You are on page 1of 29

Ore Geology Reviews 21 (2002) 127 – 155

www.elsevier.com/locate/oregeorev

Crustal lineament control on magmatism and mineralization in


northwestern Argentina: geological, geophysical, and remote
sensing evidence
Carlos J. Chernicoff a,*, Jeremy P. Richards b,1, Eduardo O. Zappettini c
a
Council for Scientific and Technical Research (CONICET), University of Buenos Aires and Argentine Geological and Mining Survey (SEGEMAR),
651 Julio A. Roca Avenue, 1322 Buenos Aires, Argentina
b
Department of Earth and Atmospheric Sciences, University of Alberta, Edmonton, AB, Canada T6G 2E3
c
Argentine Geological and Mining Survey (SEGEMAR), 651 Julio A. Roca Avenue, 1322 Buenos Aires, Argentina
Received 1 September 2001; accepted 5 June 2002

Abstract

The relationship between long-lived deep crustal lineaments and the locations of magmatic centers and associated mineral
deposits has been investigated in the Puna region of northwestern Argentina, through the analysis of regional aeromagnetic
surveys, Landsat images, and geological information. The good exposure and excellent preservation of basement and
supracrustal geology in this region makes it particularly suitable for such a study. At a regional scale, several contrasting
magnetic domains are recognized, which correlate with crustal geology. Two basement domains are separated by a NNE-
trending boundary, which is believed to correlate with a Paleozoic suture zone between the Pampia (to the southeast) and
Arequipa – Antofalla terranes (to the northwest). Locally overlying these basement terranes is the Cenozoic magmatic domain,
which is best developed in the N – S-trending volcanic arc at the western edge of the Puna (the Cordillera Occidental). In
addition, four southeast-trending volcanic zones extend for several hundred kilometers across the Puna. Many important
mineral deposits and areas of hydrothermal alteration are associated with these volcanic breakouts, and we have selected three
such areas for more detailed study: Bajo de la Alumbrera (Argentina’s largest porphyry copper deposit), Cerro Galán (the largest
ignimbrite caldera in Argentina, with associated hydrothermal alteration zones), and El Queva (a historic polymetallic district
located within a major volcanic range). A comparison of lineament maps generated from aeromagnetic and Landsat TM images
reveals broad correlation between these different remote sensing techniques, which respectively highlight subsurface magnetic
and surface geological features. In addition, the locations of magmatic and hydrothermal centers can be related to the interpreted
structural framework, and are seen to occur near the intersections of major lineament zones. It is suggested that in three
dimensions, such intersection zones form trans-lithospheric columns of low strength and high permeability during
transpressional or transtensional tectonic strain, and may thereby serve as conduits for magma ascent to the shallow crust.
Pooling of large volumes of deeply derived magma in shallow crustal magma chambers may then result in voluminous
devolatilization and the formation of hydrothermal mineral deposits. It is important to note that in this model, structural
intersections serve as facilitators for magma ascent and volatile exsolution, but do not in themselves cause this process—other

*
Corresponding author. Tel.: +54-11-4349-3148; fax: +54-11-4349-3171.
E-mail addresses: jchern@secind.mecon.gov.ar (C.J. Chernicoff), Jeremy.Richards@UAlberta.CA (J.P. Richards),
ezappe@secind.mecon.gov.ar (E.O. Zappettini).
1
Fax: +1-780-492-2030.

0169-1368/02/$ - see front matter D 2002 Elsevier Science B.V. All rights reserved.
PII: S 0 1 6 9 - 1 3 6 8 ( 0 2 ) 0 0 0 8 7 - 2
128 C.J. Chernicoff et al. / Ore Geology Reviews 21 (2002) 127–155

factors such as magma supply rate and tectonic stress are essential primary ingredients, and local magmatic and volcanic
processes affect the ultimate potential for ore formation. Nevertheless, we suggest that lineament analysis provides a valuable
framework for guiding the early stages of mineral exploration; other regional and local geological considerations must then be
applied to identify priority targets within this framework.
D 2002 Elsevier Science B.V. All rights reserved.

Keywords: Puna; Argentina; Lineaments; Mineral deposits; Aeromagnetic surveys; Landsat TM

1. Introduction detail. We follow O’Leary et al. (1976, p. 1463) in


defining a lineament as ‘‘a mappable, simple or
The Puna of northwestern Argentina connects with composite linear feature of a surface, whose parts
the Altiplano of Bolivia (Fig. 1) to form the second are aligned in a rectilinear or slightly curvilinear
largest high-altitude plateau in the world after Tibet relationship which differs distinctly from the patterns
( f 500,000 km2, average height f 3.7 km above sea of adjacent features and presumably reflects a subsur-
level, with maximum elevations f 6.7 km and local face phenomenon.’’ Lineaments may thus represent
relief >3 km; Allmendinger, 1986). Crustal shortening exposed faults, but also unexposed structures or
associated with uplift of the plateau may have begun crustal discontinuities whose existence is inferred
as early as the Late Eocene (Coutand et al., 2001) but from indirect sensing methods and their control on
was established in the Puna by the Early Miocene topography and magmatic activity. We propose that
(Allmendinger et al., 1997). Basement blocks were these deep, in some cases, trans-lithospheric, and in
uplifted along major reverse faults at this time, and the many cases, ancient (i.e., pre-Andean), discontinuities
topographic expression of this geological framework represent planes of crustal weakness that may be
has been preserved in the arid conditions that have periodically reactivated and intruded during subse-
prevailed since the Pliocene (Kleinert and Strecker, quent tectono-magmatic events (Holdsworth et al.,
2001). Well-preserved also are the andesitic –dacitic 1997). As such, these zones may serve to localize
volcanic products of the Miocene – Quaternary magmatic-related ore-forming processes, with the
Andean orogeny. As such, this region provides an implication that lineament mapping might be used
excellent opportunity to investigate the role played by as an exploration guide for porphyry- and epithermal-
crustal structures in the localization of magmatism and type ore deposits.
related mineralization.
A combination of indirect sensing techniques
(aeromagnetic data and Landsat images) and geo- 2. Geology of the Puna
logical information has been used to define crustal-
scale lineaments in the Puna. The relationship The Altiplano – Puna (Figs. 1 and 2) is a region of
between these lineaments and geological features is thickened continental crust, reaching f 80 km near
explored, and a strong control on the regional distri- the Argentina– Chile – Bolivia border, and with litho-
bution of Andean volcanic activity is noted. Three spheric thicknesses of f 150 km beneath the Alti-
areas where a particularly clear relationship can be plano and f 100 km beneath the Puna (Beck et al.,
observed between regional lineaments and lineament 1996; Allmendinger et al., 1997). Much of the crust
intersections, and the loci of volcanic centers and appears to be felsic in composition, which, combined
associated hydrothermal systems, are examined in with its anomalous thickness, accounts for the high

Fig. 1. Regional geological sketch map of northern Chile, northwestern Argentina, and southwestern Bolivia, showing the locations of regional-
scale faults and lineaments (modified from Salfity, 1985; Salfity and Gorustovich, 1998), and the extent of Central Andean Miocene –
Pleistocene volcanic rocks. Also shown are the locations of major volcanoes and porphyry Cu deposits. Note that Salfity (1985) and Salfity and
Gorustovich (1998) used mapped faults, stratigraphic discontinuities and changes, volcanic lineations, and topographic features identified from
air and satellite images in their lineament interpretation (cf. Heyl, 1972).
C.J. Chernicoff et al. / Ore Geology Reviews 21 (2002) 127–155 129
130 C.J. Chernicoff et al. / Ore Geology Reviews 21 (2002) 127–155

Fig. 2. Geological map of the Puna region, northwestern Argentina. Boxes mark areas discussed in detail (from north to south: El Queva, Cerro
Galán, Bajo de la Alumbrera).
C.J. Chernicoff et al. / Ore Geology Reviews 21 (2002) 127–155 131

elevation of the region (Allmendinger et al., 1997; Table 1


Dorbath and Masson, 2000). The plateau is bordered Summary of main tectonic events in the Central Andes of Chile and
Argentina
to the west by the Cordillera Occidental (the axis of
Time Geological event and effect in the Puna
the current magmatic arc) and to the east by the
Cordillera Oriental. Internally, the Puna features Pliocene Crustal thickening in the southern
Puna (Diaguita phase of Andean
thrust-bounded basement blocks separated by inter-
orogeny, f 1.7 Ma)
montane basins, which since the Middle Miocene Middle Miocene – Development of internal drainage
have developed internal drainage and thick sedimen- Quaternary basins in Puna – Altiplano
tary and evaporitic infillings. Miocene – Quaternary Middle Miocene Crustal thickening in the northern
volcanism produced large, often isolated, composite (15 – 10 Ma) Puna – Altiplano (Quechua phase of
Andean orogeny, f 5.3 Ma)
volcanoes, and an extensive veneer of andesitic –
Late Oligocene – Arc volcanism in Cordillera
dacitic lava flows and ignimbrites (Figs. 1 and 2). Quaternary Occidental and across Puna
The pre-Mesozoic history of the Central Andean Late Oligocene Rapid ( f 15 cm/year),
basement is not well known, with theories for its near-orthogonal plate convergence
assembly ranging from terrane accretion to intracra- Late Eocene – Deformation and crustal shortening
Early Oligocene in NE Puna (Incaic phase of Andean
tonic evolution (e.g., Coira et al., 1982; Dalziel, 1986;
orogeny, f 34 Ma). Arc volcanism
Bahlburg and Breikreuz, 1991; Bahlburg and Hervé, in Chile and western Puna
1997; Ramos, 1999; Lucassen et al., 2000). Mon and Middle Eocene Rapid oblique (NE-directed)
Salfity (1995) suggested that the various elements that convergence
make up the current basement in the Puna region were Middle – Late Arc volcanism in Chilean
Eocene Precordillera
first assembled during the Late Proterozoic Panamer-
Paleocene – Arc volcanism in Chilean
ican orogeny (700 –600 Ma; Table 1). Metamorphic Early Eocene Central Valley
protolith model ages range from 2.2 to 1.4 Ga, and Early Cretaceous Atlantic opening. Back-arc alkaline
reflect a cratonic origin with little magmatic addition syenites and granites in Puna
(Lucassen et al., 2000). The oldest basement rocks of Jurassic Back-arc granite emplacement in Puna
Late Paleozoic – Andean cycle of subduction begins;
the Central Andes (1900 – 980 Ma; Pankhurst and
Early Cretaceous arc volcanism on western seaboard
Rapela, 1998) belong to the Arequipa –Antofalla ter- of Chile (La Negra magmatic arc)
rane, which is believed to underlie the western half of Late Devonian – Intraplate granites emplaced in
the Puna. These basement rocks do not crop out in this Early Carboniferous Pampia terrane (Chanic phase of
region, but have been identified as xenoliths in Cen- Famatinian orogeny, f 355 Ma)
Ordovician Subduction-related granitoid intrusions
ozoic ignimbrites; Pb isotopic data show a link with the
in the Arequipa – Antofalla terrane
Amazonian craton to the north (Tosdal, 1996). (Faja Eruptiva de la Puna)
A major NNE-trending boundary longitudinally Early Ordovician – Final docking of Arequipa – Antofalla
divides the Puna, separating the Arequipa –Antofalla Early Carboniferous terranes (Oclóyic phase of Famatinian
terrane in the west from the Pampia terrane to the east, orogeny, f 435 Ma)
Late Proterozoic – Onset of accretion of Arequipa –
which consists of Late Proterozoic to Early Paleozoic
Cambrian Antofalla terrane (Pampean orogeny,
rocks (750 –530 Ma). Accretion of the Arequipa – 600 – 520 Ma)
Antofalla terrane is thought to have generated the Late Proterozoic First assembly of basement elements
Pampean orogeny (600 – 520 Ma; Pankhurst and (Panamerican orogeny, 700 – 600 Ma)
Rapela, 1998; Ramos, 1999). A marine sedimentary Late Proterozoic to Pampia terrane lithologies
Early Paleozoic
basin developed between these blocks in the Early
Early – Middle Arequipa – Antofalla terrane lithologies
Cambrian, and extended eastward to the present day Proterozoic (1900 – 980 Ma)
Cordillera Oriental. However, it is not clear whether
the basin formed as an intracratonic rift, a fully
developed passive margin, or an active margin (Coira
et al., 1982; Ramos, 1999). vician –Early Carboniferous; Pankhurst and Rapela,
A second orogenic phase, the Famatinian cycle, 1998; Ramos, 1999). Intraplate granites were
affected the region in the mid-Paleozoic (Early Ordo- emplaced in the Pampia terrane during the Late
132 C.J. Chernicoff et al. / Ore Geology Reviews 21 (2002) 127–155

