You are on page 1of 61

Chemie der Erde xxx (xxxx) xxx–xxx

Contents lists available at ScienceDirect

Chemie der Erde


journal homepage: www.elsevier.com/locate/chemer

Invited review

Impact cratering: The South American record – Part 1



A.P. Cróstaa, , W.U. Reimoldb,c, M.A.R. Vasconcelosd, N. Hauserb, G.J.G. Oliveiraa,
M.V. Mazivieroa, A.M. Góese
a
State University of Campinas, Brazil
b
University of Brasília, Brazil
c
Natural History Museum – Leibniz Institute for Evolution and Biodiversity Science, Berlin, Germany
d
Federal University of Bahia, Brazil
e
University of São Paulo, Brazil

A R T I C LE I N FO A B S T R A C T

Keywords: The Earth’s impact record is known to be rather limited in both time and space. There are ca. 190 impact
Impact crater structures currently known on Earth, representing a minor fraction of all the impact events that contributed to
Impact structure the initial formation of our protoplanet, and then to formation and modification of the surface of the planet.
Impact record Moreover, the distribution of impact structures on Earth is manifestly uneven. One continent that stands out for
Shock metamorphism
its relatively small number of confirmed impact structures and impact ejecta occurrences is South America. The
Meteorite
limited impact record for this large continent makes a robust case that there is a significant potential for further
Impact glasses
Tektites discoveries. Significant information on the impact record of South America is dispersed in different types of
Distal ejecta publications (journal articles, books, conferences abstracts, etc.), and in several languages, making it difficult to
K-Pg boundary access and disseminate it among the geoscientific community. We aim to present a summary of the current
Airburst knowledge of the impact record of this continent, encompassing the existing literature on the subject. It is
South America published in two parts, with the first one covering an up-to-date introduction to impact cratering processes and
to the criteria to identify/confirm an impact structure and related deposits. This is followed by a comprehensive
analysis of the Brazilian impact structures. The Brazilian impact record accounts for the totality of the large
structures of this kind currently confirmed in South America. The second part will examine the impact record of
other countries in South America, provide information about a number of proposed impact structures, and
review those that already have been discarded as not being formed by impact.

1. Introduction gas planet, which probably only has a liquid or solid core, the effects of
the impact were recorded as gigantic spots in its atmosphere, the largest
Between 16 and 22 July 1994, fragments of comet D/1993 F2 of which had a diameter of 12,000 km (Hockey, 1994).
(Shoemaker-Levy) successively impacted into the atmosphere of Jupiter This was the first collision of two planetary bodies ever observed
at a speed of some 60 km/s. The phenomenon was closely monitored by and recorded by humankind. Consequently, this event posed a para-
scientists from Earth, using telescopes at many observatories, and from digm shift for planetary scientists, as a phenomenon that had previously
space, through the Hubble Space Telescope and the Galileo Probe, been analyzed only indirectly through the study of meteorite impact
among others. The comet, which originally measured some 4 km across, craters − the scars left on planetary surfaces by ancient impact events
had broken up already in 1992 under the strong tidal forces of Jupiter. − could, for the first time, be observed, monitored, and scientifically
During the six-day impact period 21 fragments of the comet, ran- investigated in real time.
ging in size from some hundred meters to 1 kilometer, collided with Impact cratering studies have come a long way from modest be-
Jupiter’s atmosphere, causing fireballs and plumes that reached heights ginnings as far back as Robert Hooke’s crude experiments in the 17th
of up to 3000 km. The combined energy released by these impacts was century, when he dropped objects into mud and formed crater-like
calculated to have been the equivalent of 6 million megatons TNT, structures similar to those on the Moon’s surface (Koeberl, 2001;
roughly 600 times the total nuclear arsenal in the world. As Jupiter is a McHall, 2006). These beginnings also include such important


Corresponding author at: Geosciences Institute, State University of Campinas, UNICAMP, R. Carlos Gomes 250, Cidade Universitária ‘Zeferino Vaz’, Campinas, SP,
CEP 13083-855, Brazil.
E-mail address: alvaro@ige.unicamp.br (A.P. Crósta).

https://doi.org/10.1016/j.chemer.2018.06.001
Received 14 April 2018; Received in revised form 5 June 2018; Accepted 6 June 2018
0009-2819/ © 2018 Elsevier GmbH. All rights reserved.

Please cite this article as: Crósta, A.P., Chemie der Erde (2018), https://doi.org/10.1016/j.chemer.2018.06.001
A.P. Crósta et al. Chemie der Erde xxx (xxxx) xxx–xxx

benchmarks as J.K. Gilbert’s hotel room cratering experiments Russia, on 15 February 2013 (Borovička et al., 2013; Kohout et al.,
(Koeberl, 2001) and Alfred Wegener’s comparison of self-made “cra- 2013). The energy released was equivalent to 500 kilotons of TNT, and
ters” with lunar craters (Wegener, 1921). Strongly fertilized by Ralph the resulting shock waves caused human injuries (approximately 1500
Baldwin’s seminal book “The Face of the Moon”of 1946, impact studies people were injured mainly due to flying glass fragments) and extensive
finally came into their own with modern space exploration, when the property damage (such as broken window glass in approximately 7200
planetary science community realized its importance as an essential buildings).
process in shaping planetary surfaces. As well stated by this pioneer of As a consequence of the events described above, impact cratering, a
planetary impact science, Eugene Shoemaker, in his book chapter en- geological process that did not receive much attention from scientists
titled “Why study impact craters?” (Roddy et al., 1977), “the impact of until no more than a few decades ago, is now considered not only a
solid bodies is the most fundamental of all processes that have taken planetary process of fundamental, even paramount importance, but also
place on the terrestrial planets (Earth, Mars, Venus, Mercury). Indeed, the geological process that has been most ubiquitous throughout the
we can now sustain that impact processes are fundamental on all solid evolution of the Solar System. The effects of impact cratering over the
cosmic bodies, such as planets, moons, asteroids and comets” (page 1). entire timespan from planetary accretion until current times, as well as
Between the early 1980′s and the mid-1990′s, impact cratering covering the entire paleo-environmental and biological evolution on
studies experienced three decisive impulses. The first one was the re- our planet, are considered of profound significance (e.g., Melosh, 2011;
cognition of a link between a thin layer unusually enriched in platinum- papers listed in Reimold and Jourdan, 2012; Osinski and Pierazzo,
group elements (PGE), and its relation to a possible collision of a large 2013).
asteroid or comet with the Earth. It was found by Louis and Walter After decades of dedicated study of terrestrial impact structures, it is
Alvarez and colleagues in the Italian Apennines (Alvarez et al., 1980), now also known that the Earth’s known overall impact record is rather
deposited right at the boundary between the Cretaceous and the Pa- limited in both time and space. Differently from other planetary bodies,
leogene sediment deposits. This thin layer was then known as the such as the Moon, Mars or Mercury and all solid planetary bodies, in-
Cretaceous-Tertiary boundary – today it is referred to by the acronym cluding asteroids and comets, that have preserved their full impact
K-Pg (Cretaceous-Paleogene) boundary. Near-simultaneously, a Dutch records over 4.5 billion years (Fig. 1), the Earth’s crust is extremely
geologist, Jan Smit, made the same discovery in the K-Pg layer exposed dynamic, so that most of the terrestrial impact record has been ob-
in Spain (Smit and Hertogen, 1980). These authors postulated that this literated by tectonic, erosional and/or sedimentary processes. The
large-scale cosmic impact was responsible for the mass extinction event about 190 impact structures currently known on Earth (Earth Impact
that wiped out most of the life forms present at approximately 66 Database, February 2018) represent just a minor fraction of all the
million years ago, which includes the bulk of the dinosaur fauna.
However, an impact crater of compatible large-size and age remained to
be found on Earth, in order to corroborate this hypothesis.
The second boost for impact science was the actual discovery of a
very large impact structure, Chicxulub, straddling the edge of the
Yucatán Peninsula, Mexico (Hildebrand et al., 1991), with precisely the
same age as that of the K-Pg boundary, i.e., about 66 Ma. An intense
debate ensued in the geoscience community and under widespread
interest by the public, related to whether or not a single meteorite
impact, even of the magnitude of the one that produced the Chicxulub
impact structure of some 180 km diameter (Sharpton et al., 1993;
Morgan et al., 1997), would be capable of setting off a mass extinction
of this magnitude. This debate ran for nearly two decades, until a team
of 41 experts from several countries reviewed all the evidence for the
link between Chicxulub crater and the K-Pg mass extinction and con-
cluded in favor of the connection of both events (Schulte et al., 2010).
The significant role played by the Chicxulub impact and its global ef-
fects have since been strengthened (e.g. Morgan et al., 2016, on the
latest deep drilling results carried out at the peak-ring of the structure).
The third major impact discovery – or importance for impact cra-
tering studies – was then the widely televised comet impact of
Shoemaker-Levy 9 into the atmosphere of Jupiter of July 1994. Some
countries around that time invested into setting up committees and
scientific investigations analyzing the present danger from space, ac-
tively pursuing sky surveys in order to provide early warning of stray
planetary bolides approaching Earth’s airspace.
To reinforce the fact that meteorite impacts on Earth are not just a
phenomenon of the past, on 15 September 2007 a small meteorite (no
larger than 1–2 m in size) ‘touched down’ near the village of Carancas,
in Peru, forming an, admittedly, very small crater of 14 m diameter
(e.g., Borovička and Spurný, 2008; Kenkmann et al., 2009a, 2009b).
Although the local inhabitants were startled by this rare and un-
expected event, fortunately no harm was caused, which would not have
been the case if a hypothetical similar fall had happened over a more
populated area, say a large city. Fig. 1. a) Surface of Mercury exhibiting numerous impact structures. Image
Likewise, another recent phenomenon aroused serious concern from the MESSENGER probe (Credit: NASA-JPL/Johns Hopkins University
about meteorite impacts and the safety for humankind: the explosion of Applied Physics Laboratory/Carnegie Institution of Washington). b) Full view
a 20 m meteorite about 23 km above the vicinity of Chelyabinsk, a of giant asteroid Vesta (diameter 530 km), registered by the framing camera on
major city of 1 million inhabitants located in the Ural Mountains of the Dawn spacecraft (Credit: NASA/JPL-Caltech/UCAL/MPS/DLR/IDA).

2
A.P. Crósta et al. Chemie der Erde xxx (xxxx) xxx–xxx

impact events that contributed to the initial formation of our proto- offshore region of Brazil. Once again, this number of impact structures
planet, and then to formation and modification of the surface of the in sedimentary basins of Phanerozoic ages in South America, most of
planet. which have remained tectonically stable for long periods, is well below
the average when compared to similar terrains of other continents.
1.1. The South American impact record These considerations all point to the fact that there is further sig-
nificant potential for identifying new impact structures in South
The geographical distribution, as well as the density of known im- America, thus contributing to improve the overall Earth impact record.
pact structures on Earth, are remarkably uneven. The map depicting the Probably one of the main reasons why there are so few of them de-
worldwide distribution of impact structures (see Fig. 13, in Section 4 of scribed on this continent is the fact that impact cratering is a geological
this review paper) shows greater crater densities in North America, process that has not been well known amongst the geoscience com-
Europe and Australia, whereas other continents and regions exhibit a munities in South America, including universities, geological surveys,
lesser number of such structures. For example, the Earth Impact Data- and industry. With the exception of a few academics, mostly based at
base (February 2018) records 27 confirmed impact structures in Aus- Brazilian universities such as the Universities of Campinas, São Paulo,
tralia, whereas the record for the much larger continent Africa stands Brasília, and the Federal University of Bahia in Salvador, research and
only at 20. teaching covering this important field has been essentially absent in all
South America, in particular, depicts a relatively reduced number of other South-American countries. No undergraduate or graduate dis-
such structures. Whilst being among the large continents on Earth, it is cipline covering topics in planetary geology, including impact cra-
the one with the least number of confirmed impact structures, with only tering, has been taught – till now – on a regular basis as part of the
10 (see also Reimold et al., 2018). In addition, it also exhibits a very Geoscience curricula.
limited number of occurrences of materials related to the impact record, Nonetheless, interest in themes related to meteorite impact and
such as impact glasses, tektites, K-Pg boundary layer sites and reports of planetary research, in general, has been increasing in recent years
airbursts (Fig. 2). amongst the South American geoscientific community. Conferences and
This small number of confirmed impact structures in South America conference symposia, such as the International Geological Congress
is not compatible with its overall geological characteristics. The con- 2000 and the Annual Meeting of the Meteoritical Society in 2004 (both
tinent comprises two main tectonic domains: the Andean domain and held in Rio de Janeiro), as well as the 2010 AGU Fall Meeting in Foz do
the South American Platform domain (e.g., Cordani et al., 2016) Iguassu, included impact cratering in their programs, as well as field
(Fig. 3). The Andean domain is generally characterized by younger trips to some of the Brazilian impact craters. International scientific
terrains (Phanerozoic) than occurring on the rest of the continent. In cooperation is also on the rise, with scientists from South America
these younger terrains, active geological processes have not favored the frequently working in collaboration with well-known impact scientists
preservation of the impact record, in particular with respect to older from other countries, contributing to advance the knowledge about the
structures. Therefore, it comes as no surprise that the only two impact impact structures on the continent. Short courses and seminars on to-
craters known to date in the Andean domain are very young (Montur- pics related to impact research and meteoritics have been presented in
aqui – < 500–700 ka and Carancas – 10 years). However, even con- Brazil and Argentina by renowned researchers in the field, such as Klaus
sidering the limitation imposed by active geological processes in this Keil and Uwe Reimold. One of the important outcomes of these activ-
domain, this number of young craters is uncharacteristically small ities has been a significant increase in the number of articles reporting
compared to other similar regions of the world, such as North America results of impact cratering-related research in major international
(∼7 craters younger than 40 Ma), or even Africa (∼9 craters younger journals.
than 40 Ma). Maybe even more important, a growing number of undergraduate
The remainder of the continent is mostly a stable continental region, students has been exposed to these subjects in recent years – raising
which has served as the cratonic counterpart to the Andean domain expectations for further discoveries in the future.
during the Phanerozoic (Cordani et al., 2016). There are seven cratonic In this overall context, a comprehensive review of the South-
areas in South America, all of which have been tectonically stable at American impact structures, in the same way as it was done for the
least since the Neoproterozoic. These cratons, the Amazonian, São African continent by Reimold and Koeberl (2014a, 2014b), is timely
Francisco, São Luís, Parnaíba, Paranapanema, Luiz Alves, and Rio de la and, perhaps, somewhat overdue. The information currently available
Plata, are surrounded by orogenic belts related to the Neoproterozoic is either limited to Brazil (e.g., Crósta, 1987; Romano and Crósta, 2004;
Brasiliano/Pan African Orogenic Cycle (Cordani et al., 2016). There is Crósta and Vasconcelos, 2013), or dispersed across many articles, book
only one possible – still to be confirmed – impact crater in a cratonic chapters, and abstracts in conference proceedings that have been
area known in South America, the Colônia structure near the mega-city variably published in English, Portuguese, and Spanish.
of São Paulo in southeastern Brazil. Still, there is no reason why these In addition, some seriously misleading and inaccurate information
terrains, especially the cratonic areas, should not have (at least rem- has been disseminated about the impact record of South America, with
nants of) medium to large sized impact structures, as those exhibited by so-called “impact craters” having been nominated without acceptable
similar geological terrains in North America, South Africa and Aus- or even circumstantial evidence supporting a meteorite impact origin.
tralia. Such is the case of a small book by Acevedo et al. (2015), aimed at
The South American Platform has also five major intra-cratonic presenting a catalog of the impact craters in South America (see also the
sedimentary basins, which have developed from Middle Ordovician review of this book by Crósta and Reimold, 2016). However, Acevedo
times until the Cenozoic. Only two of these basins contain the majority et al. mixed scientifically proven impact structures with a number of
of the medium to large sized, confirmed impact structures of South generic and doubtful circular structures, for which no proper evidence
America: Araguainha, Vargeão, Vista Alegre, and the recently con- had been produced. Contrary to what has been extensively and widely
firmed Cerro do Jarau structure in the Paraná Basin, and Serra da accepted among the international impact crater community, i.e., the
Cangalha, Santa Marta, and Riachão in the Parnaíba Basin. need of proper scientific evidence for a structure to be considered of
Starting in the Mesozoic, the breakup of the Gondwana continent, impact origin (e.g., French, 1998; French and Koeberl, 2010; Reimold
with the opening of the South Atlantic Ocean, caused extensive rifting and Koeberl, 2014a, 2014b), Acevedo et al. (2015) erroneously stated
along the continental margin and formed a number of sedimentary that impact craters can be identified as such even in the absence of
basins. Most of these basins are located offshore and there are at least shatter cones, meteoritic material, or proven shock deformation.
two reported buried circular structures (Praia Grande and Ilha do Mel) Following this flawed line of thought, their catalog lists a number of
of possible impact origin, both in the Santos basin in the southeast uncertain structures, such as the alleged strewn fields of Bajada del

3
A.P. Crósta et al. Chemie der Erde xxx (xxxx) xxx–xxx

Fig. 2. A summary of the South American impact record, including confirmed impact structures (based partially on the Earth Impact Database, last consulted on 28
February 2018), and other impact-related materials and sites.

Diablo and Rio Cuarto, both in Argentina, the Rio Vichada structure in of impact origin, based entirely on dubious and inconclusive geophy-
Colombia, and a suite of alleged structures in Argentina, Brazil, Bolivia, sical signatures, was recently offered by Rocca et al. (2017). Alleged
Chile, Paraguay, Uruguay, Venezuela, Guyana, and Suriname. circular patterns in geophysical data are used by these authors to sug-
An even more recent example of a structure pointed out as possibly gest that a 250 km wide impact crater is present just NW of the Falkland

4
A.P. Crósta et al. Chemie der Erde xxx (xxxx) xxx–xxx

Fig. 3. Main tectonic domains of South America (after Cordani et al., 2016).

(Malvinas) Islands. They conclude that this could represent a late Pa- South America will be presented.
leozoic impact structure comparable in size to Chicxulub. No evidence We expect that one outcome of this review will be that the subject
for impact is presented, nor does this “observation” have any features will become more widely known to the South American geoscientific
comparable to the typically – although not consistently – observed community, with the expectation that a progressive number of students
geophysical or structural characteristics of large proven impact struc- and professionals may engage in activities related to searching for and
tures (Reimold et al., 2017b; McCarthy et al., 2017). studying impact structures.
Although the identification of potential impact sites is an important
step in cratering research, as a guide to future work, it is crucial that 2. The impact cratering process
experts in impact processes be involved in the confirmation of such
sites, to make sure that claims of new craters are truly based on con- Impact crater of sizable dimensions are formed when projectiles
crete evidence. Otherwise, an erroneous message may be conveyed to enters the Earth’s atmosphere at hypervelocity (i.e., velocity at impact
anybody interested in this relevant subject. This message is that it is in excess of the speed of sound) and upon impact on the Earth’s surface
possible to discover new impact structures on Earth just by perusing generates a strong compression (shock) wave that affects both the
web-based satellite image services, or even geophysical databases, to projectile and the target rock. According to French (1998), a projectile
find circular structures of unidentified origin and, then, assigning them of 50 m diameter will produce a crater with 1 km diameter. Note that
to impact events. It actually takes a lot more sound research to find smaller projectiles may be fragmented or burnt up in the atmosphere, or
proper evidence of meteorite impact events! In the following section, a – occasionally and under exceptional conditions – make it through the
concise summary of knowledge about impact structures and how to atmosphere without fragmentation and then create very small impact
recognize them is presented, before the South American impact record craters (for example – see in Part II of this review: Carancas, Peru). The
is introduced – in terms of proven and possible impact structures, and crater generation process is ultra-dynamic and characterized by shock
disproven/controversial suggestions. pressures and associated (i.e., after pressure release) shock tempera-
Part I of this review includes the current status of impact knowledge tures, as well as strain rates that are orders of magnitudes higher (up to
for South America, encompassing the Brazilian impact record. In Part II, 106 s−1) than those associated with any other geological process on
the relevant information for the remaining countries and regions of Earth (e.g., Melosh, 2011). Shock deformation of rocks and minerals is

5
A.P. Crósta et al. Chemie der Erde xxx (xxxx) xxx–xxx

achieved at impact-generated stress levels in excess of the Hugoniot


Elastic Limit. These impact generated pressures and temperatures op-
erate in the upper levels of the crust (only in the largest known ter-
restrial continental impact structures, such as the originally at least
250 km wide Vredefort impact structure are mid-crustal levels in-
volved) where these conditions exceed normal metamorphic conditions
by orders of magnitude.
It has been modelled that the formation of the currently known
largest impact structure on Earth, the complex Vredefort impact
structure in South Africa (Reimold and Koeberl, 2014a) of originally
250–300 km diameter, would have taken a mere 15 min (Henkel and
Reimold, 1998). Certainly this short time only pertains to the actual
crater-forming interval that is, then, followed by a much longer period
of post-impact crustal adjustment and equilibration. The very short
cratering interval is classically divided into the three stages of (i) con-
tact and compression, (ii) excavation, and (iii) modification and col-
lapse.
Studies of impact structures and impact cratering processes are
carried out on several fronts: (1) geological field work and geophysical
analysis of known impact structures (as well as suspected structures),
(2) laboratory investigations of samples from crater structures and
possible ejecta deposits, (3) shock experiments with known physical
conditions to obtain products (experimentally produced craters or
shock deformed rock) that can be compared with natural observations,
and (4) studies of shock features in meteorites. In addition (5), ex-
tensive numerical modeling of all parts of the cratering process on
highly variable scales, from the planetary to the micro-regime, has
created in the past decades a rather good understanding of the physical
and chemical processes taking place during such an ultra-dynamic,
short-lived cratering event. The known terrestrial impact structures, in
turn, then represent the natural observatories to test these models and,
themselves, provide input parameters for better constraining the next
generation of cratering models (see Reimold and Koeberl, 2014a, for a
more detailed introduction of impact cratering studies).
Shock experiments in specially equipped laboratories with high
velocity gas guns and explosive set-ups to simulate impacts in specific
materials or conduct micro- and meso-scale cratering experiments can
also be used to gain insight into the impact process (Deutsch et al.,
2015, and references therein). There is one drawback, however: shock
pulse durations in natural impact events are much longer than those
generated in shock experiments with projectiles orders of magnitude
smaller than planetary bolides. Consequently, the kinetics of shock
propagation are of vital importance and anybody comparing the effects
of natural and experimental impact ought to bear this in mind. Fig. 4. Stages of formation of simple impact craters (after French, 1998).

2.1. The three stages of impact cratering


strength of both the target and projectile materials, which leads to
vaporization of most of the projectile and the part of the target in the
The mechanics of impact cratering have been described in detail by,
environs of the impact point.
for example, French and Short (1968), Roddy et al. (1977), Grieve
The shock pressure becomes released when the shock wave reaches
(1987, 1991), Melosh (1989, 2011), and Melosh and Ivanov (1999),
the trailing end of the projectile and is reflected back as a so-called
and useful introductions to this topic have been provided by French
rarefaction (release) wave that travels slightly faster through the al-
(1998), French and Koeberl (2010), Collins et al. (2012), and Reimold
ready damaged target material than the shock wave did (Fig. 4-B). The
and Koeberl (2014a). In the following, the short interval of impact
target material becomes unloaded, whereby much of the projectile’s
cratering shall be discussed in terms of the processes active during the
kinetic energy is converted to thermal energy (shock temperature) that
contact and compression phase, excavation phase, and the modification
is deposited into the target.
and collapse phase (compare Fig. 4).
By far most of the projectile will be transformed to vapor, with an at
(i) Contact and compression phase: the cratering process begins
best very small remainder possibly being incorporated into impact-
with the contact of the projectile with the target (Fig. 4-A). Hyperve-
generated melt. Only in a few cases have small, solid particles from
locity impact velocities are typically of the order of tens of kilometers
projectiles been recovered, such as tiny particles at Chicxulub, on the
per second (the average velocity at asteroid impact is estimated at
sea-floor at the Eltanin impact site in the South Pacific, and – ex-
15 km/sec and for comet impacts at 25 km/sec). The exceptional kinetic
ceptionally – a 2-decimeter-sized meteorite fragment in impact melt
energy transferred by a large planetary projectile to its target is
rock of the Morokweng impact structure in South Africa. It is thought
achieved via a shock wave that at the point of contact can reach peak
that this fragment is the result of spallation off the projectile. What an
pressures of many hundreds of GPa (1 gigapascal = 10 kbar). The shock
exceedingly low probability to hit such a rare fragment in a borehole…
wave propagates hemispherically through the target and, backward,
Of the impact melt generated, much may pool in the interior of the
into the projectile. The pressures thus generated exceed the yield

6
A.P. Crósta et al. Chemie der Erde xxx (xxxx) xxx–xxx

of very large impact structures, may only last for 15–20 min from Kenkmann et al. (2014) provided a detailed review of the structural
projectile contact with the target to completion of the collapse phase geology of impact structures. This included a discussion of some first-
(e.g., Henkel and Reimold, 1998; Kenkmann et al., 2009b), the final order indicators for oblique impact, such as asymmetries of the crater
crater structures become subject to normal geological degradation. This rim, the central uplift structure, and ejecta blankets outside of an im-
involves erosion, tectonic overprint and even truncation, and finally pact structure.
burial by sediment or tectonic slices. Impact-induced thermal overprint The geophysical characteristics of impact structures have been re-
of rocks (e.g., due to the vicinity of massive and superheated impact viewed by, e.g., Pilkington and Grieve (1992), Grieve and Pilkington
melt, or prolonged residence in hot impact breccia), tectonic adjust- (1996), and Reimold and Koeberl (2014a). It suffices to introduce here
ment of the wider crustal volume, and hydrothermal overprint may that there are no unique geophysical characteristics of simple or com-
continue for many hundreds of thousands, even millions, of years. plex impact structures, as they variably depend on physical impact
Obviously, the duration of such hydrothermal activity is a function of parameters (such as angle of impact), as well as target lithological
the size of the impact structure, i.e., a function of impact energy, the composition and pre-impact geophysical anomaly patterns, and – de-
volume and composition of melt rock and its respective cooling beha- cisively – degree of erosion of an impact structure. Crater fill by low-
vior, and the availability of fluid. density impact breccias and impact-induced damage of the crater
basement and enveloping rock volume generally cause a negative
2.2. Crater morphology gravity anomaly over an impact crater, often of circular or annular
form. This, however, also depends on the distribution of rocks of dif-
Turtle et al. (2005) provided a detailed discussion of impact crater ferent densities below the crater and in the crater wall. Magnetic sig-
morphologies and the inherent nomenclature. Impact crater mor- natures of impact structures can be highly varied and depend – first-
phology ranges from simple, bowl-shaped (Fig. 4-F) to complex geo- order – on the magnetic properties of target rocks and impact-induced
metries with central uplift (peak) (Fig. 5-D), with peak ring, with peak remagnetization by thermal and chemical processes (e.g., Henkel and
ring and central basin or pit, to multi-ring basin morphologies (Hargitai Reimold, 2002; Salminen et al., 2009). Suevite generally has a stronger
et al., 2015). In contrast to volcanic crater structures, impact produces magnetic signature than impact melt rock, but the latter may be re-
generally circular, shallow (invariably only involving upper crustal le- sponsible for widespread remagnetization where its thermal overprint
vels, only in the largest known terrestrial impact structures reaching has exceeded the Curie temperatures of the magnetic carrier phases.
down to mid-crustal level – Vredefort, South Africa; Gibson and
Reimold, 2008), and rootless structures. Even for relatively low angles 3. Recognition criteria for impact structures and impactite
of impact (say, 15–30 ° from the horizontal), circular crater forms will systematics
be generated (Elbeshausen et al., 2013). Of course, erosion and post-
impact tectonic overprint may strongly affect/change the primary Morphological expression (such as circularity, crater basin forms),
morphology of a structure. A distinctive feature of fresh impact craters geophysical or regional geological anomalies may provide first-order
is – in contrast to volcanic crater structures – that they have overturned hints at the possible presence of an impact structure, but they do not
(after some erosion of the uppermost rim zone, still strongly upturned) suffice as proof thereof. Invariably, groundtruthing is required in order
rim stratigraphy. Rims of volcanic craters are generally characterized to obtain definite confirmation of such a possibility. Evidence of shock
by flat stratification or only weak upturning of the uppermost strata. metamorphism is diagnostic, and chemical evidence that demonstrates
On Earth, < 2–4 km diameter impact craters have simple, bowl- the presence of traces of an extraterrestrial projectile in breccias/melt
shape geometry. Due to wall collapse the final crater shapes, in com- rock is similarly conclusive. Only in impact structures, the rock de-
parison to the early transient crater forms, are of a somewhat wider and formation phenomenon known as shatter cones has been detected and,
flatter geometry. The onset size for formation of complex craters is a thus, constitutes a diagnostic, meso- to macroscopic recognition cri-
function of the lithological composition of the target. On Earth, the terion.
crossover from simple to complex morphology is at about 2 km for se-
dimentary targets (note that the ca. 2-km-wide BP crater of Libya has a 3.1. Shock metamorphism
distinct but small central uplift structure of complexly deformed sand-
stones – Koeberl et al., 2005), and at 4 km for crystalline targets. In the Physical expressions of shock wave compression and immediately
planetary context, these limits depend essentially on a body’s gravity; subsequent decompression are irreversible deformation effects, e.g.,
for example, on the Moon with a gravity that is only one-sixth of that of planar deformation features, formation of diaplectic glass, and high-
Earth, up to 15–20 kilometer impact structures will still display simple pressure polymorphs, in many rock-forming minerals (see e.g., French,
bowl-shape geometry. 1998; French and Koeberl, 2010; Langenhorst and Deutsch, 2012;
For complex craters, the nature of the target determines how wide Reimold and Koeberl, 2014a; Deutsch et al., 2015; Burchell, 2015;
and high a central uplift structure becomes. In the environs of variably Stöffler et al., 2017 − for reviews). These deformation and transfor-
sized crater structures, such as the 2 km wide BP impact structure in mation effects are collectively known as shock metamorphism, or im-
Libya or the ca. 10 km wide Bosumtwi structure in Ghana, conspicuous pact metamorphism if melting is also involved.
ring features may occur. It has been debated that they could have In Fig. 6 the pressure-temperature regimes of normal crustal meta-
formed due to the downward and outward directed compression vectors morphism and of impact metamorphism are compared. In cases of
during the excavation flow, or alternatively, that they are the result of impact into continental crust, impact structures are essentially confined
inward-motion of material in the course of central uplift formation to the upper crust, with only very large structures reaching into the
(Koeberl et al., 2005), or simply denote the outer edge of impact ejecta middle (such as Vredefort) or, in potentially larger yet still undetected
deposition (Wagner et al., 2002). structures, even lower crust. Normal crustal metamorphism is essen-
Very large craters may exhibit concentric ring structures, according tially limited to temperatures of 1000 °C and pressures < 1 GPa,
to which they are classified as multi-ring impact basins (Spudis, 1993). whereas impact metamorphic conditions extend to many hundreds of
There are only three terrestrial candidates for this structural type: GPa and thousands of degrees centigrade.
Vredefort in South Africa, Sudbury in Canada, and Chicxulub in Mexico Shock metamorphism in an impact crater has a “progressive”
(Grieve et al., 2008). The nature of such multiple rings is still a matter character in that there is a continuous increase of shock pressure and
of debate, with faulting or warping, décollement structures, or zones of temperature conditions in direction towards the location where the
enhanced cataclasis followed by melting of this comminuted material shock wave originated. Close to this “point of impact”, initial tem-
having been invoked. perature conditions are comparable to those on the surface of the Sun.

