extensive paleosol development in the lowermost Potoco Formation are dramatically richer in feldspar and lithic fragments (Fig. 5A). For these
(after Horton et al. 2001). Dashed lines depict approximate boundaries between feldsarenites to feldspathic litharenites, plagioclase is the dominant feldspar
partially age-equivalent strata (upper Potoco, Coniri, and Neogene–Quaternary vol- (Fig. 5B) and volcanic grains represent the main lithic component (Fig.
caniclastic units). 5C).
Well-defined upsection trends are visible in all ternary diagrams (Fig.
5). The Q-F-L plots indicate an increase in feldspar for the lower Potoco
canic tuffs, which have been dated at 25–23 Ma (Lavenu et al. 1989; Ken- followed by an increase in lithic fragments for the middle to upper Potoco
nan et al. 1995; Horton et al. 2001). (Fig. 5A). These variations are best displayed in Section 1 (Julchi site),
where the most complete stratigraphic section is preserved (Fig. 3). A de-
PROVENANCE RESULTS viation from these trends in Section 1 is marked by abrupt increases in
feldspar, conspicuously potassium feldspar, and metamorphic lithic frag-
Results are presented for 13 sandstone thin sections from the Santa Lucia
ments for samples 10 and 11 (outlined areas in Fig. 5), which were taken
Formation (mid-Paleocene), 68 thin sections from west-derived deposits of
from the 1500–2500 m level (Fig. 3). In addition to Q-F-L trends, clear
the Potoco Formation (upper Eocene–Oligocene), and 32 thin sections and
upsection increases in plagioclase (Fig. 5B) and volcanic lithic fragments
31 clast counts from east-derived deposits of the upper Potoco and Coniri
(Fig. 5C) exist for all three sections.
formations (upper Oligocene–lower Miocene).
Interpretation.—Paleocurrent data for Sections 1–3 (Horton et al. 2001)
In thin section, several framework grains have notable attributes of po-
indicate an eastern source for the Santa Lucia Formation and a western
tential use in provenance identification, including: microlitic to lathwork
source for the overlying Potoco Formation (Fig. 3). From the abundance
volcanic fragments that show mainly intersertal texture with plagioclase
of Paleozoic sedimentary and low-grade metasedimentary strata in the East-
laths set in a brown glassy matrix (Lvm, Lvl) (Fig. 4A); phyllite and quartz-
ern Cordillera (Fig. 1), the observed quartzofeldspathic Santa Lucia com-
mica schist fragments displaying a moderate to well-developed planar fab-
positions are attributed to initial erosion of mostly Paleozoic rocks to the
ric defined by elongate (commonly polycrystalline) quartz, aligned mica,
east. In contrast, the feldspathic to lithofeldspathic Potoco sandstone com-
and in some cases, aligned hornblende (Lph, Lms, Qpt) (Fig. 4B); carbon-
positions are considered the product of sediment transport from a mixed
ate lithic fragments containing allochems of bioclasts and coated grains
basement and volcanic source terrane to the west in the incipient Western
(Lca) (Fig. 4D); and microcline and perthitic potassium feldspar grains
Cordillera and regions farther west. The abrupt increase in feldspar and
exhibiting tartan (grid) twinning and albite lamellae (Kmp).
metamorphic lithic fragments for samples 10 and 11 is tentatively attributed
Although most conglomerate clasts cannot be identified confidently at
to greater contribution from potassium-feldspar-bearing Precambrian base-
the formation level, granite and gneiss clasts can be attributed to Precam-
ment rocks to the west (Fig. 1). For upper stratigraphic levels, a volcanic-
brian basement, fossiliferous carbonate and chert clasts to Permo–Carbon-
arc sediment source is inferred based on the high amount of feldspar (most-
iferous units (e.g., Copacabana Formation), laminated carbonate clasts to
ly plagioclase) and lithic fragments (mostly volcanic fragments) (Fig. 5)
the Upper Cretaceous El Molino Formation, and mudrock clasts to the
(e.g., Dickinson 1985; Marsaglia and Ingersoll 1992).
Tertiary Potoco Formation. Clasts of conglomerate may be reworked from
the Tertiary Coniri Formation, the Carboniferous Ambo Group, or the Si-
Central Region
lurian Cancañiri Formation (Sempere 1995; Suárez and Diaz 1996; Mc-
Quarrie and DeCelles 2001). The majority of sandstone and quartzite clasts Description.—West-derived Potoco deposits of Sections 5 and 6 (Chu-
are of Paleozoic origin, mostly of Silurian–Devonian age. Whereas Devo- quichambi and Turco sites) display large compositional variations in the
TERTIARY PROVENANCE HISTORY OF THE ALTIPLANO, CENTRAL ANDES 715

FIG. 3.—Generalized stratigraphic sections of Paleocene–lower Miocene deposits from the northern, central, and southern study regions (locations are shown in Fig. 1).
Depicted features include thickness (km), basic lithology, paleocurrent data, and locations of sandstone point-count samples. Stratigraphic correlations are based on (1)
distinctive lithostratigraphic elements such as the top of the El Molino Formation (0 m level, Sections 1–4, 9), the top of the Santa Lucia Formation (Sections 1–4, 9), the
basal gypsum member of the Potoco Formation (Sections 5, 7, 8A), and basal conglomerate of the Coniri Formation (Sections 7, 8A, 8B), (2) palynological ages from the
Potoco Formation, and (3) published 25–23 Ma K–Ar ages for the stratigraphically lowest tuffs in overlying volcaniclastic deposits. See Horton et al. (2001) for facies
analyses and stratigraphic nomenclature.

west but limited variations in the east (Fig. 6). In the west, 22 samples In the east, distal equivalent strata from lower to intermediate Potoco
from Section 6 (Turco site, samples 51–72) show a diverse composition levels exhibit similar, but less pronounced, compositional trends. The 15
and a well-defined upsection trend from quartzolithic to more feldspathic quartzolithic, quartzofeldspathic, and lithic arkosic samples from the lower
sandstone (Fig. 6A). Basal sandstones in the west are rich in monocrystal- 4200 m of Section 5 (Chuquichambi site, samples 29–43) are rich in mono-
line quartz (Fig. 6A, B) and contain metamorphic and sedimentary lithic crystalline quartz (Fig. 6A, B) and contain subordinate lithic fragments of
fragments (Fig. 6C). Intermediate sandstones exhibit an increase in potas- metamorphic, metasedimentary, and volcanic origin (Fig. 6C). Although
sium feldspar (Fig. 6B) and volcanic lithic fragments (Fig. 6C). These this suite shows upsection increases in feldspar (Fig. 6A, B) and volcanic
intermediate feldsarenites and lithic arkoses are from a coarse-grained in- lithic fragments (Fig. 6C), the magnitude of these variations is substantially
terval (Fig. 3; locally named Azurita conglomerate; Evernden et al. 1977; less than observed to the west. Sediment dispersal patterns for the upper
Lamb and Hoke 1997) that also contains numerous cobble–boulder con- stratigraphic levels of the eastern area (Fig. 3) indicate an eastern, rather
glomerate beds dominated by clasts of granite and gneiss (Fig. 7). Upper than western, source; therefore, compositional data for those strata are dis-
sandstones are feldsarenites to feldspathic litharenites containing greater cussed below.
amounts of plagioclase, volcanic lithic fragments, and total lithic fragments Interpretation.—In the central region, paleocurrent data indicate an ex-
(Fig. 6). clusively western sediment source for the western Potoco section and lower
716 B.K. HORTON ET AL.

FIG. 4.—Photomicrographs of sandstones from the Altiplano basin. A) Lathwork volcanic lithic fragment (Lvl). B) Phyllite lithic fragment (Lph). C) Mudrock lithic
fragment (Lmu) and monocrystalline quartz (Qm). D) Carbonate lithic fragment (Lca). Photos A and B were taken in plane-polarized light, photos C and D with crossed
polarizers.

to intermediate levels of the eastern Potoco section (Fig. 3; Horton et al.


2001; Hampton 2002). The quartzolithic sandstones (with metamorphic and
TABLE 1.—Parameters for sandstone point counts. sedimentary lithic fragments) characterizing the lowermost Potoco levels
(Fig. 6) are attributed to erosion of Paleozoic–Mesozoic sedimentary or
Symbol Grain Categories
Recalculated
Parameters
metasedimentary rocks and Precambrian gneiss to the west. Increases in
total feldspar, potassium feldspar, and volcanic lithic fragments for the
Qm Monocrystalline quartz Q–F–L:
Qp Polycrystalline quartz Q 5 Qm 1 Qp 1 Qpt 1 C
lower to intermediate Potoco section (Fig. 6) suggest increased exposure
Qpt Polycrystalline quartz with tectonic fabric F 5 P 1 Ksp 1 Kmp of Precambrian crystalline basement and growth of a volcanic arc through
C Chert L 5 Lv 1 Lm 1 Ls time. These trends coincide with the appearance of conglomerates bearing
P Plagioclase feldspar
Ksp Potassium feldspar Qm–P–K: granite and gneiss clasts (Fig. 7) derived from Precambrian basement rocks
Kmp Microcline and perthitic potassium feldspar Qm 5 Qm exposed in the Western Cordillera. Increased volcanic lithic fragments in-
Lvl Lathwork volcanic lithic fragments P5P
Lvm Microlitic volcanic lithic fragments K 5 Ksp 1 Kmp dicate that the volcanic arc progressively became an important sediment
Lms Schistose metamorphic lithic fragments source. A dominantly volcanic-arc sediment source is interpreted for the
Lph Phyllite lithic fragments Lv–Lm–Ls: upper Potoco levels in the west, where sandstones have a high proportion
Lss Sandstone lithic fragments Lv 5 Lvl 1 Lvm
Lmu Mudrock lithic fragments Lm 5 Lms 1 Lph of feldspar (mostly plagioclase) and lithic fragments (mostly volcanic frag-
Lca Carbonate lithic fragments Ls 5 Lss 1 Lmu 1 Lca ments) (Fig. 6). The proximal-to-distal change in composition from west
A Accessory minerals (dense minerals and micas)
to east, shown by more quartz-rich compositions to the east (Fig. 6), may
TERTIARY PROVENANCE HISTORY OF THE ALTIPLANO, CENTRAL ANDES 717

be related to a combination of dilution by tributary drainage networks (e.g., rity, as shown by more abundant quartz and less abundant feldspar and
DeCelles and Hertel 1989; Ingersoll et al. 1993) and rapid downstream lithic fragments, particularly volcanic fragments (Fig. 9). A general lack of
attrition by physical and chemical weathering of feldspar and lithic frag- volcanic lithic fragments further suggests that Neogene igneous rocks ob-
ments in either a tropical (e.g., Johnsson 1990) or an arid (e.g., Robinson served in the Eastern Cordillera were not emplaced or exposed at the sur-
and Johnsson 1997) climate. face by early Miocene time.