Devonian –Early Carboniferous (the Chanic orogeny; the Late Oligocene (Coira et al., 1982; Mpodozis
Page and Zappettini, 1999), whereas Pankhurst et al. and Ramos, 1989). Deformation and crustal short-
(1998) report the Early – Middle Ordovician emplace- ening began in the northeastern Puna (Protocordil-
ment of subduction-related calc-alkaline plutons to the lera Oriental) during the Late Eocene – Early
south in the Sierras Pampeanas. An important suite of Oligocene Incaic orogeny, following a period of
apparently subduction-related Ordovician granitoid rapid oblique (NE-directed) plate convergence in
intrusions crops out in a NNW-trending belt in the the mid-Eocene (Pardo-Casas and Molnar, 1987;
Arequipa – Antofalla terrane, stretching from the Hammerschmidt et al., 1992; Coutand et al.,
southwestern Puna into northern Chile (the Faja 2001). A second episode of rapid ( f 15 cm/year),
Eruptiva de la Puna; Palma et al., 1986; Damm et near-orthogonal convergence began in the Late Oli-
al., 1990). Bahlburg and Breikreuz (1991) and Bahl- gocene, and high but decreasing rates were sustained
burg and Hervé (1997) suggest that the Famatinian until the end of the Miocene (Pardo-Casas and
cycle was related to the final closure of the basin Molnar, 1987; Somoza, 1998). Crustal thickening
separating the Arequipa – Antofalla and Pampia ter- reached a maximum in the northern Puna and
ranes at the end of the Ordovician (Oclóyic orogeny; Altiplano during this period (the mid-Miocene Que-
Allmendinger et al., 1982). Welding of these terranes chua phase of the Andean orogeny, 15 – 10 Ma;
laid the foundations for the Puna, or Protopuna, which Jordan and Gardeweg, 1989; Coutand et al., 2001),
featured a west-facing active margin on its western and continued through the Pliocene in the southern
edge during the Ordovician (now the Cordillera Occi- Puna (Diaguita phase; Jordan and Alonso, 1987;
dental; Bahlburg and Breikreuz, 1991; Bahlburg and Kraemer et al., 1999). Coira et al. (1993) and Kay
Hervé, 1997). and Kay (1993) have suggested that this late Neo-
The Andean cycle of subduction of the Pacific gene deformation event was related to lithospheric
plate beneath the western margin of Gondwana was delamination beneath the southern Puna.
clearly established by Jurassic times (Coira et al., Neogene– Quaternary sedimentation in the Puna
1982), but may have begun as early as the Triassic produced fluvial and lacustrine (including evaporitic)
or even Late Carboniferous (Damm et al., 1994; deposits within closed, commonly structurally con-
Bahlburg and Hervé, 1997). This fundamental change trolled, intermontane basins (Fig. 2; Jordan et al.,
in plate interactions was related to the breakup of 1983; Jordan and Alonso, 1987; Alonso et al., 1991;
Gondwana (Dalziel, 1986). A volcanic arc was built Vandervoort et al., 1995; Allmendinger et al., 1997;
on the present-day western seaboard of Chile (La Coutand et al., 2001). Volcanism flared during the
Negra magmatic arc; Coira et al., 1982; Mpodozis Miocene in the Cordillera Occidental, and continued
and Ramos, 1989), which operated until the Early sporadically into the Quaternary (Thorpe, 1984; Jor-
Cretaceous when Atlantic opening changed plate dan and Gardeweg, 1989). Large andesitic – dacitic
motions and caused a hiatus in magmatism (Dalziel, stratovolcanoes were constructed during this period
1986; Palacios et al., 1993). During this period, along the main axis of the arc (e.g., Socompa,
magmatism in the Puna region was restricted to minor Llullaillaco), and also along southeast-trending trans-
volcanic activity with associated small alkalic intru- verse eruptive zones that cut across the Puna (e.g., the
sions (syenites, riebeckite granites, and carbonatites) El Queva, Antofalla, and Cerro Galán volcanoes; Fig.
in extensional back-arc basins (Salta Group; Salfity 1; Jordan and Gardeweg, 1989; Kraemer et al., 1999).
and Marquillas, 1999). Several centers developed collapse calderas during
Throughout the Cenozoic, the volcanic arc moved eruption of large ignimbrite sheets (e.g., Cerro Galán;
progressively eastwards in response to a shallowing Coira et al., 1982; Koukharsky, 1985; Sparks et al.,
of the angle of subduction, with axes located in the 1985; Riller et al., 2001; Siebel et al., 2001).
Central Valley of Chile during the Paleocene – Early
Eocene, the Chilean Precordillera during the Mid- 2.1. Structural history
dle – Late Eocene (Jordan and Gardeweg, 1989;
Hammerschmidt et al., 1992), and then moving to The pre-Cenozoic structural architecture of the
its present location in the Cordillera Occidental by Puna basement is dominated by the NNE-trending
C.J. Chernicoff et al. / Ore Geology Reviews 21 (2002) 127–155 133

suture zone between the Arequipa –Antofalla and the nonpenetrative, and crustal thickening processes
Pampia terranes (roughly defined by the western out- appear to have been concentrated in the lower crust.
crop limit of Precambrian rocks in Fig. 2). This Furthermore, the main structural elements in the
structural grain was reactivated during the Cenozoic, Puna—NW-, NNE-, and NE-trending faults—all
and forms a key element in the delineation of uplifted appear to have had antecedents in the Paleozoic
basement blocks and intermontane basins. Several basement.
important NW- and NE-trending transverse linea-
ments that cut across the Puna and extend westward 2.2. Lineaments and structural controls on magma-
into Chile also appear to have originated in the tism and mineral deposits
Paleozoic (Allmendinger et al., 1983; Jordan et al.,
1983; Salfity, 1985; see below). Mesozoic and Cenozoic magmatism in northern
The Cenozoic deformation history of the Puna has Chile and Argentina defines a series of approximately
been studied in detail by a number of authors (e.g., trench-parallel arcs, which young progressively east-
Allmendinger et al., 1997; Kraemer et al., 1999). wards to the currently active Cordillera Occidental.
Tectonism related to crustal thickening began during Where erosion has exposed the intrusive levels of
the Late Eocene – Early Oligocene Incaic orogeny, and these arcs (e.g., the Late Eocene– Early Oligocene
reached a peak in the Miocene, as noted above. porphyry Cu belt in northern Chile), close relation-
During this period, ESE shortening across the Puna ships between plutonism and orogen-parallel strike –
was accommodated by NNE-trending thrust faults slip faulting can be observed (Baker and Guilbert,
along which basement blocks were uplifted (Marrett 1987; Lindsay et al., 1995; Tomlinson and Blanco,
and Emerman, 1992; Kraemer et al., 1999; Coutand et 1997). More specifically, Richards et al. (2001) docu-
al., 2001). Minor amounts of N – S extension occurred mented a spatial association between porphyry Cu
at the northern and southern margins of the plateau mineralization at La Escondida, and the intersection of
(Allmendinger et al., 1997), while reactivated NW- a NW-trending structural corridor (the Archibarca
and NE-trending faults acted as transfer structures lineament) with the N – S West Fissure Zone. This
with sinistral and dextral displacements, respectively relationship is unsurprising, because plutonism and
(Allmendinger et al., 1983; Jordan et al., 1983; Salfity, faulting in such settings are interrelated processes:
1985; Matteini et al., 1997; Coutand et al., 2001). magmatic heat serves to weaken the crust and localize
Estimates of upper crustal shortening are not as strain, but preexisting structures also serve to focus
extensive as might be expected from the observed magma ascent, providing a feedback mechanism
degree of crustal thickening (e.g., only 10 – 15%; (Davidson et al., 1994; Vigneresse and Tikoff, 1999).
Coutand et al., 2001), suggesting either that the Puna At shallow crustal levels, volcanic and other surfi-
has been underthrust by cratonic material from the cial deposits commonly obscure deep-seated faults.
east (Dorbath and Masson, 2000) or by material The presence of such underlying structures can never-
tectonically eroded at the subduction zone to the west theless be inferred from alignments of secondary
(Baby et al., 1997), or that magmatic addition to the features, such as stratigraphic discontinuities or facies
crust has been volumetrically important (James and changes, linear arrays of magmatic features (intru-
Sack, 1999). sions, volcanoes), and topographic features. The
During the Diaguita deformational phase in the regional lineament map shown in Fig. 1 was produced
Pliocene, motion along the main NNE-trending struc- by Salfity (1985) on this basis using air photographs
tures changed to strike – slip, with E- or NE-directed and satellite images, as well as existing geological
shortening, and N- to NW-directed extension along maps. In addition to major orogen-parallel structural
secondary structures (Allmendinger et al., 1989; Mar- corridors, the map reveals several transverse linea-
rett and Emerman, 1992; Marrett et al., 1994; Coira et ments that can be traced from the Pacific coast of
al., 1993). Chile, across the Cordillera Occidental, and into the
In summary, although there is ample evidence for Puna (e.g., the Calama – Olacapato– El Toro, Archi-
strain in the Puna during the Andean orogeny, upper barca, and Culampajá lineaments; Allmendinger et al.,
crustal deformation was relatively thin-skinned and 1982, 1983; Jordan et al., 1983; Alonso et al., 1984;
134
Table 2
Location, characteristics, geological environment, and magnetic signature of significant mineral deposits of the Puna
Deposit examples Deposit type Age Tectonic setting Lithologic unit Lithology Magnetic signature
Santa Catalina-Rinconada district Mesothermal Ordovician Talus deposits Acoyte f. and Quarzites and Nonmagnetic to