8
A.P. Crósta et al. Chemie der Erde xxx (xxxx) xxx–xxx

of very large impact structures, may only last for 15–20 min from Kenkmann et al. (2014) provided a detailed review of the structural
projectile contact with the target to completion of the collapse phase geology of impact structures. This included a discussion of some first-
(e.g., Henkel and Reimold, 1998; Kenkmann et al., 2009b), the final order indicators for oblique impact, such as asymmetries of the crater
crater structures become subject to normal geological degradation. This rim, the central uplift structure, and ejecta blankets outside of an im-
involves erosion, tectonic overprint and even truncation, and finally pact structure.
burial by sediment or tectonic slices. Impact-induced thermal overprint The geophysical characteristics of impact structures have been re-
of rocks (e.g., due to the vicinity of massive and superheated impact viewed by, e.g., Pilkington and Grieve (1992), Grieve and Pilkington
melt, or prolonged residence in hot impact breccia), tectonic adjust- (1996), and Reimold and Koeberl (2014a). It suffices to introduce here
ment of the wider crustal volume, and hydrothermal overprint may that there are no unique geophysical characteristics of simple or com-
continue for many hundreds of thousands, even millions, of years. plex impact structures, as they variably depend on physical impact
Obviously, the duration of such hydrothermal activity is a function of parameters (such as angle of impact), as well as target lithological
the size of the impact structure, i.e., a function of impact energy, the composition and pre-impact geophysical anomaly patterns, and – de-
volume and composition of melt rock and its respective cooling beha- cisively – degree of erosion of an impact structure. Crater fill by low-
vior, and the availability of fluid. density impact breccias and impact-induced damage of the crater
basement and enveloping rock volume generally cause a negative
2.2. Crater morphology gravity anomaly over an impact crater, often of circular or annular
form. This, however, also depends on the distribution of rocks of dif-
Turtle et al. (2005) provided a detailed discussion of impact crater ferent densities below the crater and in the crater wall. Magnetic sig-
morphologies and the inherent nomenclature. Impact crater mor- natures of impact structures can be highly varied and depend – first-
phology ranges from simple, bowl-shaped (Fig. 4-F) to complex geo- order – on the magnetic properties of target rocks and impact-induced
metries with central uplift (peak) (Fig. 5-D), with peak ring, with peak remagnetization by thermal and chemical processes (e.g., Henkel and
ring and central basin or pit, to multi-ring basin morphologies (Hargitai Reimold, 2002; Salminen et al., 2009). Suevite generally has a stronger
et al., 2015). In contrast to volcanic crater structures, impact produces magnetic signature than impact melt rock, but the latter may be re-
generally circular, shallow (invariably only involving upper crustal le- sponsible for widespread remagnetization where its thermal overprint
vels, only in the largest known terrestrial impact structures reaching has exceeded the Curie temperatures of the magnetic carrier phases.
down to mid-crustal level – Vredefort, South Africa; Gibson and
Reimold, 2008), and rootless structures. Even for relatively low angles 3. Recognition criteria for impact structures and impactite
of impact (say, 15–30 ° from the horizontal), circular crater forms will systematics
be generated (Elbeshausen et al., 2013). Of course, erosion and post-
impact tectonic overprint may strongly affect/change the primary Morphological expression (such as circularity, crater basin forms),
morphology of a structure. A distinctive feature of fresh impact craters geophysical or regional geological anomalies may provide first-order
is – in contrast to volcanic crater structures – that they have overturned hints at the possible presence of an impact structure, but they do not
(after some erosion of the uppermost rim zone, still strongly upturned) suffice as proof thereof. Invariably, groundtruthing is required in order
rim stratigraphy. Rims of volcanic craters are generally characterized to obtain definite confirmation of such a possibility. Evidence of shock
by flat stratification or only weak upturning of the uppermost strata. metamorphism is diagnostic, and chemical evidence that demonstrates
On Earth, < 2–4 km diameter impact craters have simple, bowl- the presence of traces of an extraterrestrial projectile in breccias/melt
shape geometry. Due to wall collapse the final crater shapes, in com- rock is similarly conclusive. Only in impact structures, the rock de-
parison to the early transient crater forms, are of a somewhat wider and formation phenomenon known as shatter cones has been detected and,
flatter geometry. The onset size for formation of complex craters is a thus, constitutes a diagnostic, meso- to macroscopic recognition cri-
function of the lithological composition of the target. On Earth, the terion.
crossover from simple to complex morphology is at about 2 km for se-
dimentary targets (note that the ca. 2-km-wide BP crater of Libya has a 3.1. Shock metamorphism
distinct but small central uplift structure of complexly deformed sand-
stones – Koeberl et al., 2005), and at 4 km for crystalline targets. In the Physical expressions of shock wave compression and immediately
planetary context, these limits depend essentially on a body’s gravity; subsequent decompression are irreversible deformation effects, e.g.,
for example, on the Moon with a gravity that is only one-sixth of that of planar deformation features, formation of diaplectic glass, and high-
Earth, up to 15–20 kilometer impact structures will still display simple pressure polymorphs, in many rock-forming minerals (see e.g., French,
bowl-shape geometry. 1998; French and Koeberl, 2010; Langenhorst and Deutsch, 2012;
For complex craters, the nature of the target determines how wide Reimold and Koeberl, 2014a; Deutsch et al., 2015; Burchell, 2015;
and high a central uplift structure becomes. In the environs of variably Stöffler et al., 2017 − for reviews). These deformation and transfor-
sized crater structures, such as the 2 km wide BP impact structure in mation effects are collectively known as shock metamorphism, or im-
Libya or the ca. 10 km wide Bosumtwi structure in Ghana, conspicuous pact metamorphism if melting is also involved.
ring features may occur. It has been debated that they could have In Fig. 6 the pressure-temperature regimes of normal crustal meta-
formed due to the downward and outward directed compression vectors morphism and of impact metamorphism are compared. In cases of
during the excavation flow, or alternatively, that they are the result of impact into continental crust, impact structures are essentially confined
inward-motion of material in the course of central uplift formation to the upper crust, with only very large structures reaching into the
(Koeberl et al., 2005), or simply denote the outer edge of impact ejecta middle (such as Vredefort) or, in potentially larger yet still undetected
deposition (Wagner et al., 2002). structures, even lower crust. Normal crustal metamorphism is essen-
Very large craters may exhibit concentric ring structures, according tially limited to temperatures of 1000 °C and pressures < 1 GPa,
to which they are classified as multi-ring impact basins (Spudis, 1993). whereas impact metamorphic conditions extend to many hundreds of
There are only three terrestrial candidates for this structural type: GPa and thousands of degrees centigrade.
Vredefort in South Africa, Sudbury in Canada, and Chicxulub in Mexico Shock metamorphism in an impact crater has a “progressive”
(Grieve et al., 2008). The nature of such multiple rings is still a matter character in that there is a continuous increase of shock pressure and
of debate, with faulting or warping, décollement structures, or zones of temperature conditions in direction towards the location where the
enhanced cataclasis followed by melting of this comminuted material shock wave originated. Close to this “point of impact”, initial tem-
having been invoked. perature conditions are comparable to those on the surface of the Sun.

8
A.P. Crósta et al. Chemie der Erde xxx (xxxx) xxx–xxx

Fig. 6. Schematic pressure-temperature diagram for shock metamorphic processes in the context of an impact event. ‘Geotherm’ is the modelled depth-temperature
distribution from the Earth́ s upper lithosphere to the core (after Deutsch et al., 2015).

Shock pressure may reach many hundreds, perhaps thousands, of GPa. cleavage), and so-called “shock extension fractures” (also called “ver-
The affected target material will be vaporized instantaneously, as is micular microfractures” by Buchanan and Reimold, 2002) have been
much of the projectile. Strain rates achieved in impacts can be as high described from this low shock regime – all of which have doubtful
as 106 s−1, many orders of magnitude higher than the strain rates as- impact-diagnostic value.
sociated with tectonic deformation or achieved in powerful volcanic Furthermore, so-called feather features (FFs) have been described in
eruptions. In a further zone, radially somewhat more removed from the quartz in rocks from many impact structures and, additionally, have
shock epicenter (point of source of the shock wave), shock pressures rarely been produced experimentally (Poelchau and Kenkmann, 2011;
and shock temperatures are sufficiently high to cause bulk melting of Kowitz et al., 2013a, 2013b, 2016). The definition of FFs has been used
target rock. This is followed by a further zone characterized by melting differently in the literature: some refer to the combination of a longer
of distinct minerals (i.e., by so-called part melting of target rock – Keil planar fracture and short, narrow-spaced “feathers” as FFs, others only
and McCoy, 2017, and references therein). Finally, a regime of forma- refer the term FF to these short and narrow fractures. As no FFs have
tion of diaplectic glass (syn.: thetomorphic glass – a phase with long- been reported yet from tectonically deformed rock, this is a very pro-
range order of the precursor mineral partially preserved but that is mising, likely shock-characteristic deformation phenomenon. Whereas
essentially isotropic due to abundant disruption of bonds at the atomic in recent years only one-sided “feathers” had been described, Zaag
scale), of planar deformation features (PDFs), and high-pressure mi- (2013) and Zaag et al. (2016) observed in a thin section from a shatter
neral polymorphs is reached, before below the Hugoniot Elastic Limits cone from the Serra da Cangalha impact structure (Brazil) a two-sided
(HEL) of minerals only elastic (brittle) mineral deformation is possible. “feather”. Their explanation for this phenomenon was that the target
Wherever possible, quartz is the mineral of choice when searching grain must have been subject of vibrational forces in order to produce
for shock metamorphic effects. Not only is this mineral widely available shear fractures in opposite directions.
in the upper terrestrial crust, but it is also rather resistant to weathering Brazil and Dauphiné twins are generated under shock conditions in
and under moderate metamorphic overprint. Under varied conditions quartz. Twinning is also important in zircon and in shocked carbonates.
of shock pressure/temperature, quartz develops a range of different The latter show, at relatively enhanced shock pressures, melting and
shock metamorphic effects that all are shock barometers: Planar dissociation effects.
Fractures (PFs), mosaicism, Planar Deformation Features (PDFs), dia- French et al. (1974) reported tiny pockets of glass from shocked
plectic quartz glass, high-pressure polymorphs and, finally, quartz fu- sandstone samples of the BP and Oasis impact structures in Libya.
sion (formation of lechatelierite). The formation conditions of these Kowitz et al. (2013a, 2013b, 2016) observed glass/melt in experi-
shock barometers have been well calibrated by shock experiments. mentally shocked sandstone subjected to nominal shock pressures of
Crystallographic analysis of PDFs is facilitated by the uniaxial optical only 5–12.5 GPa. Such glass formations are seemingly diagnostic for
character of quartz. The mineral zircon has in recent years also become low-shock overprint of porous sedimentary rocks – however, they have
a favourite with shock workers, due to its very refractive nature and not been observed in sandstone from many impact structures – and it is
good preservation in the sedimentary environment, and the fact that likely that lithological properties may be essential in glass formation
several shock metamorphic effects (shock twinning, planar fractures, (perhaps presence of a clay component in a sandstone may facilitate
granular shock texture, and shock induced dissociation into baddeleyite melting – Ebert et al., 2017). Kowitz et al. (2013a, 2013b) also detected
– ZrO2 – and silica – SiO2) are quite readily recognizable with standard diaplectic quartz glass in their weakly shocked sandstone specimens.
optical and electron optical microscopic techniques. Numerical modeling results by Güldemeister et al. (2013; also discussed
A very large part of the impact affected target volume is not shocked in Kowitz et al., 2013b, 2016) demonstrated that interaction of a shock
above the HEL of the mineral constituents. For this reason, extensive wave with pore spaces can cause local shock temperature increase by
investigations have been carried out on weakly shocked (experimen- factors up to 4, thus reaching the conditions related to normal pressure
tally and naturally shocked at < 10 GPa) rock to detect possibly new regimes for formation of diaplectic quartz glass (> 30 GPa) and silica
diagnostic impact deformation features (e.g., Kowitz et al., 2013a, glass (> 45 GPa) even at shock pressures as low as 5–10 GPa.
2013b, 2016). To date, only compression features such as impaction of PDFs (planar deformation features) formed in various important
grains onto each other with associated radial microfracturing (so-called rock-forming minerals (including quartz, feldspars, olivine) are the
concussion fractures), kinkbanding of micaceous minerals, local cata- most widely applied recognition criterion for shock metamorphism.
clasis, planar fractures (PFs – (sub)planar fracturing akin to imperfect They are absolutely straight (planar), crystallographically controlled

9
A.P. Crósta et al. Chemie der Erde xxx (xxxx) xxx–xxx

Fig. 7. Planar deformation features (PDF) in quartz within shocked granite


from the Araguainha Dome. Note the perfectly planar features along two dis- Fig. 8. Microphotograph (in plane-polarized light) of ballen-quartz (type II of
tinct crystallographic orientations (NW-SE; WNW-ESE). Spacing between fea- Ferrière et al., 2009) in an impact melt rock from the Wanapitei structure
tures ranges from 2 to 8 μm [courtesy of Joana Sanchez, UFG, Goiânia, Brazil]. (Canada) [courtesy of Ludovic Ferrière, NHM Vienna].

features, 1–2 μm wide, and spaced at 2–10 μm (Fig. 7). They may occur
in parts of a crystal or traverse it penetratively. PDFs have been ex-
perimentally produced in quartz above 8–10 GPa (e.g., Stöffler and
Langenhorst, 1994; Huffman and Reimold, 1996). Detailed scanning
electron microscopy (e.g., Hamers and Drury, 2011 and references
therein) and transmission electron microscopy (e.g., Stöffler and
Langenhorst, 1994; Gratz et al., 1996) showed that they may represent
amorphous (glass) lamellae, high-density dislocation bands, basal Brazil
twins, Dauphiné twins, or annealed equivalents (see also Trepmann and
Spray, 2005, 2006; Wenk et al., 2005, 2011). More recently, electron
backscatter diffraction (EBSD) studies have provided a new means to
investigate the nature of planar shock deformation, and, in particular,
shock induced twinning, not only in quartz but also in zircon and other
minerals.
When PDFs are thermally overprinted (post-impact annealing), the
original glassy phase is annealed. The former PDFs are then only re- Fig. 9. Checkerboard feldspar clast in impact melt rock from the Lappajärvi
cognizable due to the straight trails of fluid inclusions exsolved from the impact structure, Finland (see C. Reimold, 1982 for further information). The
primary melt phase. These planar fluid inclusion trails are known as original feldspar crystal has been partially melted under crystallographic con-
“decorated PDFs”. “Toasted quartz” (French and Koeberl, 2010; trol. The outer narrow, compact rim is an overgrowth generated due to reaction
Ferrière et al., 2009; Whitehead et al., 2002) refers to the “dirty” tex- between the clast and the cooling, originally superheated impact melt. Width of
tured quartz of impact breccias in which fluid and solid inclusions have field of view: ca. 0.36 mm.
been widely decrepitated and their content oxidized – now staining the
host crystals to provide the textural appearance of burnt toast. (2002) in quartz. They called this crystallographically controlled micro-
At shock pressures above ca. 15 GPa a further shock effect is induced melting of quartz “vermicular quartz”. Based on the different shock
in quartz (and also in olivine): “mosaicism” is the result of micro-ro- textures produced at different shock pressure levels, shock meta-
tation of small crystal domains due to shock compression, which ob- morphic schemes have been set up for different minerals (e.g., quartz –
viously leads to a mottled appearance of the intracrystalline extinction. Stöffler and Langenhorst, 1994) – and then for different rock types (e.g.,
When shock pressures of 30–35 GPa affect quartz, the crystal Stöffler and Grieve, 2007). Additional information on the behavior of
structure of this mineral is entirely destroyed and a quasi-amorphous zircon under shock metamorphic conditions can be obtained from
phase – diaplectic quartz glass – is generated. The grains still maintain Timms et al. (2017).
the original habit and original mineral defects will still be recognizable; Some minerals may be transformed to otherwise rare or non-ex-
however, under crossed polarizers such a grain will appear amorphous. istent high-pressure polymorphs under the conditions of elevated shock.
Feldspar minerals are also known for diaplectic glass formation, with Coesite and stishovite have been found in many impact structures.
diaplectic plagioclase known as “maskelynite”, and being a frequent While coesite is known from magmatic or tectonic crustal settings as
constituent of meteorites as well. Shock metamorphism in important well (it may occur in kimberlite and in rocks associated with subduction
rock-forming minerals, including the feldspar minerals, was recently zones), stishovite is only known from impact structures. Reidite, the
reviewed by Stöffler et al. (2017). high-pressure polymorph after zircon, has also been observed widely in
Another impact-indicative thermal feature is the so-called “ballen impact structures; it is formed at shock pressures between 20 and
quartz” (Fig. 8) thought to originate from annealing of diaplectic quartz 40 GPa (Reimold et al., 2002). At even higher shock pressure/shock
glass and/or lechatelierite (fused silica glass). “Checkerboard feldspar” temperature zircon can melt and then crystallize again to so-called
(Fig. 9) can also indicate impact deformation and represents melting of “granular” or “strawberry” shock texture. Monazite shows similar be-
a feldspar crystal with typical near-rectangular habit at the microscopic havior (Deutsch and Schärer, 1994; Moser et al., 2011). At even higher
scale and controlled strictly by crystal lattice symmetry. An equivalent pressure conditions, zircon will dissociate to silica plus baddeleyite
to checkerboard feldspar was described by Buchanan and Reimold (ZrO2). A recent review by Timms et al. (2017) provides a

10
A.P. Crósta et al. Chemie der Erde xxx (xxxx) xxx–xxx

Fig. 10. Two examples of shatter cones from Brazilian impact structures. a) Formed in phyllites from Araguainha Dome; b) formed in basalt from Vista Alegre (see
Sections 5.1.1.1 and 5.1.1.3 for further information on these impact structures). Scales in cm.

comprehensive treatise of shock metamorphism of zircon. Zircon and post-shock temperatures are higher than the melting temperatures of
monazite are also very useful due to their strong refractivity; they are minerals (or mineral associations), impact-induced melting is achieved.
the minerals of choice when searching for shock deformation in old First individual minerals or mineral assemblages may melt, then at even
Proterozoic or Archean rocks (possibly subjected to polymetamorphism higher temperatures bulk rock melting may be achieved. Impact melt is,
and strong alteration). These minerals are also essential for U-Th-Pb thus, generated and will be, as a consequence of cratering, either em-
geochronology. In target rock containing graphite, impact diamond or placed in within-crater impact breccias (see below), or incorporated
lonsdaleite may be formed, as in the so-called ‘impact treasury’ of into ejecta from an impact structure. Further discussion of the products
Russia, the Popigai impact structure of northern Siberia. Shock de- of impact melting is found in the following section on impactite no-
formation in other trace minerals such as xenotime or apatite has also menclature.
been reported recently (Cavosie and Lugo-Centeno, 2014; Cavosie et al., Shock deformation is a strictly heterogeneous phenomenon (“shock
2016). When applying conventional phase diagrams to shock meta- heterogeneity”) due to the complex behavior of compression waves in a
morphic assemblages, it must be observed that the ultradynamic shock heterogeneous, multi-particle/-mineral and texturally invariably com-
processes do not allow equilibrium reactions to take place. plex – at multiple scales – rock. Waves may be reflected, scattered, or
When shock pressures are sufficiently high, so that the associated refracted on grain boundaries or defects that are always present in

11
A.P. Crósta et al. Chemie der Erde xxx (xxxx) xxx–xxx

natural rock, which results in shock attenuation or amplification. This and have been – mistaken for shatter cones (for comparison of sedi-
may lead to grains of the same mineral in any rock section showing mentary and impact cones, see, for example, Reimold and Minnitt,
highly variable deformation degrees on the grain scale, possibly from 1996; Lugli et al., 2005). Baratoux and Reimold (2016) and references
being unshocked to highly shocked. Also different minerals have dif- therein also provide a comprehensive treatise of current hypotheses
ferent affinity for shock effect generation at a specific pressure level regarding the physics on shatter cone formation. Shatter cones have
(e.g., Ogilvie et al., 2011). Due to the different behaviors of minerals been mostly described from settings in central-uplift and crater-floor
under shock conditions, different rock types also display varied re- rocks, but a few reports have related them to crater rim occurrences as
sponses to shock. Examples for progressive shock deformation of a well. In addition, they have been detected in rock clasts within impact
basaltic and a granitic rock are found in Stöffler and Grieve (2007), and breccias outside of the corresponding crater structures.
progressive shock deformation for mafic, pyroxene- and olivine-rich Small shatter cones have also been produced experimentally (Wilk
meteorites in Stöffler et al. (1991). Stöffler et al. (2017) reviewed the and Kenkmann, 2016; Kenkmann et al., 2016). The shock pressure re-
response of various rock types to shock deformation and proposed an gime for shatter cone formation has been estimated between < 5 GPa
updated classification system for terrestrial and extraterrestrial rocks. and perhaps > 30 GPa (e.g., the shatter cones occurrences in the central
When trying to obtain evidence for shock deformation in a geolo- uplift structure of the Vredefort impact structure – Nicolaysen and
gical structure suspected of being the result of impact cratering, where Reimold, 1999; Wieland et al., 2006). Zaag et al. (2016) applied com-
available, possible impact breccias (next section) should be sampled for puter tomography to investigate the 3D distribution of microfractures
investigation of evidence for shock metamorphism. Where only in a shatter cone, to attempt to understand the previously repeatedly
monomict clastic breccia or unbrecciated rock can be sampled in the published possible interrelationships between shatter cones and sub-
outer or deeply eroded parts of a suspected structure, it may still be parallel, closely set fracture systems (so-called “multipli-striated joint
possible to detect shocked minerals – albeit in comparatively lower sets” – MSJS, Nicolaysen and Reimold, 1999).
abundance, as these deformations would have been generated in the Shock metamorphic indicators should be easily observed, preferably
outermost zone of an impact structure under very weak shock condi- by widely accessible optical microscopy. However, it has been fre-
tions. A pertinent example of such a case is the Cerro do Jarau structure quently necessary to utilize sophisticated but time-consuming trans-
that will be introduced in this paper later: we recently examined some mission electron microscopy (TEM) to investigate the true nature of
250 thin sections of samples from there and only found a handful of microdeformation features not resolvable by optical microscopy alone.
specimens that displayed an altogether significant but very limited An enlarged arsenal for shock metamorphic investigations has been
amount of shocked grains. created in recent years by advances in scanning electron microscopy
Great care must be taken not to confuse impact-generated micro- (SEM) and associated techniques such as color (or color composite)
deformation with the deformation effects of tectonic processes. There cathode-luminescence (CL), electron forescatter or electron backscatter
are many erroneous reports of shock deformation in the literature, some diffraction (EBSD) techniques (e.g., Hamers, 2013; Hamers and Drury,
of which have been – very unfortunately – further promulgated due to 2011; Pittarello et al., 2015b). Electron-backscatter diffraction (EBSD)
the widely persisting lack of proper education about impact deforma- analysis and electron microscopic techniques have been very success-
tion and the requisite methodology for its investigation. fully employed for detailed shock metamorphic analysis of zircon
It should also be admitted that, in some past investigations, de- (Timms et al., 2017 and references therein) and monazite (Erickson
formation features that were only studied by normal optical microscopy et al., 2017). Scanning electron microscopic techniques now can fill the
might have been discarded as non-impact evidence, only because the gap between optical and transmission electron microscopy.
proper analysis by universal stage (or TEM analysis) was not applied.
Reassessment of such cases ought to be undertaken. Vice versa, there 3.2. The recommended impactite nomenclature
are cases of previously published supposed PDFs in the literature, the
validity of which ought to be investigated again. All such cases of al- Lithologies formed as a direct consequence of an impact event occur
leged impact structures that are based on only one or two grains of (prior to erosion) within and around impact structures, in either prox-
allegedly shocked quartz need to be readdressed with the current state- imal or distal settings. Stöffler and Grieve (2007) provided a detailed
of-the-art methodology, and where possible with investigating more nomenclature for impactites (Fig. 11) and, recently, Stöffler et al.
minerals than only quartz. As stated already above, shock deformation (2017) presented further discussion.
in porous sandstone has recently been investigated by Kowitz et al. “Distal impactites” include ejecta that can be distributed – ac-
(2013a, 2013b, 2016). An important finding from our ongoing studies cording to the magnitude of the impact event – regionally, or on con-
of naturally shocked sandstone samples is that the presence of porosity, tinental and even global scales. This includes tektites and microtektites
with or without cement, significantly affects the affinity of a rock for that may be found many hundreds to thousands of kilometers from the
shock deformation. In particular, the formation of PDF in quartz seems source structure. Four tektite strewn fields are currently known on
to be subdued. And invariably only limited formation of short and Earth: the Central European (moldavites), the Australasian (including
narrow sets of PDF has been observed – in the majority of instances at microtektite occurrences in Antarctica), the Ivory Coast, and the North
or close to grain boundaries. In such rocks, having very well-polished American strewn fields (Glass and Simonson, 2012, 2013 – for a review
thin sections is mandatory, and careful screening of them at a minimum of currently known distal impact ejecta deposits). In addition, a tektite
microscope magnification of ca. 300× is required. occurrence in Uruguay has recently been proposed (Ferrière et al.,
The only reliable mega- to macroscopic shock deformation criterion 2017), and some conference contributions have referred to possible
is represented by shatter cones (Fig. 10), centimeter- to meter-sized tektites from Colombia (see Part II of this review, for a discussion).
cone-shaped fracture phenomena with striations that diverge in a ridge- Moldavites are related to the Ries impact of southern Germany, the
and-groove pattern from a small apical area (see papers in a recent Ivory Coast tektites to the Bosumtwi crater structure in Ghana, and the
special issue of Meteoritics and Planetary Science: Baratoux and North American tektites to the Chesapeake Bay impact structure on the
Reimold, 2016; also Deutsch et al., 2015). Shatter cones are known shelf at the eastern seaboard of the USA. The Australasian tektites could
from a large number of impact structures, and have not been observed so far not be correlated with a proven impact structure but are believed
from any other geological settings. Care must be taken though, as to originate from Southeast Asia; microtektites of this occurrence have
various other fracture phenomena (e.g., plumose fractures, or cone-in- been collected as far as 6 000 km from Indochina, in Antarctica. The
cone structures of sedimentary origin, cone fractures – so-called per- proposed Uruguayan occurrence has so far not been related to a po-
cussion marks – at the base of fossil water-falls, wind-ablation features tential source crater. Distal impact ejecta also comprise regionally or
in desert environments, inter alia – Baratoux and Reimold, 2016) can – globally occurring impact ejecta, for which the best-known occurrence

12
A.P. Crósta et al. Chemie der Erde xxx (xxxx) xxx–xxx

Fig. 11. Classification of impactites according to their geological settings (proximal or distal with respect to an impact structure), as proposed by Stöffler and Reimold
(2006) and Stöffler and Grieve (2007).

is at the K-Pg boundary. In addition, the Archean and Proterozoic with the pure presence of some melt. Stöffler and Grieve (2007) defined
spherule layers (e.g., Glass and Simonson, 2013) also need to be re- this lithology as an impact breccia composed of impact-generated, co-
ferred to as distal impact ejecta from unknown, likely by now com- genetic clastic and melt particles, as a transitional lithology between
pletely degraded or subducted ancient impact structures. purely clastic lithic breccia and impact melt breccia/rock. Stöffler et al.
“Proximal impactites” (i.e., impact-generated rocks deposited closer (2013, 2017 and Artemieva et al. (2013) promoted a two-stage genetic
to their source than ca. 2 crater diameters) include the categories of scenario, in which “suevite” is generated by fallout from the impact
shocked rocks and impact breccias. Shocked rocks have maintained vapor plume over both, the crater and its surroundings, and mostly
their original texture and fabric but incorporate shock metamorphosed occurring outside of the crater, as the result of secondary phreato-
minerals. Amongst impact breccias the categories of monomict breccias magmatic-like eruptions triggered by the interaction of hot impact melt
(cataclased target rock, equivalent to tectonically produced cataclasite, and volatile matter (water, carbonate), akin to so-called molten-fluid-
but perhaps carrying shocked minerals) and polymict impact breccias coolant interaction (MFCI). Stöffler et al. (2017) inscribed a distinction
are distinguished by clast content and matrix composition. Lithic between “suevite” and “suevitic impact breccia” into their revised im-
breccias (historically termed “fragmental impact breccia”) are com- pactite classification scheme. It should be pointed out that the term
posed of clastic components only, without any melt component. Suevite “suevitic breccia” has been used for many years to describe a pre-
consists of mineral and lithic clasts of varied shock metamorphic degree dominantly clastic impact breccia with a mere tiny fraction of impact
plus a melt component. Impact melt rock is composed of a melt matrix melt (e.g., Reimold et al., 1997 – suevitic breccia from Roter Kamm
with a highly variable proportion of mineral and lithic clasts. Textures impact crater, Namibia). Osinski et al. (2016, and papers referred to
of such melt rocks can vary strongly, according to the nucleation and therein) emphasized that some impact breccias previously referred to as
cooling history of a melt. As impact melt rocks may not only texturally suevite actually contain melt material in their matrices and should,
but also chemically resemble volcanic or plutonic rocks, it is not sur- thus, be termed a “particulate impact melt breccia”. That melt particles
prising that in the past many impact melt rocks were originally con- are contained in Ries suevite was also discussed by Reimold et al.
sidered products of magmatism. Impact glass – in contrast to impact (2011, 2013a) with respect to the impact breccias from the Enkingen
melt rock not, or not fully, crystallized and either glassy or semihyaline drill core. Osinski et al. (2016) also took serious issue with the MFCI
(composed of crystallites in glassy mesostasis) – occurs in contrast to hypothesis for the origin of the Ries suevite, as they see strong differ-
tektites/microtektites in proximal settings to impact structures. Impact ences between Ries suevite and Onaping impact breccia of the Sudbury
glasses are also characterized by significant contents of water, and their Structure that they consider having been generated by this process. It is
chemical compositions may indicate origin from any type of target rock obvious from this short treatise that the origin of melt-bearing but
and depth in the transient cavity regime. In contrast, tektites only re- lithic-clast dominated impact breccia (suevite) is still a highly con-
present the geo-zone of the uppermost target and as high-temperature tentious issue in need of further conclusive investigation.
formations are comparatively dry. Impact breccias can also occur as breccia injections in the crater
There has been a lively debate about the origin of what has been floor, or as local, authigenic breccia developments in the crater floor
historically termed “suevite” over the last decades. First, the debate and walls of an impact structure. They have been summarily termed
involved the question of “how much melt material an impact breccia dike breccias in the impactite nomenclature. Dike breccia types can be
would have to contain” to deserve the label suevite. Some suggested injections of any type of crater-fill breccia (i.e., both monomict and
that 10 vol%, or 3 vol% were required, whereas others were content polymict lithic breccia, suevite, or impact melt rock). Dike breccias also