SUMMARY OF SEDIMENT SOURCE AREAS


Northern Region
Provenance data indicate three principal sediment source areas during
Description.—Sections 7, 9, and 10 (Corocoro, Berenguela, and Tia- Paleocene to early Miocene time: (1) a minor, mid-Paleocene source to the
hauanacu) contain sandstones with limited lateral or upsection variations east during Santa Lucia deposition; (2) a major western source of thrust-
in composition (Fig. 8). Two sandstone samples from the Santa Lucia For- belt-derived and arc-derived material that dominated the Altiplano during
mation in the northernmost section (Section 9, samples 101, 102) define a late Eocene–Oligocene deposition of the Potoco Formation; and (3) a major
relatively quartzofeldspathic composition for the base of the succession eastern source in the backthrust belt along the western flank of the Eastern
(Fig. 8A). Quartz grains are dominantly monocrystalline quartz (Fig. 8B), Cordillera active during deposition of upper Oligocene–lower Miocene stra-
plagioclase is the dominant feldspar (Fig. 8B), and lithic grains are mostly ta (upper Potoco and Coniri formations). A map-view summary of sand-
metamorphic (Fig. 8C). Above the Santa Lucia Formation, 14 samples from stone point-count data and paleocurrent measurements for the Potoco and
the Potoco Formation (samples 73–75, 103–113) indicate generally lithic Coniri formations (Fig. 11) reveals distinct differences in western versus
arkosic compositions for all three sections (Fig. 8A). Potoco sandstones eastern sources.
contain moderate amounts of feldspar, both plagioclase and potassium feld- (1) Paleocene Eastern Source.—For the Santa Lucia Formation, west-
spar (Fig. 8B), and lithic fragments, principally volcanic fragments (Fig. directed paleocurrents (Fig. 3; Horton et al. 2001) and generally quartzo-
8C). feldspathic sandstones containing sedimentary and metasedimentary lithic
Interpretation.—The Potoco Formation was derived from a western fragments (Figs. 5, 8) are consistent with an eastern sediment source com-
sediment source area, as indicated by regional paleocurrent measurements posed primarily of Paleozoic strata. The regional occurrence of distal-flu-
(Fig. 3; Horton et al. 2001). Compared to most other Potoco sandstones vial Santa Lucia facies across the eastern Altiplano and Eastern Cordillera
(Figs. 5, 6), the northern region contains higher amounts of potassium (Sempere et al. 1997) suggests that the ultimate source was located several
feldspar. By analogy with elevated amounts of potassium feldspar in as- hundred kilometers to the east. A lack of carbonate lithic fragments implies
sociation with granite and gneiss clasts in Section 6 (Turco site; described that carbonates of the subjacent El Molino Formation were completely
above), this relationship can be attributed to a Precambrian granitic and buried by Santa Lucia strata and therefore not exposed at the surface. One
gneissic source to the west. Supporting evidence includes reported Precam- unresolved issue concerns the presence of volcanic lithic fragments in the
brian clasts from Potoco-equivalent strata in the northern Altiplano (Ev- Santa Lucia Formation. There are no obvious candidates for volcanic
ernden et al. 1977; Jiménez de Rı́os 1992) and Precambrian gneiss and source areas to the east. Perhaps the most plausible explanation is that
granite sampled from the base of a 2.8-km-deep drill hole (Fig. 1) (Leh- volcanic grains were reworked from limited volcanic beds in Cretaceous
mann 1978; Tosdal 1996). In addition to a crystalline basement source, a clastic units of the Eastern Cordillera (e.g., Sempere et al. 1997).
volcanic-arc source is indicated by a moderate amount of volcanic lithic (2) Late Eocene–Oligocene Western Source.—For the Potoco Forma-
fragments. In contrast to the southern and central regions, there are no well- tion, east-directed paleocurrents (Fig. 3), granite and gneiss clasts (Fig. 7),
defined trends upsection, possibly suggesting a subequal importance of and sandstone compositions displaying moderate to high feldspathic and
basement and volcanic source terranes throughout Potoco deposition in the lithic components (Figs. 5, 6, 8, 11) suggest that volcanic, sedimentary,
north. and crystalline basement rocks were exposed in uplifted areas to the west.
Independent evidence for Paleogene shortening in areas west of the Alti-
East-Derived Deposits plano (discussed below) along with coarse conglomeratic facies in Section
6 (Fig. 3) suggest that a fairly proximal western thrust belt supplied detritus
Description.—East-derived Potoco and Coniri sandstones from Sections from Paleozoic–Mesozoic strata and Precambrian basement to the Altiplano
8A (Comanche) and 8B (Coniri) in the northern region and upper strati- basin. In addition, the presence of plagioclase and volcanic lithic fragments
graphic levels of Section 5 (Chuquichambi) in the central region exhibit reveals that a volcanic-arc source terrane existed to the west. An upsection
generally quartzolithic and quartzofeldspathic compositions (Fig. 9A). increase in volcanic lithic fragments and plagioclase grains (Figs. 5, 6) is
Monocrystalline quartz is the most common grain type (Fig. 9B). Lithic interpreted to indicate greater amounts of volcanic material in the source
components are dominantly sedimentary (mostly sandstone fragments) and area, due to either a larger volume of arc volcanism or eastward migration
metamorphic (mostly phyllitic fragments) (Fig. 9C). Volcanic lithic frag- of the central Andean volcanic arc.
ments are uncommon. (3) Late Oligocene–Early Miocene Eastern Source.—Paleocurrent
Conglomerate clast counts from the Coniri Formation are compatible data for upper Oligocene–lower Miocene deposits of the upper Potoco and
with observed sandstone compositions. The clasts are dominated by sand- Coniri formations (Fig. 3) indicate competing sediment source regions to
stone, quartzite, and carbonate, with subordinate amounts of vein quartz, the east and west of the basin. In the eastern part of the Altiplano basin,
chert, mudrock, and conglomerate (Fig. 10). sandstone and conglomerate of the upper Potoco and Coniri formations
Interpretation.—Paleocurrent data indicate an eastern sediment source have relatively quartz-rich and sedimentary-rich detrital compositions (Figs.
for the upper Potoco and Coniri formations in the eastern Altiplano (Fig. 9–11), consistent with erosion of a proximal source of Paleozoic sedimen-
3; LaReau and Horton 2000; Horton et al. 2001). Dominance of quartz and tary rocks to the east in the Eastern Cordillera. The appearance of east-
lithic fragments in sandstone, and sedimentary clasts in conglomerate, sug- derived detritus in the Altiplano during late Oligocene time provides a
gest that Paleozoic sedimentary and metasedimentary rocks in the Eastern minimum age for previously documented shortening (discussed below) in
Cordillera were the primary source. Quartzite, sandstone, and shale domi- the backthrust belt along the western flank of the Eastern Cordillera.
nate the Paleozoic section; chert-bearing carbonate strata are generally lim-
DISCUSSION
ited to the Upper Permian (Sempere 1995; Suárez and Diaz 1996; Mc-
Quarrie and DeCelles 2001). Relative to west-derived Potoco deposits The modern Altiplano is distinguished by its high-elevation, low-relief
(Figs. 5–8), the east-derived deposits exhibit greater compositional matu- physiographic setting between two uplifted plateau margins. Therefore,
718 B.K. HORTON ET AL.

TABLE 2.—Recalculated modal point-count data for sandstones from the Altiplano*.