C.J. Chernicoff et al. / Ore Geology Reviews 21 (2002) 127–155


(Au) (22j25VS.L., 66j17VW.L.); Au – Sb veins equivalents sandstones weakly magnetic.
Incahuasi (Au) Regional intensity
(25j27VS.L., 67j12VW.L.) generally moderate
Mina Aguilar Sedex Ordovician Syn – rift Lampazar Quarzites and Nonmagnetic to
(23j12VS.L., 65j42VW.L.); deposits and Cuarcita pelites weakly magnetic.
La Colorada (polymetallic) Aguilar ff. High regional
(23j38VS.L., 66j17VW.L.) intensity derived
from the underlying
Precambrain basement
Grupo La Mesada (W) Granite- Carboniferous Post-orogenic Llullaillaco Granites Moderately magnetic
(26j58VS.L., 66j56VW.L.) associated granites Complex and surrounded by a rim of
veins equivalents low gradient
Descubridora de Los Aparejos Skarn
(Cu) (27j42VS.L., 68j24VW.L.)
Piedra Calzada (Bi) Pegmatite
(26j48VS.L., 66j51VW.L.)
Tusaquillas and Liquinaste (W) Granite- Jurassic – Back arc Tusaquillas and Granites, Syenites Nonmagnetic to
(23j15VS.L. 66j01VW.L.) associated Cretaceous magmatism Rangel ff. weakly magnetic
veins and intraplate
alkaline
intrusions
Rangel District (Th-ETR) Carbonatites Moderately magnetic
(23j30VS.L., 66j15VW.L.) and associated with linear WNW-
veins trending fabric
Taca Taca (Cu – Mo) Porphyry Cu Paleogene Magmatic arc Santa Inés Rhyodacitic and Moderately magnetic
(24j35VS.L., 67j44VW.L.) Complex and dacitic porphyries,
equivalents dacites, andesites,
and ignimbrites
Pirquitas (Sn – Ag) Bolivian-type Middle Magmatic arc Cordón de Rhyodacitic and Moderately magnetic
(22j42VS.L., 66j30VW.L.) polymetallics Miocene Arizaro, Valle dacitic porphyries,
Ancho ff, Pozuelos dacites, andesites,
Complex and ignimbrites
and equivalents
Cerro Redondo (Au) Epithermal and
(22j23VS.L., 66j08VW.L.); transition type
Pan de Azúcar (Pb – Zn – Ag – Sb)
(22j32VS.L., 66j01VW.L.)
Bajo de La Alumbrera (Cu – Au) Porphyry Cu – Mo Upper Miocene Magmatic arc Agua Caliente f. Rhyodacitic and Moderately magnetic
(27j19VS.L., 66j38VW.L.); and Cu – Au Farallón Negro dacitic porphyries,
Agua Rica (Cu – Mo – Au) Complex and dacites, andesites,
(27j22VS.L., 66j17VW.L.); equivalents and ignimbrites
Inca Viejo (Cu)
(25j09VS.L., 66j45VW.L.)
Farallón Negro (Au – Mn) Epithermal
(27j18VS.L., 66j40VW.L.); type
Capillitas (Pb – Zn – Ag – Cu – Au)

C.J. Chernicoff et al. / Ore Geology Reviews 21 (2002) 127–155


(27j20VS.L., 66j23VW.L.);
El Queva District (Pb – Ag – Sb)
(24j16VS.L., 66j27VW.L.);
Pairique (Au – Sn)
(22j53VS.L., 66j50VW.L.);
Chinchillas (Pb – Ag – Zn)
(22j30VS.L., 66j16VW.L.);
Concordia (Pb – Ag – Zn)
(24j12VS.L., 66j25VW.L.);
La Hoyada (Cu—Pb—Ag—Zn)
(26j55VS.L., 67j50VW.L.);
Antofalla (Ag—Zn – Pb)
(25j40VS.L., 67j50VW.L.)
La Casualidad (S) Epithermal Pliocene – Magmatic arc Inca, Tecar Dacites and Magnetic, high
(25j26VS.L., 68j21VW.L.); type Holocene Voclanic andesites with magnetic gradient
Tocomar (travertine) complexes in thin interlayed
(24j11 S.L., 66j34VW.L.) and equivalents basalts. Ignimbrites
La Salteña (Fe) El Laco
(24j05VS.L., 67j20VW.L.) type iron
deposits
Cóndor (Au) Placers Pleistocene – Alluvial and Sands and gravels No signal at
(22j25VS.L., 66j08VW.L.); Holocene colluvial deposits regional coverage
Pircas (Sn – Au)
(22j43VS.L., 66j28VW.L.)

135
136 C.J. Chernicoff et al. / Ore Geology Reviews 21 (2002) 127–155

Salfity, 1985; Chernicoff and Zappettini, 2000). Sev- margin of the Puna in the transition zone to the region
eral of these lineaments coincide with southeastward- of flat-slab subduction (26 –30jS), potassic magma-
trending volcanic breakouts from the main arc axis, tism generated the large Bajo de la Alumbrera por-
which suggest a structural control on localization of phyry Cu – Au deposit, and localized smaller
magmatism, as observed in Chile (Salfity, 1985; polymetallic deposits similar to those found to the
Schreiber and Schwab, 1991; Matteini et al., 1997; north (La Hoyada, Capillitas, Farallón Negro; Guil-
Riller et al., 2001). The relationship between linea- bert, 1985; Müller and Forrestal, 1998; Sasso and
ments, structure, and mineralization was emphasized Clark, 1998).
by Garcı́a (1969), Figueroa (1971), Baldis (1980), and
Bassi (1992).
3. Aeromagnetic survey of the Puna
2.3. Metallogeny
The Puna aeromagnetic survey is the integration of
There seems to be no clear geological reason why three different regional surveys carried out by the
the mineralization potential of the Argentine Puna Servicio Geológico Minero Argentino (SEGEMAR)
should not match that of neighbouring districts in between 1996 and 1999. Together, they comprise
northern Chile, but extensive blanketing of the Puna nearly 170,000 line kilometers and cover an area of
region by young volcanic and sedimentary rocks ren- 150,000 km2. The contractors involved in the acquis-
ders exploration difficult. Thus, remote sensing meth- ition were World Geoscience, Sander Geophysics, and
ods that can be used to detect or predict underlying Sial Geosciences. The surveys were flown with fixed-
geology are of particular value. Table 2 summarizes the wing aircraft at 1 km line spacing and 150 m nominal
location, characteristics, geological environment, and terrane clearance. Due to the rugged topography, in
magnetic signatures of significant mineral deposits of many cases, the altitude above ground was signifi-
the Puna. cantly higher than the nominal terrane clearance.
The Paleozoic basement of the Puna is host to a Gamma-ray spectrometry data were also acquired
number of different mineral deposit types, including during the surveys.
Ordovician sedimentary-exhalative (SEDEX) Pb –
Zn – Ag deposits (e.g., Aguilar and La Colorada; 3.1. Processing and interpretation of the aeromag-
Sureda and Martı́n, 1990; Zappettini and Segal, netic survey
1998) and Bendigo-type mesothermal Au vein depos-
its in metamorphosed Ordovician turbiditic sequences Analysis of the aeromagnetic survey data involved
(e.g., the Santa Catalina-Rinconada and Incahuasi close inspection of a number of layers of information
districts; Zappettini and Segal, 1999). derived from the original magnetic profiles. These
However, the greatest metallogenic diversity layers included the total magnetic intensity, total
occurred in the Cenozoic, beginning with Paleogene magnetic intensity reduced to the pole (RTP; Fig. 3),
porphyry Cu mineralization at Taca Taca (Rojas et al., analytic signal (or total gradient), and first vertical
1999; Rubinstein et al., 1999), and followed by derivative (FVD; Fig. 4) of the total magnetic inten-
Miocene epithermal Au deposits associated with felsic sity reduced to the pole. Each layer focuses on differ-
volcanism (Pairique; Coira, 1999), Bolivian-type tin ent aspects of the magnetic signatures of the region.
mineralization (Pirquitas; Chayle, 1999), Pb –Ag – Zn The total magnetic intensity data were reduced
polymetallic mineralization (Pan de Azúcar, Concor- to the pole in order to reposition the magnetic
dia, Antofalla, and El Queva; Sillitoe, 1975; Caffe and anomalies above their sources, a necessary step
Coira, 1999), Cu – Au porphyry deposits (Inca Viejo; when working in areas of low magnetic latitude
Chabert, 1999), and Au porphyries (Cerro Redondo; (Silva, 1986; Keating and Zerbo, 1996). The RTP
Caffe and Coira, 1999). Historically important native and analytic signal (total gradient) layers were
sulfur deposits also occur in fumarolic alteration zones mainly used to delineate litho-magnetic domains,
within the peaks of Miocene– Quaternary stratovolca- which constitute areas of similar magnetic signature
noes (e.g., Llullaillaco, Mina Julia). Near the southern that can be interpreted to belong to a particular
C.J. Chernicoff et al. / Ore Geology Reviews 21 (2002) 127–155 137

Fig. 3. Total magnetic intensity reduced to the pole of the Puna. Data from SEGEMAR aeromagnetic surveys, 1996 – 1999. Inset: 2-km upward
continued total magnetic intensity reduced to the pole of same area.
138 C.J. Chernicoff et al. / Ore Geology Reviews 21 (2002) 127–155