13
A.P. Crósta et al. Chemie der Erde xxx (xxxx) xxx–xxx

Fig. 12. One of the best exposures of massive


pseudotachylitic breccia in the Archean grani-
toid core of the Vredefort Dome, at Leeukop
Quarry just north of the town of Parys. Note
that this seemingly steeply northward (fore-
ground) dipping melt body is exposed in sev-
eral quarrying benches, each of which is ca.
4 m high. The section shown is about 20 m
wide.

include the enigmatic pseudotachylitic breccias (PTB – Reimold, 1995, – especially – geochemical evidence from the Vredefort impact struc-
1998; Reimold and Gibson, 2006; Reimold et al., 2016, 2017a; compare ture, as most recently examined by Reimold et al. (2016, 2017a). These
Fig. 12). In the past – and still today by some workers – these breccias of authors favor decompression melting of basement rock in the course of
varied genesis have been indiscriminately termed “pseudotachylite”. central uplift formation as the likely process for the formation of mas-
This problem is rooted in history, as this term (historical spelling: sive PTB, such as those known from the large Vredefort and Sudbury
pseudotachylyte) was coined by Shand (1916) for the enigmatic melt impact structures.
breccias of the Vredefort Dome, the central part of the world’s largest One of the problems with pseudotachylitic breccias is that they can
known impact structure (Gibson and Reimold, 2008). Today, the term be generated during all stages of the cratering process. And various
“pseudotachylite” is reserved in structural geology for bona fide friction melting processes can take place – either at specific times such as de-
melt only, whereas in impact settings melt breccias of pre-, syn-, and compression melting during the modification stage, or throughout the
post-impact age and of a range of different modes of origin can occur: process as in the case of friction melting that may be a component
friction melt, cataclasite, ultramylonite that in the field cannot be dis- during shock compression and modification (for further discussion, see
tinguished from friction melt or impact melt rock, and finally bona fide e.g., Reimold, 1998; Spray, 2010; Reimold et al., 2016, 2017a).
impact melt rock (an example of multiple types of PTB in the crater In every case of observation and analysis of such dike breccias,
floor of an impact structure was reported by Reimold et al., 1999 from careful investigation in the laboratory, if necessary aided by electron
the Morokweng structure, South Africa). As discussed by, for example, microscopic methods, is required to properly classify PTB according to
Mohr-Westheide and Reimold (2010, 2011) and Reimold et al. (2016, their matrix type, clast deformation, and chemical systematics. If this is
2017a), a number of genetic processes could be responsible for PTB not possible (for instance, because the groundmass is too altered), use
formation in impact structures: compression and immediate decom- of the non-genetic term pseudotachylitic breccia is recommended until
pression during the early stage of cratering, friction melting during such time that new, possibly conclusive information may become
various impact cratering stages, the combination of compression and available for definitive classification.
friction melting, and during the modification phase PTB can be pro-
duced due to frictional movement of large blocks and/or rapid de- 3.3. Identification of extraterrestrial projectiles
compression. Only a few occurrences of massive pseudotachylite (fric-
tion melt rock) are known from tectonic occurrences, such as – at the Another way of obtaining evidence for the confirmation of the im-
meter-scale – at the Insubrian Line in the Italian Alps, and in the pact origin of a geological structure is to attempt to identify traces of
Musgrave Block of Australia (Camacho et al., 1995) where meter-scale the impactor from space. The detection and verification of an extra-
networks of cm- to dm- veins meander around blocks of precursor li- terrestrial component in impact-derived melt rocks or breccias can be of
thology. In contrast, the PTB developments of the two largest impact impact-diagnostic value, as discussed in detail by, e.g., Goderis et al.
structures known on Earth, the Sudbury and the Vredefort structures, (2013) and Koeberl (2014). Generally, a very small amount of me-
are in part orders of magnitudes larger, with the largest manifestations teoritic material is mixed with target rock-derived vapor and melt, and
in Sudbury and Vredefort attaining kilometer scale. There are some this mixture is then incorporated into impact breccia or impact glass. In
workers who consider these PTB summarily as pseudotachylite sensu most cases, the amount of extraterrestrial material within such im-
strictu, formed by shear-induced friction melting (see e.g., Deutsch pactite is < < 1% by mass, although in exceptional cases values as
et al., 2015; Spray, 2010). Others state that “macroscopic” develop- high as 10% by mass have been obtained. Detecting such generally
ments of such melt breccia “are predominantly formed by the intrusion small amounts of extraterrestrial matter is difficult. Only elements that
of impact melt into the basement of impact craters” (Stöffler et al., occur with high abundances in meteorites but at comparatively low
2017, p.15). Both of this is contrary to the geological, petrographic and abundances in terrestrial crustal rocks are used for such studies (e.g.,

14
A.P. Crósta et al. Chemie der Erde xxx (xxxx) xxx–xxx

Ni, Co, and Cr, and the platinum-group elements [PGE], cf. Koeberl, Most known impact structures are located in North America and
2014). If siderophile element contents in impact melt are distinctly Eurasia, and there is another significant cluster in Australia. Clearly,
higher than the abundances in target rock (the so-called indigenous there is a link between space exploration and related widespread
component), this may be indicative for the presence of a chondritic or knowledge about planetary impact structures (e.g., Hargitai et al.,
iron meteoritic component (e.g., McDonald et al., 2001). Complications 2015). Impact structures have been searched for and investigated ever
may, however, arise: (a) because each meteorite class has a range of since the 1950s in countries where active space exploration has been
compositions; some are better constrained than others; (b) the target pursued. Other small, active groups have made regional contributions –
rocks may have variable siderophile element concentrations; or (c) if in Scandinavia, in southern Africa, and in Australia. European efforts
the siderophile element concentrations in the impactites are very low; have been strongly supported by a European Science Network program
and (d) it may be essentially impossible to separate an achondritic for impact research in the 1990s and the Nordic countries’ educational
projectile component from the indigenous component, due to the si- program of the last decade. The large number of Australian impact
milarity of compositions of such meteorites to some mafic terrestrial structures must be accredited to the singular effort of the late Eugene
rocks. The contribution to the siderophile element budget from the Shoemaker and his wife Carolyn, in collaboration with a number of
target rock to the composition of impactites can only be understood if Australian workers.
there is either a well-constrained mixing relationship between the im- One should not forget that differential preservation of old impact
pactor and target rock compositions that results in a reliable regression structures has been governed by the varied geological histories of dif-
line and a lower intercept that reflects the average PGE concentration in ferent regions. Young oceanic crust, recently exhumed continental
the target rocks (e.g., McDonald et al., 2001), or if all contributing crust, and – to the contrary – long exhumed, stable continental plat-
target rocks have been verified and their respective contributions to the forms would obviously have very different crater accumulation and
melt mixture are known with confidence. degradation rates (this has been discussed recently in the SW
Isotopic compositions can be very good indicators for the presence Gondwana context, pertinent to this review of the South American
of a meteoritic component as well. In Re-Os isotopic analysis (see re- impact record, by Reimold et al., 2018).
view by Koeberl and Shirey, 1997), the different isotopic compositions There are several regions in the world that are underrepresented in
and significantly different abundances of the element Os in meteorites the terrestrial impact record. This includes large parts of Asia, including
and crustal terrestrial rocks are used to fingerprint a meteoritic com- eastern Siberia, Mongolia, and China, the Middle East, and Southeast
ponent. This isotope system is very sensitive for this purpose (projectile Asia. The equatorial rain forest belt of Amazonia, Central Africa and
proportions as low as 0.2% in impact melt rock have been determined – Southeast Asia has not been supportive of remote sensing investigations
e.g., Koeberl et al., 1996) but allows only detecting the presence of a for crater exploration, and Antarctica and Greenland are covered, in
chondritic or iron meteorite component, without the capability of large parts, by kilometer-thick ice carapaces.
specifying the meteorite type and identifying achondrite signatures. The terrestrial cratering record is seriously incomplete, but to esti-
The Cr isotope system is less sensitive for the detection of an extra- mate by how much is not an easy task. It was estimated by Stewart
terrestrial component but allows distinguishing between some me- (2011) that not less than 228 impact craters > 2.5 km remained to be
teorite types, including achondrites (e.g., Koeberl, 2007, 2014; Koeberl identified in the global Phanerozoic (i.e., younger than 540 Ma) sedi-
et al., 2007). Both these methods rely on the fact that isotopic com- ment record. She estimated for Africa that only some 15% of the full
positions for the elements Os and Cr are different between most me- crater population in such settings and between < 1 and > 10 km in
teorites and terrestrial rocks, and that these differences are large en- diameter had been discovered so far, and only 10% of structures <
ough to allow detection of small contributions of meteoritic Os or Cr to 210 km diameter. For craters > 10 km in diameter, about 45% (i.e., 5
impactite. The best-suited materials for analysis are impact melt rocks/ out of 11) were known. It is a complex undertaking to achieve such
breccias, preferably from a coherent impact melt body, but in some estimations that have to depend on a number of generalizations and
cases melt fragments from suevite have worked as well. assumptions, such as: how much of the Phanerozoic sediment record
has survived plate tectonics and surface degradation? Has the impact
4. The terrestrial impact cratering record flux since the end of the Precambrian remained constant? Can the
global rate of degradation be averaged? What is the effect of differential
At the time of this writing, 190 confirmed terrestrial impact struc- rates of uplift? How can one estimate differential rates of obliteration of
tures are listed in the Earth Impact Database (February 2018); their very small (< 2 km size) craters?
locations are shown schematically in Fig. 13. This number is bound to Hergarten and Kenkmann (2015) also investigated the statistical
increase further, as a few structures are usually added each year. On the validity of the current terrestrial impact crater record. They could not
other hand, it may also decrease if it would be shown that previously find any evidence for systematic incompleteness in the available record
listed structures (as happened, e.g., to the pair of Arkenu structures in for craters > 6 km in diameter and exposed at the surface. However,
Libya – see Reimold and Koeberl, 2014a) should be removed from the they noted that more than 90 craters in the range from 1 to 6 km dia-
record. There are other data bases on the internet with information meter may still have to be discovered, and even > 250 of 0.25–1 km
regarding more structures – most of which recognized only by more or diameter. They relate this computed transition from a probably com-
less circular outlines in remote sensing data sets. It must be emphasized plete inventory for craters > 6 km to a strongly incomplete record for
again that only the truly confirmed structures can be referred to as bona smaller crater sizes to the difference between simple and complex
fide confirmed impact structures, whereas others should only be re- craters.
ferred to as possible impact structures without proper verification. As a Therefore, it would be vain to try making predictions of how many
number of such doubtful records have entered even the peer-reviewed impact structures remain to be discovered in South America. However,
literature, confusion of non-experts in the field is understandable. The it is safe to anticipate that future discoveries will be made.
rules and criteria for the recognition and confirmation of impact
structures, as outlined above, are straightforward; so-called identifica- 5. The record of confirmed impact structures of South America
tions can easily be dismissed as not based on any established criteria
(e.g., Reimold et al., 2014; Crósta and Reimold, 2016). The South American continent occupies an area of ∼17.8 million
Any first-order observation – by remote sensing, geophysical or km2. Brazil corresponds to nearly half of the continent, with ∼8.5
geological data – must be followed up with ground-truthing, and where million km2. Most of the older terrains in South America, those that
the relevant expertise does not exist, interaction with established im- remained tectonically stable since the Neoproterozoic, are located in
pact cratering research groups ought to be sought. Brazil, with minor portions in neighboring countries such as Uruguay,

15
A.P. Crósta et al. Chemie der Erde xxx (xxxx) xxx–xxx

Fig. 13. World distribution of impact craters (modified after Reimold and Koeberl 2014a).

Bolivia, Peru, Colombia, Venezuela, Suriname, French Guyana and with two Phanerozoic intracratonic sedimentary basins, the Paraná
Guyana (Cordani et al., 2016, and references therein). Basin (Zalan et al., 1991; Milani et al., 2007) and the Parnaíba Basin
Consequently, the fact that the majority of large impact structures (Góes and Feijó, 1994; Santos and de Carvalho, 2009). The impact re-
known in South America are located in Brazil stems naturally from the cord of Brazil is hereby presented according to the structures’ respective
vast extent of its territory, combined with the geological nature and geological settings.
ages of its terrains. What is not to be expected is that no younger and
relatively small impact structures are at present known from Brazil. 5.1.1. Paraná Basin
However, this apparent lack of younger structures in these older geo- The Paraná Basin (Fig. 14) is a large sedimentary basin covering an
logical terrains might be partially related to factors such as the pre- area of ∼1.5 million km2. It is located in the central-eastern portion of
vailing tropical to sub-tropical climates, which do not favor the pre- South America, with ∼75% of its area occurring in Brazil and the re-
servation of small-size structures due to fast weathering rates. In mainder in eastern Paraguay, northeastern Argentina, and northern
addition, the presence of thick vegetation cover, typical of the central Uruguay.
and northern portions of Brazil, may conceal surficial or surface-near The Paraná Basin is an Ordovician-to-Cretaceous intracratonic
morphological features – thus suggesting that new discoveries of small, flexural basin, developed between 460 and 66 million years ago. Its
or even large, impact structures may well be possible, especially in genesis is related to the convergence between the former Gondwana
these regions. supercontinent and the oceanic crust of the former Panthalassa Ocean
Likewise, the Pacific margin (the Andean Belt) of South America, (e.g., Milani et al., 1994, 2007). The basin was filled by six depositional
having been tectonically active throughout the Phanerozoic (Cordani supersequences (Fig. 15), predominantly siliciclastic in nature, but also
et al., 2016, and references therein), did not favor the preservation of comprising the largest (in volume) basaltic lava flows on continental
older impact structures. This terrane corresponds, partially or entirely, crust, the Paraná-Etendeka Large Igneous Province (LIP). The maximum
to the territories of countries such as Argentina, Chile, Bolivia, Peru, thickness of the basin infill reaches 7000 m in its central area; this
Ecuador, Colombia and Venezuela. Again, one might surmise that these comprises both sedimentary and igneous strata (e.g., Zalan et al., 1991;
terrains should likely contain younger impact structures. The prevailing Peate, 1997; Milani et al., 2007).
climates in these regions are comparatively drier, which should con- There are four confirmed impact structures in the Paraná Basin:
tribute to the preservation of small impact craters and meteoritic ma- Araguainha Dome, Vargeão Dome, Vista Alegre and Cerro do Jarau
terial, such as meteorites, tektites, and micrometeorites. On the other (Fig. 14). Except for Araguainha Dome, which was formed in Permian
hand, relatively high uplift rates throughout the Andean Belt could to Devonian sedimentary strata, reaching the crystalline basement and
have been responsible for erosion of such structures. exposing Neo-Proterozoic rocks, the other three structures were formed
This section presents a review of the record of South America’s in the same stratigraphic unit, namely the Cretaceous Serra Geral For-
currently confirmed impact structures subdivided by country. It must mation, which corresponds to the volcanic rocks of the Paraná-Eten-
also be stressed that all the structures mentioned in this section exhibit deka LIP.
evidence of their impact origin, in the forms emphasized in Sections 3.1 Impact structures formed in volcanic rocks are rare on Earth. Of
to 3.3 above and also treated in detail by, for example, French (1998), those formed in basalt, the best example known is Lonar in India (e.g.,
French and Koeberl (2010), and Reimold and Koeberl (2014a), among Fredriksson et al., 1973; Fudali et al., 1980; Osae et al., 2005; Ray et al.,
others, not allowing any speculative discussion about the nature and 2017). This is a young, well preserved, simple crater of small diameter
origin of these structures. (1.8 km) that was formed at 570 ± 47 ka (40Ar/39Ar method; Jourdan
et al., 2011) ago in the Deccan Trap basalts of India. Today the crater
5.1. Brazil structure is partially filled by a lake. There are other impact structures
on Earth that encompass mafic or felsic volcanic target rocks such as
The current Brazilian impact record (Crósta, 2004, 2012; Crósta and Logancha and El'gygytgyn, both in Russia (Masaitis, 1999; and papers
Vasconcelos, 2013; Crósta and Reimold, 2016) is exclusively associated in a special issue of Meteoritics and Planetary Science, 2013). However,

16
A.P. Crósta et al. Chemie der Erde xxx (xxxx) xxx–xxx

Fig. 14. The Paraná Basin and its impact structures; inset: the Paraná Basin within South America (after Crósta et al., 2010a).

they contain no, or only very limited, exposures of shocked volcanic complex crater in mainly felsic volcanic crust (e.g., Koeberl et al., 2013;
rocks or impact breccias at the surface. Therefore, Lonar crater has Raschke et al., 2013, 2015)
remained the type locality for studying shock effects in basaltic targets, With the discovery and geological characterization of the three
and as an analog for comparative studies of impact processes on ter- large, complex impact structures formed mainly in basaltic targets in
restrial planets with basaltic crusts, such as Mars. El’gygytgyn was the the Paraná Basin of southern Brazil, there are now more alternatives for
subject of an extensive and successful drilling project by the Interna- such analog studies, with the basalts and subsidiary sandstones also
tional Continental Scientific Drilling Program (ICDP) that yielded some exhibiting different levels of shock metamorphism. Furthermore, access
200 m of impact-affected or –generated rocks from the interior of this and logistics for ground truthing are very favorable for all three of these

17
A.P. Crósta et al. Chemie der Erde xxx (xxxx) xxx–xxx

Fig. 15. Simplified stratigraphic column of the Paraná Basin showing its six depositional supersequences and respective litho-stratigraphic units (modified after
Milani, 1997).

structures. surrounded by volcanic rocks, comprising lava, breccia and tuff of


trachytic composition, as well as deformed Paleozoic sedimentary
5.1.1.1. Araguainha. The Araguainha impact structure, also known as rocks.
Araguainha Dome, has a diameter of 40 km and an area of Dietz and French (1973) reported on two probable, eroded impact
approximately 1300 km2 (Fig. 16). It is the largest known impact structures identified on satellite images of NASA’s then newly launched
structure in South America. It is centered at 16° 48′ 45″S/52° 59′ Landsat Program. These were the Araguainha Dome and the Serra da
02″W. Araguainha was the first large impact structure discovered on Cangalha structure (see Section 5.1.2.1). Their completely novel in-
this continent (Dietz and French, 1973; Crósta et al., 1981; Crósta, terpretation (at the time there were only ∼40 impact structures iden-
1982; Theilen-Willige, 1981). The structure was formed at about 253 tified on Earth) was based on the typical morphology of eroded impact
Ma ago in Paleozoic sedimentary rocks of the northeastern-most part of craters, including the rim, the central uplift, and deformed strata. A
the Paraná Basin and underlying crystalline rocks from the basement. brief exploratory visit to Araguainha by R.S. Dietz revealed distinct sets
The structure straddles the border between the Mato Grosso and Goiás of planar deformation features (shock lamellae) in several crystal-
states in central Brazil. lographic orientations in quartz grains from breccias from the central
From bottom to top, the stratigraphic units affected by the impact area of Araguainha.
are: the crystalline basement, composed of a porphyritic alkali granite, Comprehensive evidence for the impact origin of the Araguainha
and phyllites and metasandstones of the Cuiabá Group; the Devonian Dome was first provided by Crósta et al. (1981), Theilen-Willige (1981),
Paraná Group (Furnas and Ponta Grossa formations), the Carboniferous and Crósta (1982). They presented the results of geological and geo-
Tubarão Group, specifically the Aquidauana Formation, and the morphological mapping, as well as detailed petrographic and micro-
Permian Passa Dois Group that comprises the Iratí and Corumbataí structural analysis of rocks from the granitic basement, the sedimentary
formations (Zalan et al., 1991; Milani et al., 2007). The geological map strata, and impact breccias. All lithologies were shown to exhibit
of Araguainha is shown in Fig. 17. abundant evidence of shock deformation, such as shatter cones and
First reference to an anomalous circular structure in this part of the planar deformation features (PDF) in quartz. Besides these diagnostic
Paraná Basin was made by Northfleet et al. (1969), who interpreted it shock deformation features, Crósta et al. (1981) presented a collection
as the result of an igneous intrusion of alkaline nature and Cretaceous of complementary deformation features found in rocks from the central
age, which would have uplifted and deformed the Paleozoic sedimen- uplift of Araguainha. This included planar deformation features (PDF)
tary rocks – thus forming a dome structure. Because of this inter- in quartz, kinkbanding in mica, selective melting of feldspar, and dia-
pretation, and considering the economic potential of similar Cretaceous plectic silica and feldspar glass. Crósta (1982) described the widespread
alkaline intrusions in the Paraná Basin, a specific geological survey of occurrence of radial as well as concentric fault zones throughout the
Araguainha Dome was conducted by Silveira Filho and Ribeiro (1971), entire structure, from the rim to the central uplift. This faulting conveys
producing the first schematic geological map of the structure. These a multi-annular aspect to this structure.
authors referred a cryptovolcanic structure, with a granitic core be- von Engelhardt et al. (1992) presented results of a detailed petro-
longing to the crystalline basement of the Paraná Basin exposed at the graphic and geochemical study of samples from the inner portion of the
center. According to them, the central core of the structure was central uplift, focusing on the granitic basement and the impact

18
A.P. Crósta et al. Chemie der Erde xxx (xxxx) xxx–xxx

Fig. 16. View of Araguainha Dome on a Landsat Enhanced Thematic Mapper+ (ETM + ) false color image (bands 4, 5 and 3 as red, green and blue); the scene
measures about 50 × 50 km. The towns of Araguainha (∼5 km to the SW of the center) and Ponte Branca (near the northeastern rim) are shown, as well as the
Araguaia River, which meanders around the central part of the structure (central uplift) running from S to NW. It forms the border between the states of Mato Grosso
and Goiás.

breccias. They recognized three types of breccias: impact breccia with kilometer wide, lateral inward-directed movement of the supracrustal
melt matrix, polymict impact breccia, and monomict sandstone impact target rocks, accompanied by differential slip between these strata. In
breccia. The authors also identified the occurrence of red and gray comparison, vertical movement would have been restricted to distances
dykes cross-cutting the granite and ranging in thickness from < 10 to of less than a few hundred meters along radial and concentric fault
some 100 cm, and in length up to tens of meters. They were shown to zones around the crater rim. They also observed kilometer-scale folding
contain cataclastic mixtures of shocked and unshocked granitic mate- of the strata in the inner 15 km wide sector of the crater floor and
rial, as well as melts. They pointed out that these dykes must have been thickening of the strata in the collar of the central uplift. Results of
intruded into the granite under extensional stress conditions, most further structural work on the structure were reported by Lana et al.
likely during the uplift of the granitic core as a result of collapse of the (2007, 2008), who provided a structural division of the structure into a
transient crater, during the modification stage of the crater formation. central peak, an annular basin, and at least two concentric inner rings.
Lana et al. (2006) analyzed the constraints on target rock movement Hippertt et al. (2014) reported on a detailed structural study on the
during the collapse of Araguainha’s transient crater. They described the Furnas Formation in the central uplift area.
occurrence of folding and bedding-parallel shearing related to several The central uplift, about 10–12 km wide in total, includes the actual

19
A.P. Crósta et al. Chemie der Erde xxx (xxxx) xxx–xxx

Fig. 17. Geological map of Araguainha Dome and a schematic section (A-B) across the center of the structure (after Thomé-Filho et al., 2012).

20
A.P. Crósta et al. Chemie der Erde xxx (xxxx) xxx–xxx

Fig. 18. Panoramic view of part of the central uplift of Araguainha Dome. The hills to the left and center comprise mostly impact breccias and Furnas sandstone,
whereas the ones to the right comprise Furnas sandstone. Image width is ca. 4 km.

Fig. 19. Schematic lithological map of the


central part of the Araguainha Dome central
uplift. The map is based on Fig. 2 of Yokoyama
et al. (2012), modified with some observations
by the authors. Recent papers (e.g. Machado
et al., 2009; Yokoyama et al., 2012, Silva et al.,
2011, 2016), based on a statement by von
Engelhardt et al. (1992), have proposed the
occurrence of remnants of a melt sheet within
the central uplift; thus, the unit “Melt sheet” is
shown on this map. However, the authors of
this review do not want to imply that there
ever was a melt sheet covering this area, or
that the veins and blobs of melt rock observed
in the field represent remnants thereof. Only
small, < 15 m2 outcrops of melt rock have ever
been observed.

central peak and the innermost part of the annular basin (Lana et al., rocks have experienced three metamorphic phases: the first, a regional
2007, 2008; Yokoyama et al., 2012) (Fig. 18). The outer limit is set metamorphism linked to the Brasiliano-Pan African Cycle
where the Ponta Grossa Formation partially overlaps the Furnas strata, (Neoproterozic), the second, a thermal metamorphism related to the
with upright to overturned attitude, and comes into contact with the intrusion of the Serra Negra granite (locally also called Araguainha
Aquidauana Formation. This is well observed along the Araguaia River, granite), and the last, shock metamorphism related to the impact. The
about 5 km to the south of Ponte Branca town. The schematic geological best outcrops are to the west of the road between Ponte Branca and
map of the inner part of the central uplift is shown in Fig. 19. Araguainha town, to the north of the granite plug, where also well-
The Cuiabá Group, composed of phyllites and metasandstones, is the developed shatter cones can be observed in extensive road-cuts
oldest stratigraphic unit within the central uplift of Araguainha. These (Fig. 20).

21
A.P. Crósta et al. Chemie der Erde xxx (xxxx) xxx–xxx

Fig. 20. Three examples of shatter cones from the western sector of the
Araguainha central uplift, where excellent outcrops of phyllites from the Cuiabá
Group occur. In a) and c), pens for scale are ca. 13 cm long. Coin in b) is 2.5 cm
across.

The central peak includes a 4–5 km wide inner plug of porphyritic


alkali granite that displays different degrees of cataclasis (Fig. 21). It is
surrounded by occurrences of polymict breccia and, in some parts, by a
1–2 km wide collar of highly deformed rocks of the Furnas Formation.
(caption on next page)
The granite is composed of 0.5–2 cm long, euhedral megacrysts of K-
feldspar in a coarse-grained matrix of plagioclase, biotite, quartz, and
accessory zircon, titanite, magnetite, hematite, and tourmaline. Xeno-
liths of metamorphosed sandstone and siltstone, probably from the host

22
A.P. Crósta et al. Chemie der Erde xxx (xxxx) xxx–xxx

Fig. 21. a) Typical aspect of the porphyritic alkali granite of the Araguainha formation to the granite, the Cuiabá Group, have been observed.
Dome; megacrysts of K-feldspar up to 6 cm long, immersed into a coarse- to On the latest geological map of the structure, created by the
medium-grained matrix composed by K-feldspar, quartz, biotite and albite. b) Institute of Geosciences of the University of Brasilia (UnB-IG, 2012,
Local deformation and melting in the granite are frequently observed. c) and d) unpublished), sedimentary rocks belonging to the Río Ivaí Group
Aspects of the transitional, partially impact-melted granite, after Fischer
(Ordovician) were identified for the first time. They occur locally be-
(2015). Pink schlieren composed of feldspar are visible. Cores of quartz are
tween the Furnas Formation and the alkali granite, with the best out-
surrounded by dark lines that represent neoblastic biotite. In a), hammer for
scale ca 35 cm long. Coin in b) is 2.8 cm in diameter, and coins in the other crops being in the northern and southern sectors, just inside of the
figures are 2.5 cm in diameter. collar-like occurrences of Furnas Formation quartzite and sandstone.
The Furnas Formation is the more prominent unit in the central
uplift (Fig. 22). It constitutes a 400–800 m thick pile of clastic quartzose
strata. They are segmented into fault-bounded blocks that have ex-
perienced varied rotation, resulting in a megablock zone around the
granite core (Lana et al., 2007, 2008; Hippertt et al., 2014). The most
prominent exposures occur in the northeastern parts, in the Serra da
Arnica, and to the southeast, on the Morro Pelado hill. To explain the
large-scale folding of the Furnas sandstone strata around the central
uplift within the short time frame of an impact event (seconds/minutes)
and the absence of significant friction melts, Hippertt et al. (2014) re-
cently drew on a combination of microbrecciation, cataclasis, and fluid
expansion.
The annular basin, between the central peak and the outer annular
ring, comprises strata of the Ponta Grossa Formation and the lower part
of the Aquidauana Formation. The Ponta Grossa Formation is best ob-
served along the Araguaia River, where its strata are seen to be strongly
folded (Fig. 22-B).
The annular ring between the annular basin and the crater rim
comprises two main ring features at 10–12 km and 14–18 km from the
center of the structure (Lana et al., 2007). The inner ring structure
marks the extent of the deformation related to the outward movement
of the target rocks (Lana et al., 2008). A 700–800 m thick section of
sandstones and conglomerates, overlain by stratified siltstones and
sandstones belonging to the Aquidauana Formation of the Tubarão
Group (Carboniferous) are the more conspicuous sedimentary units in
the annular ring. The Passa Dois Group, comprising the Iratí and Cor-
umbataí formations, is well preserved as lenses in the areas of normal
faults (in the form of graben or halfgraben structures) (Fig. 23). About
20–40 m thick Permian siltstone, chert and carbonate of the Iratí For-
mation and 80–100 m thick Permian-Triassic siltstones and subsidiary
sandstones and chert of the Corumbataí Formation were identified.
The rocks of the Araguainha Dome that were generated by the im-
pact occur in the central uplift (Fig. 24). von Engelhardt et al. (1992)
distinguished impact breccia with melt matrix (impact melt rock), as
well as red and gray dikes (Fig. 24B, C, D) and polymict and monomict
breccias. The dikes are characterized by angular to subrounded clasts of
granite and schlieren of mobilised feldspar that are immersed in a fine-
grained groundmass (Fig. 24D). von Engelhardt et al. (1992) showed a
Fig. 22. a) Folded and recrystallized sandstone of the Furnas Formation, large occurrence of monomict breccia described as cataclasites, mainly
Araguainha Dome; b) folded siltstone and shale of the Ponta Grossa Formation.
in the south-eastern sector of their schematic map of the central uplift.
Later workers have not confirmed this (e.g., Yokoyama et al., 2012),
and we can not substantiate these extensive occurrences of monomict
breccia either. von Engelhardt et al. (1992) also alleged that the impact

Fig. 23. Strongly deformed strata of the Passa Dois Formation (Irati and Corumbataí members) along the road MT-100 that cuts the Araguainha impact structure in
the southwestern sector.