Q–F–L % Qm–P–K % Lv–Lm–Ls %


No. Sample m Level Q F L Qm P K Lv Lm Ls
1) Julchi: 198489S, 678079W
1 P11 288 74 13 13 83 16 1 44 14 42
2 P12 295 77 21 2 77 23 0 0 20 80
3 P14 334 88 12 0 88 12 0 0 0 100
4 P15 374 64 28 7 68 32 0 68 0 32
5 P16 422 39 41 20 47 53 0 93 5 1
6 P18 443 17 30 53 33 67 0 82 9 9
7 P22 616 18 39 43 30 63 6 94 2 5
8 P25 874 39 50 11 42 57 1 86 2 12
9 P28 1162 34 44 21 41 57 2 78 7 15
10 P31 1640 36 56 8 37 54 9 68 11 22
11 P34 2205 57 37 6 60 33 7 12 50 38
12 P39 2976 37 33 30 50 34 16 95 1 4
13 P41 3263 2 32 66 6 91 3 99 0 1
14 P44 3530 7 28 65 20 73 7 99 0 1
2) Maicoma: 198399S, 678049W
15 99SL1 82 77 12 11 85 14 1 40 20 40
16 99SL6 236 84 12 4 86 14 0 71 18 12
17 99SL7 264 83 12 4 86 13 2 21 37 42
18 99SL8 298 82 14 4 85 15 0 12 24 65
19 99PT2 410 73 24 3 74 26 0 57 29 14
20 99PT3 458 58 31 10 63 36 1 88 2 10
21 99PT4 482 51 35 14 58 41 1 92 3 5
22 99PT5 513 30 29 40 47 52 0 96 1 4
3) Andamarca: 188479S, 678319W
23 R99SL5 155 83 9 8 90 9 1 34 49 17
24 R99SL7 222 94 4 2 96 4 0 0 63 38
25 R99C2 227 92 5 3 95 5 0 0 46 54
26 R99P3 260 75 17 8 81 19 0 77 9 14
27 R99P4 292 60 34 6 63 36 1 79 11 11
28 ANQPT001 299 32 62 6 33 66 1 84 8 8
5) Chuquichambi: 178559S, 678509W
29 SS-5 390 86 7 8 93 4 3 0 70 30
30 TP425 575 85 9 6 90 4 6 0 46 54
31 SS-8 850 86 3 11 96 2 2 0 72 28
32 SS-9 1395 88 3 9 96 2 2 11 86 3
33 SS-12 1673 84 5 11 94 3 3 0 80 20
34 SS-13 2023 84 5 12 95 2 3 0 93 7
35 SS-16 2432 86 6 9 94 3 3 8 92 0
36 SS-19B 2942 80 6 14 93 3 4 0 100 0
37 SS-19 2947 89 4 8 95 2 2 0 100 0
38 SS-21 3667 80 14 6 83 10 7 26 52 22
39 SS-22B 3680 81 8 11 91 6 4 0 5 95
40 SS-22M 3682 63 15 22 79 11 10 8 38 54
41 SS-22NT 3695 62 18 20 74 14 13 76 4 20
42 SS-3870.EO 4020 84 11 5 87 8 5 0 90 10
43 SS-23 4190 81 11 9 88 5 8 18 76 6
44 SS-25 4554 81 13 6 85 4 12 0 88 13
45 SS-29 5316 77 11 12 83 7 11 0 82 18
46 SS-30 5696 77 8 16 89 0 10 0 97 3
47 SS-31A 5962 65 24 12 69 13 17 7 15 78
48 SS-33 6341 80 14 7 84 5 11 0 100 0
49 SS-35 6656 83 7 11 91 3 6 0 95 5
50 SS-36 6817 87 66 8 94 1 5 0 100 0
6) Turco: 188079S, 688109W
51 T1 8 84 2 14 97 2 2 0 67 33
52 T3 118 77 6 17 91 3 6 12 42 46
53 T6 297 73 11 16 86 10 5 13 71 17
54 T8 420 48 21 31 65 26 9 49 38 13
55 T10 616 54 21 25 69 24 7 30 55 15
56 9T11 728 72 19 9 77 13 9 44 38 18
57 T12 860 74 13 12 83 12 5 16 66 18
58 T14 1070 66 27 7 68 25 7 61 29 10
59 9T16 1260 59 32 8 63 27 9 77 8 15
60 T17 1380 59 28 13 64 31 5 45 43 12
61 T19 1648 61 25 14 68 23 10 51 43 6
62 T20 1850 58 26 17 67 28 5 58 26 16
63 9T21 1945 64 28 8 68 17 15 53 11 37
64 T22 2090 44 48 8 45 33 22 36 28 36
65 T24 2430 28 59 13 31 42 26 37 17 47
66 T25 2498 43 42 16 48 25 27 64 3 33
67 T28 2750 14 54 32 15 67 18 81 7 13
68 T30 3085 51 33 16 57 22 21 46 25 29
69 T32 3310 47 31 22 55 32 13 75 13 12
70 T34 3540 17 60 23 17 77 7 74 6 20
71 T36 3770 14 68 18 15 59 26 81 2 16
72 T38 4000 6 36 58 11 75 14 96 0 4
TERTIARY PROVENANCE HISTORY OF THE ALTIPLANO, CENTRAL ANDES 719

TABLE 2.—Continued.

Q–F–L % Qm–P–K % Lv–Lm–Ls %


No. Sample m Level Q F L Qm P K Lv Lm Ls
7) Corocoro: 178139S, 688289W
73 PT00NB 50 66 26 8 71 23 6 50 25 25
74 PT001 364 63 27 10 68 23 9 57 25 18
75 PT002 493 29 46 25 35 55 10 96 1 3
8) Comanche: 168589S, 688259W
76 COM2 585 93 4 4 96 3 1 0 12 88
77 COM3 730 84 12 4 87 11 2 22 17 61
78 COM4 823 88 6 6 92 2 6 8 35 58
79 COM5 955 86 4 10 94 5 0 5 23 72
80 COM6 1145 78 4 18 95 5 0 2 17 81
8B) Coniri: 168479S, 688259W
81 Hy-1 1 78 11 11 87 3 10 37 6 57
82 Hy-75 75 67 14 19 82 4 14 34 4 62
83 C1-18 504 79 5 16 94 1 5 0 25 75
84 C1-63 549 79 6 15 93 1 5 0 19 81
85 C1-91 577 80 5 15 94 2 5 0 33 67
86 C1-230 716 75 4 20 94 1 4 0 21 79
87 C1-286 772 76 4 19 95 2 3 0 24 76
88 C1-386 872 80 4 16 95 0 4 0 30 70
89 C1-475 961 79 5 16 94 1 4 0 26 74
90 C1-494 980 77 7 17 92 2 6 0 26 74
91 C1-567 1053 76 5 19 93 2 5 5 28 67
92 C2-650 1136 75 5 20 94 2 4 0 26 74
93 C2-735 1221 78 5 17 93 4 3 0 30 70
94 C2-963 1449 85 3 12 96 3 1 0 62 38
95 C2-1025 1511 83 3 14 96 3 1 4 43.1 53
96 C2-1095 1581 83 5 12 94 4 2 0 53.3 47
97 C2-1300 1786 83 5 12 94 4 1 0 58.7 41
98 C3-1581 2067 82 4 13 95 3 2 0 76.5 24
99 C3-1775 2261 82 4 14 95 3 2 1 62.9 36
100 C3-1795 2281 86 3 11 97 2 1 0 56.1 44
9) Tiahuanacu: 168369S, 688559W
101 TSL1 20 80 15 5 81 19 0 27 64 9
102 997134 40 88 8 4 91 9 0 37 44 19
103 THC2 215 49 37 14 56 33 10 85 3 11
104 THC3 412 54 36 10 59 27 13 87 4 9
105 THC4 505 58 26 16 68 24 8 90 0 10
106 THC5 801 53 36 11 59 26 15 80 0 20
107 THC6 1095 54 27 20 66 25 9 69 1 30
108 THC7 1345 54 26 20 67 24 9 71 3 26
109 THC8 1453 52 30 18 62 22 16 74 1 25
10) Berenguela: 178179S, 698129W
110 BER1 5 57 20 23 73 4 23 78 3 19
111 BER2 30 71 15 14 80 6 14 60 27 13
112 BER3 70 60 27 13 67 22 11 88 5 7
113 BER4 80 58 25 17 69 9 22 76 3 21
* See Appendix 1 (archived) for raw point-count data.

determination of paleoelevation and plateau-margin tectonics will provide 3, 5, 8) indicate that distal-fluvial Santa Lucia strata were derived from a
a sound basis for interpreting plateau development. Although we mention distant eastern source composed predominantly of Paleozoic clastic rocks
the paleoelevation issue below, our study focuses on the topographic evo- (Fig. 12A). Although regionally extensive, the Santa Lucia Formation is
lution of plateau margins. We present a tectonic reconstruction (Fig. 12) quite thin (50–300 m) in the Altiplano. We suggest that the Santa Lucia
that draws on the provenance data presented here and structural and ther- and El Molino formations were deposited in a zone of regional Late Cre-
mochronological data from previous studies. We find that magmatic-arc taceous to mid-Paleocene subsidence (Fig. 12A). Accumulation occurred
and probable thrust-belt source terranes were of temporally and spatially within a broad zone that may have been located between a flexurally in-
variable importance during basin evolution. Hence, we conclude that the duced forebulge to the west and the Brazilian shield to the east (Horton
present geomorphic simplicity of the plateau belies the complexity exhib- and DeCelles 1997; Horton et al. 2001). Alternatively, deposition may have
ited in the Tertiary record of provenance and presumably sediment accu- been above a formerly active Cretaceous rift system (Welsink et al. 1995).
mulation in the Altiplano basin. Support for the flexural loading hypothesis is provided by evidence of Late
Cretaceous–Paleocene shortening and possible foredeep conditions to the
Initial Conditions west in the Precordillera (Fig. 12A) (discussed below and by Horton et al.
2001). In either case, it is unlikely that there was significant Late Creta-
Lower Tertiary strata provide information on the initial tectonogeo- ceous–Paleocene topography or mountain building associated with the dis-
morphic conditions of the Altiplano region prior to plateau development. tal eastern sediment source.
Given the existence of marine fossils in the regionally extensive El Molino
Formation (Sempere et al. 1997), the Altiplano and Eastern Cordillera were Western Plateau Margin
approximately at sea level during Late Cretaceous–early Paleocene time.
We consider the El Molino and overlying mid-Paleocene Santa Lucia For- Provenance data for the western plateau margin indicate initial unroofing
mation to be good records of pre-plateau conditions. Provenance data (Figs. of a composite sediment source consisting of a probable thrust belt with
720 B.K. HORTON ET AL.