Fig. 4. First vertical derivative of the total magnetic intensity reduced to the pole of the Puna.
C.J. Chernicoff et al. / Ore Geology Reviews 21 (2002) 127–155 139

lithological unit (Fig. 5). The litho-magnetic of the region. Because these rocks are typically
domains are generally larger than the known out- exposed at the surface, this domain is well constrained
crop areas of the lithological units to which they by outcrop geology. The main axis of Cenozoic
relate, because the magnetic basement is extensively magmatism lies along the border between Chile and
covered by nonmagnetic supracrustal sediments. Argentina, and is only partially imaged here. How-
The first vertical derivative (Fig. 4) layer was ever, four narrow ( f 50 km wide) belts of volcanic
found to be particularly valuable for interpreting the rocks extend southeast from the Cordillera Occidental
basement structure of the area because this process across the Puna (Fig. 5). A similar situation has been
highlights contrasts and changes in magnetic intensity. observed in Chile (Behn, 1998), where magnetic
Various angles of illumination were applied to the ‘‘cross-anomalies’’ (i.e., transverse to the Andean
FVD and RTP layers to produce shaded relief images orogen) have been interpreted to be the track of the
that emphasized different sets of structures. In addi- eastward migration of the segmented Andean mag-
tion, several composite images (e.g., RTP/FVD, RTP/ matic arc, or the axes of structurally controlled upper
analytic signal, FVD/analytic signal) were also used to crustal intrusive complexes.
constrain the interpretations shown in Fig. 5.
3.3. Magnetic lineament interpretation
3.2. Litho-magnetic domains
The magnetic lineament framework of the study
A map of the total magnetic intensity reduced to area, as interpreted from the aeromagnetic data (Figs.
the pole of the Puna region is shown in Fig. 3. Three 3 and 4), is shown in Fig. 5. Lineaments were defined
major litho-magnetic domains can be defined within as breaks in the magnetic pattern and are regarded as
the study area (Chernicoff et al., 1996; Chernicoff and dislocations, either shallow (as shown by mapped
Zappettini, 2000): Precambrian – Early Cambrian faults/fractures) or deep-seated discontinuities (as evi-
basement, Paleozoic basement, and the Cenozoic denced by their presence in deeper ‘‘slices’’ of the
magmatic arc (Fig. 5). aeromagnetic data, as well as their length; see inset in
Precambrian – Early Cambrian basement rocks of Fig. 3). Four main groups of lineaments can be
the inferred Pampia terrane occupy the southeastern recognized from this analysis:
part of the region, which is characterized by a regional
magnetic high with low gradient; this area is trans- 3.3.1. NNE- to NE-trending lineaments
ected by three belts that belong to the Cenozoic This group is the most conspicuous in terms of
magmatic arc. lineament frequency and length, and corresponds to
The Paleozoic basement domain (Arequipa – Anto- major Cenozoic structures (faults, thrusts, fractures)
falla terrane) lies immediately to the west of the characteristic of the Argentine northwest. The corre-
Precambrian – Early Cambrian domain (Fig. 5), and spondence of several NNE-trending lineaments with
is characterized by a predominantly low magnetic the boundary between the Precambrian – Early Cam-
gradient derived from Ordovician sedimentary rocks. brian and Paleozoic domains suggests antecedents in
It is distinguished from the older basement domain to the structure of this suture zone.
the southeast by the absence of a regional magnetic
high. The Paleozoic basement domain contains all of 3.3.2. WNW- and ENE-trending lineaments
the Bendigo-type gold vein deposits (Rinconada and This group comprises structures of a broad range of
Incahuasi districts) and SEDEX-type Pb – Ag – Zn ages, from pre-Andean zones of weakness (e.g., in the
deposits (Aguilar and La Colorada mines) known in Nevados de Palermo range) to late Cenozoic trans-
the Puna, many of which are in the vicinity of long current faulting (as evidenced by the WNW twist of
NNE lineaments. the N –S-trending faults in the Olacapato area). These
The Cenozoic magmatic arc is distinguished from lineaments are interpreted to represent deep crustal
the other domains by its conspicuously high magnetic structures that have been selectively reactivated in the
gradient (Fig. 3), and an average total magnetic Cenozoic (Allmendinger et al., 1983; Alonso et al.,
intensity that is lower than the average for the rest 1984; Salfity, 1985). The intervening shorter linea-
140 C.J. Chernicoff et al. / Ore Geology Reviews 21 (2002) 127–155

Fig. 5. Litho-magnetic domains and magnetic lineaments interpreted from SEGEMAR aeromagnetic surveys of the Puna (modified from
Chernicoff and Zappettini, 2000).
C.J. Chernicoff et al. / Ore Geology Reviews 21 (2002) 127–155 141

Fig. 6. Bajo de la Alumbrera district. (A) Total magnetic intensity reduced to the pole. (B) First vertical derivative of the total magnetic intensity
reduced to the pole. (C) Analytic signal (total gradient) of the total magnetic intensity. Contours of analytic signal shown to emphasize structure.
142 C.J. Chernicoff et al. / Ore Geology Reviews 21 (2002) 127–155

ments are likely to be younger (mostly Cenozoic), mán Transfer Zone and a conspicuous set of N – S-
shallower structures. trending lineaments.

3.3.3. N – S-trending lineaments


This group is restricted mainly to the southeast of 4. Example districts
the area (within the Precambrian Pampia terrane), and
is likely to be late Cenozoic in age, as evidenced by The use of lineament distributions, and particularly
disrupted alluvial fans to the west of La Quiaca. lineament intersections, to explain or predict the
locations of magmatic or hydrothermal systems is a
3.3.4. Circular anomalies (5– 20 km diameter) well-established but not foolproof method. As
Several of these anomalies correspond to major reviewed by Richards (2000), magmatism and hydro-
Cenozoic eruptive centers. Smaller circular anomalies thermal activity may be focused to shallow crustal
are related to domes and hydrothermal systems. levels by deep-seated structures, which may in turn be
The transverse volcanic belts host the majority of recognized as lineaments or lineament corridors in
mineral deposits in the Puna, and in many cases, the geophysical or remotely sensed images. However, the
locations of these deposits coincide closely with converse is not necessarily true, i.e., an identified
structural elements of the major WNW lineament lineament does not necessarily reflect the presence
corridors (Fig. 5). For example: the Concordia, El of a major structure, and even where such structures
Queva, and La Poma epithermal districts lie along the exist, they do not cause the emplacement of magmas
Calama – Olacapato – El Toro lineament; the Mina or the passage of hydrothermal fluids—they merely
Julia, Antofalla, and Cerro Galán fumarolic and epi- offer high permeability structural pathways that may
thermal systems lie along the Archibarca lineament; be invaded by magmas or volatiles if such fluids are
and deposits in the Bajo de la Alumbrera district occur present. Furthermore, magmas and hydrothermal flu-
near the intersections between the Culampajá linea- ids are not restricted to following such structural
ment, ENE-trending lineaments related to the Tucu- pathways, but may be emplaced away from any

Fig. 7. Aeromagnetic lineament interpretation of the Bajo de la Alumbrera district. Small black dots indicate the locations of known mineral
deposits.
C.J. Chernicoff et al. / Ore Geology Reviews 21 (2002) 127–155 143

Fig. 8. Landsat TM image (A) and lineament interpretation (B) of the Bajo de la Alumbrera district, showing the locations of major mineral
deposits.
144 C.J. Chernicoff et al. / Ore Geology Reviews 21 (2002) 127–155

obvious lineament association. Nevertheless, although tens of kilometer-scale. At these scales, the relation-
a perfect correlation between identified lineaments ship is to lineament corridors (i.e., structural corri-
and the locations of volcanic centers or mineral dors), rather than individual structural elements.
deposits is rarely observed, good examples exist that Three such example districts have been chosen
display clear spatial associations at the kilometer- to from the Puna study area: Bajo de la Alumbrera,

Fig. 9. Cerro Galán district. (A) Total magnetic intensity reduced to the pole. (B) First vertical derivative of the total magnetic intensity reduced
to the pole. (C) Analytic signal (total gradient) of the total magnetic intensity. Contours of analytic signal shown to emphasize structure.
C.J. Chernicoff et al. / Ore Geology Reviews 21 (2002) 127–155 145

Cerro Galán, and El Queva. In these districts, close mented by ENE-trending lineaments, the latter repre-
spatial relationships can be seen between lineaments senting traces of the Tucumán Transfer Zone (Fig. 1).
defined by geophysical surveys and on Landsat This structural assemblage is intersected by N – S- to
images, and the locations of magmatic centers and NNW-trending lineaments, along which a smaller
hydrothermal deposits. In illustrating these examples, number of mineral deposits are aligned. Unambiguous
Landsat TM and selected geophysical images are cross-cutting relationships are not clear, although both
presented separately from lineament interpretations NNW- and ENE-trending lineaments delimit or cross-
drawn at the same scale. Note that only selected cut volcanic features in the map area, and are therefore
geophysical images are shown, but several other interpreted to represent faults active in the late Cen-
composite layers, or layers illuminated from different ozoic.
angles, were employed in the magnetic lineament A similar framework is suggested in the interpre-
interpretations. Landsat lineament interpretations were tation of the Landsat image shown in Fig. 8, where the
made at 1:250,000-scale, using spectral bands 7, 4, surface topography and the spectral response more
and 1 (assigned to red, green, and blue, respectively) clearly define traces of the NE-trending Tucumán
from scenes 231-079 (dated 13/10/86), 232-078 Transfer Zone. The locations of the Bajo de la
(dated 02/09/86), and 232-077 (dated 10/03/86). Alumbrera, Capillitas, and Agua Rica mineral depos-
its near intersections of this zone with NW- and
4.1. Bajo de la Alumbrera district WNW-trending lineaments are particularly evident.
A number of other large circular features (up to 15
The Bajo de la Alumbrera district is located at the km diameter) occur to the north of the main WNW
eastern end of the southern belt of Cenozoic magma- corridor (Fig. 7), where they are bounded and/or
tism (Fig. 5), f 200 km east of the main volcanic arc transected by WNW-, ENE-, and N –S-trending linea-
and f 500 km east of the trench where the Nazca ments. The sources for these features are covered by
plate is subducting beneath South America. This young alluvial sediments, but they may be older,
transverse volcanic belt coincides with, and locally
defines, part of the trans-Andean Culampajá linea-
ment (Figs. 1 and 5). Economic porphyry Cu –Au
mineralization at Bajo de la Alumbrera is centered on
a dacite porphyry ( f 6.9 Ma) emplaced into coeval
Late Miocene (12.6 –5.1 Ma) high-K andesites and
shoshonites of the Farallon Negro volcanic complex
(Guilbert, 1985; Müller and Forrestal, 1998; Sasso
and Clark, 1998). Sasso and Clark (1998) suggested
that magmatism was localized in a dextral pull-apart
basin along the NE-trending Tucumán Transfer Zone
(Figs. 1 and 5), which is interpreted to follow a major
basement terrane boundary (see also de Urreiztieta et
al., 1996). Lineament control on the porphyry-type
mineralization in the district was analysed by Ramos
(1977).
Inspection of the magnetic pattern of the Bajo de la
Alumbrera district (Fig. 6) reveals the occurrence of
several circular features (up to 6 km diameter) corre-
sponding to eruptive centers within a WNW-trending
corridor ( f 50 km long, 8 km wide), which hosts
most of the known mineral occurrences in the area
(Fig. 7). The structural framework of this corridor Fig. 10. Aeromagnetic lineament interpretation of the Cerro Galán
comprises a prominent WNW-trending lineament seg- district.
146 C.J. Chernicoff et al. / Ore Geology Reviews 21 (2002) 127–155