23
A.P. Crósta et al. Chemie der Erde xxx (xxxx) xxx–xxx

Fig. 24. Araguainha Dome. a) The contact between the porphyritic alkali granite (spotted aspect) and the impact melt rock (more homogeneous). Inset: a porphyritic,
fluidal-textured impact melt rock, in which some granite clasts are recognized in the red, aphanitic melt matrix. b) and c) Examples of two red dykes, in b) of 20 cm
and in c) of 10 cm width. d) Macroscopically, the melt rocks in the dykes are characterized by fluidal texture, where K-feldspar and granite fragments may be well
aligned. e) and f) Two examples of pseudotachylitic breccia (PTB) – images from Preuss (2012). Scale bars in cm, pens are ca. 13 cm long and coin has a diameter of
2.5 cm.

melt rocks form at least two big continuous outcrops on the top of the demonstrated that there is no continuous outcrop and that there is no
hills in the south and central part of the central uplift, and three smaller direct relationship between occurrences of impact melts rock and the
ones in the north. tops of hills – as suggested by the work of von Engelhardt et al. (1992)
On the basis of field observation and petrographic studies, the im- and Machado et al. (2009).
pact melt rocks were classified variably by different authors (von Several petrographic and geochemical studies have been carried out
Engelhardt et al., 1992; Machado et al., 2009; Yokoyama et al., 2012, on the granite and impact melt rocks (von Engelhardt et al., 1992;
Preuss, 2012; Fischer, 2015; Silva et al., 2011, 2016). Machado et al. Machado et al., 2009; Yokoyama et al., 2012, Preuss, 2012; Fischer,
(2009) divided the impact melt rocks into melt sheet and melt vein 2015; Silva et al., 2016; Hauser et al., 2017). They showed that the
occurrences, whereby we do not feel confident about the assumption impact melt rocks and the granite are very similar in composition. Some
that a melt sheet, or remnants thereof, is actually in evidence (as also workers stated that seemingly no sedimentary rocks were involved in
questioned by Fischer, 2015). It appears that the sketch by von the production of melt (von Engelhardt et al., 1992; Machado et al.,
Engelhardt et al. (1992) that shows impact melt breccia overlying other 2009). However, there are some slight differences between granite and
rocks of the central uplift, has caused later assumption that remnants of impact melt rock abundances in terms of some lithophile and side-
a melt sheet stratigraphically occurring above shocked alkali granite rophile trace elements, and this observation, coupled with some pet-
existed in the dome. Preuss (2012) introduced the term pseudotachy- rographic findings, could be interpreted that a minor proportion of
litic breccia for selected, cm-wide veinlets of breccia (Fig. 24E, F) cut- Cuiabá Group material may have been mixed into impact melt as well.
ting across the alkali granite that contained a polymict clast population Some PGE analyses were carried out by Hippert and Lana (1998) and
(in contrast to the impact melt veins that essentially only show rem- Machado et al. (2009). These data were interpreted to indicate that the
nants of alkali granite precursor, besides some comparatively rare, Araguainha impact melt rock does not carry an identifiable meteoritic
oxidized and possibly melted clasts that might be derived from the component.
Cuiabá Group). Fischer (2015) introduced the term transitional granite Large outcrops of polymict impact breccia occur in the western,
(Fig. 21C, D) for partially impact-melted alkali granite. For comparison, northern and northeastern sectors of the central uplift (von Engelhardt
and in order to see possible equivalents, a compilation of the varied et al., 1992), and more restricted in the southern part (Yokoyama et al.,
nomenclatures applied in the past is presented in Table 1. 2012; UnB-IG 2012) (Fig. 25A). The main components of the polymict
Yokoyama et al. (2012) and UnB-IG (2012) reported individual breccia were summarized by von Engelhardt et al. (1992) and include
occurrences of impact melt rock within the larger areas mapped by sandstone clasts of different grades of metamorphism, with or without
earlier workers – demonstrating that these do not represent continuous alteration, whereby some of these are completely fused; often squarish
bodies. Fischer (2015) carried out extensive field work in the southern clasts of red pelites that are probably derived from the Aquidauana
part of the central uplift, where in theory there should have been some Formation; rounded quartzite pebbles with fractures – likely from the
of the large outcrops referred to by von Engelhardt et al. (1992). She Furnas Formation (Fig. 25B); strongly deformed and laminated, folded

24
A.P. Crósta et al. Chemie der Erde xxx (xxxx) xxx–xxx

Impact melt rock including transitional granite in the form of


Metamorphic basement (Cuiabá Group)

Shocked porphyritic alkali granite

Suevite and lithic impact breccia


melt sheet and melt dykes(?)
Fischer (2015)

Pseudo-tachylitic breccia (Type 2


Impact melt rock (Type 1

Suevite and lithic breccia


Monomictic breccia
Preuss (2012)

breccia)

breccia)
Granite
Partially molten granite

Monomictic breccia
Silva et al. (2011)
Comparison of rock nomenclatures applied to Araguainha lithologies by various authors. Modified from Fischer (2015).

Polymict breccia
Shocked granite

Melt dykes?

Melt sheets

Fig. 25. a) Aspect of the polymict breccia in the northern part of the
Araguainha central uplift. Some of the laminated sedimentary clasts are plas-
Dikes and veins filled with molten

Rocks with a lot of relict material

tically deformed. Section of a hammer-head shown, for scale, ca. 14 cm wide. b)


Rocks with a few relict material
Metamorphic basement (Cuiabá

Multi-faulted quartz pebble from coarse conglomerate, probably derived from


either the Furnas Formation (Devonian) or the Rio Ivaí Group (Ordovician-
Silurian); float collected in the southern part of the central uplift in an area with
Granite (porphyritic)

extensive occurrence of polymict breccia. Note that the individual segments of


Polymictic breccia
Yokoyama (2008)

the pebble have been displaced against each other. Scale bar in cm.

and purple/yellow clasts probably from the Passa Dois Group; and
material
Group)

minor proportions of clasts derived from alkali granite, as well as


phyllite from the Cuiabá Group.
Preuss (2012) showed that there are distinct melt rock clasts in the
Machado et al. (2009)

polymict breccia from the northern sector – which therefore should be


Monomictic breccia
Polymictic breccia
Crystalline alkali

classified as a suevite. Some of these melt clasts are quartzitic in com-


position. De Marchi et al. (2016) studied the polymict breccia in a thick
Melt sheet
Melt veins

sequence in the northwestern region. Her observations indicate that in


granite

terms of composition, size and shape of clasts there is no general trend


along a vertical axis through this package. However, alkali granite and
likely Cuiabá Group derived clasts are significantly more abundant in
Monomictic impact breccia from
Impact breccia with melt matrix

the lower part of this breccia package than higher up.


von Engelhardt et al. (1992)

5.1.1.1.1. Geophysical characteristics of the Araguainha Dome. The


Polymictic impact breccia

region where the Araguainha Dome is located was covered by a


Red dykes/gray veins

regional aerial geophysical survey in the 1970s, the Alto Garças


Project (PROSPEC, 1972). It comprised gamma spectrometry and
magnetic methods. At the time, the structure was still interpreted as
sandstone

an alkaline igneous intrusion, with potential for uranium


Granite

mineralization. The main goal of the survey was to identify areas


with anomalous radioactive signals, and the Araguainha Dome stood
out in the region as a conspicuous circular and concentric radiometric
Melt breccia
Basement

anomaly (Fig. 26). A circular potassium anomaly occurs in the area


Table 1

where the granitic core is exposed at the center of the structure,


Dikes

surrounded by a circular area of low gamma signal, corresponding to

25
A.P. Crósta et al. Chemie der Erde xxx (xxxx) xxx–xxx

Fig. 26. Regional gamma spectrometric survey


over the Araguainha Dome, showing the an-
nular and concentric patterns of the total count
(TC) signal, comprising the intensities for po-
tassium, uranium, and thorium. This bulĺs eye
pattern exhibits remarkable coincidence with
the local geology (see Fig. 17 for comparison).
(Gamma image courtesy of Johann Lambert
Silva, Unicamp).

Fig. 27. a) Landsat ETM+ color composite image showing Vargeão Dome with its multi-annular inner features. The outer white dashed line indicates the rim of the
structure, whereas the inner circle marks the limit of the central uplift. b) A 3D perspective view of the structure from SRTM (Shuttle Radar Topographic Mission)
digital elevation data. Image width is ca. 20 km.

the occurrence of the Furnas sandstones. This area is, in turn, correlation with the regional geological map of the structure (Fig. 17).
surrounded by another circular anomaly of high potassium, thorium The interpretation of the magnetic data from this aerial survey
and uranium, corresponding to the area of occurrence of the Ponta provided information about the depth of the basement underneath the
Grossa Formation. Continuing towards the border of the structure, there structure, which ranges from a maximum of 1500 to 1700 m at the rim
is an ample area of very low gamma signal with sparse, small areas of to progressively lower values towards the center, where the basement is
low potassium signal. At approximately 20 km from the center, and exposed. A highly disturbed magnetic pattern was also identified within
surrounding almost the entire structure, a striking circular anomaly of the structure, contrasting with the predominatly NE-SW oriented pat-
high potassium marks the rim of the structure, corresponding to the tern typical of the region surrounding the Dome.
areas of occurrence of the deformed strata of the Passa Dois Group. Masero et al. (1994) and Fischer and Masero (1994) employed the
These gamma anomalies confer an overall appearance of a bull’s eye magnetotelluric method to estimate the depth of the basement and the
pattern to Araguainha, as seen on Fig. 26, which shows a remarkable total displacement of the central uplift. They concluded that in the

26
A.P. Crósta et al. Chemie der Erde xxx (xxxx) xxx–xxx

Fig. 28. Schematic cross sections along the N-S (top) and W-E (bottom) directions across the center of Vargeão Dome. Note the morphology of the structure, with
topographic gradients of up to 150 m between the rim and the interior of the structure, as well as the remarkable structural control of the topography related to high-
angle faulting (after Crósta, 2012).

Fig. 29. Interpreted seismic section for Vargeão Dome and the surrounding area (Kazzuo-Vieira et al., 2009; Crósta et al., 2011).

annular area located between 9 and 20 km from the center, the base- wide conductive source near the center of the structure based on
ment forms a symmetric ring at 1 km depth. From 9 km distance to- magnetotelluric data, which would coincide with the area where the
wards the center of the structure, the top of the basement rises up, granitic core is exposed.
reaching the surface at 1.5–2 km from the center. Masero et al. (1997) Aeromagnetic data allowed characterizing the magnetic signature of
performed 2-D and 3-D modeling of the magnetotelluric data, sug- the granite at the center of the structure. Vasconcelos (2007) estimated
gesting the existence of an elliptic zone within the granite lying at the depth of the basement through power spectrum analysis of the
depths between 3 and 7 km, characterized by resistivity values below aeromagnetic data. The analysis pointed out a deeper source lying at
those of the upper crust. They interpreted this anomaly as the result of around 1 km depth, on average, which reinforces the magnetotelluric
impact-induced deformation (brecciation and faulting). Schnegg and interpretation.
Fontes (2002) also suggested the existence of a ∼1 km deep and 4 km Electrical resistivity tomography data were acquired and

27
A.P. Crósta et al. Chemie der Erde xxx (xxxx) xxx–xxx

Fig. 30. a) 3-D gravity model of the Vargeão Dome showing the central uplift as a pronounced gravimetric low, with an approximate diameter of 3 km; b) Depth map
of the of the Pirambóia/Botucatu sandstone in relation to the topographic surface; c) an enlarged view of the central uplift showing the location of the sandstone
outcrops (in red); the dashed line shows the position of the central uplift established in previous studies using remote sensing data, which is off-set in relation to the
gravimetric low (Ferreira et al., 2015).

interpreted by Tong et al. (2010) for characterizing the subsurface Araguainha samples. According to these authors, this method provided
geometry of the central uplift. They distinguished granitic from sedi- a paleomagnetic record as reliable as those obtained from volcanic
mentary rocks based on resistivity values. In addition, zones with low rocks.
resistivity values within the granite were interpreted as being due to 5.1.1.1.2. Dating the Araguainha impact event. Crósta (1982)
higher porosity related to deformation. presented the first radiometric age dating for the deformed granitic
Studies using magnetic anisotropy conducted by Yokoyama et al. basement. He obtained a K/Ar age of 283.6 ± 17.2 Ma for K-feldspar
(2012) allowed to track the orientation of microscopic deformation and pointed out that this age was likely a mixed age between the
features (PDF and PF) in minerals from the granitic core, which are original, pre-Devonian age of the granite and the impact event, as the
decorated by magnetite and hematite. The orientation of magnetic shock level of the dated sample was probably not high enough for a
lineation and magnetic foliation shows a regular pattern at the border complete loss of inherited Ar.
of the central peak, consistent with the fabric of sedimentary strata at A 40Ar/39Ar cooling age was obtained on biotite from the alkali
the structure's inner collar. In contrast, they found a complex pattern at granite by Hammerschmidt and von Engelhardt (1995), which sug-
the center of the structure. The authors pointed out that the cataclastic gested that the granite passed through the 300 °C isotherm before 481
flow obtained from microstructural observations and the structural Ma. A younger Rb-Sr cooling age of 449 Ma was also obtained by
pattern of the magnetic anisotropy match the predictions from nu- Deutsch et al. (1992). On the basis of geochemistry and age, Lacerda
merical modeling of complex impact structures. Yokoyama et al. (2014) Filho et al. (2004) related the Araguainha granite to the Serra Negra
used samples from the shocked granite of Araguainha, together with Suite, a group of post-tectonic granites with U-Pb ages between 462 and
samples from impact melt and the alleged melt sheet, to establish a new 503 Ma.
magnetic pole for the Pangea supercontinent in the Permo-Triassic. Hammerschmidt and von Engelhardt (1995) established the age of
Alternating field and thermal demagnetization indicate a stable re- the impact event using the 40Ar/39Ar method on melt rock. They ob-
manent magnetization carried by magnetite and hematite from the tained two ages of 245.5 ± 3.5 Ma and 243.3 ± 3.0 Ma, thus

28
A.P. Crósta et al. Chemie der Erde xxx (xxxx) xxx–xxx

Fig. 31. Simplified geological map of the Vargeão Dome impact structure (after Crósta et al., 2011).

indicating that the impact event had occurred around the Permo- this context, attention turned to the Araguainha structure, as the ages
Triassic (P-Tr) boundary. These results were subsequently re-calculated obtained since Hammerschmidt and von Engelhardt (1995) indicated
by Jourdan et al. (2009) to 248 ± 4 Ma. an age of formation, within error, very close to the P-T boundary.
In terms of Sr-Nd isotopes, Deutsch et al. (1992) obtained a Rb-Sr Mutter et al. (2008) examined the potential role that the Araguainha
isochron (whole rock + albite + biotite) of 449 ± 9 Ma for shocked impact event might have had in the P-Tr extinction, based on pa-
granite and a second isochron of 243 ± 19 Ma (whole rock and cor- leontological data collected in outcrops of the strata belonging to the
dierite) on a sample of impact melt rock. The Sm-Nd results for the Permian Passa Dois Group, around and at the rim of the impact struc-
impact melt rock show that the whole rock, cordierite and magnetite ture. They pointed out that the magnitude of the Araguainha impact
data form a regression line of 476 ± 110 Ma, similar to the assigned event would not have been large enough to produce the global mass
age of the shocked granite, demonstrating that the Sm-Nd system was extinction event of the P-Tr boundary, as it is largely known based on
somewhat disturbed. modeling of large impact events, but would rather be compatible with a
Tohver et al. (2012) dated samples of impact-generated melts using regional extinction event. However, these authors did not offer a con-
the U-Th-Pb methods on neocrystallized monazite and 40Ar/39Ar on clusion, only stating that it was unclear whether the Araguainha impact
quartz inclusions from impact melt. They obtained a 254.7 ± 2.5 Ma predated, coincided, overlapped with, or postdated the P-Tr extinction.
weighted mean age for the impact event. Recently, Erickson et al. A different hypothesis linking Araguainha and the P-Tr extinction
(2017) obtained a similar U-Pb concordia age of 259 ± 5 Ma on neo- was proposed by Tohver et al. (2013). In their view, even a moderate
blastic monazite from impact melt rock. size impact event, such as the one that formed the Araguainha impact
Hauser et al. (2017) made some Sr-Nd isotope analyses on target structure, could be responsible for a large-scale mass extinction, if a
rocks (granite and Cuiabá Group samples), impact melt rock, and large enough amount of methane would be released by the impact. In
transitional granite. Their results show that the impact melt rocks, the the case of Araguainha, the authors associated a possible release of
transitional granite and the granite are homogeneous in terms of great amounts of methane to three possible mechanisms: volatilization
87
Sr/86Sr and 143Nd/144Nd, whereas a melt clast found in suevite has an of the bolide itself, volatilization of the target rocks, and methane re-
isotopic composition more similar to the Cuiabá Group samples. leased by impactogenic seismicity. They pointed out the C-rich rocks of
5.1.1.1.3. A possible relationship between the Araguainha impact event the Permian Iratí Formation that occur throughout the Paraná basin as
and the Permo-Triassic boundary. The mass extinction event that the potential source of the methane for the latter two mechanisms.
happened at the P-Tr boundary occurred about 252 Ma ago and Tohver et al. (2013) then used mass balance calculations to estimate the
caused the disappearance of > 90% of all marine species and 70% of total amount of methane released by these three mechanisms to ca.
the terrestrial vertebrate species (Erwin, 1990; Shen et al., 2011). The 1600 Gt. According to their interpretation, this massive release of me-
causes for what is known as the ‘Great Dying’ has been under severe thane would have had immediate consequences for the biosphere, with
scrutiny for decades and are still a matter of intense debate. climate perturbation effected by a rapid spike in greenhouse gases.
Based on what is currently known about the younger mass extinc- Their hypothesis, albeit interesting, still needs to be scrutinized and
tion event at the Cretaceous-Paleogene (K-Pg) boundary, and its direct tested, before any conclusion as to the direct connection between the
relationship with the Chicxulub impact event (Schulte et al., 2010), Araguainha impact and the P-Tr mass extinction can be substantiated.
researchers started to look for a similar direct link between the P-Tr Besides, the recent dating by Erickson et al. (2017) has poised the
extinction event and a large impact structure of such age on Earth. In question whether the P-Tr boundary and Araguainha impact are really

29
A.P. Crósta et al. Chemie der Erde xxx (xxxx) xxx–xxx

Fig. 32. Vargeão Dome a) Brecciated tholeiitic basalt with intense fracturing; the fractures (red veinlets) and the texture resemble a pseudotachylitic breccia –
however, due to the secondary overprint on the veins it is uncertain what breccia type this material originally represented. b) Brecciated volcanic rock (BVR) with
originally possibly glassy material showing fluidal texture. Scale bars are in cm.

coeval. geosite by SIGEP in 2002 (Crósta, 2002), but this did not result in
5.1.1.1.4. Geoheritage of Araguainha. The Araguainha impact significant heritage development action.
structure has been among the first significant geological features of This first declaration was complemented by a more recent project
Brazil to be included in the geoheritage database coordinated by the conducted by the Brazilian Geological Survey (CPRM), in which some
Brazilian Commission of Geological and Paleontological Sites – SIGEP. geosites of the SIGEP database were selected as potentially important
This is part of a geoheritage project sponsored by UNESCO and the for the establishment of Brazilian Geoparks. CPRM published a book
International Union of Geological Sciences (IUGS) in an effort to with the proposals for the creation of Brazilian Geoparks, with detailed
establish an international network of protected sites of geological geological, environmental, and heritage-related information for each
interest (geosites and geoparks). Araguainha was approved as a site. This publication does include Araguainha (Thomé-Filho et al.,

30
A.P. Crósta et al. Chemie der Erde xxx (xxxx) xxx–xxx

Fig. 34. a) Polymict breccia from the center of Vargeão; b) clast of sandstone
with a shatter cone in polymict breccia. Scale bars are in cm.
Fig. 33. Vargeão Dome a) Monomict sandstone breccia. b) Shatter cones (note
the undulating nature of the lower terminations indicating the curvature on the
striated fracture planes) in sandstone. Scale bars are in cm. in the basalts of the Jurassic-Cretaceous Serra Geral Formation, as well
as in sandstone outcrops in the central part of the structure that they
associated with the Jurassic Botucatu Formation. Owing to the fact that
2012). So far, the Araguainha Geopark has not been created, due to the
these sandstones occur stratigraphically underneath the Serra Geral
lack of potential local partners, such as state and municipal govern-
lava flows, actually at depths of about 1 km in this region, the authors
ments, and non-governmental organizations.
suggested that the structure might have been formed by a buried ig-
Sanchez (2006) studied the geo-touristic aspects of Araguainha
neous intrusion that uplifted the strata, thus exposing the Botucatu
Dome, whereas Sanchez and Brilha (2017) assessed the scientific value
sandstone at the center of the structure.
of the Brazilian impact structures as geoheritage sites. They concluded
The structure was subsequently analyzed in detail by Barbour and
that the Araguainha Dome, together with Serra da Cangalha (see below,
Corrêa (1981), with the purpose of assessing its potential for oil and gas
Section 5.1.2.1), were the ones to rank higher in their assessment,
resources. They produced the first geological map of Vargeão Dome,
calling for the development of geoconservation strategies and adequate
showing the occurrence of four volcanic flows between the rim and the
management of these sites.
central portion of the structure. The lower three flows have a basaltic
composition, whereas the fourth and uppermost flow represents a more
5.1.1.2. Vargeão. The Vargeão Dome structure of 12.4 km diameter is acidic lava type of porphyritic texture. The authors pointed out the
located in the western portion of Santa Catarina State (26°49′S and apparent tectonic nature of the sandstone outcrops near the center of
52°10′W). The structure is named after the homonymous town, which is the structure, with the contacts between them and the basalts marked
located on the inside of the southern rim of the structure (Fig. 27). by faults. They also found sandstone monomict breccias in the interior
Vargeão Dome is a complex impact structure with a central uplift, of the structure, which they related to structural deformation due to the
which was formed in igneous and sedimentary rocks of the São Bento uplift of the sandstone. As to the origin of the structure, they suggested
Group (Pirambóia, Botucatu and Serra Geral formations – Figs. 14 and four possible causes: vertical faulting, cryptovolcanic explosion caused
15) of the Paraná Basin. by either volcanic gases or meteorite impact, or a buried igneous in-
The presence of a circular geomorphological feature at this location trusion.
was first noted by Paiva Filho et al. (1978), through interpretation of Crósta (1987) pointed to a number of similarities with respect to
airborne radar images; they coined the name Vargeão Dome. These deformation of Vargeão Dome with other meteorite impact structures,
authors identified a remarkable pattern of annular and radial fractures including Araguainha Dome that was, at that time, the only confirmed

31
A.P. Crósta et al. Chemie der Erde xxx (xxxx) xxx–xxx

geophysical signatures. The characteristics of the structure presented


next are based on the findings of these authors.
Vargeão Dome was formed in lava flows of mafic and felsic com-
position that belong to the Serra Geral Formation. Some of the re-
markable aspects of this structure are the circular and multi-annular
morphology, as shown by remote sensing images and digital elevation
data (Fig. 27). This morphology is due to structural control by sub-
vertical faulting, related to the modification stage of crater formation,
with the uplift of the sandstone in the central area and rim collapse,
followed by the gravitational collapse of the central uplift. The struc-
tural framework of Vargeão Dome is shown in the schematic cross-
section of Fig. 28. Current topographic escarpments of up to 150 m
elevation between the rim and the inner portion of the structure cor-
respond to offsets along the planes of listric faults that mark the col-
lapse of the rim. Other annular features observed in the interior of the
structure are believed to also represent the surface expression of high-
angle collapse-related faults.
In the central uplift area there are occurrences of highly deformed
sandstone that are in structural contact with brecciated volcanic rocks
(mostly basalt). These sandstones are currently related to the
Pirambóia-Botucatu succession of the Late Permian to Early Cretaceous
São Bento Group (Donatti et al., 2001). The Botucatu eolian system
rests abruptly on the fluvio-eolian deposits of the Pirambóia system,
with both systems being covered by the Serra Geral volcanic flows. The
degree of cataclasis exhibited by the sandstones in the central portion of
the Vargeão structure does not allow differentiating between these two
stratigraphic units, which are therefore referred to only as the Pir-
ambóia-Botucatu sandstones (Crósta et al., 2011).
Kazzuo-Vieira et al. (2009), Crósta et al. (2011), and Ferreira et al.
(2015) conducted geophysical analysis of Vargeão Dome. Using seismic
data acquired for oil exploration in the 1980′s, Kazzuo-Vieira et al.
(2009) and Crósta et al. (2011) interpreted the subsurface stratigraphy
and structures in the interior of the structure and its surrounding region
(Fig. 29). Their results depict Vargeão Dome as a shallow crustal
structure, as one would expect for an impact crater, with the rim
Fig. 35. a) Shatter cones on basalt collected in the central uplift of Vargeão; it is bounded by listric faults – likely related to the collapse phase of cra-
evident that they originate from a fracture; b) shatter cones on fine-grained tering, and a marked central uplift characterized by a conical structure,
tholeiitic basalt. in which sedimentary strata have been uplifted by up to 4 km and are
deformed. Ferreira et al. (2015) produced a 3D gravity model, based on
gravity data acquired in the field. Their model (Fig. 30) shows the
impact structure in Brazil. Among the deformation features mentioned
position of the central uplift, which differs slightly from the inter-
were planar fractures in quartz grains from sandstone in the central
pretation from satellite images presented by Kazzuo-Vieira et al.
portion of the structure.
(2009). The model of these authors also shows the depths of the Pir-
Kazzuo-Vieira et al. (2009), Crósta et al. (2005, 2011), Crósta
ambóia-Botucatu sandstone strata in relation to the present topographic
(2012), and Ferreira et al. (2015) analyzed the geological, morpholo-
surface. This, in turn, helped to explain the current position of the fault-
gical and deformation features of Vargeão Dome, as well as its

Fig. 36. a) Feather features (FF) in quartz from a conglomeratic level of the Pirambóia/Botucatu sandstone exposed at the central uplift of Vargeão; b) curvilinear
microfractures of a FF in quartz from another sample from the same location as the previous figure (Crósta et al., 2011).

32
A.P. Crósta et al. Chemie der Erde xxx (xxxx) xxx–xxx

Fig. 37. A 3-D perspective view of the Vista Alegre impact structure, obtained by combining data from Landsat-7/ETM+ with a digital elevation model (GDEM)
generated from Terra/ASTER data. Scene shown is ca. 14 km wide.

Fig. 38. Panoramic view of the Vista Alegre structure taken from the northern rim, looking towards the south. The near-vertical escarpments that mark the outer rim
of the structure can be seen well on the right-hand side. The width of this view is approximately 6 km.

bounded sandstone outcrops at the central uplift. rhyodacite corresponding to the uppermost flow. Faulting along the rim
The simplified geology of Vargeão Dome is presented in Fig. 31, resulted in large blocks of undeformed volcanic rocks from these two
showing the main lithologies found in the interior of the structure. Most units slumping from the upper portions of the rim towards the interior
of the rocks that occur within the structure show some degree of de- of the structure, where they are currently found in tilted positions.
formation (ranging from brecciation to melting). Rock types found in the interior of Vargeão Dome comprise un-
Two types of volcanic rocks occur in the dome that belong to the shocked/slightly to moderately fractured basalt and rhyodacite in outer
Serra Geral Formatuion: tholeiitic basalt of the Alto Uruguai Unit, and parts of the central uplift (the so-called collar) and along its rim, and
porphyritic rhyodacite of the Ácidas Chapecó Unit. On the escarpments brecciated/ fractured/melted rocks. The latter includes the two vol-
of the southern rim, these volcanics are well exposed, with the tholeiitic canic rock types as well as sandstones. There are also monomict brec-
basalt corresponding to the lower three lava flows and the porphyritic cias made up of any of the three major rock types, and polymict

33
A.P. Crósta et al. Chemie der Erde xxx (xxxx) xxx–xxx

Fig. 39. Residual Bouguer anomaly map for the Vista Alegre impact structure (after Ferreira et al., 2013); a) plan view; b) perspective view with the rim of the
structure indicated by a black line.

breccias containing fragments of all these lithologies. In the geological dominant type of BVR in Vargeão, although at some locations such as in
map of Fig. 31, these breccias – with the exception of brecciated the central-southern part of the structure there are BVR made of brec-
sandstones – have been grouped into a single unit called ‘brecciated ciated rhyodacite. The degree of brecciation varies considerably at
volcanic rocks and polymict breccias’ (BVR), because the lack of con- different locations. All the interstices and fractures among the clasts are
tinuous exposures and the current level of mapping of these breccia filled with an intensely oxidized matrix, thus making the more brec-
types is not sufficient to separate distinct types. ciated varieties appear more reddish compared to the less brecciated
The BVR contain clasts of dark gray basalt or rhyodacite in a poorly varieties, which appear light gray. This alteration also precludes iden-
sorted and heavily oxidized matrix of intense red color. They are found tification of the original breccia matrix type. Locally, the BVR show
in scattered outcrops at several locations in the interior of the structure. fluidal texture (dark color) comprising glassy material (Fig. 32b).
The volcanic clasts, in turn, are strongly fractured, with the fractures Geochemical classification (major and trace elements) of the BVR in
being filled by a reddish material, often in vein form. These red veinlets comparison with the target rocks (basalt, rhyodacite and sandstone
comprise finely comminuted fragments in an oxidized matrix, similar to collected outside the structure) showed that they fall between the basalt
the matrix of the BVR. The appearance of the BVR is that of pseudo- and the rhyodacite compositional clusters, but slightly closer to the
tachylitic breccia (Fig. 32a, b). Brecciated tholeiitic basalt forms the basalt end (Crósta et al., 2011).

34
A.P. Crósta et al. Chemie der Erde xxx (xxxx) xxx–xxx

Fig. 40. Schematic geological map of the Vista Alegre impact structure (after Crósta et al., 2010a).