FIG. 5.—Ternary diagrams for 11 Santa Lucia


and 17 Potoco sandstone samples from the
southern region. Note the distinct compositional
difference between the Santa Lucia and Potoco
samples, and clear upsection trends (arrows)
toward greater feldspar (particularly plagioclase)
and lithic fragments (particularly volcanic
fragments). Outlined samples for Section 1
(Julchi site) denote abrupt increase in feldspar
(particularly potassium feldspar) and
metamorphic lithic fragments (see text for
discussion). Stratigraphic levels for samples are
shown in Fig. 3 and Table 2.

minor magmatic-arc contributions, followed by unroofing of a single mag- We propose that a Paleogene east-vergent thrust belt in the Precordillera
matic-arc source (Figs. 3, 5–8, 11). Late Eocene–Oligocene denudation in and Western Cordillera acted both as a sediment source and a topographic
this region resulted in deposition of up to 6.5 km of fluvial strata (Potoco load to an early foreland basin in the Altiplano (Fig. 12B) (Mortimer 1981;
Formation) in the Altiplano basin (Fig. 12B). Continued Neogene erosion Sempere et al. 1990a; Horton et al. 2001). This interpretation predicts that
along the western margin facilitated deposition of several more kilometers the low-temperature cooling history related to Paleogene denudation should
of strata. be recorded in areas west of the Altiplano. However, this record may be
Evidence for Late Cretaceous–Eocene shortening in a probable fold–thrust largely obscured by subsequent arc magmatism. On the basis of increasing
belt in the Precordillera region (presently 50 km west of the Western Cor- amounts of arc-derived detritus in Altiplano basin fill (Figs. 5, 6, 8), we
dillera axis) includes mapped contractional structures (Chong 1977; Chong speculate that thrusting west of the Altiplano diminished by late Oligocene
and Reutter 1985; Hammerschmidt et al. 1992; Scheuber and Reutter 1992; time. This hypothesis is supported by the majority of paleomagnetic data,
Scheuber et al. 1994; Amilibia et al. 1999; Nicolas et al. 1999) and complex, which indicate limited Neogene rotation in areas west of the Altiplano (e.g.,
paleomagnetically determined (vertical-axis) block rotations attributable to Kuhn and Reuther 1999; Arriagada et al. 2000; Roperch et al. 2000a; Ro-
shortening and related (possibly conjugate) strike-slip deformation (Hartley perch et al. 2000b), although Hartley et al. (1992) report post–early Mio-
et al. 1992; Kuhn and Reuther 1999; Arriagada et al. 2000; Roperch et al. cene vertical-axis rotations locally in the Precordillera.
2000a; Roperch et al. 2000b). Identification of shallow ramp–flat thrust sys- Long-term eastward migration of the Mesozoic–Cenozoic magmatic-arc
tems (Hartley et al. 1992) and broad-wavelength folds with inferred detach- system is well documented in the central Andes (James 1971; Scheuber and
ments at 8–10 km depth (Chong 1977; Chong and Reutter 1985; Scheuber Reutter 1992; Scheuber et al. 1994; Allmendinger et al. 1997; Wörner et al.
and Reutter 1992) are similar to structural geometries observed in the Andean 2000a). Dated volcanic and intrusive igneous rocks indicate that the arc was
thrust belt east of the Altiplano (Roeder 1988; Kley et al. 1997; Horton 1998). centered on the Precordillera during Paleogene time (Hammerschmidt et al.
Precambrian basement is involved in most structures and both east-vergent 1992; Scheuber and Reutter 1992; Scheuber et al. 1994). From a summary
and west-vergent geometries are present. Similar east-vergent and west-ver- of central Andean magmatic activity by Allmendinger et al. (1997, their fig.
gent, basement-involved structures imaged seismically in the Western Cor- 9), arc magmatism commenced in the Western Cordillera during late Oli-
dillera and western Altiplano (Lamb and Hoke 1997; Rochat et al. 1998; gocene time. Late Oligocene–early Miocene igneous rocks are widely dis-
McQuarrie and DeCelles 2001) may also have been active during this time. tributed from the Precordillera to Eastern Cordillera, suggesting a broad zone
Thermochronological data from Precambrian basement in the Western Cor- of arc magmatism at this time (Fig. 12B) (Jordan and Gardeweg 1989; Scheu-
dillera include U–Pb zircon ages and 40Ar/39Ar and K–Ar hornblende ages ber and Reutter 1992; Allmendinger et al. 1997).
indicative of Middle Proterozoic and early Paleozoic metamorphism in this
part of the Arequipa–Antofalla cratonic block (Tosdal 1996; Wörner et al. Eastern Plateau Margin
2000b). To our knowledge, the low-temperature cooling history of the crys- For the eastern margin of the Altiplano, provenance data indicate middle
talline basement has not been studied. to late Tertiary unroofing of a contractional belt composed principally of
TERTIARY PROVENANCE HISTORY OF THE ALTIPLANO, CENTRAL ANDES 721

FIG. 7.—Clast composition data for the Potoco conglomeratic interval in the cen-
tral region, 1730–2470 m level of Section 6 (Turco site). See Figures 1 and 3 for
location.

contributed to Altiplano subsidence as early as middle–late Eocene time


(Fig. 12). However, our provenance data for the Altiplano basin do not
show evidence of an eastern sediment source until late Oligocene time. For
this reason, we consider late Oligocene time to be a minimum age of initial
uplift in the Eastern Cordillera. This age represents the arrival of westward-
prograding alluvial deposystems in the Altiplano study area. Although pro-
gradation rates are unknown, rates of ;30 km/Myr have been determined
for case studies in the Himalayan foreland basin (Burbank et al. 1988;
Brozovic and Burbank 2000). If similar rates were to apply to the Altiplano
basin, then the required 50–100 km of westward progradation may have
taken as little as 2–3 million years. Alternatively, if progradation rates were
substantially lower, then the estimate of initial Eastern Cordillera uplift
could be pushed back to Eocene time. More chronological data are required
to assess fully the age of earliest deformation in the Eastern Cordillera.
Concerning a possible igneous provenance, we note that the youngest
FIG. 6.—Ternary diagrams for 37 Potoco sandstone samples from the central re- deposits analyzed in our study are of early Miocene age. Because the prov-
gion. For Section 6 (Turco), note the clear upsection trends (arrows) toward greater enance data do not show any distinctive appearance of an igneous source
feldspar (particularly plagioclase) and lithic fragments (particularly volcanic frag- area to the east, the erosional record of Neogene volcanic and intrusive
ments). Distal-equivalent strata in Section 5 (Chuquichambi) to the east exhibit a rocks in the western part of the Eastern Cordillera (Gubbels 1993; Lamb
minor trend toward more feldspathic and lithic compositions. Stratigraphic levels et al. 1997) must be mostly post–early Miocene.
for samples are shown in Fig. 3 and Table 2.
Modes of Basin Development
Paleozoic clastic rocks (Figs. 3, 9, 10). Late Oligocene–early Miocene un- On the basis of the variable topographic evolution of the Altiplano pla-
roofing of this belt generated over 3 km of fluvial and alluvial-fan strata teau margins, we suggest that different modes of basin development (i.e.,
(upper Potoco and Coniri formations) in the Altiplano (Fig. 12C). This mechanisms of basin subsidence) may have been in operation during the
erosion was largely contemporaneous with continued erosion along the Tertiary (Fig. 12). For early to mid-Tertiary time, we interpret a causal
western plateau margin. relationship between plateau-margin shortening and basin subsidence. Late
A west-vergent backthrust belt (the ‘‘Huarina fold–thrust belt’’) in the Cretaceous–Eocene shortening in a probable east-vergent thrust belt in the
western part of the Eastern Cordillera has been recognized by several work- Precordillera and Western Cordillera would have produced a thrust load
ers (e.g., Roeder 1988; Sempere et al. 1990b). The amount of Tertiary along the western margin of the Altiplano (Fig. 12B). Similarly, late Oli-
shortening in this belt is estimated to be 100–150 km (Roeder and Cham- gocene–early Miocene (and possibly earlier) shortening in a west-vergent
berlain 1995; McQuarrie and DeCelles 2001). The timing of shortening is backthrust belt in the western part of the Eastern Cordillera would have
not well constrained, but thermochronological evidence for rapid middle– produced a thrust load along the opposite side of the Altiplano (Fig. 12C).
late Eocene cooling (Benjamin et al. 1987; McBride et al. 1987; Farrar et Although not a major contributor to Altiplano loading, primarily Neogene
al. 1988; Masek et al. 1994; Lamb and Hoke 1997; McQuarrie and De- shortening and uplift in the east-vergent Andean thrust belt (east of the
Celles 2001) suggests that uplift-related denudation was underway by 45– backthrust belt) (Roeder 1988; Sheffels 1990; Roeder and Chamberlain
40 Ma. Syndepositional shortening during late Oligocene–early Miocene 1995; Kley 1996; Baby et al. 1997; Kley et al. 1997) was concurrent with
time is indicated by seismically imaged growth structures in basin fill of continued Altiplano deposition. Given the spatial and temporal link be-
the eastern Altiplano (Lamb and Hoke 1997, their fig. 9; McQuarrie and tween basin development and upper-crustal shortening, we suggest that the
DeCelles 2001, their fig. 12). On the basis of the ages of overlapping, long-term history of the Altiplano basin be viewed as the development of
slightly deformed strata, shortening in the backthrust belt terminated during a crustal-scale piggyback basin (e.g., Rochat et al. 1999; McQuarrie and
early Miocene time (Sempere et al. 1990b; McQuarrie and DeCelles 2001). DeCelles 2001).
From available timing constraints, a west-directed thrust load could have Given that shortening drives the greatest component of crustal thickening
722 B.K. HORTON ET AL.