Fig. 11. Landsat TM image (A) and lineament interpretation (B) of the Cerro Galán district. Fault locations are taken from the Provincia de
Catamarca 1:500,000 geological map sheet (Servicio Geológico, 1995). The location of a prominent hydrothermal alteration zone at the caldera
margin is highlighted.
C.J. Chernicoff et al. / Ore Geology Reviews 21 (2002) 127–155 147

eroded magmatic centers. If so, they might be pro- the largest Cenozoic ignimbrite caldera in Argentina
spective for intrusion-hosted porphyry mineralization. (Fig. 5). The Cerro Galán crater (Sparks et al.,
1985) is clearly visible on the aeromagnetic maps
4.2. Cerro Galán district (Fig. 9), and is interpreted to be a nested caldera,
as indicated by three intersecting circular features.
The Cerro Galán district is located within the Additional circular features, interpreted to be small
south-central Cenozoic magmatic belt, and features satellite eruptive centers, are located around the

Fig. 12. El Queva district. (A) Total magnetic intensity reduced to the pole. (B) First vertical derivative of the total magnetic intensity reduced to
the pole. (C) Analytic signal (total gradient) of the total magnetic intensity. Contours of analytic signal shown to emphasize structure.
148 C.J. Chernicoff et al. / Ore Geology Reviews 21 (2002) 127–155

main caldera (Fig. 10). Extensive NE-, NNE-, and range is clearly depicted as an area of high mag-
NW-trending lineaments can be discerned on both netic gradients in the southeastern quadrant of the
the magnetic and Landsat images (Fig. 11), and are aeromagnetic map (Fig. 12). A prominent circular
consistent with regional structural trends. In partic- shaped area of very high magnetic gradient in the
ular, the NW-trending lineaments comprise elements right-central part of the map derives from the Aguas
of the major trans-Andean Archibarca lineament Calientes caldera.
(Fig. 1). The Landsat and magnetic lineament interpreta-
Sparks et al. (1985) reported that eruption of the tions shown in Figs. 13 and 14, respectively, are
f 1000 km3 rhyodacitic Cerro Galán ignimbrite at broadly in agreement with the mapped structural
2.1 Ma was the culmination of a volcanic episode framework, and reveal WNW-trending lineaments
that began in the Middle Miocene, and involved a intersected by NE, NW, ENE, and N –S lineaments.
significant amount of deep crustal melting (Francis The NE lineaments probably relate to the boundary
et al., 1989). Sparks et al. (1985) also conjectured between the Precambrian and Ordovician basement
that the location of the caldera was controlled by blocks, whereas the WNW and NW lineaments are
N –S-trending faults in the late Paleozoic metamor- interpreted to represent individual structural ele-
phic basement. Although some N – S lineaments are ments of the Calama –Olacapato – El Toro lineament.
indeed recognized, the caldera lies more noticeably Most of the mineral occurrences of the district fall
in the zone of intersection of regionally extensive within the zone of intersection between the Cal-
NNE- and NW-trending lineaments and mapped ama – Olacapato – El Toro corridor and NE linea-
faults (Figs. 10 and 11; Martı́nez, 1995). Hydro- ments, and many deposits are spatially related to
thermal alteration is extensively developed in the individual lineaments or eruptive centers (Figs. 13
volcanic rocks (Sparks et al., 1985), and several and 14).
mining companies have conducted mineral explora-
tion programs in the area. One particularly prom-
inent alteration zone, highlighted in Fig. 11, occurs
at the intersection of a NW-trending lineament and
the caldera rim, itself apparently controlled by a
major NNE-trending fault.

4.3. El Queva district

The El Queva district is located within the north-


central Cenozoic magmatic belt (Fig. 5), which
defines part of the SE-trending trans-Andean Cal-
ama – Olacapato– El Toro lineament (Fig. 1). Exten-
sive hydrothermal alteration, fumarolic sulfur
deposits, and minor Pb –Ag mineralization (worked
from 1968 to 1973 at the El Queva mine; Sillitoe,
1975) occur within edifices of the Nevados de
Pastos Grandes volcanic range, which have been
partially excavated by minor Pleistocene glaciation.
The volcanoes are constructed of andesitic to dacitic
lava flows and pyroclastic deposits of Upper Mio-
cene to Pliocene age. Sillitoe (1975) proposed a
spatial relationship between the locations of individ-
ual volcanic vents and the projection of a zone of Fig. 13. Aeromagnetic lineament interpretation of the El Queva
N – S-trending faults exposed in the Ordovician district. Small black dots indicate the locations of known mineral
basement. The Nevados de Pastos Grandes volcanic deposits.
C.J. Chernicoff et al. / Ore Geology Reviews 21 (2002) 127–155 149

Fig. 14. Landsat TM image (A) and lineament interpretation (B) of the El Queva district, showing the locations of prominent hydrothermal
alteration zones within the volcanic edifices of the Nevados de Pastos Grandes volcanic range.
150 C.J. Chernicoff et al. / Ore Geology Reviews 21 (2002) 127–155

5. Discussion orientation and extent of these breakouts has been


suspected for some time (e.g., Jordan et al., 1983;
In the foregoing sections, we have firstly estab- Salfity, 1985), but such structures are hard to identify
lished the regional-scale geophysical and interpreted in the field because they have largely been covered by
structural framework of the Puna, and have then the volcanic rocks that they have helped to localize.
focused on three detailed areas within that region at Blanketing of the pre-Cenozoic basement by such
a scale at which spatial relationships between linea- materials is not complete, however, and on a regional
ments and individual magmatic and hydrothermal scale, the trends of basement structures become visi-
centers can be observed. This combination of ble and can be projected beneath the younger succes-
regional- and local-scale analyses is essential, because sions. Thus, in Fig. 5, strong WNW structural trends
local details are lost on the small-scale regional maps, can be recognized, several of which correlate with the
whereas the significance of regional trends and linea- transverse belts of late Cenozoic volcanism. At least
ment corridors is often not clear within the bounds of three of these structural zones are of such intensity
the large-scale local maps. The combination of differ- and lateral persistence that they have been named: the
ent methodologies (geophysical and Landsat) is also Calama –Olacapato – El Toro, Archibarca, and Culam-
important because these techniques record different pajá lineaments (Figs. 1 and 5).
characteristics and contrasts of the surface and subsur- Studies of exposed fault strands along these WNW
face geology. In particular, Landsat images record lineaments suggest that, at least during the late Cen-
only features visible at the Earth’s surface, whereas ozoic, they have behaved as strike –slip faults, mostly
the aeromagnetic method is sensitive to magnetic with limited sinistral offsets ( f 20 km for the El Toro
contrasts throughout the crust. Therefore, lineaments fault; Allmendinger et al., 1982, 1983; Jordan et al.,
visible in both data sets, and in particular, those of 1983; Salfity, 1985; Matteini et al., 1997). Coutand et
significant linear extent (i.e., comparable to the crustal al. (2001) have referred to these structures as transfer
thickness of 50 –70 km), are interpreted to represent faults, accommodating shear strain between large
major, through-going crustal structures. crustal blocks during Neogene crustal shortening.
Richards (2000) and Tosdal and Richards (2001) Other structural trends are also visible in the geo-
recently reviewed evidence for a relationship between physical and Landsat images, including a prominent
deep crustal structures and the emplacement of deeply corridor of NNE-trending lineaments that broadly
derived magmas, and concluded that while such corresponds to the suture zone between the Are-
structures may provide weak or permeable pathways quipa –Antofalla and Pampia terranes, and numerous
for magma ascent, their mere existence does not in but less extensive N –S structures.
itself result in intrusion. Instead, the tectono-magmatic In order to explore in more detail the spatial
history of the region also needs to be considered in relationship between magmatism, mineralization, and
order to satisfy both the necessity for an active magma deep crustal structure, three areas in which this
source, and for the existence of a conducive stress association is clear both at a regional and local scale
field acting on the structural framework to promote were selected for closer examination: Bajo de la
permeability. Alumbrera (containing the largest porphyry Cu –Au
Supracrustal magmatism in each of the three study deposit in Argentina), Cerro Galán (the largest late
areas is related to the late Cenozoic Andean orogeny, Cenozoic caldera complex in Argentina), and El
which occurred in response to relatively rapid and Queva (a historically important polymetallic district
shallow subduction of the Nazca plate beneath South hosted by a major volcanic complex). In each case,
America. The main focus of arc magmatism during these centers are found to lie within the zones of
this period was, and continues to be, the Cordillera intersection of major crustal lineaments, most notably
Occidental, which forms the western border of the the WNW trans-Andean set (all locations), the NNE-
Puna. However, coeval breakouts of volcanism and to NE-suture-related set (El Queva and Cerro Galán),
associated shallow-level plutonism extended as fin- and the ENE-trending Tucumán Transfer Zone (Bajo
gers to the SE across the Puna; four such fingers are de la Alumbrera). Relatively discontinuous N – S-
identified in Fig. 5. A deep structural control on the trending lineaments are also identified in several
C.J. Chernicoff et al. / Ore Geology Reviews 21 (2002) 127–155 151

