Download as pdf or txt
Download as pdf or txt
You are on page 1of 19

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/275353187

Chemostratigraphy of Neoproterozoic banded iron formation (BIF): types, age


and origin

Chapter · February 2015

CITATIONS READS

0 1,350

1 author:

A. N. Sial
Federal University of Pernambuco
377 PUBLICATIONS   8,151 CITATIONS   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Chemostratigraphy of Neoproterozoic successions in Argentina View project

Petrologia e geoquímica do pluton Rio Formoso, Domínio Pernambuco-Alagoas, Província Borborema View project

All content following this page was uploaded by A. N. Sial on 27 April 2015.

The user has requested enhancement of the downloaded file.


CHAPTER

CHEMOSTRATIGRAPHY OF
NEOPROTEROZOIC BANDED
IRON FORMATION (BIF):
TYPES, AGE AND ORIGIN
17
 1,   2, !" #3
1Departamento de Geología, Facultad de Ciencias, Universidad de la República, Montevideo, Uruguay;
2NEG-LABISE, Department of Geology, Federal University of Pernambuco, Recife, Pernambuco, Brazil;
3Department of Geosciences and Natural Resource Management, Geology Section, University of Copenhagen and
Nordic Center for Earth Evolution (NordCEE), Copenhagen, Denmark

17.1 INTRODUCTION
T$% Banded iron formations (BIFs) are nonactualistic rocks deposited in the Archean, Paleoprotero-
zoic, and Neoproterozoic. They are characterized by rhythmic alternations of iron-rich layers (“bands”)
and chert layers. Based on their iron mineralogy, BIFs can be classified as oxide-, carbonate-, silicate-,
or sulfide-facies (James, 1954)& among which, the oxide-facies is most important. Based on deposi-
tional and geotectonic setting, Gross (1980) '(*+(-./(*$%' 0%+2%%- 35.6789 8-' :8;% </=%>(6>9+?=%
BIFs. Accordingly, the Lake Superior-type BIFs were deposited on stable continental platforms, and
Algoma-type BIFs are associated with volcanic rocks and were deposited in arc or rift settings. A third
BIF type, exclusively Neoproterozoic in age, is the Rapitan type (i.e., Beukes and Klein, 1992)& and is
characterized by a glacially influenced depositional environment.
Geochemistry has played a central role in unraveling the origin of BIFs, and many proxies such as
rare earth element (REE) concentrations, have been widely used (e.g., Kato et al., 2006)@ However,
chemostratigraphic studies are scarce, probably because traditional stable and radiogenic isotope stud-
ies such as δ13C, δ18O, δ34S and 87Sr/86Sr are impossible to apply for the more widespread oxide-facies
BIF. Isotope chemostratigraphy has played a significant role in the study of Neoproterozoic successions
since the pioneer reports by Knoll et al. (1986) 8-' Kaufman et al. (1991)@ T$% %*+805(*$7%-+ 6A +$%
“snowball Earth” hypothesis was based mainly on chemostratigraphic and paleomagnetic evidences
(Hoffman et al., 1998)@ Curves of global secular variations of δ13C and 87Sr/86Sr in the Neoproterozoic
(e.g., Jacobsen and Kaufman, 1999; Melezhik et al., 2001; Halverson et al., 2010) $8B% 0%%- (7=6>+8-+
tools for correlation and geochronology, mostly in combination with biostratigraphy and U–Pb radio-
chronology. The caveat of all these studies is that, for analytical reasons, they can only be applied to the
carbonate successions or, more rarely, to the organic-rich shales (e.g., Johnston et al., 2010)& but not to
the oxide-facies Neoproterozoic iron formations.
The development of new techniques, such as Fe speciation, Cr and Fe isotopes (δ53Cr, δ56Fe), and
Mo concentrations and isotopes (δ98Mo) opened new horizons and enabled thorough chemostratigraphic
C
 y. http://dx.doi.org/10.1016/B978-0-12-419968-2.00017-0 4
yright © 2015 Elsevier Inc. All rights reserved.
DED CHAPTER 17 CHEMOSTRATIGRAPHY OF NEOPROTEROZOIC BIF