FIG. 8.—Ternary diagrams for 2 Santa Lucia


and 14 Potoco sandstone samples from the
northern region. Note the compositional
difference between Santa Lucia and Potoco
samples. Potoco data do not exhibit well-defined
upsection or lateral variations in composition.
Stratigraphic levels for samples are shown in
Fig. 3 and Table 2.

in the central Andes (Isacks 1988; Sheffels 1990), maintenance of Airy generate 12 km of Tertiary flexural subsidence in the Altiplano. Based on
isostatic balance (as presently observed across the Altiplano and its mar- simple 2-D calculations (Turcotte and Schubert 1982), two extensive to-
gins; Beck et al. 1996) would require that surface uplift accompanied Ter- pographic loads (each 2.25–3 km high by 100–150 km long in cross sec-
tiary shortening. Because the bulk of shortening along plateau margins was tion) on opposing sides of the basin could explain 7 km of late Eocene to
apparently complete by early Miocene time (McQuarrie and DeCelles early Miocene sediment accumulation for a low- to moderate-strength crust
2001), most surface uplift should have taken place by that time. This simple (13.5–20 km effective elastic thickness) (Hampton 2002). In lieu of purely
rationale, however, conflicts with recent paleoelevation analyses based on flexurally driven subsidence, we hypothesize that additional modes of sed-
paleobotanical data (Gregory-Wodzicki 2000) that suggest greater than half iment accumulation have played a role in basin development. One obvious
(;2 km) of the surface uplift in the Altiplano region since 10 Ma. Clearly, possibility is that phases of Tertiary extension may have affected the Al-
further investigation of the Tertiary paleoelevation record is warranted. tiplano (e.g., Martinez 1980; Sébrier et al. 1988; Rochat et al. 1998; All-
Finally, we call attention to the extreme topographic loading required to mendinger et al. 1997). However, the structural evidence for normal fault-

FIG. 9.—Ternary diagrams for 14 east-derived


Potoco and 18 east-derived Coniri sandstone
samples from the central region (upper level of
Section 5) and northern region (Sections 8A,
8B). Note the difference in composition from the
majority of west-derived sandstone samples
(Figs. 5, 6, 8). Data do not exhibit well-defined
upsection variations in composition. Stratigraphic
levels for samples are shown in Fig. 3 and Table 2.
TERTIARY PROVENANCE HISTORY OF THE ALTIPLANO, CENTRAL ANDES 723

ing is not compelling. In fact, contractional deformation appears to have


dominated the middle to upper crust of the Altiplano throughout the Ter-
tiary (e.g., Lamb and Hoke 1997; McQuarrie and DeCelles 2001). Another
possible mechanism is thermal subsidence (e.g., Jordan and Alonso 1987),
possibly related to cooling after extensive Miocene magmatism across the
region. We propose that a third alternative, simple sediment ponding in a
closed depositional area due to drainage blockage by structural sills (e.g.,
Coney et al. 1996; Métivier et al. 1998; Rochat et al. 1999), can explain
accumulation and preservation of several hundred to a few thousand meters
of sediment in the Altiplano. We consider this a particularly viable expla-
nation for deposition during middle Miocene–Quaternary time when there
were apparently few major faults active along plateau margins capable of
driving significant tectonic subsidence (Lamb and Hoke 1997; McQuarrie
and DeCelles 2001). Because our provenance data indicate topographic
expression of an eastern sediment source by late Oligocene time, we pro-
pose that growth of an eastern structural sill may have triggered internal
FIG. 10.—Clast composition data for the Coniri conglomerate in the northern drainage by ;25 Ma. This mode of sediment accumulation would have
region, Section 8B (Coniri site). Note the difference in composition from west- supplemented any flexurally driven subsidence related to plateau-margin
derived conglomerate (Fig. 7). See Figures 1 and 3 for location. thrust loading (Fig. 12C). Without a more exhaustive sedimentologic and
provenance dataset spanning the entire Altiplano, however, we cannot rule

FIG. 11.—Map-view summary of sandstone


point-count data and paleocurrent data for the
Potoco and Coniri formations. The map outline
corresponds approximately to the map in Fig. 1.
In the Q-F-L ternary diagrams, irregular filled
areas represent the approximate range of
compositions for sandstone samples from the
indicated stratigraphic section (Figs. 5, 6, 8, 9).
Note the substantial compositional difference
between west-derived and east-derived deposits.
Arrows represent mean paleocurrent directions
for entire measured sections or indicated
stratigraphic intervals. Dotted lines define
Paleocene–lower Miocene outcrop belts, and
dashed lines indicate the present margins of the
Altiplano (from Fig. 1).
724 B.K. HORTON ET AL.

FIG. 12.—Schematic reconstructions of Altiplano basin development. The top row shows map-view reconstructions based on provenance data (the map outline corresponds
approximately to the map in Fig. 1). The bottom row shows interpretive, unrestored regional cross sections based on provenance, structural, and thermochronological data
(see text for discussion). A) Mid-Paleocene deposition of a thin succession of east-derived Santa Lucia Formation in eastern Altiplano and Eastern Cordillera concurrent
with contractional deformation and foredeep development in the Precordillera. At this time, depositional systems derived from a probable thrust belt to the west did not
reach the Altiplano. B) Late Eocene–Oligocene deposition across Altiplano of the thick Potoco Formation derived from a thrust belt and volcanic arc to the west (Western
Cordillera and Precordillera). Question marks and dashed thrust faults convey uncertainty in initial age of shortening-related uplift in Eastern Cordillera. C) Late Oligocene–
Miocene deposition of uppermost Potoco and Coniri formations derived primarily from a west-vergent backthrust belt in the Eastern Cordillera. Coeval west-derived
volcaniclastic units are limited to the westernmost Altiplano and signify erosion of a volcanic arc to the west (Western Cordillera). Neogene deposition in the Altiplano is
synchronous with shortening in the east-vergent Andean thrust belt (central Eastern Cordillera to Subandean Zone) and foreland-basin development in the Chaco Plain.

out the possibility that a few drainage systems may have escaped eastward Neogene–Quaternary volcanic arc, both arc-related and non-arc-related
from the Altiplano since late Oligocene time. Nevertheless, late Oligocene (thrust-belt) sediment sources were involved in the early to mid-Tertiary
to Quaternary internal drainage in the Altiplano is an attractive possibility evolution of the western plateau margin. Sandstone and conglomerate com-
because it could generate a net increase in mean surface elevation, possibly positions provide evidence for erosion of Paleozoic–Mesozoic sedimentary
on the order of several hundred to over a thousand meters, by simple and metasedimentary strata and Precambrian granitic and gneissic base-
sediment accumulation. ment. Erosion of these rocks suggests early to mid-Tertiary uplift of a
probable thrust belt to the west, consistent with previously reported evi-
CONCLUSIONS dence for Paleogene shortening in the region. Upsection increases in pla-
Temporal and spatial variations in sandstone and conglomerate compo- gioclase and volcanic lithic fragments indicate progressively greater arc
sition within the Altiplano basin help define the Tertiary history of basin influence during middle to late Tertiary time. Increased volcanic input may
filling in relationship to the topographic development of eastern and west- represent larger-magnitude volcanism or an eastward migration of the vol-
ern plateau margins. After limited mid-Paleocene deposition (50–300-m- canic arc.
thick Santa Lucia Formation), major sediment accumulation commenced East-derived fluvial and alluvial-fan deposits from the upper Potoco For-
during roughly late Eocene time (3–6.5-km-thick Potoco Formation). Pa- mation and ;3-km-thick Coniri Formation of the Altiplano basin contain
leocurrent and stratigraphic data indicate that late Eocene–Oligocene basin relatively quartz-rich sandstone with sedimentary and metasedimentary lith-
filling occurred by deposition of fluvial systems derived from western sed- ic fragments, consistent with derivation from deformed Paleozoic strata in
iment sources in the incipient Western Cordillera and Precordillera. Al- the west-vergent backthrust belt along the western flank of the Eastern
though the present topography of the Western Cordillera is formed by a Cordillera. Judging from the age of these strata, initial shortening, crustal
TERTIARY PROVENANCE HISTORY OF THE ALTIPLANO, CENTRAL ANDES 725