locations, and may be conjugate to WNW- or NE- unfavourable for other, predictable reasons, such
trending structures. as that they lie off the magmatic axis, that the local
The favourability of lineament intersections for structural framework was misoriented with respect
focusing magma ascent has frequently been noted in to the prevailing stress field at the time of potential
the past (e.g., Rehrig and Heidrick, 1972; Bussell, mineralization, or that local conditions were inap-
1976; Pitcher, 1979; Sylvester and Linke, 1993; propriate (e.g., nonprospective magma chemistry,
Rowley et al., 1998). In particular, development of nonideal depth of emplacement, catastrophic erup-
extensional pull-apart volumes that may serve to tion, etc.). In addition, the region might not yet
focus magma emplacement is favoured at step- have been explored in sufficient detail (much of the
overs, bends, and structural intersections along Puna has received only reconnaissance exploration).
strike –slip faults during transtensional or transpres- Thus, more geological information must be brought
sional deformation (Bellier and Sébrier, 1994; Tom- to bear in making exploration decisions, rather than
masi et al., 1994; Tobisch and Cruden, 1995; simply looking for lineaments and their intersec-
Román-Berdiel et al., 1997; Benn et al., 2000; tions.
Riller et al., 2001). From a metallogenic perspec- Finally, it is important to remember that the
tive, areas of enhanced magmatic flux also represent definition of mineralization, and particularly of
zones of increased prospectivity for magmatic- ore, is an economic one: potentially useful minerals
related hydrothermal ore deposits, and recognition may be present, but either at concentrations too low
of such centers therefore has significance for min- to be considered interesting, or at depths too great
eral exploration. to be recovered economically today. In the young
Our analysis of three major magmatic centers in volcanic belts of the Puna, the latter problem is
the Puna, each associated with an ore deposit and/or particularly serious because erosion has been mini-
large hydrothermal alteration zone, suggests a spa- mal, such that subsurface porphyry and even near-
tial relationship between magmatic – hydrothermal surface epithermal mineral deposits have barely
activity and the intersections of deep-seated base- been exposed. Surficial fumarolic alteration zones
ment structures. This relationship may be further are widespread in these volcanic edifices, however,
rationalized in terms of structural weakness and and may overlie mineralized systems at depth (e.g.,
permeability, permitting focused ascent of magma at Cerro Galán). From an exploration perspective,
to shallow crustal levels where devolatilization may therefore, it will be helpful to identify systems that
produce hydrothermal ore deposits. However, con- have been sufficiently dissected either by erosion or
version of this relationship into an exploration edifice collapse (e.g., Sillitoe, 1994) to bring poten-
model is complicated by the fact that structural tially mineralized zones within range of surface
architecture does not in itself generate or localize exploration and development.
magmatism—as noted above, it merely provides a
favourable environment for magma ascent when a
magma source is present within a suitably oriented 6. Conclusions
regional stress field. Furthermore, magmatism and
mineralization are not precluded from forming away Aeromagnetic and Landsat TM data have been
from such intersections, although it might be used to investigate a spatial relationship between
expected that the largest centers would indeed be lineaments, lineament intersections, and large mag-
localized by zones of maximum structural favour- matic and hydrothermal alteration centers. Many of
ability. As can be seen in Fig. 5, many intersections the identified lineaments and lineament zones are
between individual lineament strands are evident interpreted to represent faults, some of trans-crustal
throughout the Puna region, and many are not extent. Observation at, and correlation between,
directly associated with recorded mineralization. both regional- and local-scale perspectives is impor-
This lack of perfect correlation could be viewed tant for the recognition of significant controlling
as a limitation of the model for exploration pur- structures in magma emplacement and hydrothermal
poses. However, such ‘‘barren’’ loci might be activity, because the dimensions of regional linea-
152 C.J. Chernicoff et al. / Ore Geology Reviews 21 (2002) 127–155

ments are such that they are commonly not clearly Canada. We thank Daniel Knepper and Victor Ramos
resolved within a large-scale local map area. For for helpful and constructive reviews.
example, the extensive trans-Andean lineaments are
rarely defined by single structural elements, but
instead by zones of faulting, often tens of kilo- References
meters wide. Thus, ENE-trending lineaments that
are interpreted as elements of the Tucumán Transfer Allmendinger, R.W., 1986. Tectonic development, southeastern bor-
Zone at Bajo de la Alumbrera are spaced over a der of the Puna Plateau, northwestern Argentine Andes. Geo-
width of f 35 km, and WNW-trending elements of logical Society of America Bulletin 97, 1070 – 1082.
Allmendinger, R.W., Jordan, T.E., Palma, M., Ramos, V.A., 1982.
the Calama – Olacapato – El Toro lineament at El
Perfil estructural en la Puna Catamarqueña. Quinto Congreso
Queva form a corridor f 40 km wide. In terms Latinoamericano de Geologia, Buenos Aires, Argentina, Actas
of localizing magmatic or hydrothermal centers, it is 1, pp. 499 – 518.
these broad zones of intersection that are important Allmendinger, R.W., Ramos, V.A., Jordan, T.E., Palma, M., Isacks,
rather than the intersections of specific structural B.L., 1983. Paleogeography and Andean structural geometry,
northwest Argentina. Tectonics 2, 1 – 16.
elements, which individually may have limited lat-
Allmendinger, R.W., Strecker, M., Eremchuk, J.E., Francis, P., 1989.
eral and vertical extent. Neotectonic deformation of the southern Puna Plateau, north-
Use of lineament intersection relationships for western Argentina. Journal of South American Earth Sciences 2,
mineral exploration gains validity when defined by 111 – 130.
multi-technique approaches, such as combinations Allmendinger, R.W., Jordan, T.E., Kay, S.M., Isacks, B.L., 1997.
The evolution of the Altiplano – Puna plateau of the Central
of remote sensing, geophysical, and geological
Andes. Annual Review of Earth and Planetary Sciences 25,
methods. In addition, applicability can be tested 139 – 174.
by inclusion of known metallogenic information, Alonso, R.N., Viramonte, J., Gutiérrz, R., 1984. Puna Austral—
which may help to identify favourable structural bases para el subprovincialismo geológico de la Puna. Actas,
settings in a given regional context. Crustal-scale Noveno Congreso Geológico Argentino. S.C. Bariloche, Argen-
tina, pp. 43 – 63.
structural controls on magma emplacement are just
Alonso, R.N., Jordan, T.E., Tabbutt, K.T., Vandervoort, D.S., 1991.
one of a number of independent factors that affect Giant evaporite belts of the Neogene central Andes. Geology 19,
the potential to form an ore deposit, and so the 401 – 404.
identification of an apparently favourable lineament Baby, P., Rochat, P., Mascle, G., Hérail, G., 1997. Neogene short-
intersection does not guarantee prospectivity. Sim- ening contribution to crustal thickening in the back arc of the
Central Andes. Geology 25, 883 – 886.
ilarly, this model does not preclude the occurrence
Bahlburg, H., Breikreuz, C., 1991. Paleozoic evolution of active
of mineral deposits unrelated to lineament intersec- margin basins in the southern Central Andes (northwestern Ar-
tions. Nevertheless, large deposits typically require gentina and northern Chile). Journal of South American Earth
optimum conditions for their formation, and well- Sciences 4, 171 – 188.
defined intersections that offer structural favourabil- Bahlburg, H., Hervé, F., 1997. Geodynamic evolution and tectonos-
tratigraphic terranes of northwestern Argentina and northern
ity for magmatism will have an increased probability
Chile. Geological Society of America, Bulletin 109, 869 – 884.
for hosting significant ore deposits. We conclude that Baker, R.C., Guilbert, J.M., 1987. Regional structural control of
lineament analysis is a useful tool for first-pass target porphyry copper deposits in northern Chile. Geological Society
selection, but priority rankings should include addi- of America, Abstracts with Programs 19, 578.
tional geological and metallogenic information. Baldis, B., 1980. Control megaestructural de los Distritos metal-
ı́feros en el Noroeste de Argentina. In Metalogénesis en Latino-
américa, Publicación IUGS 5.
Bassi, H.G.L., 1992. Hypothesis concerning a regmagenic network
Acknowledgements controlling metallogenic and other geologic events in the
South American austral cone. Geologische Rundschau 77,
The authors wish to thank the Servicio Geológico 491 – 511.
Beck, S.L., Zandt, G., Myers, S.C., Wallace, T.C., Silver, P.G.,
Minero Argentino (SEGEMAR) for financial support
Drake, L., 1996. Crustal-thickness variations in the central An-
and for permission to use the geophysical data. JPR des. Geology 24, 407 – 410.
acknowledges the support of a grant from the Natural Behn, G., 1998. Respuestas geofı́sicas de yacimientos tipo cobre
Sciences and Engineering Research Council of porfı́rico. X Congreso Latinoamericano de Geologı́a, Simposio
C.J. Chernicoff et al. / Ore Geology Reviews 21 (2002) 127–155 153