*+/'(%* 6A +$(s; FGH */ss%**(6-* ICanfield et al., 2008; Frei et al., 2009; Halverson et al., 2011; Baldwin
et al., 2013)@ J%-s%& A6> +$% K>*+ +(7%& +$% *+/'? 6A *%s/58> B8>(8+(6-* 2(+$(- +$% (>6- A6>78+(6-* 0%s87%
possible, enabling insights into the basin evolution and changes in the redox conditions. Studies using
REE concentrations changed from only reporting REE distributions for whole iron formations (e.g.,
Graf et al., 1994L Kato et al., 2006) +6 >%=6>+(-. *+>8+(.>8=$(s B8>(8+(6-* 6A s%>+8(- MNN =8>87%+%>*& */s$
as Ce and Eu anomalies (e.g., Stern et al., 2013)@ T$(* '%B%56=7%-+ $8' 8''%' +$% +(7% '(7%-*(6- +6
fundamental parameters such as the importance of hydrothermal processes, redox conditions and sources
of iron, which need not be constant troughout deposition of individual iron formations.
The new techniques, coupled with a growing interest on the Neoproterozoic iron formations, show
a more complex picture than previously thought. We review here the occurrence of Rapitan, Algoma,
and Lake Superior BIF types in the Neoproterozoic, and present recent chemostratigraphic data that
enable their characterization.

17.2 AGE OF NEOPROTEROZOIC BIFS


G+ has been recently proposed that all the Neoproterozoic iron formations were deposited during the
middle Cryogenian (“Sturtian”) glaciation at c. 715 Ma (Macdonald et al., 2010aL Cox et al., 2013)@
This is certainly the case for BIF of the Rapitan Group (Sayunei Formation, Canada), which is con-
strained by an U–Pb age of a felsic tuff of 716.5 ± 0.24 Ma (Macdonald et al., 2010b)@ Similarly, gla-
cially influenced BIF of the Fulu Formation and associated diamictites (Jiangkou Group, South China)
directly overlie a tuffaceous siltstone that yielded an U–Pb age of 725 ± 10 Ma. The Chuos Formation
of the Otavi Group (NW-Namibia) includes BIF associated to glacial diamictites, which rests atop a
porphyry lava dated U-Pb at 746 ± 2 Ma (Hoffman et al., 1996)@ This, combined with the fact that there
are more than 500 m of intervening carbonates between the BIF and the end-Cryogenian Ghaub Forma-
tion, dated at 636 ± 1 Ma (Hoffmann et al., 2004)& makes an age around 715 Ma for the Chuos Forma-
tion unlikely, but it cannot be completely ruled out. The Holowilena Ironstone of the Flinders Ranges,
South Australia, is an interesting example that illustrates the problems of global correlations of Neopro-
terozoic ice ages. Fanning and Link (2008) >%=6>+%' OPQ0 <JMGRQ S(>s6- 8.% 6A UUV ± 5 Ma for the
volcaniclastic intercalations in the partially glaciogenic lower Wilyerpa Formation, which is part of the
Sturtian type section. The Holowilena Ironstone occurs some 250 m stratigraphically beneath the Wily-
erpa Formation (Le Heron et al., 2011)@ Thus, although the Holowilena Ironstone is undoubtedly Stur-
ian, because it occurs at the Sturtian type section in Australia, it is probably much younger than the
Rapitan glacials.
BIF occurring at Wadi Karim, Wadi El Dabbah, and Um Anab in Egypt (Basta et al., 2011&
Fig. 1) 8-' +$% s6>>%58+(B% <8282(- FGH (- -6>+$2%*+%>- <8/'( 3>80(8 8>% s6-*+>8(-%' 0%+2%%-
underlying metavolcanics dated U–Pb at 759 ± 17 Ma and intrusive rocks yielding U–Pb age of
710 ± 5 Ma (Stern et al., 2013& 8-' >%A%>%-s%* +$%>%(-)@ T$%>%A6>%& +$%*% FGH* 8>% '(*+(-s+(B%5?
older than the Rapitan iron formation, and they are not related to glacial processes (Basta et al.,
2011)@ T$% WV979+$(s; N>S(- FGH (- *6/+$%>- T/B8 IM/**(8- H%'%>8+(6-) 8-' -6>+$%>- R6-.65(8
occurs within the Mugur Formation, and rests on the top of the metavolcanics of the same unit
dated U–Pb at 767 ± 15 Ma (Ilyin, 2009 )@ T$% <$(5/ X>6/= (- <6/+$ Y$(-8 I Fig. 1) $6*+* ZV +6
400-m-thick, high-grade iron formation deposited in the Tonian between 960 and 830 Ma, as
shown, respectively, by U–Pb ages of detrital zircons and the age of metamorphic overprint (Xu
17.2 AGE OF NEOPROTEROZOIC BIFS [\]