The sandstone outcrops are restricted to the central portion of the to the melt found on the surface of some shatter cones from Vista Alegre
structure, where they occur in the form of large blocks (up to several (Pittarello et al., 2015a).
hundred meters in size), arranged in a semi-annular pattern around the In addition to the diagnostic evidence for impact provided by the
center; they are bounded by faults (Fig. 31). The sandstones are related shatter cone finds, microscopic shock deformation features have also
to the Pirambóia-Botucatu succession (Crósta et al., 2005). Logging been described in rocks from Vargeão Dome, in particular in the
data from an oil exploration well (1RCH-0001-SC) from 22 km to the monomict sandstone breccia from the center of the structure (Crósta
NE of the rim of Vargeão Dome places the Botucatu Fm. at 980 m depth et al., 2011). This includes abundant planar fractures (PF) and some
and the Pirambóia Formation at 1100 m depth (Crósta et al., 2011). The rare feather features (FF) (Fig. 36). The former may give a hint at shock
sandstones on the dome are highly deformed and recrystallized, and deformation, and can be assumed to be shock induced when they occur
abundantly form monomict breccias (Fig. 33a). At a few locations, they abundantly and in sets of multiple orientations per host grain. However,
exhibit shatter cones (Fig. 33b). single sets of such features are also known from tectonic deformation
Polymict breccias at Vargeão have significant contributions from and, thus, should be taken as possible shock indicators with caution.
volcanic rocks as well as from sandstone. Yet, this type of breccia is not Feather features to date have only been described from impact de-
common, having been found at just two locations near the center of the formed rock (e.g., French, 1998, 2004; French and Koeberl, 2010;
structure, in weathered outcrops of only a few meters extent. The French et al., 2004; Poelchau and Kenkmann, 2011). Both features to-
polymict breccias are generally composed of angular, at maximum gether indicate a likely maximum shock pressure of 7–10 GPa, in the
centimeter-sized basalt clasts and minor amounts of sandstone and absence of planar deformation features (PDFs). However, according to
calcite clasts, all set in a fine-grained red matrix (Fig. 34a). The matrix Kowitz et al. (2013a, 2013b, 2016), these values should be considered
is composed of comminuted basalt and rhyodacite, as well as quartz and with caution and – likely – as an upper shock pressure limit, as these
calcite clasts, besides some unidentified fine-grained material. A single microdeformations formed in porous sandstone could be indicative of
fragment of shatter cone has been found in polymict breccia (Fig. 34b). even lower shock pressure.
Shatter cones formed in fine-grained tholeiitic basalt were also Yokoyama et al. (2015) investigated centimeter wide breccia veins
found in the central portion of Vargeão Dome, at a single location near in the shocked basalts of Vargeão from the inner collar surrounding the
exposures of polymict breccia (Fig. 35). They comprise aggregates of central area, and identified remnants of glass, newly-formed Fe-oxy-
small nested cones, with individual cones ranging in size from 2 to hydroxides and secondary phases (calcite, phyllosilicates, quartz, zeo-
7 cm. They contain volcanic glass, with incipient crystallization and lites). Based on the results of textural and mineralogical studies, the
devitrification, and randomly distributed, millimeter-sized, euhedral, authors suggested a hydrothermal origin for them. The hydrothermal
opaque minerals (Crósta et al., 2011). A red layer decorates the striated processes would have been prompted by the disturbance of the local
surfaces, possibly constituting alteration or an altered melt film, similar aquifer following the impact.

35
A.P. Crósta et al. Chemie der Erde xxx (xxxx) xxx–xxx

Fig. 41. View into the abandoned quarry near the village of Vista Alegre that displays the best exposures of the polymict impact breccia of this impact structure. As a
result of Vista Alegre having been declared a geological site of interest by the government of the State of Paraná, a set of posters with information about the Vista
Alegre impact structure was installed in various places, including this quarry. Alvaro P. Crósta, seen in this picture by the poster, provided the contents.

Vargeão Dome, like the other two impact structures formed in the Bento Group (Pirambóia, Botucatu and Serra Geral formations) of the
Paraná-Etendeka LIP (i.e., Vista Alegre and, as we show here convin- Paraná Basin, similarly to Vargeão Dome. Vista Alegre was named after
cingly for the first time, Cerro do Jarau), has a maximum age de- the rural village of this name located within the structure, which is part
termined by the age of the regional Serra Geral volcanism, which oc- of Coronel Vivida County. Crósta et al. (2010a) presented the geology of
curred during the interval from 137 to 127 Ma (Turner et al., 1994). the Vista Alegre impact structure in detail.
More recently, Thiele and Vasconcelos (2010) obtained a 40Ar-39Ar age Vista Alegre is a complex impact structure, with a pronounced (up
of 134.6 ± 0.6 Ma for the Serra Geral flood basalts, which points to- to 150 m high) circular rim along nearly 70% of its perimeter and steep
wards rapid eruption of the Paraná-Etendeka LIP. In the case of the escarpments along the inner slopes of the rim (Fig. 38). The interior of
felsic volcanic rocks of the Ácidas Chapecó Unit, which is the upper- the structure is a topographically depressed area in comparison with the
most stratigraphic unit at Vargeão Dome and that was also affected by surrounding terrain. The morphology involves gently rolling hills, with
the impact, there is a precise U-Pb age of 134.3 ± 0.8 Ma on badde- topographic gradients of not more than 50 m. In contrast, in the area
leyite crystals by Janasi et al. (2011). surrounding the structure the relief is very uneven, with much higher
Specifically for Vargeão Dome, there has only been one dating study and steeper topographic gradients (typically over 200 m) between
reported by Nédélec et al. (2013) who provided a laser-ablation ICP-MS hilltops and river valleys.
age of 123.0 ± 1.4 Ma (Early Aptian) for zircon. However, this age The presence of a central uplift is not noticeable, but it is inferred
should be taken with some caution, as the sample from which the zir- based on the occurrence of blocks of deformed sandstone near the
cons were extracted, a brecciated rhyodacite with veinlets containing center of the structure, as well as from gravimetric data. These sand-
lithic fragments in what could have been former melt or a cataclasite, stones are possibly related to the Pirambóia-Botucatu succession that is
was collected on the outside of the southern rim of the structure. As the normally at depths of about 700–800 m below the present surface in
impact-induced deformation along the rim that is characterized by this portion of the Paraná Basin, where they are covered by the multiple
block-faulting and perhaps locally occurring cataclasis must be ex- basaltic flows of the Serra Geral Formation. Thus, the specific outline
pected to have been almost entirely brittle, it cannot be expected that and structure of the central uplift is not yet clear due to the lack of
melt resulting from the impact would have penetrated the acidic vol- subsurface information such as drilling or geophysical data.
canic flows at such a distance from the center. Thus, this age may not The escarpments along the border of the Vista Alegre structure ex-
represent the time of impact. hibit step-like features, formed by differential erosion along the con-
tacts between different lava flows. Four to five distinct flows are re-
cognized along the edge of the Vista Alegre structure. Most of the
5.1.1.3. Vista Alegre. The Vista Alegre impact structure, located at
central portion of the structure is covered with up to a few meter thick
25°57′S, 52°41′W in the western portion of Paraná State, was first
soil, which is intensely used for growing crops; rock outcrops are scarce.
proposed as a meteorite impact structure by Crósta et al. (2004), after
Ferreira et al. (2013) conducted a gravimetric survey at Vargeão
ground truthing a circular feature interpreted from a satellite image
and at Vista Alegre. The resulting Bouguer anomaly map of Vista Alegre
(Fig. 37). The structure is 9.5 km in diameter and is formed in tholeiitic
exhibits remarkable alternation of circular and concentric gravimetric
basalt and sandstones of the Late Permian to Early Cretaceous São

36
A.P. Crósta et al. Chemie der Erde xxx (xxxx) xxx–xxx

Fig. 42. a) View of a boulder of polymict impact breccia from the quarry at
Vista Alegre village (compare Fig. 4.15). b) Detail of a slab cut from the breccia,
showing mostly basalt (brownish, predominant) and sandstone (mottled, light Fig. 43. Vista Alegre a) Polymict lithic impact breccia with a shatter cone
grey, lower left) clasts of varied sizes and shapes in a medium-gray, fine-grained bearing basalt clast (white arrow). Scale bar is in cm. b) Shatter cone aggregate
matrix. Scale bars are in cm. of ca. 15 cm width formed in basalt – as commonly found at the quarry near
Vista Alegre town.
low and high anomalies (Fig. 39). A gravimetric low of approximately
−2 mGal occurs at the center of the structure, surrounded by an an-
nular gravity high reaching up to 1.4 mGal. This is, in turn, surrounded
by a less defined, but still circular, gravity low of −0.4 to −2 mGal.
This conspicuous bull’s eye anomaly pattern is completed by a semi-
circular gravity high, which coincides with the rim of the structure. This
gravimetric signature is very similar to the Bouguer gravity signature of
the Vargeão Dome structure (Fig. 30A), suggesting similar character-
istics of their morphology, geology, and structural deformation.
The schematic geological map of the Vista Alegre structure shown in
Fig. 40 was drawn based on occurrence of only a limited number of
outcrops that are sometimes located several hundred meters apart.
Thus, considerable extrapolation was required.
Along the escarpments on the inside of the rim of the structure, and
in an annular area that occupies just over one third of the distance
between the center of the structure and its edge, the only lithology to
occur is fractured basalt. Although fracturing is not uncommon in the
basalts of the Serra Geral Formation outside of the structure, and also
elsewhere in the Paraná Basin, the fracturing within the Vista Alegre
structure seems to be more intense in this annular zone than elsewhere.
However, no diagnostic shock deformation features have been found in
basalt samples from this zone so far. Fig. 44. Shocked quartz grain in a sandstone clast from the polymict breccia at
Polymict impact breccias occur in the central portion of the struc- the Vista Alegre quarry exhibiting one set of PDF and one of PF (after Crósta
ture. Outcrops of this breccia, although very scarce, occur along the two et al., 2010a).
main drainage channels in the structure, the Quieto and Surubim

37
A.P. Crósta et al. Chemie der Erde xxx (xxxx) xxx–xxx

Fig. 45. a) The circular Cerro do Jarau structure seen on a perspective view of the Shuttle Radar Topographic Mission shaded-relief DEM; the central elevated semi-
circular ridges are represented in red, with the circular drainage patterns enhanced in green and blue (corresponding to lower elevations). b) Perspective view of the
Cerro do Jarau structure from Terra/ASTER sensor data, combined with the ASTER/GDEM digital elevation data (after Crósta, 2012).

Fig. 46. View across the rough terrain of the core of Cerro do Jarau, showing the semi-annular ridges formed by silicified sandstones. The photo was taken from the
top of the northern-most ridge, looking towards the south. In the interior of the central ring a basin occurs that is drained by the Nhanduvai stream. The drainage
shows a centripetal pattern of radial drainage flowing towards the opening of the ring-form in the south and then into the Quaraí-Mirim creek.

38
A.P. Crósta et al. Chemie der Erde xxx (xxxx) xxx–xxx

Fig. 47. Schematic geological map of Cerro do Jarau (modif. from Crósta et al., 2010b).

streams, and also in a small quarry near the village of Vista Alegre. This impact breccia are known to occur, for example, at the Ries Crater in
quarry (Fig. 41) represents the best exposure of polymict breccia. It is southern Germany (see French, 1998 and references therein). In Vista
located at the side of the paved road that links the village to the town of Alegre, they appear to have been formed in significant amounts, as they
Coronel Vivida, located 15 km from Vista Alegre. Notably, there is a have been found quite often at this quarry over the years. A similar
geosite information board within this quarry providing introductory occurrence was also found in polymict impact breccia at Vargeão Dome
information about the formation of the Vista Alegre impact structure. (Fig. 34B). Raschke et al. (2013) also reported three shatter coned clasts
The Vista Alegre breccia is a parautochthonous type of impact in rhyodacite clasts that are part of the suevite interval from the ICDP
breccia (French, 1998) with a matrix composed of a very fine (pelitic), drill core from El’gygytgyn. These are the only reported cases of shatter
gray material with a powdery texture, which is mostly derived from cones formed in volcanic rocks to date. Geochemical compositions of
basalt of the Serra Geral Formation. The clasts vary in size from a few three shatter cone samples obtained by Crósta et al. (2010a, 2010b) are
millimeters to several decimeters. Their shape is generally angular and nearly identical to the Paraná low-Ti, high-Fe tholeiite – typical within-
their compositions include volcanic (basalt) and sedimentary (sand- plate basalt – from the southern part of the Paraná LIP (Peate et al.,
stone, siltstone) litho-types, with predominance of the former (Fig. 42). 1992; Peate, 1997).
The fact that only parautochthonous breccia has been found in the in- Pittarello et al. (2015a) reported the occurrence of melt on the
terior of Vista Alegre suggests that the structure is relatively well pre- striated surfaces of some of these shatter cones, as well as fine-grained
served from erosion, in contrast to Vargeão Dome and Cerro do Jarau cataclasites filling fractures in them. The melt phase represents a con-
that must be quite deeply eroded. tinuous, quenched melt film, comprising a so far unidentified micro-
Notably, some of the basalt clasts found in the breccia contain crystalline phase, mica, and amorphous material. Ultracataclasites,
shatter cone aggregates (Fig. 43). The shocked basalt with shatter cones containing subrounded pyroxene clasts in an ultrafine-grained matrix
was fragmented and incorporated into the breccia, and subsequently occur subparallel to the striated surface. The characterization of the
mixed with other fragmented rocks. Currently, there is a strong debate crystalline phase in the melt was not conclusive, but suggested a rare or
in the impact cratering community whether polymict impact breccias in new type of clinopyroxene (Pittarello et al., 2015a, p. 1239).
and around impact structures are the result of ejecta deposition, or Shocked quartz was reported by Crósta et al. (2010a) to occur at
whether this material in large part never left the crater where it may Vista Alegre. Samples of the polymict breccia collected at the quarry
have been formed due to collapse of a high-density mass flow (ground contain also sandstone clasts. Some of these contain quartz grains with
surge or suspension flow; e.g., Stöffler et al., 2013; Osinski and planar deformation features (PDF); most of these grains have one or two
Pierazzo, 2013; Osinski et al., 2016; Siegert et al., 2017). Another sets of PDFs (Fig. 44).
possibility is that several processes might have combined to form and The age of the Vista Alegre impact is still an open question. Crósta
emplace these breccia packages. et al. (2012) reported the only isotopic dating available. They con-
Although rather rare, some examples of clasts with shatter cones in ducted four 40Ar-39Ar step-heating experiments on different aliquots of

39
A.P. Crósta et al. Chemie der Erde xxx (xxxx) xxx–xxx

Fig. 48. a) Residual Bouguer gravity anomaly map of the Cerro do Jarau structure. b) 3D density models calculated for profiles A, B and C. c) Geological map
(Fig. 47) over the residual Bouguer gravity map. d) Geological interpretation of the density model shown in section ‘A’ (Fig. 48B), a N-S section near the center of the
structure (Giacomini et al., 2017).

a melt fragment recovered from polymict impact breccia and obtained by terrain with lower elevation and with equally circular drainage
an apparent weighted mean age of 115 ± 4 Ma for the formation of patterns. The outer part comprises rather flat ‘pampa’ plains typical of
this melt (and, by implication, the impact event). However, these au- southern Brazil, as well as neighboring regions of Uruguay and
thors recommended to take this result with caution, as they noted Argentina. Therefore, Cerro do Jarau stands out conspicuously in
pervasive alteration of this material. The age of this alteration was these flattish lands (Fig. 45A). The entire structure is ∼13 km wide.
calculated at 13 ± 6 Ma, which may have biased the results. Conse- It has been known to geologists since the late 1960s, when it was first
quently, they offer a more conservative age estimate for the impact identified through photo-interpretation. Its origin, however, has been
event between ∼111 Ma (the apparent minimum age obtained for the the subject of debate ever since.
melt) and 134 Ma, the current age constraint for the Serra Geral vol- First mention of this structure was made by Grehs (1969), who
canism according to Thiele and Vasconcelos (2010). described the occurrence of quartzites that form the ridges and which,
in turn, are surrounded by basalt. He interpreted the structure as a
5.1.1.4. Cerro do Jarau. The Cerro do Jarau structure is centered at structural dome, with the lower strata of the Botucatu Formation
30°12′S, 56°32′W. It is located in Quaraí county, in the southwestern having been uplifted and exposed at surface. Lisboa et al. (1987) de-
portion of the state of Rio Grande do Sul, near the border between scribed local occurrences of brecciation in the interior of Cerro do
Brazil and Uruguay. The structure comprises a series of semi-circular Jarau. They suggested as possible causes for the origin of this structure
ridges in its central parts, which rise up to 200 m higher than the tectonic deformation or – for the first time – meteorite impact.
surrounding terrains. The central elevated ridges are composed of Hachiro et al. (1995) interpreted the elevated sandstone ridges as
folded and silicified sandstone that is likely related to the Botucatu the rim of an eroded impact structure, assigning a maximum diameter
Formation. In contrast, the surrounding areas correspond to basaltic of 5 km to it. They mentioned the occurrence of microscopic shock
lava flows of the Serra Geral Formation. This inner part is surrounded deformation features, without being more specific or providing

40
A.P. Crósta et al. Chemie der Erde xxx (xxxx) xxx–xxx

Fig. 49. Types of breccia of possible impact origin at Cerro do Jarau: a)


sandstone breccia; b) basalt breccia. In a) hammer head segment for scale is
10 cm. In b) segment of pen is 9 cm.
Fig. 50. Cerro do Jarau a) Multiply-striated joint surfaces (MSJS) in sandstone
documented evidence such as photos of these alleged features. of the Guará Formation; b) a similar striated deformation feature in basalt of the
The morphology and geology of Cerro do Jarau were investigated in Serra Geral Formation.
detail by Crósta et al. (2010b). Different from previous works, they
interpreted the outer limit of the structure, which is not evident and with cross-stratification at the meter scale. These sandstones were re-
does not stand out in the topography, to coincide with the anomalous lated by these authors to the Upper Jurassic Botucatu Formation. The
semi-circular drainage patterns of the Guarupá stream to the north and basalts of the Serra Geral Formation then surround the mentioned
northwest, and the Quaraí-Mirim Creek in the south and southwest. sandstone units, forming a bull’s eye pattern with the oldest strata at the
These drainage channels join to form the Quaraí River, which delimits center and the younger ones in the outermost zone.
the structure in its western portion (Fig. 45a, b). Within this extent, at a Giacomini et al. (2017) presented the results of a ground gravi-
diameter of 13.5 km, the two rock types that occur within the structure metric survey of Cerro do Jarau. This was conducted with the objective
– basalt and sandstone – exhibit some degree of deformation, usually in to establish the gravimetric signature and to better define the internal
the form of brecciation, whereas outside of this limit the basalt does not structures and the outer limit of the structure. They found a positive
show deformation. The core of the structure, composed of silicified Bouguer gravity anomaly (> 6 mGal) forming a semi-circular pattern
sandstones (Fig. 46), was described by these authors as having an el- surrounding the central part of the structure, as well as another positive
liptical shape, with a diameter of 6.5 km in north-south direction, and anomaly (> 7 mGal) in the NE portion of the structure, extending to-
5.2 km in east-west direction. It forms an inner basin, drained by the wards its center (Fig. 48A, C ). The inner core of the structure, which
Nhanduvai stream and its tributaries (also compare Fig. 47). corresponds to the occurrence of sandstone, exhibits a more complex
The geology of Cerro do Jarau is shown in the schematic geological gravity signature in the form of a negative anomaly surrounding a small
map of Fig. 47. The structure comprises two different types of sandstone positive anomaly.
in its central part, which is surrounded by fractured and brecciated These authors developed 3D models based on their gravimetric data,
basalt. In the innermost part, with outcrops along the Nhanduvai Creek in order to explore the distribution of the different types of rocks pre-
and its tributary drainage channels at lower topographic levels, there sent on the surface and in the subsurface (sandstone and basalt). They
are sandstones showing characteristics of fluvial deposition, such as found their models (Fig. 48B) to be in accordance with the impact
parallel lamination and pelitic layers. Crósta et al. (2010b) attributed hypothesis for the formation of Cerro do Jarau, with the gravity low in
them to the Middle Jurassic Guará Formation of the São Bento Group. the central area denoting the uplifted and more deformed sandstones
In the environs of these fluvial sandstones, and still within the limits of and the Serra Geral basalt producing the gravity high surrounding the
the inner basin of Cerro do Jarau, are exposures of eolian sandstones central portion of the structure (Fig. 48D). The two types of sandstones

41
A.P. Crósta et al. Chemie der Erde xxx (xxxx) xxx–xxx

shown in Fig. 48D would represent a more deformed sandstone


(“Sandstone I”, exhibiting lower density, 2.2–2.3 g/cm3), and a less
deformed sandstone (“Sandstone II”) with relatively higher density
(2.4–2.6 g/cm3).
The debate regarding the origin of Cerro do Jarau has been ongoing
since the first reference to its possible impact origin by Lisboa et al.
(1987). However, their suggestion was based solely on the circular
nature of the structure, combined with an unlikely possibility of it
having been created by endogenous processes. The reference to the
hypothetical shock features by Hachiro et al. (1995) remained un-
substantiated.
The first evidence to strongly support an impact origin for Cerro do
Jarau was reported by Crósta et al. (2010b). This included a report of
brecciation of sandstone and basalt (Fig. 49), as well as deformation
features in these two rock types (Fig. 50). They interpreted these stri-
ated fracture surfaces as multiply-striated joint surfaces (MSJS – after
Nicolaysen and Reimold, 1999); these fractures are apparently related
to the shock-diagnostic shatter cone phenomenon. MSJS were also de-
scribed by Milton et al. (1996) from the Australian Gosses Bluff impact
structure and by Nicolaysen and Reimold (1999) from Sudbury (Ca-
nada).
Crósta et al. (2010b) also reported multiple observations of micro-
deformation features in quartz within such breccias, including quite
frequently the occurrence of planar fractures. By themselves, planar
fractures in quartz are not recognized as a diagnostic shock indicator,
but if they occur abundantly, they may provide a hint. In any case, if
these PF in Cerro do Jarau samples had been produced by shock, they
would likely represent low-shock pressure deformation (< 5 GPa).
Subsequent microscopic analysis of Cerro do Jarau samples by Philipp
et al. (2010) confirmed the occurrence of the deformation phenomena
found by Crósta et al. (2010b), and also alleged the occurrence of
planar deformation features (PDF) in quartz from sandstone. However,
scrutiny of their results indicates that these hypothetical PDFs are likely
planar fractures as well.
Sanchez et al. (2014), based mostly on structural measurements of
the strata exposed in the Cerro do Jarau structure, suggested that the
basalt flows are structurally and stratigraphically overlain by the
sandstones, and not the other way around as suggested by all previous
studies of the structure. According to their interpretation, the two types
of sandstones (medium-coarse stratified sandstones and laminated
sandstones) would represent a stratigraphic unit younger than the Serra
Geral lava flows and, therefore, were originally deposited on top of the
basalt. They also observed that the dip of the sandstone strata was to-
wards the center of the structure, suggesting a concave configuration,
and not a convex one. Based on these assumptions, they interpreted the
whole Cerro do Jarau structure as an impact structure, but shaped as a
basin with a central depression, instead of a domal structure with an
uplifted central portion.
In our view, the interpretation presented by Sanchez et al. (2014)
does not offer adequate answers to the overall geologic and geophysical
data available for Cerro do Jarau, which point to the existence of a
central uplift (as shown in Fig. 48). Besides, a basin configuration
would suggest a simple crater rather than a complex one, and this
would be incompatible with the likely diameter of 10–13 km, according
to the current knowledge about impact craters on Earth (Melosh, 1989).
Another weak aspect of the interpretation offered by Sanchez et al.
(2014) is the complete lack of support – both, in the literature and in
the regional stratigraphic record – for such a sandstone unit with the
characteristics described for the sedimentary strata that occur within
Cerro do Jarau that would overlie the volcanic strata of the Serra Geral
Fig. 51. Examples of microscopic deformation features in quartz from Cerro do Formation.
Jarau sandstone. a) Grain with two sets of PDF (0001 and 10-14 orientations) in Recently, bona fide shock deformation features were identified in a
quartz grain in sandstone. b) Two sets of PDF at the margins of two adjoining number of samples from Cerro do Jarau in the course of a thorough
quartz grains in sandstone (red arrows). c) Another example of PDF in quartz optical microscopic study by one of us, WUR, of circa 250 thin sections
grain in sandstone (red arrow). d) Basalt breccia with thin veinlets of entirely covering all lithologies sampled throughout the extent of the structure.
altered material.
This comprises the extensive thin section collection of J. Sanchez

42
A.P. Crósta et al. Chemie der Erde xxx (xxxx) xxx–xxx

Fig. 52. Location of the confirmed impact structures of the Parnaíba Basin, in northern Brazil (MA = Maranhão State; PI = Piauí State; TO = Tocantins State).

(Sanchez, 2013), the specimens in A.P. Crósta’s collection, and a host of filled with matrix of dark-red color but indefinable character (Fig. 51D).
new samples obtained in the course of a structural geological pilot Some veins contain a polymict clast population, with quartz and
project by P.T. Zaag (formerly of the Museum für Naturkunde Berlin) in quartzite clasts in addition to the dominant basalt clast population.
2016. However, all these occurrences of polymict clast populations are devoid
Amongst these numerous thin sections, about a dozen sandstone of any intra-clast microdeformation. Whether this brecciation can be
samples show microdeformation including PF, FF, and in four cases, related to impact deformation, or should be completely, or partially,
PDF. PF occur in up to three sets of different orientations in selected considered the result of volcanic processes, as favored by Sanchez
quartz grains. FF occur in up to 3 orientations per host grain as well, (2013), remains uncertain at this time. Further field observations to
and PDF were noted in 2 grains in 2 orientations each, whereas a dozen better constrain the geological context for these breccias are required.
other grains altogether showed only single sets of PDF. Invariably, these The unambiguous finding of the shock deformation features re-
shock-deformed grains are rare but PF and FF occur abundantly in se- ported here, coupled with previous data especially from Crósta et al.
lected samples. PDF are generally confined to grain margins, where (2010b) and Giacomini et al. (2017), does no longer leave any doubt
they occur in sets of very short features – likely the reason why they had about the impact origin of Cerro do Jarau. The age of the impact is still
not been observed earlier. Notably, Kowitz et al. (2013a, 2013b, 2016) undetermined. It is only defined by an upper limit provided by the Serra
did not find optically identifiable PDF in their experimentally shocked Geral volcanism, at ca. 134 Ma (Thiele and Vasconcelos, 2010). The
sandstone specimens. Fig. 51A, B, C gives a few examples of these shock dimensions of the structure are not yet entirely established, with a
deformation features in Cerro do Jarau samples. diameter estimated currently at ca. 13.5 km. This number is based on
Some of J. Sanchez’ thin sections of basalt specimens do indeed geomorphological features, such as the circular drainage patterns of the
show brecciation. Invariably all these breccia zones and veinlets are Garupa and Quarai streams marking the supposed outer rim of the

43
A.P. Crósta et al. Chemie der Erde xxx (xxxx) xxx–xxx

Brasiliano orogenic cycle (late Proterozoic and early Paleozoic) and is


bordered by amalgamated fold belts (Almeida et al., 1981).
The sedimentary filling of the Parnaíba Basin is divided into three
main sedimentary sequences that correspond, from bottom to top, to
the Serra Grande, Canindé, and Balsas groups, the ages of which range
from the Silurian to the Triassic (Góes and Feijó, 1994). These mega-
cycles record multiple transgressive-regressive events that indicate a
progressive tendency to contraction of the basin area, strong con-
tinentalization, and climatic desertification (Góes and Feijó, 1994;
Santos and de Carvalho, 2009). Additional Mesozoic volcanosedimen-
tary successions overlap the mentioned megacycles and represent the
final stage of the evolution of the Parnaíba Basin (Goés and Feijó,
1994).
The Parnaíba Basin has been known to contain several circular
structures. Some of these are related to igneous intrusions, including
kimberlites (Tompkins, 1994), whereas others have been shown to be
Fig. 53. ASTER image combined with ASTER/GDEM digital elevation model the result of meteorite impact (this work). There still are a number of
giving a perspective view of the Serra da Cangalha impact structure. Image other structures in the basin of so far unknown origin.
width is ca. 15 km. The impact record of the Parnaíba Basin comprises three confirmed
meteorite impact structures: Serra da Cangalha, Riachão, and Santa
structure, the geophysical (gravity) signature, as shown by Giacomini Marta (Fig. 52). A fourth structure, São Miguel do Tapuio, is currently
et al. (2017), and, finally, the important fact that no apparent de- under investigation by some of the authors of this paper (A.P. Crósta,
formation was found to affect the basalt outside the perimeter estab- M.A.R. Vasconcelos, W.U. Reimold), with promising partial results re-
lished by the two streams. Therefore, the rim of the structure seems to garding its impact origin, as presented in more detail in Part II of this
have been almost completely obliterated by erosion. What has re- review.
mained is mostly the deeper part of the central uplift, represented by
the silicified sandstone (meta-sandstone) that forms the highest eleva-
5.1.2.1. Serra da Cangalha. The Serra da Cangalha (SdC) – “Horse
tions in the northern and northeastern parts of the central uplift of the
Collar Range” in Portuguese – structure is a complex impact structure
structure.
centered at 8°0′ S, 46°52′ W, in the northeastern-most part of Tocantins
Future investigations of Cerro do Jarau shall focus on a better
State (Fig. 52). In the early 1970s, as part of a regional exploration
characterization of shock deformation, as well as establishing its di-
program for industrial diamonds, 3 boreholes were drilled at the center
mensions and age.
of a circular structure in order to investigate its possible igneous
intrusive origin, then suspected to be a possible kimberlite (DNPM,
5.1.2. Parnaíba Basin 1972).
The intracratonic Parnaíba Basin covers an area of ∼600,000 km2 First reference to a possible impact origin for SdC was made by Dietz
in central and northern Brazil. The basin is filled with up to 3500 m and French (1973), based on the circular morphology of the structure
thick sedimentary sequences deposited between the Silurian and Early documented by some early images of the Landsat satellite. This was
Triassic (Góes and Feijó, 1994; Góes, 1995; Pedreira da Silva et al., followed by reconnaissance fieldwork by McHone (1979) and Santos
2003; Santos and de Carvalho, 2009). The basin was formed in a crustal and McHone (1979). McHone (1986) reported on the anomalous local
domain that underwent a major tectono-magmatic event during the deformation of the strata and on low-shock pressure features, such as

Fig. 54. Aerial view of the inner ring feature, the so-called ‘collar’, of the central uplift of the Serra da Cangalha impact structure. It comprises massive, silicified and
deformed sandstones of the Poti Fm. that rise up to 350 m above the surrounding terrain. View towards the south (Photo courtesy of A. Bartorelli).

44
A.P. Crósta et al. Chemie der Erde xxx (xxxx) xxx–xxx

Fig. 55. Serra da Cangalha a) Gamma-ray data: K, Th, U in red, green and blue, respectively, showing a well-defined anomaly of the three radioelements, especially
for potassium, over the annular basin around the central uplift. b) Bouguer signature for Serra da Cangalha, expressed in mGal, with a weak positive anomaly in the
center. c) Forward gravity model along the N-S direction across the structure – emphasizing the central uplift of the structure (adapted from Vasconcelos et al., 2012a,
2012b).

intense grain fracturing and curvi-linear features in quartz. However, sedimentary rocks of the Parnaíba Basin. The resulting impact structure
only recently bona fide evidence for the impact origin of SdC was is eroded to a considerable degree but still exhibits some remarkable
presented by Kenkmann et al. (2011) and Vasconcelos et al. (2013). geomorphological features – including several prominent ring features
The bolide that impacted at SdC targeted Paleozoic/Mesozoic (Fig. 53). Almeida-Filho et al. (2005) and Reimold et al. (2006)

45
A.P. Crósta et al. Chemie der Erde xxx (xxxx) xxx–xxx

Fig. 56. Geological map of the Serra da Cangalha impact structure (diameter is ca.12 km) in Tocantins State, Brazil (adapted from Kenkmann et al., 2011; see also
Vasconcelos et al., 2013).