III. Geofı́sica aérea y Geoquı́mica en la prospección geológico- tectonics. Geological Society of London Special Publication 19,
minera, Buenos Aires, Servicio Geológico-Minero Argentino, pp. 389 – 404.
Anales, vol. 31, pp. 9 – 20. Damm, K.-W., Pichowiak, S., Harmon, R.S., Todt, W., Kelley, S.,
Bellier, O., Sébrier, M., 1994. Relationship between tectonism and Omarini, R., Niemeyer, H., 1990. Pre-Mesozoic evolution of the
volcanism along the Great Sumatran Fault Zone deduced by central Andes; the basement revisited. In: Kay, S.M., Rapela,
SPOT image analysis. Tectonophysics 233, 215 – 231. C.W. (Eds.), Plutonism from Antarctica to Alaska. Boulder,
Benn, K., Odonne, F., Lee, S.K.Y., Darcovich, K., 2000. Analogue Colorado, Geological Society of America Special Paper 241,
scale models of pluton emplacement during transpression in pp. 101 – 126.
brittle and ductile crust. Transactions of the Royal Society of Damm, K.-W., Harmon, R.S., Kelley, S., 1994. Some isotopic and
Edinburgh, Earth Sciences 91, 111 – 121. geochemical constraints on the origin and evolution of the Cen-
Bussell, M.A., 1976. Fracture control of high-level plutonic contacts tral Andean basement (19j – 24jS). In: Reutter, K.J., Scheuber,
in the Coastal Batholith of Peru. Proceedings of the Geologists E., Wigger, P.J. (Eds.), Tectonics of the Southern Central Andes;
Association 87, 237 – 246. Structure and Evolution of an Active Continental Margin.
Caffe, P., Coira, B.L., 1999. Complejos de domos volcánicos del Springer-Verlag, Berlin, pp. 263 – 276.
Mioceno Medio de Puna Norte. Un modelo geológico y metal- Davidson, C., Schmid, S.M., Hollister, L.S., 1994. Role of melt
ogenético para yacimientos epitermales de metales de base during deformation in the deep crust. Terra Nova 6, 133 – 142.
ricos en plata (estaño). In: Zappettini, E.O. (Ed.), Recursos de Urreiztieta, M., Gapais, D., Le Corre, C., Cobbold, P.R., Rossel-
Minerales de la República Argentina, Anales, vol. 35. SEGE- lo, E., 1996. Cenozoic dextral transpression and basin develop-
MAR, Instituto de Geologı́a y Recursos Minerales, Buenos ment at the southern edge of the Puna Plateau, northwestern
Aires, pp. 1569 – 1578. Argentina. Tectonophysics 254, 17 – 39.
Chabert, M.R., 1999. El pórfiro cuprı́fero Inca Viejo, Salta. In: Dorbath, C., Masson, F., 2000. Composition of the crust and upper-
Zappettini, E.O. (Ed.), Recursos Minerales de la República Ar- mantle in the Central Andes (19j30VS) inferred from P wave
gentina, Anales, vol. 35. SEGEMAR, Instituto de Geologı́a y velocity and Poisson’s ratio. Tectonophysics 327, 213 – 223.
Recursos Minerales, Buenos Aires, pp. 1429 – 1436. Figueroa, L.A., 1971. Fotolineamientos y mineralización en el Nor-
Chayle, W., 1999. Mina Pirquitas, Jujuy. In: Zappettini, E.O. (Ed.), oeste Argentino. 1er Simposio Nacional de Geologı́a Económ-
Recursos Minerales de la República Argentina, Anales, vol. 35. ica, San Juan 1, pp. 107 – 124.
SEGEMAR, Instituto de Geologı́a y Recursos Minerales, Bue- Francis, P.W., Sparks, R.S.J., Hawkesworth, C.J., Thorpe, R.S.,
nos Aires, pp. 1593 – 1598. Pyle, D.M., Tait, S.R., Mantovani, M.S., McDermott, F.,
Chernicoff, C.J., Zappettini, E.O., 2000. Interpretación geológica – 1989. Petrology and geochemistry of volcanic rocks of the Cer-
metalogénica del levantamiento aeromagnétio de la Puna, Ar- ro Galan caldera, northwest Argentina. Geological Magazine
gentina. 9j Congreso Geológico Chileno, Puerto Varas, Actas, 126, 515 – 547.
Vol. 2, Simposio Nacional No. 3. Sociedad Geológica de Chile, Garcı́a, H., 1969. Consideraciones sobre algunas alineaciones de
Puerto Varas, Chile, pp. 277 – 280. desarrollos hidrotermales tipo pórfido cuprı́fero en el noroeste
Chernicoff, C.J., Garea, G., Hongn, F., Seggiaro, R., Zappettini, E., argentino. Dirección Nacional de Geologı́a y Minerı́a, Revista
Coira, B., Caffe, P., Chayle, W., Gutierrez, G., Perez, A., Soler, 18, 55 – 66.
M., Rankin, L., 1996. Interpretación geológica del relevamiento Guilbert, J.M., 1985. Geology, alteration, mineralization, and gen-
aeromagnético de la Puna septentrional, Jujuy y Salta. Serie esis of the Bajo de la Alumbrera porphyry copper – gold de-
Contribuciones Técnicas, Geofı́sica, vol. I. Dirección Nacional posit, Catamarca province, Argentina. In: Pierce, F.W., Bolm,
del Servicio Geológico, Buenos Aires, 46 pp. J.G. (Eds.), Porphyry copper deposits of the American Cor-
Coira, B.L., 1999. Potencialidad minera de sistemas megacaldéricos dillera. Tucson, Arizona Geological Society Digest 20, pp.
miocenos en Puna norte, Jujuy. In: Zappettini, E.O. (Ed.), Re- 646 – 656.
cursos Minerales de la República Argentina, Anales, vol. 35. Hammerschmidt, K., Döbel, R., Friedrichsen, H., 1992. Implica-
SEGEMAR, Instituto de Geologı́a y Recursos Minerales, Bue- tions of 40Ar/39Ar dating of Early Tertiary volcanic rocks from
nos Aires, pp. 1557 – 1567. the north-Chilean Precordillera. Tectonophysics 202, 55 – 81.
Coira, B., Davidson, J., Mpodozis, C., Ramos, V., 1982. Tectonic Heyl, A.V., 1972. The 38th parallel lineament and its relationship to
and magmatic evolution of the Andes of northern Argentina and ore deposits. Economic Geology 67, 879 – 894.
Chile. Earth-Science Reviews 18, 303 – 332. Holdsworth, R.E., Butler, C.A., Roberts, A.M., 1997. The recogni-
Coira, B., Kay, S.M., Viramonte, J., 1993. Upper Cenozoic mag- tion of reactivation during continental deformation. Journal of
matic evolution of the Argentine Puna—a model for chang- the Geological Society (London) 154, 73 – 78.
ing subduction geometry. International Geology Review 35, James, D.E., Sack, I.S., 1999. Cenozoic formation of the Central
677 – 720. Andes: a geophysical perspective. In: Skinner, B.J. (Ed.), Geol-
Coutand, I., Cobbold, P.R., de Urreiztieta, M., Gautier, P., Chauvin, ogy and ore deposits of the Central Andes. Society of Economic
A., Gapais, D., Rossello, E.A., López-Gamundı́, O., 2001. Style Geologists, Special Publication 7, pp. 1 – 25.
and history of Andean deformation, Puna plateau, northwestern Jordan, T.E., Alonso, R.N., 1987. Cenozoic stratigraphy and basin
Argentina. Tectonics 20, 210 – 234. tectonics of the Andes mountains, 20j – 28j south latitude.
Dalziel, I.W.D., 1986. Collision and Cordilleran orogenesis: an An- American Association of Petroleum Geologists Bulletin 71,
dean perspective. In: Coward, M.P., Ries, A.C. (Eds.), Collision 49 – 64.
154 C.J. Chernicoff et al. / Ore Geology Reviews 21 (2002) 127–155

Jordan, T.E., Gardeweg, M., 1989. Tectonic evolution of the Late and Mineral Resources. Houston, Texas, Circum-Pacific Coun-
Cenozoic central Andes (20j – 33jS). In: Ben-Avraham, Z. cil for Energy and Mineral Resources Earth Science Series 11,
(Ed.), Oxford Monographs on Geology and Geophysics 8. pp. 59 – 90.
The Evolution of the Pacific Ocean Margins. Clarendon Press, Müller, D., Forrestal, P., 1998. The shoshonite porphyry Cu – Au
Oxford, pp. 193 – 207. association at Bajo de la Alumbrera, Catamarca Province, Ar-
Jordan, T.E., Isacks, B.L., Allmendinger, R.W., Brewer, J.A., Ra- gentina. Mineralogy and Petrology 64, 47 – 64.
mos, V.A., Ando, C.J., 1983. Andean tectonics related to geom- O’Leary, D.W., Friedman, J.D., Pohn, H.A., 1976. Lineament, lin-
etry of subducted Nazca plate. Geological Society of America ear, lineation: some proposed new standards for old terms. Geo-
Bulletin 94, 341 – 361. logical Society of America Bulletin 87, 1463 – 1469.
Kay, R.W., Kay, S.M., 1993. Delamination and delamination mag- Page, S., Zappettini, E., 1999. El magmatismo gondwánico Provin-
matism. Tectonophysics 219, 177 – 189. cias de Jujuy, Salta, Tucumán y Catamarca, República Argenti-
Keating, P., Zerbo, L., 1996. An improved technique for reduction na. XIV Congreso Geológico Argentino, Relatorio, Salta, pp.
to the pole at low latitudes. Geophysics 61, 131 – 137. 120 – 142.
Kleinert, K., Strecker, M.R., 2001. Climate change in response to Palacios, C.M., Townley, B.C., Lahsen, A.A., Egaña, A.M., 1993.
orographic barrier uplift: paleosol and stable isotope evidence Geological development and mineralization in the Atacama seg-
from the late Neogene Santa Marı́a basin, northwestern Argen- ment of the South American Andes, northern Chile (26j15V –
tina. Geological Society of America Bulletin 113, 728 – 742. 27j25VS). Geologische Rundschau 82, 652 – 662.
Koukharsky, M., 1985. Caracterización petrológica de ignimbritas Palma, M.A., Parica, P.D., Ramos, V.A., 1986. El granito Archibar-
Cenozoicas de la Puna Argentina (Petrologic characterization of ca: Su edad y significado tectonico, Provincia de Catamarca.
Cenozoic ignimbrites of Puna, Argentina). Actas, IV Congreso Associacion Geologica Argentina, Revista 41, 414 – 419.
Geologico Chileno, Antofagasta, Chile, Aug. 19 – 24. Universi- Pankhurst, R.J., Rapela, C.W., 1998. The proto-Andean margin of
dad del Norte Chile, Antofagasta, Chile 4, pp. 254 – 274. Gondwana: an introduction. In: Pankhurst, R.J., Rapela, C.W.
Kraemer, B., Adelmann, D., Alten, M., Schnurr, W., Erpenstein, K., (Eds.), The Proto-Andean Margin of Gondwana. Geological
Kiefer, E., van den Bogaard, P., Görler, K., 1999. Incorporation Society of London, Special Publications 142, pp. 1 – 9.
of the Paleogene foreland into the Neogene Puna plateau: the Pankhurst, R.J., Rapela, C.W., Saavedra, J., Baldo, E., Dahlquist, J.,
Salar de Antofalla area, NW Argentina. Journal of South Amer- Pascua, I., Fanning, C.M., 1998. The Famatinian magmatic
ican Earth Sciences 12, 157 – 182. arc in the central Sierras Pampeanas: an early to mid-Ordovi-
Lindsay, D.D., Zentilli, M., Rojas de la Rivera, J., 1995. Evolution cian continental arc on the Gondwana margin. In: Pankhurst,
of an active ductile to brittle shear system controlling mineral- R.J., Rapela, C.W. (Eds.), The proto-Andean margin of Gond-
ization at the Chuquicamata porphyry copper deposit, northern wana. Geological Society of London, Special Publications 142,
Chile. International Geology Review 37, 945 – 958. pp. 343 – 367.
Lucassen, F., Becchio, R., Wilke, H.G., Franz, G., Thirlwall, M.F., Pardo-Casas, F., Molnar, P., 1987. Relative motion of the Nazca
Viramonte, J., Wemmer, K., 2000. Proterozoic – Paleozoic de- (Farallon) and South American plates since Late Cretaceous
velopment of the basement of the Central Andes (18 – 26jS)—a time. Tectonics 6, 233 – 248.
mobile belt of the South American craton. Journal of South Pitcher, W.S., 1979. The nature, ascent and emplacement of gran-
American Earth Sciences 13, 697 – 715. itic magmas. Journal of the Geological Society (London) 136,
Marrett, R.A., Emerman, S.H., 1992. The relations between faulting 627 – 662.
and mafic magmatism in the Altiplano – Puna plateau (central Ramos, V., 1977. Basement tectonics from Landsat imagery in min-
Andes). Earth and Planetary Science Letters 112, 53 – 59. ing exploration. Journal of the Royal Geological and Minera-
Marrett, R.A., Allmendinger, R.W., Alonso, R.N., Drake, R.E., logical Society (The Netherlands) 56 (3), 243 – 256.
1994. Late Cenozoic tectonic evolution of the Puna Plateau Ramos, V., 1999. Rasgos estructurales del territorio Argentino.
and adjacent foreland, northwestern Argentine Andes. Journal In: Caminos, R. (Ed.), Geologı́a Argentina, Anales, vol. 29.
of South American Earth Sciences 7, 179 – 207. Instituto de Geologı́a y Recursos Minerales, Buenos Aires,
Martı́nez, L., 1995. Mapa geológico de la Provincia de Catamarca: pp. 715 – 784.
1:500,000 map sheet. Servicio Geológico-Minero Argentino. Rehrig, W.A., Heidrick, T.L., 1972. Regional fracturing in Laramide
Matteini, M., Mazzuoli, R., Omarini, R., 1997. The Volcanism stocks of Arizona and its relationship to porphyry copper min-
Along the Calama – Olacapato – El Toro Transversal Fault Sys- eralization. Economic Geology 67, 198 – 213.
tem in the Central Andes: The Tultul, del Dedio and Pocitos Richards, J.P., 2000. Lineaments revisited. Society of Economic
Volcanoes (Puna, Argentina). Universidad Católica del Norte, Geologists Newsletter 42 (1), 14 – 20.
VIII Congreso Geológico Chileno, pp. 159 – 163. Richards, J.P., Boyce, A.J., Pringle, M.S., 2001. Geological evolu-
Mon, R., Salfity, J.A., 1995. Tectonic evolution of the Andes of tion of the Escondida area, northern Chile: a model for spatial
northern Argentina. In: Tankard, A.J., Suárez, R., Welsink, H.J. and temporal localization of porphyry Cu mineralization. Eco-
(Eds.), Petroleum basins of South America. AAPG Memoir 62, nomic Geology 96, 271 – 305.
pp. 269 – 283. Riller, U., Petrinovic, I., Ramelow, J., Strecker, M., Oncken, O.,
Mpodozis, C., Ramos, V., 1989. The Andes of Chile and Argentina. 2001. Late Cenozoic tectonism, collapse caldera and plateau
In: Ericksen, G.E., Cañas Pinochet, M.T., Reinemund, J.A. formation in the central Andes. Earth and Planetary Science
(Eds.), Geology of the Andes and its Relation to Hydrocarbon Letters 188, 299 – 311.
C.J. Chernicoff et al. / Ore Geology Reviews 21 (2002) 127–155 155