thickening, and uplift of the Eastern Cordillera must have been in progress determination from trough cross-stratification: Journal of Sedimentary Petrology, v. 53, p.
629–642.
by late Oligocene time. Deformation could have started earlier if alluvial DICKINSON, W.R., 1970, Interpreting detrital modes of graywacke and arkose: Journal of Sed-
deposystems took substantial time to prograde westward into studied parts imentary Petrology, v. 40, p. 695–707.
of the Altiplano basin. We propose that late Oligocene time may also rep- DICKINSON, W.R., 1985, Interpreting provenance relations from detrital modes of sandstones,
in Zuffa, G.G., ed., Provenance of Arenites: Dordrecht, The Netherlands, Reidel, p. 333–
resent the initiation of internally drained, plateau-like conditions in the 361.
Altiplano basin. Provided that closed-basin conditions have persisted since EISBACHER, G.H., CARRIGY, M.A., AND CAMPBELL, R.B., 1974, Paleodrainage pattern and late-
that time, topographic infilling due to structural damming may have played orogenic basins of the Canadian Cordillera, in Dickinson, W.R., ed., Tectonics and Sedi-
mentation: Society of Economic Paleontologists and Mineralogists, Special Publication 22,
a role in accommodating several hundred to a few thousand meters of upper p. 143–166.
Oligocene to Quaternary basin fill. Nevertheless, structural development of EVERNDEN, J.F., KRIZ, S.J., AND CHERRONI, C., 1977, Potassium–argon ages of some Bolivian
both plateau margins synchronous with early to mid-Tertiary basin filling rocks: Economic Geology, v. 72, p. 1042–1061.
suggests that flexural subsidence related to shortening and topographic FARRAR, E., CLARK, A.H., KONTAK, D.J., AND ARCHIBALD, D.A., 1988, Zongo–San Gabán zone:
Eocene foreland boundary of the Central Andean orogen, northwest Bolivia and southeast
loading along the margins of a crustal-scale piggyback basin may have Peru: Geology, v. 16, p. 55–58.
been the principal subsidence mechanism responsible for accumulation of GAZZI, P., 1966, Le arenarie del flysch sopracretaceo dell’Appennino modenese; correlazioni
the thick (up to 12 km) Tertiary succession in the Altiplano. con il flysch di Monghidoro: Mineralogica et Petrographica Acta, v. 12, p. 69–97.
GRAHAM, S.A., HENDRIX, M.S., WANG, L.B., AND CARROLL, A.R., 1993, Collisional successor
basins of western China: Impact of tectonic inheritance on sand composition: Geological
ACKNOWLEDGMENTS Society of America, Bulletin, v. 105, p. 323–344.
GREGORY-WODZICKI, K.M., 2000, Uplift history of the central and northern Andes: A review:
Geological Society of America, Bulletin, v. 112, p. 1091–1105.
This research was supported by National Science Foundation grant EAR-9908003. GUBBELS, T.L., 1993, Tectonics and geomorphology of the eastern flank of the Central Andes,
Compositional data for the Chuquichambi and Coniri sections represent part of the 188 to 238 south latitude [Ph.D. thesis]: Cornell University, Ithaca, New York, 211 p.
M.S. research by Hampton and LaReau, respectively. We thank Enrique Diaz and HAMMERSCHMIDT, K., DÖBEL, R., AND FRIEDRICHSEN, 1992, Implication of 40Ar/39Ar dating of
Thierry Sempere for helpful discussions concerning Altiplano stratigraphy and prov- Early Tertiary volcanic rocks from the north-Chilean Precordillera: Tectonophysics, v. 202,
enance. Field work was facilitated by logistical assistance from Sohrab Tawackoli, p. 55–81.
Reinhard Rössling, Carlos Riera, Juan Huchani, and Pedro Churata of Sergeomin HAMPTON, B.A., 2002, Early–middle Tertiary deposition in the Corque syncline, Altiplano pla-
(La Paz). Richard Fink, Danielle Horton, and Jason Barnes assisted in the field. We teau, Bolivia [unpublished M.S. thesis]: Baton Rouge, Louisiana, Louisiana State University,
124 p.
thank Tim Lawton, Steve Graham, Mark Johnsson, and Ray Ingersoll for construc- HARTLEY, A.J., JOLLEY, E.J., AND TURNER, P., 1992, Paleomagnetic evidence for rotation in the
tive reviews of the manuscript. Precordillera of northern Chile: structural constraints and implications for the evolution of
The appendix cited in this paper has been archived, and is available in digital the Andean forearc: Tectonophysics, v. 205, p. 49–64.
form, at the World Data Center-A for Marine Geology and Geophysics, NOAA/ HENDRIX, M.S., 2000, Evolution of Mesozoic sandstone compositions, southern Junggar, north-
NGDC, 325 Broadway, Boulder, CO 80303, U.S.A; telephone 303-497-6339; fax ern Tarim, and western Turpan basins, northwest China: A detrital record of the ancestral
303-497-6513; E-mail: wdcamgg@ngdc.noaa.gov; URL http://www.ngdc.noaa.gov/ Tian Shan: Journal of Sedimentary Research, v. 70, p. 520–532.
mgg/sepm/archive/index/html. HORTON, B.K., 1998, Sediment accumulation on top of the Andean orogenic wedge: Oligocene
to late Miocene basins of the Eastern Cordillera, southern Bolivia: Geological Society of
America, Bulletin, v. 110, p. 1174–1192, and p. 1513.
REFERENCES HORTON, B.K., AND DECELLES, P.G., 1997, The modern foreland basin system adjacent to the
Central Andes: Geology, v. 25, p. 895–898.
AHLFELD, F., AND BRANISA, L., 1960, Geologı́a de Bolivia: La Paz, Instituto Boliviano del HORTON, B.K., AND DECELLES, P.G., 2001, Modern and ancient fluvial megafans in the foreland
Petroleo, 245 p. basin system of the central Andes, southern Bolivia: implications for drainage network
ALLMENDINGER, R.W., JORDAN, T.E., KAY, S.M., AND ISACKS, B.L., 1997, The evolution of the evolution in fold–thrust belts: Basin Research, v. 13, p. 43–63.
Altiplano–Puna plateau of the Central Andes: Annual Review of Earth and Planetary Sci- HORTON, B.K., HAMPTON, B.A., AND WAANDERS, G.L., 2001, Paleogene synorogenic sedimen-
ences, v. 25, p. 139–174. tation in the Altiplano plateau and implications for initial mountain building in the central
AMILIBIA, A., SÀBAT, F., CHONG, G., MUÑOZ, J.A., ROCA, E., AND RODRIGUEZ-PEREA, A., 1999, Andes: Geological Society of America, Bulletin, v. 113, p. 1387–1400.
Evolution of Domeyko range, northern Chile, in Extended Abstracts, Fourth International HUNGERBÜHLER, D., STEINMANN, M., WINKLER, W., SEWARD, D., EGÜEZ, A., HELLER, F., AND FORD,
Symposium on Andean Geodynamics, Göttingen, Germany: Paris, IRD, p. 25–29. M., 1995, An integrated study of fill and deformation in the Andean intermontane basin of
ARRIAGADA, C., ROPERCH, P., AND MPODOZIS, C., 2000, Clockwise block rotations along the Nabón (Late Miocene), southern Ecuador: Sedimentary Geology, v. 96, p. 257–279.
eastern border of the Cordillera de Domeyko, northern Chile (228459–238309S): Tectono- INGERSOLL, R.V., BULLARD, T.F., FORD, R.L., GRIMM, J.P., PICKLE, J.D., AND SARES, S.W., 1984,
physics, v. 326, p. 153–171. The effect of grain size on detrital modes: A test of the Gazzi-Dickinson point-counting
BABY, P., ROCHAT, P., MASCLE, G., AND HÉRAIL, G., 1997, Neogene shortening contribution to method: Journal of Sedimentary Petrology, v. 54, p. 103–116.
crustal thickening in the back arc of the Central Andes: Geology, v. 25, p. 883–886. INGERSOLL, R.V., KRETCHMER, A.G., AND VALLES, P.K., 1993, The effect of sampling scale on
BABY, P., SEMPERE, T., OLLER, J., BARRIOS, L., HÉRAIL, G., AND MAROCCO, R., 1990, A late actualistic sandstone petrofacies: Sedimentology, v. 40, p. 937–953.
Oligocene–Miocene intermontane foreland basin in the southern Bolivian Altiplano: Aca- ISACKS, B.L., 1988, Uplift of the central Andean plateau and bending of the Bolivian orocline:
démie des Sciences [Paris], Comptes Rendus, série II, v. 311, p. 341–347. Journal of Geophysical Research, v. 93, p. 3211–3231.
BECK, S., ZANDT, G., MYERS, S.C., WALLACE, T.C., SILVER, P.G., AND DRAKE, L., 1996, Crustal- JAMES, D.E., 1971, Plate-tectonic model for the evolution of the central Andes: Geological
thickness variations in the central Andes: Geology, v. 24, p. 407–410. Society of America, Bulletin, v. 82, p. 3325–3346.
BENJAMIN, M.T., JOHNSON, N.M., AND NAESER, C.W., 1987, Recent rapid uplift in the Bolivian JIMÉNEZ DE Rı́OS, G., 1992, Acerca del basamento Precámbrico en el bloque Andino de Bolivia
Andes: Evidence from fission-track dating: Geology, v. 15, p. 680–683. (abstract): Sociedad Geológica Boliviana, Boletı́n (X Congreso Geológico Boliviano, La
BROZOVIC, N., AND BURBANK, D.W., 2000, Dynamic fluvial systems and gravel progradation in Paz), v. 27, p. 88–93.
the Himalayan foreland: Geological Society of America, Bulletin, v. 112, p. 394–412. JOHNSSON, M.J., 1990, Tectonic versus chemical-weathering controls on the composition of
BURBANK, D.W., BECK, R.A., RAYNOLDS, R.G.H., HOBBS, R., AND TAHIRKHELI, R.A.K., 1988, fluvial sands in tropical environments: Sedimentology, v. 37, p. 713–726.
Thrusting and gravel progradation in foreland basins: A test of post-thrusting gravel dis- JOHNSSON, M.J., 2000, Tectonic assembly of east-central Alaska: Evidence from Cretaceous–
persal: Geology, v. 16, p. 1143–1146. Tertiary sandstones of the Kandik River terrane: Geological Society of America, Bulletin,
CHASE, C.G., GREGORY-WODZICKI, K.M., PARRISH, J.T., AND DECELLES, P.G., 1998, Topographic v. 112, p. 1023–1042.
history of the western Cordillera of North America and controls on climate, in Crowley, JORDAN, T.E., AND ALONSO, R.N., 1987, Cenozoic stratigraphy and basin tectonics of the Andes
T.J., and Burke, K.C., eds., Tectonic Boundary Conditions for Climate Reconstructions: Mountains, 208–288 south latitude: American Association of Petroleum Geologists Bulletin,
New York, Oxford University Press, p. 73–99. v. 71, p. 49–64.
CHONG, G., 1977, Contribution to the knowledge of the Domeyko Range in the Andes of JORDAN, T.E., AND GARDEWEG, P., 1989, Tectonic evolution of the late Cenozoic central Andes
northern Chile: Geologische Rundschau, v. 66, p. 374–404. (208–338S), in Ben-Avraham, Z., ed., The Evolution of the Pacific Ocean margins: New
CHONG, G., AND REUTTER, K.J., 1985, Fenómenos de tectónica compresiva en las Sierras de York, Oxford University Press, p. 193–206.
Varas y de Argomedo, Precordillera Chilena, en el ámbito del paralelo 258 sur: IV Congreso KAY, R.F., MACFADDEN, B.J., MADDEN, R.H., SANDEMAN, H., AND ANAYA, F., 1998, Revised age
Geologico Chileno, Actas, v. 2, p. 219–238. of the Salla beds, Bolivia, and its bearing on the age of the Deseadan South American Land
CONEY, P.J., MUÑOZ, J.A., MCCLAY, K.R., AND EVENCHICK, C.A., 1996, Syntectonic burial and Mammal ‘‘Age’’: Journal of Vertebrate Paleontology, v. 18, p. 189–199.
post-tectonic exhumation of the southern Pyrenees foreland fold-thrust belt: Geological So- KENNAN, L., LAMB, S., AND RUNDLE, C., 1995, K–Ar dates from the Altiplano and Cordillera
ciety of London, Journal, v. 153, p. 9–16. Oriental of Bolivia: implications for Cenozoic stratigraphy and tectonics: Journal of South
DECELLES, P.G., AND HERTEL, F., 1989, Petrology of fluvial sands from the Amazonian foreland American Earth Sciences, v. 8, p. 163–186.
basin, Peru and Bolivia: Geological Society of America, Bulletin, v. 101, p. 1552–1562. KLEY, J., 1996, Transition from basement-involved to thin-skinned thrusting in the Cordillera
DECELLES, P.G., LANGFORD, R.P., AND SCHWARTZ, R.K., 1983, Two new methods of paleocurrent Oriental of southern Bolivia: Tectonics, v. 15, p. 763–775.
726 B.K. HORTON ET AL.