Fig. 57. a) Shatter cones in sandstone from the


central depression of Serra da Cangalha. b)
Feather features in quartz of a shatter cone
flound in the inner basin of ther central uplift.
c) Planar fractures developed in a quartz grain
in polymict lithic breccia. d) Planar deforma-
tion features in a quartz grain of a shatter cone.
Note the distinct planarity and sharpness of
many of these features. However, there are
some features that are comparatively broader
and “grainy”. These are likely deformation
bands with countless fluid inclusions that in-
tersect the thin section plane at a rather low,
oblique angle.

conducted remote sensing analysis of SdC, distinguinshing, from the approximately 6 km diameter, followed by an incomplete second in-
center to the rim, the following morpho-structural features: (1) a pro- termediate ring structure with 11 km diameter; and (3), the outer ring
minent inner ring with 3.2 km diameter, with a central depression of with 13.4 km diameter, representing the rim of the structure.
2.2 km diameter; (2) a first intermediate ring feature with A prominent collar built up by massive, silicified sandstone

46
A.P. Crósta et al. Chemie der Erde xxx (xxxx) xxx–xxx

Fig. 58. a) Numerical modeling results showing Serra da Cangalha shortly after impact. The model shows that temperatures (left-hand side) of up to1000 oC were
attained around the center, and pressures (right-hand side) of around 15 GPa. Pressure distribution: red (25 GPa), orange (15 GPa), yellow (10 GPa), green (5 GPa),
magenta (1 GPa), blue (< 1 GPa). Dashed line represents the current erosion level. b) Megafolds at the collar of SdC, representing examples of inverted strata shown
in the numerical model.

surrounds the central depression. This collar has a nearly-quadrangular


shape and is up to 350 m high, forming the most conspicuous mor-
phological feature of SdC (Fig. 54). The shape of the folded sandstone
strata that form the collar of this structure has inspired its name: Horse
Collar Range. This intensely silicified/recrystallized sandstone collar
has resisted erosion comparatively better than the surrounding target
rocks, thus forming a formidable erosion remnant that stands out in the
regional landscape.
Based on the analysis of magnetotelluric data, Adepelumi et al.
(2005) estimated the deformation at SdC to be restricted to 2 km depth.
An airborne gamma-ray survey showed a well-defined anomaly of the
Fig. 59. A 3D perspective view of Riachão. Landsat 7/ETM+ combined with three radioelements, particularly for potassium, over the annular basin
SRTM digital elevation data. Image width is ca. 6 km. around the central uplift (Fig. 55), more conspicuous over the rim and
central uplift (Vasconcelos et al., 2012a). Estimation of the average

47
A.P. Crósta et al. Chemie der Erde xxx (xxxx) xxx–xxx

Fig. 60. a) Perspective view of gammaspectrometric maps of the Riachão impact structure showing the approximate locations of the outcrops and samples shown in
the following pictures of this figure; the maps show (bottom to top): thorium concentrations, ternary gamma-ray (K in red, Th in green and U in blue), and silica
abundance. ER1 (upper left): Outcrop of ‘Pedra da Lapa’, a lithic sandstone that forms the largest outcrop of the elevated rim of the structure. ER1 (upper right): detail
of the lithic sandstone showing intraclasts. CU1 (middle left): Sandstone with kaolinite cement. ER3 (middle right): Microscopic view of the sandstone with in-
traclasts. CU3 (lower left): Quartz grain with planar deformation features (PDF). CU3 (lower right): Detail of the previous photo showing PDF along 0001 crys-
talographic plane. (ER = Elevated Rim; CU = Central Uplift; Out = Outside. Kln = kaolinite; Ms = muscovite; Qtz = quartz). (after Maziviero et al., 2013a, 2013b).

depth to basement was also attempted by Vasconcelos et al. (2012a), the northeastern and western parts of the annular basin, the strata of
based on evaluation of regional aeromagnetic data, resulting in a depth the Pedra de Fogo Formation exhibit centimeter- to meter-scale open
of 2.4 km. High resolution magnetic data presented by Vasconcelos folds. The rim exhibits flat-topped plateaus with high topographic ele-
et al. (2012b) showed magnetic sources lying at 200–500 m depth vations that mark the outer limit of SdC. The rim exhibits mostly rocks
below the center of the structure, and deeper magnetic sources below of the Pedra de Fogo Formation with some occurrences of the under-
the entire structure, at ∼1900 m depth. The Bouguer anomaly map lying sandstones of the Piauí Formation in the NW and E sectors. The
shows a slight positive anomaly (1 mGal) over the central area, and a Pedra de Fogo Formation consists of red/gray laminated mudstone with
low gravity signature of 2 mGal at the northwestern rim (Fig. 55B). The abundant chert and subordinate coarse- to medium-grained and well
gravity models show disturbed basement in the innermost part of the sorted, pink sandstone.The Piauí Formation shows frequent intercala-
crater, down to about 2–3 km depth (Fig. 55C). tions of purple/gray mudstone and fine- to medium-grained, red
Vasconcelos et al. (2013) divided the structure into several domains sandstone, usually with parallel bedding or with incipient sigmoidal
based on the intensity of deformation that decreases from the center cross-bedding. Beyond the rim the attitude of the strata is nearly hor-
toward the rim: central uplift, annular basin, crater rim, and external izontal and rocks do not display significant deformation.
terrain. The stratigraphic units that occur at the SdC structure comprise, In the northwestern sector of the structure, at a road cut that in-
from the bottom to the top, the Devonian Longá Formation (shales, tersects the rim, a prominent concentrically trending monocline in
siltstones), the Carboniferous Poti and Piauí formations (predominantly sandstones of the Piauí Formation can be seen, whose inner limb dips
sandstones), and the Permian Pedra de Fogo Formation (mudstone, gently toward the crater center.
carbonate and chert) (Fig. 56). Shock features were found in specific locations within SdC.
The central uplift was defined by Kenkmann et al. (2011) and Kenkmann et al. (2011) reported on the occurrence of shatter cones
Vasconcelos et al. (2012a) as the innermost zone of 5.8 km diameter. It of < 1 decimeter size in coarse sandstone found at two locations in the
comprises an outer peripheral zone with steeply dipping strata of the central depression, likely in Longá Formation mudstone (Fig. 57A).
Piauí Formation, the prominent collar with silicified sandstone strata of They display a characteristic cone geometry with convergent grooves
the Poti Formation, and a central depression underlain by strata of the and ridges pointing to an apex and cone angles betweeen 50° and 60°.
Longá Formation. The lithologies that form the Longá Formation, Rare monomict and polymict breccias were also found in the central
mostly shales and siltstones, are more susceptible to weathering and depression and in the periphery of the central uplift.
erosion than the other siliciclastic rocks, thus favoring erosion and the Vasconcelos et al. (2013) also presented shock microdeformation
formation of the central depression, where there is a relative lack of evidence from a few samples. They described the existence of feather
outcrops. Information about the strata of the Longá Formation comes features (FF) in quartz grains of sandstones from the Longá and Poti
mostly from drill cores obtained in the course of diamond exploration in formations, including quartz grains from shatter cones (Fig. 57B).
the 1970s. The three boreholes were drilled vertically through strongly Planar fractures (PF) were found in quartz grains from the same stra-
compacted and deformed, dark gray and black shales, and poorly sorted tigraphic units, in polymict breccias (Fig. 57C) and in shatter cones.
sandstones (Vasconcelos et al., 2013). These authors also found planar deformation features (PDF) in coarse
The annular basin comprises the area between the central uplift and sandstone from the drill cores into Longá Formation, and in shatter
the rim. Here, strata of the Pedra de Fogo Formation occur. The strata cones (Fig. 57D), most of them with sets oriented parallel to the c
comprise mainly brecciated (cataclased) chert layers, as well as silici- (0001) axis.
fied sandstone, with subordinate mudstone. In some areas, such as in Vasconcelos et al. (2012b) employed numerical modeling to

48
A.P. Crósta et al. Chemie der Erde xxx (xxxx) xxx–xxx

Fig. 61. a) Simplified geological map of the Riachão impact structure. The red line represents the interpreted profile shown in B. b) Northwest-southeast profile
through the southeastern sector of the Riachão structure (y axis in meters; x axis in kilometers). c) Schematic stratigraphic column (modified after Maziviero et al.,
2013a). See text for discussion.

estimate the dimensions of the bolide that formed SdC at ∼1.4 km McHone and Dietz (1978) and McHone (1979), based on photographs
diameter, and the total energy released by the impact at approximately taken by the astronauts of the 1975 Apollo-Soyuz Test Project. A first
2.74 × 1020 J. According to their results, the uplift of the strata below geological map of Riachão was produced by McHone (1986). He de-
the center of the structure was ∼500 m. The peak pressure in the rocks scribed impact breccia with chert clasts occurring on the rim and quartz
currently exposed at the central uplift would have been ∼15 GPa grains with sets of planar fractures; however, bona fide evidence for an
(Fig. 58), which is consistent with peak pressures necessary for the impact origin was not reported at the time. Only recently, Maziviero
formation of shock features as found in the target rocks of that area, et al. (2012, 2013a) provided conclusive evidence for Riachão’s origin
such as relatively rare PDF. Inverted strata shown in the numerical by impact.
models are similar to the overturned beds observed in the rocks of the A flat to gently rolling relief, reflecting a dendritic, radial fluvial
collar. drainage pattern, characterizes Riachão. In general, soil, vegetation,
and alluvial sediments extensively cover the area, and there are only a
5.1.2.2. Riachão. The Riachão impact structure is located in Maranhão few scattered outcrops of sandstone and sedimentary breccia of the
State of northeastern Brazil, centered at 7°42′S, 46°38′W. This semi- Pedra de Fogo Formation in and around the structure.
circular, complex structure is nearly 4 km in diameter. It was formed on Maziviero et al. (2013a) conducted a detailed remote sensing in-
flat-lying Upper Carboniferous to Permian sedimentary rocks of the vestigation using Landsat 7/ETM+ combined with SRTM digital ele-
Parnaíba Basin. vation data for assessing the geometry and morphology of the structure
The occurrence of a circular structure was first recognized by Ojeda (Fig. 59). Using these data, the outer limit and the morphostructural
and Bembom (1966), who described the anomalous radial drainage zones were interpreted. The raised rim zone indicates an apparent
pattern in this region and interpreted the structure as being of crypto- diameter of 4.1 km for the structure; it resembles a horseshoe shape that
volcanic origin. A possible impact origin was then suggested by is open to the northwest. The rim rises 50 m above the terrain outside

49
A.P. Crósta et al. Chemie der Erde xxx (xxxx) xxx–xxx

Fig. 62. a) The topography of the Santa Marta impact structure, in terms of color-coded elevation values (meters above sea level) and main morphological features.
b) Perspective view (3.5 x vertical exaggeration) showing the morpho-structural domains (outer edge, annular basin with inner rings, and central elevation). The
central elevation is composed of the central uplift (CU) and the overlying central elevated plateau (CEP); D = diameter of the structure. Images were prepared from a
GoogleEarth® grayscale mosaic draped over elevation data from SRTM (after Oliveira et al., 2017).

50
A.P. Crósta et al. Chemie der Erde xxx (xxxx) xxx–xxx

Fig. 63. a) Geological map of the Santa Marta impact structure. The A-B line in purple represents the geological profile shown in B). b) Geological profile (AB section)
across one-half of the Santa Marta impact structure. The “Central Elevation” includes both the deformed Paleozoic rocks of the central uplift (CU) and the non-
deformed detrital cover of the central elevated plateau (CEP). Note that Datum 0 relates to the actual altitude of 500 m above sea level, which corresponds
approximately to the average altitude of the outcrops in the annular basin.

51
A.P. Crósta et al. Chemie der Erde xxx (xxxx) xxx–xxx

Fig. 64. Bouguer anomaly map of the Santa Marta structure. The outer dashed line corresponds to the outer boundary of the structure, and the inner dashed circle
corresponds to the central elevated plateau (CEP). A conspicuous NE-SW oriented gravity anomaly trend suggests a contrast between higher density rocks in the NW
sector of the crater and lower density rocks in the SE sector. It may also be linked to a deeper crustal trend (Oliveira et al., 2014).

the structure and the maximum topographic elevation of the central anomaly pattern is also consistent with a 4.1 km diameter for Riachão.
uplift is 30 m above the inner ring basin. Magnetic data show an elongated signature along a NW-SE axis across
A study involving ASTER thermal infrared (TIR) data, combined the central uplift (Vasconcelos et al., 2010, 2012c).
with gamma-ray spectrometric data (Maziviero et al., 2013b), allowed The sedimentary strata in the Riachão area are part of the deposi-
to produce a quartz index map (areas with high silica content) and the tional sequence of the Balsas Group of the Parnaíba Basin (Góes, 1995).
identification of areas characterized by high values of thorium and The Riachão structure was subdivided by Maziviero et al. (2013a) into
uranium (Fig. 60A). 3 morpho-structural zones, from the periphery to the center: (1) an
Aerial gamma-spectrometry, gravity, and magnetic data were ana- elevated rim; (2) an annular depression; and (3) a central uplift. Fig. 61
lyzed by Vasconcelos et al. (2010) and Maziviero et al. (2013a, 2013b). shows a geological map, together with the local stratigraphic column.
A well-defined gamma-ray signature outlines the elevated rim and the The rock types on the elevated rim include predominantly lithic
central uplift. The rim exhibits low concentrations of radioactive ele- sandstone with sandstone intraclasts. This apparently massive litho-
ments, whereas the core shows high K values, probably related to the facies is medium- to coarse-grained, poorly to moderately sorted, and
phyllosilicate cement of the Piauí sandstones (Maziviero et al., 2013b; matrix-supported with 20–30% subangular intraclasts of silty sand-
Fig. 3 – CU1). Riachão shows a positive large-wavelength gravity stone. Locally, medium-scale tabular and trough cross-bedding were
anomaly of about 1 mGal that is roughly coincident with the entire observed. This rock is tilted with dip angles > 40° and dip directions
structure, as well as negative anomalies outside the structure. This usually toward the opening of the rim in the northwest. In some places

52
A.P. Crósta et al. Chemie der Erde xxx (xxxx) xxx–xxx

Fig. 65. View toward the SE across the central parts of the Santa Marta structure. The prominent escarpment of ca 50 m elevation, at the location known as Sinésio
Hill (Fig. 62A), indicates where the subhorizontal cover overlaps the strata of the central uplift. The base of this hill consists of monomict breccia formed by pelitic
rock fragments and is overlain by the young polymict sedimentary breccia forming the CEP.

on the rim there is a clastic breccia with sandstone clasts of 0.2–10 cm could be potentially caused by the Riachão impact, which would put
sizes. McHone (1986) interpreted this breccia to be of impact nature, this event at the Permo-Triassic transition (ca. 251 Ma). This is ob-
but the complete lack of shock indicators suggests that it is a sedi- viously a speculative suggestion.
mentary breccia of the Pedra de Fogo Formation. Currently, it is not clear whether the Riachão impact is related to the
In the annular depression occur several outcrops of laminated silt- formation of the Serra da Cangalha impact structure, located some
stone and calcilutite, with dip angles of up to 10°–20° toward 40 km south-southwest of Riachão in the neighboring state of Tocantins
ENE–WSW and NW–SE. At the transition to the central uplift, brec- – a suggestion made originally by McHone (1986).
ciated chert is observed. This morpho-structural zone exposes the Pedra
de Fogo Formation, as already proposed by McHone (1986).
The central uplift comprises fine- to medium-grained, apparently 5.1.2.3. Santa Marta. The Santa Marta impact structure (Fig. 62) is
massive sandstone, with 55°–80° dip angles and dips in various direc- located in the southwestern portion of Piauí State, northeastern Brazil.
tions. According to Maziviero et al. (2013a), this sandstone represents It is centered at 10°10′S, 45°14′W. It is a complex impact structure of
the lowermost stratigraphic unit in the Riachão structure, namely the 10 km diameter that is reasonably well preserved. The structure
Piauí Formation. This is a reasonable assumption given the strati- comprises a partially exposed raised rim, as well as a central uplift
graphic uplift of ca. 0.4 km estimated from the diameter of the structure with a diameter of 3.2 km (Oliveira et al., 2014, 2017).
using the formula proposed by Grieve and Pilkington (1996), namely Master and Heymann (2000) first mentioned Santa Marta as a po-
SU = 0.086D1.03, where SU is the amount of stratigraphic uplift un- tential impact structure. Based on the geomorphological circular
dergone by the deepest strata now exposed at the surface and D is the anomaly observed by remote sensing, they suggested the impact nature
rim diameter. The Riachão structure is deeply eroded as suggested by of the structure that they named “Gilbués” after a nearby town. This
its subdued topographic expression. name was later revised to “Santa Marta structure” by Vasconcelos et al.
No shatter cones have been found at Riachão. Microscopic de- (2010) according to the name of the village located within the structure
formation features observed in thin sections of sandstone from the and also to the name of the local hills formed by the sedimentary strata
central uplift reveal only features formed in the low shock deformation disrupted by the deformation processes associated with the structure.
regime. Only one quartz grain from a sample of Piauí sandstone exhibits These authors identified distinctive geophysical signatures, similar to
PDF of basal orientation (Fig. 60 − CU3). Other lower pressure shock those observed at other impact structures formed in sedimentary target
features have also been observed, including feather features (FF) and rocks. However, confirmation of this origin was only recently presented
planar fractures (PF) (Maziviero et al., 2013a). by Oliveira et al. (2014), based on abundant occurrence of shock fea-
Direct radiometric age determination for the formation of the tures in rocks from Santa Marta, including the widespread occurrence
Riachão structure is not possible, due to the absence of suitable (melt- of shatter cones and the presence of diagnostic shock deformation
bearing) material. Therefore, the age of the Permian Pedra de Fogo features in quartz.
Formation, attributed to the Cisuralian Epoch (299–271 Ma) by The Santa Marta structure was formed in sedimentary strata of the
Iannuzzi et al. (2016), provides an unconstrained upper limit for the Parnaíba Basin and of the Sanfranciscan Basin, in a region where se-
age of this impact event. quences of these two different sedimentary basins overlap (Lima and
Abrantes (2016), studying the Permo-Triassic sequences of the Leite, 1978). The structure is composed of rocks belonging to the Or-
Parnaíba Basin in a region that includes the Riachão structure, identi- dovician-Silurian Serra Grande and Devonian Canindé groups (Pi-
fied deformational structures in the contact zone between the strata of menteiras and Cabeças formations), both related to the Parnaíba Basin,
the Motuca and Sambaíba formations. These structures comprise in- and the Cretaceous Areado (Abaeté, Quiricó and Três Barras forma-
traformational breccias, folded strata, convoluted lamination and tions) and Urucuia (Posse Formation) groups related to the San-
faults/microfaults with clay injections. He also noticed that, near Ria- franciscan Basin (Fig. 63). An undeformed detrital cover un-
chão, there was more plastic deformation affecting the top of the Mo- conformably overlies these sequences mainly to the west of the
tuca Formation, and interpreted these structures as seismic-induced. structure where it forms a prominent plateau, but extends also into its
More interestingly, he also identified metal anomalies (Cr, Co, Cu, Mn, interior. This cover can be regionally correlated with the Chapadão
Au, Pd and Pt) in the deformed strata at the contact zone between these Formation of Cenozoic age, comprising the unconsolidated alluvial,
two formations, together with metallic micro-particles within the clay colluvial and elluvial sediments that cover the Late Cretaceous Urucuia
matrix of the deformed sandstones. He then suggested that the combi- Group (Campos and Dardenne, 1997a, 1997b). The informal name
nation of deformation features, metal anomalies and metallic particles given to this plateau is “Central Elevated Plateau” or “CEP”.
The results of a ground gravimetric survey of Santa Marta are

53
A.P. Crósta et al. Chemie der Erde xxx (xxxx) xxx–xxx

Fig. 66. Locations of impact breccia and shatter cone occurrences in the Santa Marta structure (Oliveira et al., 2014).

presented in Fig. 64, showing a NE-SW oriented negative gravity morpho-structural domains within the structure (Figs. 62 and 65): outer
anomaly partially coinciding with the CEP, and a positive anomaly of edge/outer rim, annular basin, and central elevation. The latter was
∼2mGal in the central-eastern region. The transition between these divided into two distinct sub-domains: central elevated plateau (CEP)
two anomalies has a NE-SW orientation, with an inflection around the and central uplift (CU). The CEP detrital cover partially overlies the
center of the structure. The eastern rim of Santa Marta is well marked deformed (e.g., shatter-cone bearing) Paleozoic units in the CU. These
by a positive anomaly, whereas the western rim, which is partially sub-domains do not exactly coincide spatially.
concealed by the detrital cover correlated to the Chapadão Formation, The central uplift region is the area where Paleozoic strata of the
does not show up well in the gravity data. Parnaíba Basin are exposed. Considering a diameter of 10 km for the
A variety of lithotypes in the interior of the Santa Marta structure is impact structure, Oliveira et al. (2014) estimated an uplift of around
responsible for the different responses to impact-generated stress and 1000 m of the strata in the central part of the structure. This number is
results in different styles of deformation from the center to the crater in accordance with the estimated depth to the crystalline basement
rim. The deformation is conditioned not only by the respective shock- made by Vasconcelos et al. (2010) on the basis of modeling of regional
pressure zones, but also by these lithological characteristics. The oc- aerogeophysical data (magnetic and gravity), thus suggesting that al-
currence of inner rings in the Santa Marta impact structure was inter- most the entire stratigraphic sequence of the Parnaíba Basin was af-
preted by Oliveira et al. (2017) as a direct consequence of the variation fected by the impact.
in lithologies. These authors recognized the existence of four distinct The CU is made up of strongly deformed Paleozoic strata. The

54
A.P. Crósta et al. Chemie der Erde xxx (xxxx) xxx–xxx

widely distributed, and monomict breccias of less frequent occurrence.


The locations of impact breccia occurrences are shown in Fig. 66.
Among the different types of breccias interpreted as impact breccias,
only the monomict sandstone breccia and one of the numerous samples
of polymict breccias that occur at the top of Sinésio Hill contain shock
features. It is thought that these shocked particles are detritus that was
incorporated into the fluvial cover deposits. Other breccias are inter-
preted as products of impact based on stratigraphic and structural re-
lations observed in the field.
A sedimentary (not impact-related) breccia also occurs on the CEP.
This breccia forms a cover that extends laterally beyond the crater rim
(Fig. 62). The best exposures are in the Sinésio Hill area (Figs. 62 and
65), in the eastern portion of the CEP, where they exhibit a thickness of
approximately 50 m (Figs. 62 and 65). This cover bears no evidence of
deformation. According to Oliveira et al. (2017), it may be correlated
with the Chapadão Formation.
In the Santa Marta structure, shatter cones (Fig. 67A) occur in
sandstones and clasts of monomict sandstone breccia possibly related to
the Cretaceous Areado Group at several locations in the central part of
the structure (Fig. 65). Shatter cones are also developed within mostly
well-rounded quartzite pebbles and blocks that occur as large clasts in
conglomerates possibly related to the Cretaceous Abaeté Formation, in
the annular basin near the outer edge of the structure (Fig. 67B) (Uchôa
et al., 2013; Oliveira et al., 2014). A total of 19 sites with shatter cone
occurrences were reported by Oliveira et al. (2014).
Shock microdeformation includes planar deformation features
(PDFs) in quartz grains found in a number of impact breccia samples
from the central part of the structure, as well as feather features (FFs)
that commonly occur associated with PDFs in intensely deformed
sandstones (Fig. 68). The same microdeformation features were ob-
served in quartz grains of shatter cones. Oliveira et al. (2014, 2017)
pointed out that these occurrences increase in frequency towards the
center of the structure. The rocks currently exposed in the central uplift
region were likely subjected to peak shock pressures (at the grain scale)
of around 20–25 GPa, as indicated by the crystallographic orientations
of PDFs in quartz grains (Oliveira et al., 2014).
The maximum age for the Santa Marta impact event is estimated
Fig. 67. Multiple shatter cones on sandstone: a) pink sandstone likely from the
between 93 and 100 Ma, based on the age of the sandstones of the Posse
Cabeças Fm., from the center of the Santa Marta structure; b) shatter cones in Formation, Urucuia Group, that occur at the rim of the structure and
the interior of a well-rounded quartzite block from conglomerate of the Abaeté represent the youngest unit affected by the impact event (Oliveira et al.,
Fm. (sample G25, location E-81C – see Fig. 66). Scale bars are in cm. 2017). The minimum age is constrained by the undeformed sedimen-
tary cover, correlated to the Cenozoic Chapadão Formation. This cover
has likely been a key factor that contributed to the preservation of the
detrital cover of the central elevated plateau (CEP) hinders a more
structure from more intense erosion. This makes Santa Marta one of the
comprehensive understanding of the stratigraphy in the central portion
better preserved impact structures in Brazil and, despite its consider-
of the structure. From the edge of the CEP outwards, there are Devonian
able age, is a still very well preserved impact structure in the global
sandstones and pelites of the Cabeças and Pimenteiras formations, re-
context.
spectively (Fig. 63). The area of occurrence of these two units extends
up to the concentric and discontinuous inner rings formed by Silurian
6. Concluding remarks
conglomeratic sandstones and conglomerate of the Serra Grande Group
(Oliveira et al., 2017).
In this first part of the “South American impact record“ we have
The Mesozoic units are exposed in the annular basin and along the
emphasized the universal importance of impact cratering, dealt with
inner side of the outer rim of Santa Marta. In the annular basin, towards
the scientific background related to impact cratering studies and how to
the center of the structure, tectonically induced lateral juxtaposition of
identify/confirm an impact structure or impact deposit, and finally
Paleozoic and Cretaceous strata can be observed, with both sequences
introduced the Brazilian impact record.
being intensely deformed: the juxtaposition occurs along shear zones,
The latter represents the bulk of the South American impact struc-
along which the sedimentary strata of the two basins have been brought
ture record – which is natural, as impact cratering studies have mostly
into contact, as evidenced by overturned top-base indicators, drag folds,
been performed in South America at Brazilian universities. However, it
and steeply-dipping fault planes. The distinction between Cretaceous
remains to be seen in Part II of this treatise that there are highly in-
rudaceous and the Silurian-Devonian pelitic strata is based on contact
teresting sites and discoveries in several other South American coun-
relations with pelitic units and on structural observations, i.e., evidence
tries. We do not want, with this first part of our review, discourage
for normal or overturned bedding. Deformed sandstones of the Posse
Brazilian researchers to continue to investigate the already known
Formation and the Três Barras Formation occur preferentially along the
impact structures. To the contrary, we hope that this review of Brazilian
outer edge of the structure (Oliveira et al., 2017).
structures has highlighted where there are deficiencies in the scientific
There are two types of impact breccias in the Santa Marta structure:
knowledge about them, and which types of geological terrains still hold
polymict breccias, which are comparatively more abundant and more
good potential to possibly unveil further impact structures: for example,

55
A.P. Crósta et al. Chemie der Erde xxx (xxxx) xxx–xxx

Fig. 68. Santa Marta impact structure. a)


Shocked quartz with two sets of PDF, one set of
PF (NW-SE), and shock extension fractures
(sample SM-4 – for sample locations see
Fig. 66). In addition, there is a set of seemingly
broad and decorated features thought to re-
present a low-angle intersection of decorated
PDF with the thin section surface. b) A shocked
quartz grain in a quartzite clast with a single
set of PDF, two sets of PF and, additionally, a
set of broad, brownish inclusion-decorated
curved features (sample SM-1). c) Sub-planar
features and at least two sets of dense PDF in a
quartzite clast (sample SM-4). d) A brecciated
deformation band in sandstone with angular
quartz fragments resulting from intense cata-
clasis (sample SM-3A). e) Single set of PDF in
adjacent quartz grains of a quartzite clast
(sample SM-1). f) Low-angle intersections of
two parallel planar fractures that are part of
feather feature arrays. In the uppermost central
part of the grain, a set of decorated PDF is also
cut at a low-angle to the section plane (sample
E-23A) (Oliveira et al., 2014).

there could be potential, buried impact structures in the many Acknowledgments


Phanerozoic sedimentary basins, and older, deeply eroded structures
may occur in Proterozoic terrains. Foremost, there is a dearth of age Authors’ acknowledgements are given in Part II of the review, to be
constraints on these impact events. We have discussed that it has not published in a subsequent issue of Chemie der Erde/Geochemistry.
been possible for not less than 6 of the known impact structures in
Brazil to identify datable, impact-generated melt rock. Maybe paleo- References
magnetic and/or thermochronological studies may be able to further
constrain the formation ages of these structures. Further detailed Abrantes Jr., F.R., 2016. O Permo-Triássico da Bacia do Parnaíba, norte do Brasil:
ground studies are also required and could, for example, yield valuable implicações paleoambientais, paleoclimáticas e paleogeográficas para o Pangea
Ocidental. Ph.D. thesis. Universidade Federal do Pará 130 p.
structural geological information. The investigation of polymict impact Acevedo, R.D., Rocca, M.C.L., Ponce, J.F., Stinco, S.G., 2015. Impact Craters in South
breccias in Brazil’s largest impact structure, Araguainha Dome, has – America. Springer, Heidelberg 104 p.
with the exception of some earlier work on the melt rocks – only re- Adepelumi, A.A., Fontes, S.L., Schnegg, P.A., Flexor, J.M., 2005. An integrated magne-
totelluric and aeromagnetic investigation of the Serra da Cangalha impact crater,
cently been begun in earnest and will contribute much to the widely Brazil. Phys. Earth Planet. Inter. 150, 159–181.
conducted debate about the origin and emplacement of clastic-com- Almeida, F.F.M., Hasui, Y., Brito Neves, B.B., Fuck, R.A., 1981. Brazilian structural pro-
ponent dominated impact breccias. vinces: an introduction. Earth-Sci. Rev. 17, 1–29.
Almeida-Filho, R., Moreira, F.R.S., Beisl, C.H., 2005. The Serra da Cangalha astrobleme as
Part II of our review will examine the impact record of other
revealed by ASTER and RTM orbital data. Int. J. Remote Sens. 26, 833–838.
countries in South America, provide information about a number of Alvarez, L.W., Alvarez, W., Asaro, F., Michel, H.V., 1980. Extraterrestrial cause for the
proposed impact structures, and review those that already have been Cretaceous-Tertiary extinction. Science 208, 1095–1108.
Artemieva, N.A., Wünnemann, K., Krien, F., Reimold, W.U., Stöffler, D., 2013. Ries crater
discarded as not being formed by impact. There are valuable lessons to
and suevite revisited − observations and modeling. part II: modelling. Meteorit.
be learnt from this – namely, that it is no dilettante task to detect and Planet. Sci. 48, 590–627.
confirm an impact structure. The easy pickings have, most seemingly, Baratoux, D., Reimold, W.U., 2016. The current state of knowledge about shatter cones:
been made – but the record is still far from complete and detailed introduction to the special issue. Meteorit. Planet. Sci. 51, 1389–1434.
Barbour Jr., E., Corrêa, W.A.G., 1981. Geologia da Estrutura de Vargeão, SC. Consórcio
geoscientific detective work employing state-of-the-art techniques CESP/IPT, São Paulo 33 p. (Unpublish. report, Paulipetro).
should reap reward. Borovička, J., Spurný, P., 2008. The Carancas meteorite impact –encounter with a
monolithic meteoroid. Astron. Astrophys. 485, L1–L4.
Borovička, J., Spurný, P., Brown, P., Wiegert, P., Kalenda, P., Clark, D., Shrbený, L., 2013.