Rojas, N., Drobe, J., Lane, R., Bonafede, D., 1999. El pórfiro cu- building in the central Andean region. Journal of South Amer-
prı́fero de Taca Taca Bajo, Salta. In: Zappettini, E.O. (Ed.), ican Earth Sciences 11, 211 – 215.
Recursos Minerales de la República Argentina, Anales, vol. Sparks, R.S.J., Francis, P.W., Hamer, R.D., Pankhurst, R.J., O’Cal-
35. SEGEMAR, Instituto de Geologı́a y Recursos Minerales, laghan, L.O., Thorpe, R.S., Page, R., 1985. Ignimbrites of the
Buenos Aires, pp. 1321 – 1331. Cerro Galán caldera, NW Argentina. Journal of Volcanology and
Román-Berdiel, T., Gapais, D., Brun, J.-P., 1997. Granite intrusion Geothermal Research 24, 205 – 248.
along strike – slip zones in experiment and nature. American Sureda, R.J., Martı́n, J.L., 1990. El Aguilar mine: an Ordovician
Journal of Science 297, 651 – 678. sediment-hosted stratiform lead – zinc deposit in the central An-
Rowley, P.D., Cunningham, C.G., Steven, T.A., Mehnert, H.H., des. In: Fontboté, L., Amstutz, G.C., Cardozo, M., Cedillo, E.,
Naeser, C.W., 1998. Cenozoic igneous and tectonic setting of Frutos, J. (Eds.), Stratabound Ore Deposits in the Andes.
the Marysvale volcanic field and its relation to other igneous Springer-Verlag, Berlin, pp. 161 – 174.
centers in Utah and Nevada. In: Friedman, J.D., Huffman, A.C. Sylvester, H., Linke, M., 1993. Structural control of intrusions and
(coordinators), Laccolith Complexes of Southeastern Utah: hydrothermal alteration zones by intersecting fault systems in
Time of Emplacement and Tectonic Setting—Workshop Pro- the Cretaceous magmatic arc of the southern Central Andes at
ceedings. U.S. Geological Survey Bulletin 2158, pp. 167 – 201. 27jS: III. Region, Chile. Zentralblattes für Geologie und Pal-
Rubinstein, N.A., Segal, S.J., Zappettini, E.O., 1999. El pórfiro äontologie Teil I, Heft 1/2, 361 – 376.
cuprı́fero de Taca Taca Alto, Salta. In: Zappettini, E.O. (Ed.), Thorpe, R.S., 1984. The tectonic setting of active Andean volcan-
Recursos Minerales de la República Argentina, Anales, vol. 35. ism. In: Harmon, R.S., Barreiro, B.A. (Eds.), Andean Magma-
SEGEMAR, Instituto de Geologı́a y Recursos Minerales, Bue- tism Chemical and Isotopic Constraints. Nantwich, Cheshire,
nos Aires, pp. 1333 – 1336. Shiva, pp. 4 – 8.
Salfity, J.A., 1985. Lineamientos transversales al rumbo andino en Tobisch, O.T., Cruden, A.R., 1995. Fracture-controlled magma con-
el noroeste Argentino. IV Congreso Geológico Chileno. Anto- duits in an obliquely convergent continental magmatic arc.
fagasta, Chile, Part 2, pp. 119 – 137. Geology 23, 941 – 944.
Salfity, J.A., Gorustovich, S.A., 1998. The geological evolution of Tomlinson, A.J., Blanco, N., 1997. Structural Evolution and Dis-
the province of Salta (Argentina) and neighboring regions. Min- placement History of the West Fault System, Precordillera,
istry of Production and Employment, Secretariat of Mining, Chile: Part 1. Synmineral History. Universidad Católica del
Industry and Energy Resources, Salta, Argentina, atlas of 67 Norte, VIII Congreso Geológico Chileno, pp. 1873 – 1877.
Figures with explanations. Tommasi, A., Vauchez, A., Fernandes, L.A.D., Porcher, C.C.,
Salfity, J., Marquillas, R., 1999. La Cuenca Cretácico-Terciaria del 1994. Magma-assisted strain localization in an orogen-parallel
Norte Argentino. In: Caminos, R. (Ed.), Geologı́a Argentina, transcurrent shear zone of southern Brazil. Tectonics 13,
Anales, vol. 29. Instituto de Geologı́a y Recursos Minerales, 421 – 437.
Buenos Aires, pp. 613 – 626. Tosdal, R.M., 1996. The Amazon – Laurentian connection as viewed
Sasso, A.M., Clark, A.H., 1998. The Farallón Negro Group, north- from the Middle Proterozoic rocks in the Central Andes western
west Argentina: magmatic, hydrothermal and tectonic evolution Bolivia and Northern Chile. Tectonics 15, 827 – 842.
and implications for Cu – Au metallogeny in the Andean back- Tosdal, R.M., Richards, J.P., 2001. Magmatic and structural controls
arc. SEG Newsletter 34, 1 and 8 – 18. on the development of porphyry Cu F Mo F Au deposits. In:
Schreiber, U., Schwab, K., 1991. Geochemistry of quaternary Richards, J.P., Tosdal, R.M. (Eds.), Structural Controls on Ore
shoshonitic lavas related to the Calama – Olacapato – El Toro Genesis. Society of Economic Geologists, Reviews in Economic
Lineament, NW Argentina. Journal of South American Earth Geology 14, pp. 157 – 181.
Sciences 4, 73 – 85. Vandervoort, D.S., Jordan, T.E., Zeitler, P.K., Alonso, R.N., 1995.
Siebel, W., Schnurr, W.B.W., Hahne, K., Kraemer, B., Trumbull, Chronology of internal drainage development and uplift,
R.B., van den Bogaard, P., Emmermann, R., 2001. Geochemis- southern Puna plateau, Argentina central Andes. Geology 23,
try and isotope systematics of small- to medium-volume Neo- 145 – 148.
gene – Quaternary ignimbrites in the southern central Andes: Vigneresse, J.L., Tikoff, B., 1999. Strain partitioning during partial
evidence for derivation from andesitic magma sources. Chem- melting and crystallizing felsic magmas. Tectonophysics 312,
ical Geology 171, 213 – 237. 117 – 132.
Sillitoe, R.H., 1975. Lead – silver, manganese, and native sulfur Zappettini, E.O., Segal, S.J., 1998. El depósito polimetálico La
mineralization within a stratovolcano, El Queva, northwest Ar- Colorada (Salta, Argentina): Un cuerpo de sulfuros masivos
gentina. Economic Geology 70, 1190 – 1201. de filiación sedex. X Congreso Latinoamericano de Geologı́a
Sillitoe, R.H., 1994. Erosion and collapse of volcanoes: causes of y VI Congreso Nacional de Geologı́a Económica, Buenos Aires,
telescoping in intrusion-centered ore deposits. Geology 22, Actas 3, pp. 200 – 206.
945 – 948. Zappettini, E.O., Segal, S.J., 1999. Los depósitos aurı́feros veti-
Silva, J.B.C., 1986. Reduction to the pole as an inverse problem formes de la Sierra de Rinconada, Jujuy. In: Zappettini, E.O.
and its application to low-latitude anomalies. Geophysics 51, (Ed.), Recursos Minerales de la República Argentina, Anales,
369 – 382. vol. 35. SEGEMAR, Instituto de Geologı́a y Recursos Miner-
Somoza, R., 1998. Updated Nazca (Farallon) – South America rela- ales, Buenos Aires, pp. 507 – 514.
tive motions during the last 40 My: implications for mountain

You might also like