KLEY, J., MÜLLER, J., TAWACKOLI, S., JACOBSHAGEN, V., AND MANUTSOGLU, E., 1997, Pre-Andean tectonosedimentary model of the northern Bolivian Altiplano: Académie des Sciences
and Andean-age deformation in the Eastern Cordillera of southern Bolivia: Journal of South [Paris], Comptes Rendus, série II, v. 327, p. 769–775.
American Earth Sciences, v. 10, p. 1–19. ROCHAT, P., HÉRAIL, G., BABY, P., AND MASCLE, G., 1999, Crustal balance and control of the
KRAEMER, B., ADELMANN, D., ALTEN, M., SCHNURR, W., ERPENSTEIN, K., KIEFER, E., VAN DEN erosive and sedimentary processes on the Altiplano formation: Académie des Sciences [Par-
BOGAARD, P., AND GÖRLER, K., 1999, Incorporation of the Paleogene foreland into the Neo- is], Comptes Rendus, série II, v. 328, p. 189–195.
gene Puna plateau: The Salar de Antofalla area, NW Argentina: Journal of South American ROEDER, D., 1988, Andean-age structure of Eastern Cordillera (province of La Paz, Bolivia):
Earth Sciences, v. 12, p. 157–182. Tectonics, v. 7, p. 23–39.
KUHN, D., AND REUTHER, C.D., 1999, Strike-slip faulting and nested block rotations: structural evidence ROEDER, D., AND CHAMBERLAIN, R.L., 1995, Structural geology of Sub-Andean fold and thrust
from the Cordillera de Domeyko, northern Chile: Tectonophysics, v. 313, p. 383–398. belt in northwestern Bolivia, in Tankard, A.J., Suárez, R., and Welsink, H.J., eds., Petroleum
LAMB, S., AND HOKE, L., 1997, Origin of the high plateau in the Central Andes, Bolivia, South Basins of South America: American Association of Petroleum Geologists, Memoir 62, p.
America: Tectonics, v. 16, p. 623–649. 459–479.
LAMB, S., HOKE, L., KENNAN, L., AND DEWEY, J., 1997, Cenozoic evolution of the Central Andes ROPERCH, P., ARRIAGADA, C., AND COUTAND, I., 2000a, Tectonic rotations in the central Andes:
in Bolivia and northern Chile, in Burg, J.P., and Ford, M., eds., Orogeny through Time: Implications for the geodynamic evolution of the Altiplano–Puna plateau (abstract): Eos,
Geological Society of London, Special Publication 121, p. 237–264. Transactions, American Geophysical Union, v. 81, p. 1083.
LAREAU, B.N., AND HORTON, B.K., 2000, Stratigraphy of the middle Tertiary Coniri Formation, ROPERCH, P., FORNARI, M., HÉRAIL, G., AND PARRAGUEZ, G.V., 2000b, Tectonic rotations within
northern Altiplano plateau, Central Andes, Bolivia (abstract): Geological Society of America, the Bolivian Altiplano: Implications for the geodynamic evolution of the central Andes
Abstracts with Programs, v. 32, no. 7, p. 458. during the late Tertiary: Journal of Geophysical Research, v. 105, p. 795–820.
LAVENU, A., BONHOMME, M.G., VATIN-PERIGNON, N., AND DE PACHTERE, P., 1989, Neogene mag- ROPERCH, P., HÉRAIL, G., AND FORNARI, M., 1999, Magnetostratigraphy of the Miocene Corque
matism in the Bolivian Andes between 168S and 188S: Stratigraphy and K/Ar geochronol- basin, Bolivia: Implications for the geodynamic evolution of the Altiplano during the late
Tertiary: Journal of Geophysical Research, v. 104, p. 20,415–20,429.
ogy: Journal of South American Earth Sciences, v. 2, p. 35–47.
RUSSO, R.M., AND SILVER, P.G., 1996, Cordillera formation, mantle dynamics, and the Wilson
LEHMANN, B., 1978, A Precambrian core sample from the Altiplano/Bolivia: Geologische Rund-
cycle: Geology, v. 24, p. 511–514.
schau, v. 67, p. 270–278. SCHEUBER, E., BOGDANIC, T., JENSEN, A., AND REUTTER, K.J., 1994, Tectonic development of the
MACFADDEN, B.J., ANAYA, F., PEREZ, H., NAESER, C.W., ZEITLER, P.K., AND CAMPBELL, K.E., north Chilean Andes in relation to plate convergence and magmatism since the Jurassic, in
1990, Late Cenozoic paleomagnetism and chronology of Andean basins of Bolivia: Evidence Reutter, K.J., Scheuber, E., and Wigger, P.J., eds., Tectonics of the Southern Central Andes:
for possible oroclinal bending: Journal of Geology, v. 98, p. 541–555. New York, Springer-Verlag, p. 121–139.
MACFADDEN, B.J., ANAYA, F., AND SWISHER, C.C., III, 1995, Neogene paleomagnetism and or- SCHEUBER, E., AND REUTTER, K.J., 1992, Magmatic arc tectonics in the Central Andes between
oclinal bending of the central Andes of Bolivia: Journal of Geophysical Research, v. 100, 218 and 258S: Tectonophysics, v. 205, p. 127–140.
p. 8153–8167. SÉBRIER, M., LAVENU, A., FORNARI, M., AND SOULAS, J.P., 1988, Tectonics and uplift in central
MAROCCO, R., LAVENU, A., AND BAUDINO, R., 1995, Intermontane late Paleogene–Neogene basins Andes (Peru, Bolivia and northern Chile) from Eocene to present: Géodynamique, v. 3, p.
of the Andes of Ecuador and Peru: Sedimentologic and tectonic characteristics, in Tankard, 85–106.
A.J., Suárez, R., and Welsink, H.J., eds., Petroleum Basins of South America: American SEMPERE, T., 1995, Phanerozoic evolution of Bolivia and adjacent regions, in Tankard, A.J.,
Association of Petroleum Geologists, Memoir 62, p. 597–613. Suárez, R., and Welsink, H.J., eds., Petroleum Basins of South America: American Asso-
MARSAGLIA, K.M., AND INGERSOLL, R.V., 1992, Compositional trends in arc-related, deep-marine ciation of Petroleum Geologists, Memoir 62, p. 207–230.
sand and sandstone: A reassessment of magmatic-arc provenance: Geological Society of SEMPERE, T., BUTLER, R.F., RICHARDS, D.R., MARSHALL, L.G., SHARP, W., AND SWISHER, C.C.,
America, Bulletin, v. 104, p. 1637–1649. 1997, Stratigraphy and chronology of Late Cretaceous–early Paleogene strata in Bolivia and
MARSH, S.P., RICHTER, D.H., LUDINGTON, S., SORIA-ESCALANTE, E., AND ESCOBAR-DIAZ, A., 1992, northwest Argentina: Geological Society of America, Bulletin, v. 109, p. 709–727.
Geologic map of the Altiplano and Cordillera Occidental, Bolivia, in Geology and Mineral SEMPERE, T., HÉRAIL, G., OLLER, J., BABY, P., BARRIOS, L., AND MAROCCO, R., 1990a, The Alti-
Resources of the Altiplano and Cordillera Occidental, Bolivia: U.S. Geological Survey and plano: A province of intermontane foreland basins related to crustal shortening in the Bolivia
Servicio Geológico de Bolivia, U.S. Geological Survey Bulletin 1975, plate 1. orocline area: International Symposium on Andean Geodynamics, Grenoble, p. 167–170.
MARSHALL, L.G., SEMPERE, T., AND BUTLER, R.F., 1997, Chronostratigraphy of the mammal- SEMPERE, T., HÉRAIL, G., OLLER, J., AND BONHOMME, M.G., 1990b, Late Oligocene–early Miocene