56
A.P. Crósta et al. Chemie der Erde xxx (xxxx) xxx–xxx

The trajectory, structure and origin of the Chelyabinsk asteroidal impactor. Nature Ebert, M., Kowitz, A., Schmitt, R.T., Reimold, W.U., Mansfeld, U., Langenhorst, F., 2017.
503, 235–237. Localized shock-induced melting of sandstone at low shock pressures(< 17 .5 GPa):
Buchanan, P.C., Reimold, W.U., 2002. Planar deformation features and impact glass in an experimental study. Meteorit. Planet. Sci. https://doi.org/10.1111/maps.12948.
inclusions from the Vredefort Granophyre, South Africa. Meteorit. Planet. Sci. 37, Elbeshausen, D., Wünnemann, K., Collins, G.S., 2013. The transition from circular to
807–822. elliptical impact craters. J. Geophys. Res. 118, 2295–2309.
Burchell, M.J., 2015. Shocked rocks: impacts from the laboratory to the Solar System. In: Erickson, T.M., Timms, N.E., Kirkland, E.L., Tohver, E., Cavosie, A.J., Pearce, M.A.,
Lee, M.R., Leroux, H. (Eds.), Planetary Mineralogy. EMU Notes in Mineralogy 15. Reddy, S.M., 2017. Shocked monazite chronometry: integrating microstructural and
European Mineralogical Union and the Mineralogical Society of Great Britain and in situ isotopic age data for determining precise impact ages. Contrib. Mineral. Petrol.
Ireland, London, pp. 227–251. 172, 11. https://doi.org/10.1007/s00410-017-1328-2.
Camacho, A., Vernon, R.H., Fitzgerald, J.D., 1995. Large volumes of anhydrous pseudo- Erwin, D.H., 1990. The end-Permian mass extinction. Annu. Rev. Ecol. Evol. Syst. 21,
tachylite in the Woodroffe Thrust, eastern Mus grave Ranges, Australia. J. Struct. 69–91.
Geol. 17, 371–383. Ferreira, J.C., Leite, E.P., Vasconcelos, M.A.R., 2013. Assinatura gravimétrica e magnética
Campos, J.E.G., Dardenne, M.A., 1997a. Estratigrafia e sedimentação da Bacia das estruturas de impacto de Vargeão e Vista Alegre, Brasil. In: 13th International
Sanfranciscana: Uma Revisão. Rev. Bras. Geoc. 27, 257–282. Congress of the Brazilian Geophysical Society XIII. Rio de Janeiro, Brazil. pp.
Campos, J.E.G., Dardenne, M.A., 1997b. Origem e evolução tectônica da Bacia 401–406. https://doi.org/10.1190/sbgf2013-085.
Sanfranciscana. Rev. Bras. Geoc. 27, 283–294. Ferreira, J.C., Leite, E.P., Vasconcelos, M.A.R., Crósta, A.P., 2015. 3D gravity modeling of
Cavosie, A.J., Lugo-Centeno, C., 2014. Shocked apatite from the Santa Fe impact struc- impact structures in basaltic formations in Brazil: part 1–Vargeão Dome, Santa
ture. Lunar Planet. Sci. XXXXV. Lunar and Planetary Institute, Houston, Texas Catarina. Braz. J. Geoph. 33, 319–332. https://doi.org/10.22564/rbgf.v33i2.723.
Abstract #1691. Ferrière, L., Koeberl, C., Reimold, W.U., Hecht, L., Bartosova, K., 2009. The origin of
Cavosie, A.J., Montalvo, P.E., Timms, N.E., Reddy, S.M., 2016. Nanoscale deformation toasted quartz in impactites revisited. Lunar Planet. Sci. XL. Lunar and Planetary
twinning in xenotime, a new shocked mineral, from the Santa Fe impact structure Institute, Houston, Texas (Abstract #1751).
(New Mexico, USA). Geology 44, 803–806. https://doi.org/10.1130/G38179.1. Ferrière, L., Barrat, J.-A., Giuli, G., Koeberl, C., Schulz, T., Topa, D., Wegner, W., 2017. A
Collins, G.S., Melosh, H.J., Osinski, G.R., 2012. The impact-cratering process. Elements 8, new tektite strewn field discovered in Uruguay. 80th Annual Meeting of the
25–30. Meteoritical Society Abstract #6195.
Cordani, H.G., Ramos, V.A., Fraga, L.M., Cegarra, M., Delgado, I., Souza, K.G., Gomes, Fischer, G., Masero, W., 1994. Rotational properties of the magnetotelluric impedance
F.E.M., Schbbenhaus, C., 2016. Tectonic map of South America – Explanatory Notes, tensor −the example of the Araguainha impact crater, Brazil. Geophys. J. Int. 119,
2nd edition. scale 1:5,000,000. Commission for the Geological Map of the World, 548–560.
Paris 13 pp. Fischer, S., 2015. Ein Beitrag über die Impaktgesteine der Araguainha Impaktstruktur
Crósta, A.P., Reimold, W.U., 2016. Book review: Impact Craters in South America, by (Brasilien). M. Sc thesis. Freie Universität, Berlin.
Acevedo et al., 2015. Meteorit. Planet. Sci. 51, 996–999. Fredriksson, K., Dube, A., Milton, D.J., Balasundaram, M.S., 1973. Lonar lake, India: an
Crósta, A.P., Vasconcelos, M.A.R. 2013. Update on the current knowledge of the Brazilian impact crater in basalt. Science 180, 862–864.
impact craters. In: Lunar Planet. Sci. XXXXIV. Abstract #1318. Lunar and Planetary French, B.M., Koeberl, C., 2010. The convincing identification of terrestrial meteorite
Institute, Houston, Texas. impact structures What works, what doesn't, and why. Earth-Sci. Rev. 98, 123–170.
Crósta A.P., Gaspar J.C., Candia M.A.F., 1981. Feições de metamorfismo de impacto no Shock Metamorphism of Natural Materials. In: French, B.M., Short, N.M. (Eds.), Mono
Domo de Araguainha. Rev. Brasil. Geoc. 11, 139–146. Book, Baltimore 644 pp.
Crósta, A.P., Kazzuo-Vieira, C., Schrank, A., 2004. Vista Alegre: a newly discovered im- French, B.M., Underwood Jr., J.R., Fisk, E.P., 1974. Shock metamorphic features in two
pact crater in Southern Brazil. Meteorit. Planet. Sci. 39, A–28. meteorite impact structures, southeastern Libya. Geol. Soc. Am. Bull. 85, 1425–1428.
Crósta, A.P., Kazzuo-Vieira, C., Choudhuri, A., Schrank, A., 2005. Vargeão Dome, State of French, B.M., Cordua, W.S., Plescia, J.B., 2004. The Rock Elm meteorite impact structure,
Santa Catarina: A meteoritic impact record on volcanic rocks of the Paraná 64 Basin. Wisconsin: geology and shock-metamorphic effects in quartz. Geol. Soc. Am. Bull.
In: Winge, M., Schobbenhaus, C., Berbert-Born, M., Queiroz, E.T., Campos, D.A. 116, 200–218.
(Eds.), Sítios Geológicos e Paleontológicos do Brasil. v. 2. DNPM/CPRM/SIGEP French, B.M., 1998. Traces of Catastrophe: A Handbook of Shock-Metamorphic Effects in
Brasília, pp. 23–34. Terrestrial Meteorite Impact Structures. LPI Contribution 954, Lunar and Planetary
Crósta, A.P., Koeberl, C., Furuie, R.A., Kazzuo-Vieira, C., 2010a. Vista Alegre, southern Institute, Houston 120 pp.
Brazil: a new impact structure in the Paranó flood basalts. Meteorit. Planet. Sci. 45, French, B.M., 2004. The importance of being cratered: the new role of meteorite impact as
181–194. a normal geological process. Meteorit. Planet. Sci. 39, 169–197.
Crósta, A.P., Lourenço, F.S., Priebe, G.H., 2010b. Cerro Jarau, Rio Grande do Sul: apos- Fudali, R.F., Milton, D.J., Fredriksson, K., Dube, A., 1980. Morphology of Lonar crater,
sible new impact structure in Southern Brazil. In: Gibson, R.L., Reimold, W.U. (Eds.), India: comparison and implications. Moon Planets 23, 493–515.
Large Meteorite Impacts and Planetary Evolution IV. The Geological Society of Góes, A.M., Feijó, F.J., 1994. Bacia do Parnaíba. Bol. Geoc. Petrobras 8, 57–67.
America Special Paper SPE465, GSA, Boulder. Góes, A.M., 1995. A Formaçao Poti (Carbonífero Inferior) da Bacia do Parnaíba. Ph.D.
Crósta, A.P., Kazzuo-Vieira, C., Pitarello, L., Koeberl, C., Kenkmann, T., 2011. Geology thesis. University of São Paulo.
and impact features of Vargeóo Dome, southern Brazil. Meteorit. Planet. Sci. 47, Güldemeister, N., Wünnemann, K., Durr, N., Hiermaier, S., 2013. Propagation of impact-
51–71. induced shock waves in porous sandstone using mesoscale modeling. Meteorit.
Crósta, A.P., Jourdan, F., Koeberl, C., 2012. 40Ar/39Ar dating of the Vista Alegre crater, Planet. Sci. 48, 115–133.
Brazil. 34th International Geological Congress, Brisbane, Australia. Giacomini, B.B., Leite, E.P., Crósta, A.P., 2017. 3D gravimetric investigation of the Cerro
Crósta A.P., 1982. Mapeamento geológico do Domo de Araguainha utilizando técnicas de do Jarau structure, Rio Grande do Sul, Brazil. Meteorit. Planet. Sci. 52, 565–583.
sensoriamento remoto. M.Sc. thesis. Instituto Nacional de Pesquisas Espaciais (INPE). https://doi.org/10.1111/maps.12813.
Crósta, A.P., 1987. Impact structures in Brazil. In: Pohl, J. (Ed.), Research in Terrestrial Gibson, R.L., Reimold, W.U., 2008. Geology of the Vredefort Impact Structure: A Guide to
Impact Structures. Springer Vieweg Wiesbaden, pp. 30–38. Sites of Interest. Memoir 97, Council for Geoscience, Pretoria, 181 pp.
Crósta, A.P., 2002. Domo de Araguainha (GO-MT): O maior astroblema da América do Glass, B.P., Simonson, B.M., 2012. Distal impact ejecta layers: spherules and more.
Sul. In: Schobbenhaus, C., Campos, D.A., Queiroz, E.T., Winge, M., Berbert-Born, M. Elements 8, 15–60.
(Eds.), Sítios Geológicos e Paleontológicos do Brasil. v. 1. DNPM/CPRM/SIGEP Glass, B.P., Simonson, B.M., 2013. Distal Ejecta Layers: A Record of Large Impacts in
Brasília, pp. 531–540. Sedimentary Deposits. Springer, Heidelberg 71 6pp.
Crósta, A.P., 2004. Impact craters in Brazil: how far we’ve gotten. Meteorit. Planet. Sci. Goderis, S., Paquay, F., Claeys, P., 2013. Projectile identification in terrestrial impact
39, A-27. structures and ejecta material. In: Osinski, G., Pierazzo, E. (Eds.), Impact Cratering:
Crósta, A.P., 2012. Estruturas de Impacto e Astroblemas Brasileiros. In: Hasui, Y., Processes and Products, pp. 223–239.
Carneiro, C.D.R., Almeida, F.F.M., Bartorelli, A. (Eds.), Geologia do Brasil. Beca-Ball, Gratz, A.J., Fisler, D.K., Bohor, B.F., 1996. Distinguishing shocked from tectonically de-
São Paulo, São Paulo, pp. 673–708. formed quartz by the use of the SEM and chemical etching. Earth Planet. Sci. Lett.
DNPM, 1972. Relatório de pesquisa de diamante industrial na região da Serra da 142, 513–521.
Cangalha. Estado de Goiás, n. 805.015/70-805.019/70. Grehs, S.A., 1969, Aspectos geológicos e geomorfológicos do Cerro do Jarau, Rio Grande
De Marchi, L., Hauser, N., Reimold, W.U., Crósta, A.P., 2016. Geological and petro- do Sul, Brasil. In: Congresso Brasileiro de Geologia XXIII, v 1, Salvador, Brazil,
graphical characterization of the polymict impact breccia of the Araguainha Dome, Sociedade Brasileira de Geologia, p. 365–375.
Brazil. Lunar Planet. Sci. XXXXVII. Lunar and Planetary Institute, Houston, Texas. Grieve, R.A.F., Pilkington, M., 1996. The signature of terrestrial impacts. AGSO J. Aust.
Deutsch, A., Schärer, U., 1994. Dating terrestrial impact events. Meteoritics 29, 301–322. Geol. Geophys. 16, 399–420.
Deutsch, A., Buhl, D., Langenhorst, F., 1992. On the significance of crater ages: new ages Grieve, R.A.F., Reimold, W.U., Morgan, J., Riller, U., Pilkington, M., 2008. Observations
for Dellen (Sweden) and Araguainha (Brazil). Tectonophysics 216, 205–218. and interpretations at Vredefort, Sudbury and Chicxulub: towards a composite ki-
Deutsch, A., Poelchau, M.H., Kenkmann, T., 2015. Impact metamorphism in terrestrial nematic model of terrestrial impact basin formation. Meteorit. Planet. Sci. 43,
and experimental cratering events. In: Lee, M.R. Leroux, H. (eds.), Planetary 855–882.
Mineralogy. EMU Notes in Mineralogy 15, European Mineralogical Union and the Grieve, R.A.F., 1987. Terrestrial impact structures. Annu. Rev. Earth Planet. Sci. 15,
Mineralogical Society of Great Britain and Ireland, London. pp.89-127. 245–270.
Dietz, R.S., French, B.M., 1973. Two probable astroblemes in Brazil. Nature 244, 562. Grieve, R.A.F., 1991. Terrestrial impact: the record in the rocks. Meteoritics 26, 175–194.
Donatti, L.M., Sawakuchi, A.O., Giannini, P.C.F., Fernandes, L.A., 2001. The Pirambóia- Hachiro, J., Coutinho, J.M.V., Coimbra, A.M., 1995. The Cerro do Jarau astrobleme (Rio
Botucatu succession (Late Permian - Early Cretaceous, Paraná Basin, São Paulo and Grande do Sul), Brazil: a Cretaceous cryptoexplosive structure. Academia Brasileira
Paraná states): Two contrasting eolian systems. Na. Acad. Bras. Ciênc. 73, 465 (ab- de Ciências, Resumo das Comunicações 67, 4.
stract). DOI 10.1590/S0001-37652001000300020. Hamers, M.F., Drury, M.R., 2011. Scanning electron microscope-cathodoluminescence
Earth Impact Database (http://www.unb.ca/passc/ImpactDatabase), last visited February (SEM-CL) imaging of planar deformation features and tectonic deformation lamellae
2018. in quartz. Meteorit. Planet. Sci. 46, 1814–1831.

57
A.P. Crósta et al. Chemie der Erde xxx (xxxx) xxx–xxx

Hamers, M.F., 2013. Identifying Shock Microstructures in Quartz from Terrestrial Koeberl, C., Pittarello, L., Reimold, W.U., Raschke, U., Brigham-Grette, J., Melles, M.,
Impacts: New Scanning Electron Microscopy Methods. Ph.D. Thesis. Utrecht Minyuk, P., 2013. El'gygytgyn impact crater, Chukotka, Arctic Russia: impact cra-
University, The Netherlands. tering aspects of the 2009 ICDP drilling project. Meteorit. Planet. Sci. 48, 1108–1129.
Hammerschmidt, K., von Engelhardt, W., 1995. Ar/Ar dating of the Araguainha impact https://doi.org/10.1111/maps.12146.
structure, Mato Grosso, Brazil. Meteoritics 30, 227–233. Koeberl, C., 2001. Craters on the Moon from Galileo to Wegener: a short history of the
Hargitai, H., Reimold, W.U., Bray, V.J., 2015. Impact structure. In: In: Hargitai, H., impact hypothesis, and implications for the study of terrestrial impact craters. Earth
Kereszturi, Á. (Eds.), Encyclopedia of Planetary Landforms, vol. 2. Springer, New Moon Planets 85–86, 209–224.
York, pp. 988–1023. https://doi.org/10.1007/978-1-4614-3134-3. Koeberl, C., 2007. The geochemistry and cosmochemistry of impacts. In: Davis, A. (Ed.),
Hauser, N., Reimold, W.U., Voll, K., Chaves, J.G.S., Alves de Mattos, B., Dantas, E.L., Treatise of Geochemistry. Elsevier, New York pp. 1.28. 1–1.28.52.
2017. Are the impact melt rocks from the Araguainha impact structure, Brazil, Koeberl, C., 2014. The geochemistry and cosmochemistry of impacts. In: Holland, H.D.,
homogeneous? Evidence from geochemistry and Sr-nd isotopes. In: 80th Annual Turekian, K.K. (Eds.), Treatise on Geochemistry. Elsevier, Oxford, pp. 73–118.
Meeting of the Meteoritical Society. Santa Fe, USA. (Abstract #6215). Kohout, T., Gritsevich, M., Grokhovsky, V.I., Yakovlev, G.A., Haloda, J., Halodova, P.,
Henkel, H., Reimold, W.U., 1998. Integrated geophysical modelling of a giant complex Michallik, R.M., Penttilä, A., Muinonen, K., 2013. Mineralogy, reflectance spectra,
impact structure: anatomy of the Vredefort structure, South Africa. Tectonophysics and physical properties of the Chelyabinsk LL5 chondrite, insight into shock induced
287, 1–20. changes in asteroid regoliths. Icarus 228, 78–85.
Henkel, H., Reimold, W.U., 2002. Magnetic model of the central uplift of the Vredefort Kowitz, A., Schmitt, R.T., Reimold, W.U., Hornemann, U., 2013a. The MEMIN shock
impact structure, South Africa. J. Appl. Geophys. 51, 43–62. recovery experiments: first results on shock effects in dry, porous Seeberger sand-
Hergarten, S., Kenkmann, T., 2015. The number of impact craters on Earth: any room for stone from the low shock pressure range (5–12.5 GPa). Meteorit. Planet. Sci. 48,
further discoveries? Earth Planet. Sci. Lett. 425, 187–192. 99–114.
Hildebrand, A.R., Penfield, G.T., Kring, D.A., Pilkington, M., Zanoguera, A.C., Jacobsen, Kowitz, A., Güldemeister, N., Reimold, W.U., Schmitt, R.T., Wünnemann, K., 2013b.
S.B., Boynton, W.V., 1991. Chicxulub Crater; a possible Cretaceous/Tertiary Diaplectic quartz glass and SiO2 melt experimentally generated at only 5 GPa shock
boundary impact crater on the Yucatan Peninsula, Mexico. Geology 19, 867–871. pressure in porous sandstone: laboratory observations and meso-scale numerical
Hippert, J., Lana, C., 1998. Aerial crystallization of hematite in impact bombs from the modeling. Earth Planet. Sci. Lett. 384, 17–26.
Araguainha astrobleme, Mato Grosso, central Brazil. Meteorit. Planet. Sci. 33, Kowitz, A., Güldemeister, N., Schmitt, R.-T., Reimold, W.U., Wünnemann, K., Holzwarth,
1303–1309. A., 2016. Revision and recalibration of existing shock classifications for quartzose
Hippertt, J.P., Lana, C., Weinberg, R.F., Tohver, E., Schmieder, M., Scholz, R., Gonçalves, rocks using low-shock pressure (2.5–20 GPa) recovery experiments and mesoscale
L., Hippertt, J.F., 2014. Liquefaction of sedimentary rocks during impact crater de- numerical modeling. Meteorit. Planet. Sci. 51, 1741–1761. https://doi.org/10.1111/
velopment. Earth Planet. Sci. Lett. 408, 285–295. maps.12712.
Hockey, T., 1994. The Shoemaker-Levy 9 spots on Jupiter: their place in history. Earth Kruppa, C., 2016. Petrographic Characterization of a Large Impact Melt Bomb from
Moon Planets 66, 1–9. Otting, Nördlinger Ries Impact Structure, Germany. M.Sc. Thesis. Freie Universität
Huffman, A.R., Reimold, W.U., 1996. Experimental constraints on shock-induced mi- Berlin.
crostructures in naturally deformed silicates. Tectonophysics 256, 165–217. Lacerda Filho, J.V., Abreu Filho, W., Valente, C.R., Oliveira, C.C., Albuquerque, M.C.,
Iannuzzi, R., Cisneros, J.C., Conceição, D.M., 2016. A provável idade Permiana inicial da 2004. Geologia e recursos minerais do estado de Mato Grosso. Geologia do Brasil
Formação Pedra de Fogo, Bacia do Parnaíba, e suas consequências (abstract.). CPRM/MME/SICME, 252p.
Congresso Brasileiro de Geologia XXXVIII. http://cbg2017anais.siteoficial.ws/ste01/ Lana, C., Romano, R., Reimold, W.U., Hippertt, J.P., 2006. Collapse of large complex
ID7734_112520_52_Resumo_Iannuzzi_et_al_48_CBG.pdf. impact structures: implications from the Araguainha impact structure. Geology 34,
Ivanov, B.A., 2005. Numerical modeling of the largest terrestrial meteorite craters. Solar 9–912.
Syst. Res. 39, 381–409. Lana, C., Souza-Filho, C.R., Marangoni, Y.R., Yokoyama, E., Trindade, R.I.F., Tohver, E.,
Janasi, V.A., Freitas, V.A., Heaman, L.H., 2011. The onset of flood basalt volcanism, Reimold, W.U., 2007. Insights into the morphology, geometry, and post-impact
northern Paraná Basin, Brazil: a precise U-Pb baddeleyite/zircon age for a Chapecó- erosion of the Araguainha peak-ring structure central Brazil. Geol. Soc. Am. Bull. 119,
type dacite. Earth Planet. Sci. Lett. 302, 147–153. 1135–1150.
Jourdan, F., Renne, P.R., Reimold, W.U., 2009. An appraisal of the ages of terrestrial Lana, C., Souza-Filho, C.R., Marangoni, Y.R., Yokoyama, E., Trindade, R.I.F., Tohver, E.,
impact structures. Earth Planet. Sci. Lett. 286, 1–13. Reimold, W.U., 2008. Structural evolution of the 40 km wide Araguainha impact
Jourdan, F., Moynier, F., Koeberl, C., Eroglu, S., 2011. 40Ar/39Ar age of the Lonar crater structure, central Brazil. Meteorit. Planet. Sci. 43, 701–716.
and consequence for the geochronology of planetary impacts. Geology 39, 671–674. Langenhorst, F., Deutsch, A., 2012. Shock metamorphism of minerals. Elements 8, 31–36.
Kazzuo-Vieira, C., Crósta, A.P., Gamboa, F., Tygel, M., 2009. Caracterização geofísica da https://doi.org/10.2113/gselements.8.1.31.
estrutura de impacto do Domo de Vargeão, Brasil. Rev. Brasil. Geof. 27, 375–388. Lima, M.A.E., Leite, F.J., 1978. Estudo global dos recursos minerais da Bacia Sedimentar
Keil, K., McCoy, T.J., 2017. Acapulcoite-lodranite meteorites: ultramafic asteroidal partial do Parnaíba. Companhia de Pesquisa de Recursos Minerais, Recife, pp. 50–115.
melt residues. Chem. Erde – Geochem. 78, 153–203. https://doi.org/10.1016/j. Lisboa, N.A., Schuck, M.T.G.O., Tramontina H.C., 1987. Reconhecimento geológico da
chemer.2017.04.004. região do Jarau, Quaraí-RS. In Simpósio Sul-Brasileiro de Geologia, 3, Curitiba, v. 1,
Kenkmann, T., Artemieva, N.A., Wünnemann, K., Poelchau, M.H., Elbeshausen, D., Núñez p. 319–332.
del Prado, H., 2009a. The Carancas meteorite impact crater, Peru: geologic surveying Lugli, S., Reimold, W.U., Koeberl, C., 2005. Silicified cone-in-cone structures from Erfoud
and modeling of crater formation and atmospheric passage. Meteorit. Planet. Sci. 44, (Morocco): a comparison with impact-generated shatter cones. In: Koeberl, C.,
985–1000. Henkel, H. (Eds.), Impact Tectonics. Springer-Verlag, Berlin-Heidelberg, pp. 81–110.
Kenkmann, T., Collins, G.S., Wittmann, A., Wünnemann, K., Reimold, W.U., Melosh, H.J., Machado, R., Lana, C., Stevens, G., Filho, C.R.S., Reimold, W.U., McDonald, I., 2009.
2009b. A model for the formation of the Chesapeake Bay impact crater as revealed by Generation, mobilization and crystallization of impact-induced alkali-rich melts in
drilling and numerical simulation. In: Gohn, G.S., Koeberl, C., Miller, K.G., Reimold, granitic target rocks Evidence from the Araguainha impact structure, central Brazil.
W.U. (Eds.), The ICDP-USGS Deep Drilling Project in the Chesapeake Bay Impact Geochim. Cosmochim. Acta 73, 7183–7201.
Structure: Results from the Eyreville Core Holes. Geol. Soc. Amer., pp. 571–585 Masaitis, V.L., 1999. Impact structures of northeastern Eurasia: the territories of Russian
Special Paper 458. and adjacent countries. Meteorit. Planet. Sci. 34, 691–711.
Kenkmann, T., Vasconcelos, M.A.R., Crósta, A.P., Reimold, W.U., 2011. The complex Masero, W., Schenegg, P.A., Fontes, S.L., 1994. A magnetotelluric investigation of the
impact structure Serra da Cangalha, Tocantins State, Brazil. Meteorit. Planet. Sci. 46, Araguainha impact structure in Mato Grosso-Goiás, central Brazil. Geophys. J. Intl.
875–889. 116, 366–376.
Kenkmann, T., Poelchau, M.H., Wulf, G., 2014. Structural geology of impact craters. J. Masero, W., Fischer, G., Schnegg, P.A., 1997. Electrical condutivity and crustal de-
Struct. Geol. 62, 156–182. formation from magnetotelluric results in the region of the Araguainha impact,
Kenkmann, T., Hergarten, S., Kuhn, T., Wilk, J., 2016. Formation of shatter cones by Brazil. Phys. Earth Planet. Int. 101, 271–289.
symmetric fracture bifurcation: phenomenological modeling and validation. Master, S., Heymann, J., 2000. A possible new impact structure near Gilbués in Piauí
Meteorit. Planet. Sci. 51, 1519–1533. Province, Northeastern Brazil. Meteorit. Planet. Sci. 35, 5.
Kenkmann, T., Deutsch, A., Thoma, K., et al., 2018. Experimental impact cratering: a Maziviero, M.V., Vasconcelos, M.A.R., Góes, A.M., Crósta, A.P., Reimold, W.U., 2012. The
summary of the major results of the MEMIN research unit. Meteorit. Planet. Sci. Riachão Ring impact structure, northeastern Brazil: Re-evaluation of its stratigraphy
https://doi.org/10.1111/maps.13048. and evidence for impact. Lunar Planet. Sci. XXXXIII. Lunar and Planetary Institute,
Kieffer, S.W., Phakey, P.P., Christie, J.M., 1976. Shock processes in porous quartzite: Houston, Texas Abstract #1511.
transmission electron microscope observations and theory. Contrib. Mineral. Petrol. Maziviero, M.V., Vasconcelos, M.A.R., Crósta, A.P., Góes, A.M., Reimold, W.U., Carneiro,
59, 41–93. C.C., 2013a. Geology and impact features of Riachão structure northern Brazil.
Koeberl, C., Shirey, S.B., 1997. Re-Os systematics as a diagnostic tool for the study of Meteorit. Planet. Sci. 48, 2044–2058.
impact craters and distal ejecta. Palaeogeogr. Palaeoclimatol. Palaeoecol. 132, Maziviero, M.V., Carrino, T.A., Crósta, A.P., Góes, A.M., Vasconcelos, M.A.R., Reimold,
25–46. W.U., 2013b. Integrating ASTER, geologic and airborne gamma-ray spectrometric
Koeberl, C., Reimold, W.U., Shirey, S.B., 1996. A Re-Os isotope study of the Vredefort data as an aid to mapping silica-bearing rocks in the Riachão impact structure,
Granophyre: clues to the origin of the Vredefort Structure, South Africa. Geology 24, northeastern Brazil. Latin American Remote Sensing Week (LARS). Santiago, Chile
913–916. (abstract).
Koeberl, C., Reimold, W.U., Plescia, J., 2005. BP and Oasis impact structures, Libya: re- McCarthy, D., Aldiss, D., Arsenikos, S., Stone, P., Richards, P., 2017. Comment on
mote sensing and field studies. In: Koeberl, C., Henkel, H. (Eds.), Impact Tectonics. Geophysical evidence for a large impact structure on the Falkland (Malvinas) Plateau.
Springer-Verlag, Berlin-Heidelberg, pp. 161–190. Terra Nova 29, 411–415. https://doi.org/10.1111/ter.12285.
Koeberl, C., Shukolyukov, A., Lugmair, G.W., 2007. Chromium isotopic studies of ter- McDonald, I., Andreoli, M.A.G., Hart, R.J., Tredoux, M., 2001. Platinum-group elements
restrial impact craters Identification of meteoritic components at Bosumtwi, in the Morokweng impact structure, South Africa: evidence for the impact of a large
Clearwater East, Lappajärvi, and Rochechouart. Earth Planet. Sci. Lett. 256, 534–546. ordinary chondrite projectile at the Jurassic-Cretaceous boundary. Geochim.

58
A.P. Crósta et al. Chemie der Erde xxx (xxxx) xxx–xxx

Cosmochim. Acta 65, 299–309. PROSPEC, 1972. Levantamento gama-espectrométrico e magnetométrico na área do
McHall, G.J.H., 2006. Meteorite cratering: Hooke, Gilbert, Barringer and beyond. In: In: Projeto Alto Garças, estados de Goiás e Mato Grosso. Unpubl. Report 76 p.
McHall, G.J.H., Bowden, A.J., Howarth, R.J. (Eds.), The History of Meteorites and Paiva Filho, A., Andrade, C.A.V., Scheibe, L.F., 1978. Uma janela estratigráfica no oeste
Key Meteorite Collections: Fireballs, Falls and Finds, vol. 256. Geological Society de Santa Catarina: o Domo de Vargeão. In: Congresso Brasileiro de Geologia XXX.
London, pp. 443–470 Special Publ. Recife.
McHone Jr., J.F., Dietz, R.S., 1978. Astroblemes in Brazil. Proceedings, 74th GSA Peate, D.W., Hawkesworth, C.J., Mantovani, M.S.M., 1992. Chemical stratigraphy of the
Cordilleran Section Meeting. Paraná lavas (South America): Classification of magma types and their spatial dis-
McHone Jr., J.F., 1979. Riachão Ring, Brazil: a possible meteorite crater discovered by tribution. Bull. Volcanol. 55, 119–139.
the Apollo astronauts. Apollo-Soyuz Test Project Summary Science Report, vol. 412. Peate, D.W., 1997. The Paraná-Etendeka Province. In: In: Mahoney, J.J., Coffin, M.F.
pp. 193–202. (Eds.), Large Igneous Provinces: Continental, Oceanic, and Planetary Flood
McHone Jr., J.F., 1986. Terrestrial Impact Structures: Their Detection and Verification Volcanism. Geophysical Monograph, vol. 100. American Geophysical Union,
with Two New Examples from Brasil. Ph.D. Thesis. University of Illinois at Urbana- Washington, DC, pp. 217–245.
Champaign, Urbana, Illinois, USA. Pedreira da Silva, A.J., Cunha, L.R., Vasconcelos, M., Bahia, R.B.C., 2003. Bacias sedi-
Melosh, H.J., Ivanov, B.A., 1999. Impact crater collapse. Annu. Rev. Earth Planet. Sci. 27, mentares Paleozóicas e Meso-Cenozóicas interiores. In: Bizzi, L.A., Schobbenhaus, C.,
385–415. Vidotti, R.M., Gonçalves, J.H. (Eds.), Geologia, tectônica e recursos minerais do
Melosh, H.J., 1989. Impact Cratering: A Geologic Process. Oxford University Press, New Brasil, pp. 55–85 CPRM, Brasilia.
York 245 pp. Philipp, R.P., Rolim, S.B.A., Sommer, C.A., Souza Filho, C.R., Lisboa, N.A., 2010. A es-
Melosh, H.J., 2011. Planetary Surface Processes. Cambridge Planetary Sciences No. 13. trutura de impacto de Cerro do Jarau. Rev. Bras. Geoc. 40, 468–483.
Cambridge University Press, New York 534 pp. Pilkington, M., Grieve, R.A.F., 1992. The geophysical signature of terrestrial impact
Milani, E.J., França, A.B., Schneider, R.L., 1994. Bacia do Paraná. Bol. Geoc. Petrobras 8, craters. Rev. Geophys. 30, 161–181.
69–82. Pittarello, L., Nestola, F., Viti, C., Crósta, A.P., Koeberl, C., 2015a. Melting and cataclastic
Milani, E.J., Melo, J.H.G., Souza, P.A., Fernandes, L.A., França, A.B., 2007. Bacia do features in shatter cones in basalt from the Vista Alegre impact structure, Brazil.
Paraná. Bol. de Geoc. Petrobras 15, 265–287. Meteorit. Planet. Sci. 50, 1228–1243. https://doi.org/10.1111/maps.12466.
Milani, E.J., 1997. Evolução Tectono-Estratigráfica da Bacia do Paraná e seu Pittarello, L., Roszjar, J., Mader, D., Debaille, V., Claeys, P., Koeberl, C., 2015b.
Relacionamento com a Geodinâmica Fanerozóica do Gondwana Sul-Ocidental. Ph.D. Cathodoluminescence as a tool to discriminate impact melt, shocked and unshocked
Thesis. Universidade Federal do Rio Grande do Sul, Brazil 254 p. volcanics: a case study from the El’gygytgyn impact structure. Meteorit. Planet. Sci.
Milton, D.J., Glickson, A.Y., Brett, R., 1996. Gosses Bluff—a late jurassic impact structure, 50, 1954–1969.
central Australia: part 1: geological structure, stratigraphy, and origin. AGSO J. Poelchau, M.H., Kenkmann, T., 2011. Feather features: a low-shock-pressure indicator in
Austral. Geol. Geophys. 16, 453–486. quartz. J. Geophys. Res. 116. https://doi.org/10.1029/2010JB007803.
Mohr-Westheide, T., Reimold, W.U., 2010. Microchemical investigation of small-scale Preuss, J., 2012. Characterization of Impact-Related Melt Breccias from the Impact
pseudotachylitic breccias from the Archean gneiss of the Vredefort dome, South Structure Araguainha (Brazil). M. Sc Thesis. 92pp.. University of Potsdam, Berlin.
Africa. In: Gibson, R.L., Reimold, W.U. (Eds.), Large Meteorite Impacts and Planetary Raschke, U., Reimold, W.U., Zaag, P.T., Pittarello, L., Koeberl, C., 2013. Lithostratigraphy
Evolution IV. Geol. Soc. Amer., pp. 619–643 (Special Paper 465). of the impactite and bedrock section of ICDP drill core D1c from the El’gygytgyn
Mohr-Westheide, T., Reimold, W.U., 2011. Formation of pseudotachylitic breccias in the impact crater, Russia. Meteorit. Planet. Sci. 48, 1143–1159. https://doi.org/10.1111/
central uplifts of very large impact structures: scaling the melt formation. Meteorit. maps.12072.
Planet. Sci. 46, 543–555. Raschke, U., Schmitt, R.T., McDonald, I., Reimold, W.U., Mader, D., Koeberl, C., 2015.
Morgan, J., Warner, M., the Chicxulub Working Group Brittan, J., Buffler, R., Camargo, Geochemical studies of impact breccias and country rocks from the El'gygytgyn im-
A., Christeson, G., Denton, P., Hildebrand, A., Hobbs, R., Macintyre, H., Mackenzie, pact structure, Russia. Meteorit. Planet. Sci. 50, 1071–1088. https://doi.org/10.
G., Maguire, P., Marin, L., Nakamura, Y., Pilkington, M., Sharpton, V., Snyder, D., 1111/maps.12455.
Suarez, G., Trejo, A., 1997. Size and morphology of the Chicxulub impact crater. Ray, D., Updhyay, D., Misra, S., Newsom, H.E., Ghosh, S., 2017. New insights on petro-
Nature 390, 472–476. graphy and geochemistry of impactites from the Lonar crater, India. Meteorit. Planet.
Morgan, J.V., Gulick, S.P.S., Bralower, T., Chenot, E., Christeson, G., Claeys, P., Cockell, Sci. https://doi.org/10.1111/maps.12881.
C., Collins, G.S., Coolen, M.J.L., Ferrière, L., Gebhardt, C., Goto, K., Jones, H., Kring, Reimold, W.U., 1982. The Lappajärvi meteorite crater, Finland: petrography, Rb-Sr,
D.A., Le Ber, E., Lofi, J., Long, X., Lowery, C., Mellett, C., Ocampo-Torres, R., Osinski, major and trace element geochemistry of the impact melt and basement rocks.
G.R., Perez-Cruz, L., Pickersgill, A., Poelchau, M., Rae, A., Rasmussen, C., Rebolledo- Geochim. Cosmochim. Acta 46, 1203–1225.
Vieyra, M., Riller, U., Sato, H., Schmitt, D.R., Smit, J., Tikoo, S., Tomioka, N., Urrutia- Reimold, W.U., Gibson, R.L., 2006. The melt rocks of the Vredefort Impact Structure
Fucugauchi, J., Whalen, M., Wittmann, A., Yamaguchi, K.E., Zylberman, W., 2016. –Vredefort Granophyre and pseudotachylitic breccias: implications for impact cra-
The formation of peak rings in large impact craters. Science 354, 878–882. https:// tering and the evolution of the Witwatersrand Basin. Chem. Erde 66, 1–35.
doi.org/10.1126/science.aah6561. Reimold, W.U., Jourdan, F., 2012. Impact! –bolides, craters and catastrophes. Elements 8,
Moser, D.E., Cupelli, C.L., Barker, I.R., Flowers, R.M., Bowman, J.R., Wooden, J., Hart, 19–24.
R.J., 2011. New zircon shock phenomena and their use for dating and reconstruction Reimold, W.U., Koeberl, C., 2014a. Impact structures in Africa: a review. J. Afr. Earth Sci.
of large impact structures revealed by electron nanobeam (BSD CL, EDS) and isotopic 93, 57–175.
U-Pb and (UTh)/He analysis on the Vredefort dome. Can. J. Earth Sci. 48, 117–139. Reimold, W.U., Koeberl, C., 2014b. Reply to comment on impact structures in Africa: a
Mutter, R.J., Tomassi, H.Z., Do Carmo, D.A., 2008. In pursuit of causes for the greatest review (Short Note) by Acevedo R.D. et al. J. Afr. Earth Sci. 100, 757–758.
mass extinction: the Permo-Triassic Boundary in the Southern Hemisphere – part II. Reimold, W.U., Minnitt, R.C.A., 1996. Impact-induced shatter cones or percussion marks
Vierteljahresschr. Naturforsch. Ges. Zürich 153, 81–91. on quartzites of the Witwatersrand and Transvaal supergroups? S. Afr. J. Geol. 99,
Nédélec, A., Paquette, J.-L., Yokoyama, E., Trindade, R.I.F., Aigouy, T., Baratoux, D., 299–308.
2013. In situ U/Pb dating of impact-produced zircons from the Vargeão Dome Reimold, W.U., Koeberl, C., Brandt, D., 1997. Suevite from the Roter Kamm impact crater,
(Southern Brazil). Meteorit. Planet. Sci. 48, 420–431. Namibia. Meteorit. Planet. Sci. 32, 431–437.
Nicolaysen, L.O., Reimold, W.U., 1999. Vredefort shatter cones revisited. J. Geophys. Res. Reimold, W.U., Koeberl, C., Brandstätter, F., Kruger, F.J., Armstrong, R.A., Bootsman, C.,
104, 4911–4930. https://doi.org/10.1029/1998JB900068. 1999. The Morokweng impact structure, South Africa: geological, petrographical, and
Northfleet, A.A., Medeiros, R.A., Muhlmann, H., 1969. Reavaliação dos dados geológicos isotopic results, and implications for the size of the structure. In: Dressler, B.O.,
da Bacia do Paraná. Bol. Técn. Petrobrás 12, 291–346. Grieve, R.A.F., Sharpton, V.L. (Eds.), Large Meteorite Impacts and Planetary
Ogilvie, P., Gibson, R.L., Reimold, W.U., Deutsch, A., Hornemann, U., 2011. Experimental Evolution. Geological Society of America, pp. 61–90 Special Paper 339.
investigation of shock metamorphic effects in a metapelitic granulite: the importance Reimold, W.U., Leroux, H., Gibson, R.L., 2002. Shocked and thermally metamorphosed
of shock impedance contrast between components. Meteorit. Planet. Sci. 46, zircon from the Vredefort impact structure, South Africa : A transmission electron
1565–1586. microscopic study. Eur. J. Mineral. 14, 859–868.
Ojeda, H.O., Bembom, F.C., 1966. Mapeamento geológico em semi-detalhe do sudoeste de Reimold, W.U., Cooper, G.R.J., Romano, R., Cowan, D.R., Koeberl, C., 2006. Investigation
Riachão. Petrobras Internal Report, Rio de Janeiro Brazil. 74p. of Shuttle Radar Topography Mission data of the possible impact structure at Serra da
Oliveira, G.J.G., Vasconcelos, M.A.R., Crósta, A.P., Reimold, W.U., Góes, A.M., Kowitz, A., Cangalha, Brazil. Meteorit. Planet. Sci. 41, 237–243.
2014. Shatter cones and planar deformation features confirm Santa Marta in Piauí Reimold, W.U., Hansen, B.K., Jacob, J., Artemieva, N.A., Wünnemann, K., Meyer, C.,
State Brazil, as an impact structure. Meteorit. Planet. Sci. 49, 1915–1928. 2011. Petrography of the impact breccias of the Enkingen (SUBO 18) drill core,
Oliveira, G.J.G., Chamani, M.A.C., Góes, A.M., Crósta, A.P., Vasconcelos, M.A.R., southern Ries crater, Germany: new estimate of impact melt volume. Geol. Soc. Am.
Reimold, W.U., 2017. Geological investigation of the central portion of the Santa Bull. 124, 104–132.
Marta impact structure, Piauí State. Brazil. Braz. J. Geol. 47, 673–692. https://doi. Reimold, W.U., McDonald, I., Schmitt, R.T., Hansen, B., Jacob, J., Koeberl, C., 2013a.
org/10.1590/2317-4889201720160095. Geochemical studies of the SUBO 18 (Enkingen) drill core and other impact breccias
Osae, S., Misra, S., Koeberl, C., Sengupta, D., Gosh, S., 2005. Target rocks, impact glasses, from the Ries crater, Germany. Meteorit. Planet. Sci. 48, 1531–1571.
and melt rocks from the Lonar impact crater, India: petrography and geochemistry. Reimold, W.U., Ferriére, L., Deutsch, A., Koeberl, C., 2014. Impact controversies: impact
Meteorit. Planet. Sci. 40, 1473–1492. recognition criteria and related issues. Meteorit. Planet. Sci. 49, 723–731.
Osinski, G.R., Pierazzo, E. (Eds.), 2013. Impact Cratering: Processes and Products. Wiley- Reimold, W.U., Hoffmann, M., Schmitt, R.T., Hauser, N., Zaag, P.T., Mohr-Westheide, T.,
Blackwell, Chichester 316 pp. 2016. A geochemical contribution to the discussion about the genesis of impact-re-
Osinski, G.R., Tornabene, L.L., Grieve, R.A.F., 2011. Impact ejecta emplacement on ter- lated pseudotachylitic breccias: studies of PTB in the Otavi and Kudu quarries of the
restrial planets. Earth Planet. Sci. Lett. 310, 167–181. Vredefort Dome support the in situ formation hypothesis. S. Afr. J. Geol. 119,
Osinski, G.R., Grieve, R.A.F., Chanou, A., Sapers, H.M., 2016. The suevite conundrum, 453–472. https://doi.org/10.2113/gssajg.119.3.453.
part 1: the ries suevite and sudbury onaping formation compared. Meteorit. Planet. Reimold, W.U., Hauser, N., Hansen, B.T., Thirlwall, M., Hoffmann, M., 2017a. The impact
Sci. 51, 2316–2333. pseudotachylitic breccia controversy: insights from first isotope analysis of Vredefort

59
A.P. Crósta et al. Chemie der Erde xxx (xxxx) xxx–xxx

PTB and impact melt rock (Vredefort Granophyre). Geochim. Cosmochim. Acta 214, Stöffler, D., Langenhorst, F., 1994. Shock metamorphism of quartz in nature and ex-
266–281. periment: I: basic observations and theory. Meteoritics 29, 155–181.
Reimold, W.U., Crósta, A.P., Koeberl, C., Hauser, N., 2017b. Comment on Geophysical Stöffler, D., Keil, K., Scott, E.R.D., 1991. Shock metamorphism of ordinary chondrites.
evidence for a large impact structure on the Falkland (Malvinas) Plateau. Terra Nova Geochim. Cosmochim. Acta 55, 3845–3867.
29, 409–410. https://doi.org/10.1111/ter.12284. Theilen-Willige, B., 1981. The Araguainha impact structure, Central Brazil. Rev. Bras.
Reimold, W.U., Hauser, N., Crósta, A.P., 2018. The impact record of SW gondwana. In: Geoc. 11, 91–97.
Siegesmund, S., et, al. (Eds.), Geology of Southwest Gondwana. Regional Geology Thiele, D.S., Vasconcelos, P.M., 2010. Paraná flood basalts: Rapid extrusion hypothesis
Reviews, Springer, pp. 677–688. confirmed by new 40Ar/39Ar results. Geology 38, 747–750.
Reimold, W.U., 1995. Pseudotachylites in impact structures –generation by friction Thomé-Filho, J.J., Crósta, A.P., de Paula, T.L.F., 2012. Astroblema de Araguainha-Ponte
melting and shock brecciation? A review and discussion. Earth-Sci. Rev. 39, 247–265. Branca (GO/MT). In: Schobbenhaus, C., Silva, C.R. (Eds.), Geoparques do Brasil:
Reimold, W.U., 1998. Exogenic and endogenic breccias: a discussion of major proble- Propostas. CPRM/Serviço Geológico do Brasil, Rio de Janeiro, pp. 151–182.
matics. Earth-Sci. Rev. 43, 25–47. Thompson, L.M., Spray, J.G., 2017. Dynamic interaction between impact melt and frag-
Rocca, M.C.L., Rampino, M.R., Presser, J.L.B., 2017. Geophysical evidence for a large mented basement at Manicouagan: the suevite connection. Meteorit. Planet. Sci. 52,
impact structure on the Falkland (Malvinas) Plateau. Terra Nova 29, 233–237. 1300–1329.
Roddy, D.J., Pepin, R.O., Merril, R.B. (Eds.), 1977. Impact and Explosion Cratering. Timms, N., Erickson, T., Pearce, M., Cavosie, A., Schmieder, M., Tohver, E., Reddy, S.,
Pergamon Press, New York 1301 pp. Zanetti, M.R., Nemchin, A.A., Wittmann, A., 2017. A pressure-temperature phase
Romano, R., Crósta, A.P., 2004. Brazilian impact structures: a review. Lunar Planet. Sci. diagram for zircon at extreme conditions. Earth-Sci. Rev. 165, 185–202.
XXXV. Lunar and Planetary Institute, Houston, Texas Abstract #1546. Tohver, E., Lana, C., Cawood, P.A., Fletcher, I.R., Jordan, F., Jourdan, F., Sherlock, S.,
Salminen, J., Pesonen, L.J., Reimold, W.U., Donadini, F., Gibson, R.L., 2009. Rasmussen, B., Trinidade, R.I.F., Yokoyama, E., Souza Filho, C.R., Marangoni, Y.,
Paleomagnetic and rock magnetic study of the Vredefort impact structure and the 2012. Geochronological constraints on the age of a Permo–Triassic impact event:
Johannesburg Dome Kaapvaal Craton, South Africa −implications for the paleor- U–Pb and 40Ar/39Ar results for the 40 km Araguainha structure of central Brazil.
econstructions of continents during the Mesoproterozoic. Precambr. Res. 168, Geochim. Cosmochim. Acta 86.
167–184. Tohver, E., Cawood, P.A., Riccomini, C., Trindade, R.I.F., 2013. Shaking a methane fizz:
Sanchez, J.P., Brilha, J.B.R., 2017. Terrestrial impact structures as geoheritage: an as- seismicity from the Araguainha impact event and the Permian–Triassic global carbon
sessment method of their scientific value and its application to Brazil. Ann. Acad. isotope record. Palaeogeogr. Palaeoclimatol. Palaeoecol. 387, 66–75. https://doi.
Bras. Ciênc. 89, 825–834. org/10.1016/j.palaeo.2013.07.010.
Sanchez, J.P., Simões, L.S.A., Martins, L.E.B., 2014. Estratigrafia e estrutura do Cerro do Tompkins, L.A., 1994. Tectono-structural environments of primary diamond source rocks
Jarau: nova proposta. Braz. J. Geol. 44, 265–276. in Brazil. In: Meyer, H.O.A., Leonardos, O.H. (Eds.), Proceedings of the Fifth
Sanchez, J.P., 2006. Mapeamento 1:25 000 do núcleo do Astroblema Domo de International Kimberlite Conference, vol. 2. Diamonds: Characterization, Genesis and
Araguainha e aspectos geoturísticos da região. Diploma thesis. Universidade Estadual Exploration. CPRM Special Publication 1B/94, Brasilia, pp. 259–267.
Paulista 57 p. Tong, C.H., Lana, C., Marangoni, Y.R., Elis, V.R., 2010. Geoelectric evidence for cen-
Sanchez, J.P., 2013. Mapeamento geológico-estrutural do astroblema de Cerro do Jarau tripetal resurge of impact melt and breccias over central uplift of Araguainha impact
(RS), Brasil. Ph.D. thesis. Universidade Estadual Paulista 173 p. structure. Geology 38, 91–94.
Santos, U.P., McHone Jr., J.F., 1979. Field report on Serra da Cangalha Riachão circular Trepmann, C.A., Spray, J.G., 2005. Planar microstructures and Dauphine twins in shocked
feature. Instituto de Pesquisas Espaciais (INPE) Relatório 1458-NTE/153, 13 p. quartz from the Charlevoix impact structure, Canada. In: Kenkmann, T., Horz, F.,
Santos, M.E.C.M., de Carvalho, M.S.S., 2009. Paleontologia das Bacias do Parnaíba, Deutsch, A. (Eds.), Large Meteorite Impacts III. Geol. Soc. Am., pp. 315–328 Spec.
Grajaú e São Luís. Reconstituições Paleobiológicas. CPRM/Serviço Geológico do Paper 384.
Brasil, Rio de Janeiro 215 p. Trepmann, C.A., Spray, J.G., 2006. Shock-induced crystal-plastic deformation and post-
Schnegg, P.A., Fontes, S.L., 2002. Feasibility study of the geoelectric structure of the shock annealing of quartz: microstructural evidence from crystalline target rocks of
Araguainha impact, Brazil. Earth Planets Space 54, 597–606. the Charlevoix impact structure, Canada. Eur. J. Mineral. 18, 161–173. https://doi.
Schulte, P., Alegret, L., Arenillas, I., et al., 2010. The Chicxulub asteroid impact and mass org/10.1127/0935-1221/2006/0018-0161.
extinction at the Cretaceous-Paleogene boundary. Science 327, 1214–1218. https:// Turner, S., Regelous, M., Kelley, S., Hawkesworth, C.J., Mantovani, M.S.M., 1994.
doi.org/10.1126/science.1177265. Magmatism and continental break-up in the South Atlantic: high precision 40Ar-39Ar
Shand, S.J., 1916. The pseudotachylyte of Parijs (Orange Free State) and its relation to geochronology. Earth Planet. Sci. Lett. 121, 333–348.
trapshotten gneiss and flinty crush-rock. Q. J. Geol. Soc. London 72, 198–221. Turtle, E.P., Pierazzo, E., Collins, G.S., Osinski, G.R., Melosh, H.J., Morgan, J.V., Reimold,
Sharpton, V.L., Burke, K., Camargo-Zanoguera, A., Hall, S.A., Lee, D.S., Marín, L.E., W.U., 2005. In: Kenkmann, T., Hörz, F., Deutsch, A. (Eds.), Impact Structures: What
Suárez-Reynoso, G., Quezada-Muñeton, J.M., Spudis, P.D., Urrutia-Fucugauchi, J., Does Crater Diameter Mean? Geol. Soc. Amer., pp. 1–24 Special Paper 384.
1993. Chicxulub multiring impact basin: size and other characteristics derived from Uchôa, E.B., Vasconcelos, M.A.R., Crósta, A.P., 2013. Santa Marta crater: macroscopic
gravity analyses. Science 261, 1564–1567. and petographic evidences of a new confirmed impact struture in Northeastern Brazil.
Shen, S.Z., Crowley, J.L., Wang, Y., Bowring, S.A., Erwin, D.H., Sadler, P.M., Cao, C.Q., Lunar Planet. Sci. XXXXIV. Lunar and Planetary Institute, Houston, Texas Abstract
Rothman, D.H., Henderson, C.M., Ramezani, J., Zhang, H., Shen, Y., Wang, X.D., #1316.
Wang, W., Mu, L., Li, W.Z., Tang, Y.G., Liu, X.L., Liu, L.J., Zeng, Y., Jiang, Y.F., Jin, UnB-IG, 2012. Final Mapping Project by the Graduate Class of 2012. Institute of
Y.G., 2011. Calibrating the end-Permian mass extinction. Science 334, 1367–1372. Geosciences, University of Brasilia, Brasilia, Brasil Unpubl. report.
https://doi.org/10.1126/science.1213454. Vasconcelos, M.A.R., Crósta, A.P., Molina, E.C., 2010. Geophysical characteristics of four
Siegert, S., Branney, M., Hecht, L., 2017. Density current origin of a melt-bearing impact possible impact structures in the Paraíba Basin, Brazil. In: Gibson, R.L., Reimold
ejecta blanket (Ries suevite, Germany). Geology 45, 855–858. (Eds.), Large Meteorite Impacts and Planetary Evolution IV. Geol. Soc. Amer., pp.
Silva, D., Lana, C.C., Stevens, G., Souza-Filho, C.R., 2011. Effects of shock-induced in- 201–217 Special Paper 465.
congruent melting within Earth́ s crust: the case of biotite melting. Terra Nova 23, Vasconcelos, M.A.R., Leite, E.P., Crósta, A.P., 2012a. Contributions of gamma-ray spec-
225–231. trometry to terrestrial impact crater studies: the example of Serra da Cangalha,
Silva, D., Lana, C., Souza-Filho, C.R., 2016. Petrographic and geochemical characteriza- northeastern Brazil. Geophys. Res. Lett. 39, pL04306.
tion of the granitic rocks of the Araguainha impact crater. Brazil. Meteorit. Planet. Vasconcelos, M.A.R., Wünnemann, K., Crósta, A.P., Molina, E.C., Reimold, W.U.,
Sci. 51, 443–467. Yokoyama, E., 2012b. Insights into the morphology of the Serra da Cangalha impact
Silveira Filho, N.C., Ribeiro, C.L., 1971. Informações geológicas preliminares sobre a structure from geophysical modeling. Meteorit. Planet. Sci. 47, 1659–1670.
estrutura vulcânica de Araguainha, Mato Grosso. Int. Report. Departamento Nacional Vasconcelos, M.A.R., Maziviero, M.V., Crósta, A.P., 2012c. The Riachão impact structure:
de Pesquisas Minerais, Brasília 50 pp. evidence of an oblique impact. In: 75th Annual Meeting of the Meteoritical Society.
Smit, J., Hertogen, J., 1980. An extraterrestrial event at the Cretaceous-tertiary boundary. Cairns, Queensland, Australia. Abstract #5127.
Nature 285, 158–200. Vasconcelos, M.A.R., Crósta, A.P., Reimold, W.U., Goes, A.M., Kenkmann, T., Poelchau,
Spray, J.G., 2010. Frictional melting of planetary materials: from hypervelocity impact to M.H., 2013. The Serra da Cangalha impact structure, Brazil: Geological, stratigraphic
earthquakes. Annu. Rev. Earth Planet. Sci. 38, 221–254. and petrographic aspects of a recently confirmed impact structure. J. South Am. Earth
Spudis, P.D., 1993. The Geology of Multi-Ring Basins: The Moon and Other Planets. Sci. 45, 316–330.
Cambridge University Press, Cambridge 263pp. Vasconcelos, M.A.R., 2007. Caracterização geofísica da estrutura de impacto de
Stöffler, D., Reimold, W.U., 2006. Geologic setting, properties, and classification of ter- Araguainha. M. Sc. Thesis. University of São Paulo (157 p.).
restrial impact formations. In: First Int. Conf. on Impact Cratering in the Solar System. von Engelhardt, W., Matthäi, S.K., Walzebuck, J., 1992. Araguainha impact crater, Brazil.
European Space Agency, Nordwijk, The Netherlands. pp. 205–207 Abstracts. 1. The interior part of the uplift. Meteoritics 27, 442–457.
Stöffler, D., Reimold, W.U., Jacob, J., Hansen, B.K., Summerson, I.A.T., Artemieva, N.A., Wagner, R., Reimold, W.U., Brandt, D., 2002. Bosumtwi impact crater, Ghana: a remote
Wünnemann, K., 2013. Ries crater and suevite revisited – observations and modeling: sensing investigation. In: Plado, J., Pesonen, L.J. (Eds.), Meteorite Impacts in
part I: observations. Meteorit. Planet. Sci. 48, 515–589. Precambrian Shields. Springer-Verlag, Berlin-Heidelberg, pp. 189–210.
Stöffler, D., Hamann, C., Metzler, K., 2017. Shock metamorphism of planetary silicate Wegener, A., 1921. Die Entstehung der Mondkrater. Friedrich Vieweg & Sohn
rocks and sediments: proposal for an updated classification system. Meteorit. Planet. Braunschweig 48 pp.
Sci. 53, 5–49. https://doi.org/10.1111/maps.12912. Wenk, H.-R., Lanardelli, I., Vogel, S.C., Tullis, J., 2005. Dauphiné twinning as evidence
Stewart, S.A., 2011. Estimates of yet-to-find impact crater population on Earth. J. Geol. for an impact origin of preferred orientation in quartzite: an example from Vredefort,
Soc. London 168, 1–14. South Africa. Geology 33, 273–276.
Stöffler, D., Grieve, R.A.F., 2007. Impactites. In: Fettes, D., Desmons, J. (Eds.), Wenk, H.-R., Janssen, C., Kenkmann, T., Dresen, G., 2011. Mechanical twinning in quartz:
Metamorphic Rocks: A Classification and Glossary of Terms, Recommendations of the shock experiments, impact, pseudotachylites and fault breccias. Tectonophysics 510,
International Union of Geological Sciences. Cambridge University Press, Cambridge, 69–79.
U.K (pp. 82–92 111–125, and 126–242). Whitehead, J., Grieve, R.A.F., Spray, J.G., 2002. Mineralogy and petrology of melt rocks

60
A.P. Crósta et al. Chemie der Erde xxx (xxxx) xxx–xxx

from the Popigai impact structure, Siberia. Meteorit. Planet. Sci. 37, 629–647. Yokoyama, E., Nédélec, A., Baratoux, D., Trindade, R.I.F., Fabre, S., Berger, G., 2015.
Wieland, F., Reimold, W.U., Gibson, R.L., 2006. New observations on shatter cones in the Hydrothermal alteration in basalts from Vargeão impact structure South Brazil, and
Vredefort impact structure, South Africa, and an evaluation of current models for implications for recognition of impact-induced hydrothermalism on Mars. Icarus 252,
shatter cone formation. Meteorit. Planet. Sci. 41, 1737–1759. 347–365. https://doi.org/10.1016/j.icarus.2015.02.001.
Wilk, J., Kenkmann, T., 2016. Formation of shatter cones in MEMIN impact experiments. Yokoyama, E., 2008. Trama magnética do núcleo soerguido da estrutura de impacto de
Meteorit. Planet. Sci. 51, 1477–1496. Araguainha (MT/GO). M. Sc thesis. University of São Paulo 173 p.
Yokoyama, E., Trindade, R.I.F., Lana, C.R., Souza-Filho, C.R., Baratoux, D., Marangoni, Zaag, P.T., Reimold, W.U., Hipsley, C.A., 2016. Microcomputed tomography and shock
Y.R., Tohver, E., 2012. Magnetic fabric of Araguainha complex impact structure microdeformation studies on shatter cones. Meteorit. Planet. Sci. 51, 1435–1459.
(Central Brazil): Implications for deformation mechanisms and central uplift forma- Zaag, P.T., 2013. Petrographic and μ-Computed Tomography Investigations of Shatter
tion. Earth Planet. Sci. Lett. 331–332, 347–359. Cones. Diploma Thesis. 104pp.. Freie Universität Berlin, Germany.
Yokoyama, E., Brandt, D., Tohver, E., Trindade, R.I.F., 2014. Palaeomagnetism of the Zalan, P.V., Wolf, S., Astolfi, M.A.M., Vieira, I.S., Conceição, J.C., Appi, V.T., Santos Neto,
Permo-Triassic Araguainha impact structure (Central Brazil) and implications for E.V., Cerqueira, J.R., Marques, A., 1991. The Paraná Basin, Brazil. In: Leighton, M.W.,
Pangean reconstructions. Geophys. J. Internat. 198, 154–163. https://doi.org/10. Kolata, D.R., Oltz, D.F., Eidel, J.J. (Eds.), Interior Cratonic Basins. American Assoc.
1093/gji/ggu125. Petrol. Geol., Tulsa, OK, pp. 707–708 AAPG Memoir 51.

61

You might also like