bearing Paleocene of South America: Journal of South American Earth Sciences, v. 10, p. major tectonic crisis and related basins in Bolivia: Geology, v. 18, p. 946–949.
49–70. SHAO, L., STATTEGGER, K., AND GARBE-SCHOENBERG, C.D., 2001, Sandstone petrology and geo-
MARSHALL, L.G., SWISHER, C.C., III, LAVENU, A., HOFFSTETTER, R., AND CURTIS, G.H., 1992, chemistry of the Turpan basin (NW China): Implications for the tectonic evolution of a
Geochronology of the mammal-bearing late Cenozoic on the northern Altiplano, Bolivia: continental basin: Journal of Sedimentary Research, v. 71, p. 37–49.
Journal of South American Earth Sciences, v. 5, p. 1–19. SHEFFELS, B.M., 1990, Lower bound on the amount of crustal shortening in the central Bolivian
MARTINEZ, C., 1980, Structure et évolution de la chaı̂ne hercynienne et da la chaı̂ne andine Andes: Geology, v. 18, p. 812–815.
dans le nord de la Cordillère des Andes de Bolivie: Paris, Office de la Recherche Scientifique STEINMANN, M., HUNGERBÜHLER, D., SEWARD, D., AND WINKLER, W., 1999, Neogene tectonic
et Technique Outre-Mer (ORSTOM), Travaux et Documents, v. 119, 352 p. evolution and exhumation of the southern Ecuadorian Andes: A combined stratigraphy and
MASEK, J.G., ISACKS, B.L., GUBBELS, T.L., AND FIELDING, E.J., 1994, Erosion and tectonics at fission-track approach: Tectonophysics, v. 307, p. 255–276.
the margins of continental plateaus: Journal of Geophysical Research, v. 99, p. 13,941– SUÁREZ, R., AND DIAZ, E., 1996, Léxico estratigráfico de Bolivia: Revista Técnica de Yaci-
13,956. mientos Petrolı́feros Fiscales Bolivianos, v. 17, no. 1–2, p. 7–227.
MCBRIDE, S.L., CLARK, A.H., FARRAR, E., AND ARCHIBALD, D.A., 1987, Delimitation of a cryptic TOSDAL, R.M., 1996, The Amazon–Laurentian connection as viewed from the Middle Prote-
Eocene tectono-thermal domain in the Eastern Cordillera of the Bolivian Andes through K–Ar rozoic rocks in the central Andes, western Bolivia and northern Chile: Tectonics, v. 15, p.
827–842.
dating and 40Ar–40Ar step-heating: Geological Society of London, Journal, v. 144, p. 243–255.
TROËNG, B., SORIA, E., CLAURE, H., MOBAREC, R., AND MURILLO, F., 1994, Descubrimiento de
MCQUARRIE, N., AND DECELLES, P.G., 2001, Geometry and structural evolution of the central basamento Precámbrico en la Cordillera Occidental Altiplano de los Andes Bolivianos (ab-
Andean backthrust belt, Bolivia: Tectonics, v. 20, p. 669–692. stract): XI Congreso Geológico de Bolivia, Santa Cruz, Memorias, p. 231–237.
MÉGARD, F., NOBLE, D.C., MCKEE, E.H., AND BELLON, H., 1984, Multiple pulses of Neogene TROP, J.M., AND RIDGWAY, K.D., 1997, Petrofacies and provenance of a Late Cretaceous suture
compressive deformation in the Ayacucho intermontane basin, Andes of central Peru: Geo- zone thrust-top basin, Cantwell basin, central Alaska Range: Journal of Sedimentary Re-
logical Society of America, Bulletin, v. 95, p. 1108–1117. search, v. 67, p. 469–485.
MÉTIVIER, F., GAUDEMER, Y., TAPPONNIER, P., AND MEYER, B., 1998, Northeastward growth of TURCOTTE, D.L., AND SCHUBERT, G., 1982, Geodynamics: Applications of Continuum Physics
the Tibet plateau deduced from balanced reconstruction of two depositional areas: The Qai- to Geological Problems: New York, John Wiley & Sons, 450 p.
dam and Hexi Corridor basins, China: Tectonics, v. 17, p. 823–842. VANDERVOORT, D.S., 1993, Non-marine evaporite basin studies, southern Puna plateau, central
MORTIMER, C., 1981, Drainage evolution in the Atacama Desert of northernmost Chile: Revista Andes: [Ph.D. thesis]: Cornell University, Ithaca, New York, 177 p.
Geológica de Chile, v. 11, p. 3–28. VANDERVOORT, D.S., JORDAN, T.E., ZEITLER, P.K., AND ALONSO, R.N., 1995, Chronology of in-
NICOLAS, C., WILKE, H.-G., SCHNEIDER, H., AND LUCASSEN, F., 1999, New constraints on the ternal drainage development and uplift, southern Puna plateau, Argentine central Andes:
Cretaceous–Tertiary deformation in Sierra de Moreno, Precordillera of the Segunda region Geology, v. 23, p. 145–148.
de Antofagasta, northern Chile, in Extended abstracts, Fourth International Symposium on WELSINK, H.J., MARTINEZ, E., ARANIBAR, O., AND JARANDILLA, J., 1995, Structural inversion of a
Andean Geodynamics, Göttingen, Germany: Paris, IRD, p. 543–546. Cretaceous rift basin, southern Altiplano, Bolivia, in Tankard, A.J., Suárez, R., and Welsink,
PAREJA, J., VARGAS, C., SUÁREZ, R., BALLÓN, R., CARRASCO, R., AND VILLAROEL, C., 1978, Mapa H.J., eds., Petroleum Basins of South America: American Association of Petroleum Geol-
geológico de Bolivia y memoria explicativa: Yacimientos Petrolı́feros Fiscales Bolivianos ogists, Memoir 62, p. 305–324.
y Servicio Geológico de Bolivia, 1:1,000,000 scale, 27 p. WÖRNER, G., HAMMERSCHMIDT, K., HENJES-KUNST, F., LEZAUN, J., AND WILKE, H., 2000a, Geo-
REUTTER, K.J., DÖBEL, R., BOGDANIC, T., AND KLEY, J., 1994, Geological map of the Central chronology (40Ar/39Ar, K–Ar and He-exposure ages) of Cenozoic magmatic rocks from
Andes between 208 and 268 S, in Reutter, K.J., Scheuber, E., and Wigger, P.J., eds., Tec- northern Chile (18–228S): implications for magmatism and tectonic evolution of the central
tonics of the Southern Central Andes: New York, Springer-Verlag, Enclosure 1. Andes: Revista Geológica de Chile, v. 27, p. 205–240.
WÖRNER, G., LEZAUN, J., BECK, A., HEBER, V., LUCASSEN, F., ZINNGREBE, E., RÖSSLING, R., AND
RICKETTS, B.D., EVENCHICK, C.A., ANDERSON, R.G., AND MURPHY, D.C., 1992, Bowser basin,
WILKE, H.G., 2000b, Precambrian and Early Paleozoic evolution of the Andean basement
northern British Columbia: Constraints on the timing of initial subsidence and Stikinia– at Belen (northern Chile) and Cerro Uyarani (western Bolivia Altiplano): Journal of South
North America terrane interactions: Geology, v. 20, p. 1119–1122. American Earth Sciences, v. 13, p. 717–737.
ROBINSON, R.S., AND JOHNSSON, M.J., 1997, Chemical and physical weathering of fluvial sands YIN, A., AND HARRISON, T.M., 2000, Geologic evolution of the Himalayan–Tibetan orogen:
in an Arctic environment: Sands of the Sagavanirktok River, North Slope, Alaska: Journal Annual Review of Earth and Planetary Sciences, v. 28, p. 211–280.
of Sedimentary Research, v. 67, p. 560–570.
ROCHAT, P., BABY, P., HÉRAIL, G., MASCLE, G., AND ARANIBAR, O., 1998, Geometric analysis and Received 1 June 2001; accepted 7 February 2002.

You might also like