FIGURE 1
bcdefgdhidcjhikel imn oipcqg iggfpcidcfm fo rqfkhfdqhftfcp chfm ofhuidcfmgv wxy zhfm ofhuidcfm id deq {f|mnihl
between the Suyunei and Shezal formations, Rapitan Group, Canada (From Baldwin et al. (2013)}~. (B) Wadi
El Dabbah iron formation, Egypt (From Ali et al. (2009)}~. (C) Jucurutu Formation, Seridó Belt, NE Brazil (From
Van Schmus et al. (2003) Nascimento et al. (2007)}~. Eq. Fm.: Equador Formation. (D) Shilu Group, South
China (From Xu et al. (2013b)}~. Indicated are maximum thicknesses observed. S. Fm.: Shihuiding Formation.

et al., 2013a)@ R6>%6B%>& +$% <$(5/ FGH* 8>% /->%58+%' +6 +$% .58s(85 =>6s%**%* 8-' 8>% s6-*('%>%'
to be Lake Superior type by Xu et al. (2013bL Fig. 1)@
Several occurrences of Neoproterozoic BIFs are demonstrably younger than the middle Cryogenian
glaciation, including the largest one, the >300-m-thick BIF and associated manganese formations of the
Jacadigo Group (Brazil) and its correlative Boqui Group (Bolivia). The glacially influenced BIF (Figs
2(A)–(B) and 3(C)) ?(%5'%' ?6/-.%*+OPQ0 '%+>(+85 S(>s6- 8.%A>67 (-+%>0%''%' *8-'*+6-%* 6A U^Z ± 17 Ma
(Døssing et al., 2010)@ 3>P3> '8+(-. 6- s>?=+67%58-% 2(+$(- +$% R- 6>%* >%B%85* 8 7(-(7/7 '%=6*(-
tional age of 587 ± 7 Ma (Piacentini et al., 2013)& 2$(s$ 78? 0% (-+%>=>%+%' 8* +$% 8.% 6A %8>5? '(8.%-%*(*@
Considering that the metamorphic braunite yielded Ar–Ar ages between 547 and 513 Ma (Piacentini
et al., 2013)& 2$(s$ 78+s$%* 2(+$ +$% 8.% 6A +$% Q8- 3A>(s8-9F>8*(5(8-6 6>6.%-? IZ_` ± 7 Ma) in the
region (McGee et al., 2012)& (+ s6/5' 2%55 8==>6a(78+% +$% '%=6*(+(6-85 8.% 8-' (* 2(+$(- +$% %>>6> 6A +$%
€‚ CHAPTER 17 CHEMOSTRATIGRAPHY OF NEOPROTEROZOIC BIF

(A) (B)

(C) (D)

(E) (G)

(F)

FIGURE 2
ƒ|dphfk imn kf„cgeqn g„i{ kefdfjhikeg fo rqfkhfdqhftfcp {imnqn chfm ofhuidcfm w…z†yv wxy ‡„fmjidq jmqcgg
dropstone with long axis at high angle to bedding in BIF of the Jacadigo Group (W Brazil). (B) Granite
dropstone in pure BIF of the Jacadigo Group. Note the separation between the dropstone and the gritty layer
below. (C) Basement dropstone in BIF of the Jakkalsberg Member, Numees Formation (Namibia). (D) BIF
of the Yerbal Formation (Uruguay) as exposed in a shallow exploration trench. (E) Polished slab of magne-
tite + hematite BIF of the Yerbal Formation, showing thin chert bands (From Gaucher et al. (2004)}~. Field of
view is 9 cm wide. (F) BIF of the Jucurutu Formation, Seridó Belt, at Riacho Fundo section (NE Brazil). (G)
Same unit as previous (Bonito Mine section), but showing thicker and more frequent chert bands. Scale in B,
F, and G is 8 cm long.
FIGURE 3
ˆequfgdhidcjhikel fo rqfkhfdqhftfcp chfm ofhuidcfmg ohfu ‰f|de xuqhcpiv wxy x„jfuiŠdlkq {imnqn chfm ofhuidcfm
(BIF), Jucurutu Formation, Seridó Belt (NE Brazil; Sial et al., submittedyv rfdq ecje ‡|‹‡|Œ imn ˆh pfmpqmdhidcfmg
and δ53Cr entirely within the high-temperature field (colored area). (B) Lower Arroyo del Soldado Group, Uruguay
(Frei et al., 2013yv ‰einqn ihqig uih pequcpi„ gqncuqmdg peihipdqhctqn {l ecje δ53Cr, Eu/Eu*, and FeHR/FeTot,
and low εNd. Dashed line in Fe speciation column marks boundary between oxic and anoxic conditions. (C)
Jacadigo Group (W Brazil, Døssing et al., 2010yv rfdq gqp|„ih Žihcidcfmg cm δ53Cr within the iron formation.
‘ CHAPTER 17 CHEMOSTRATIGRAPHY OF NEOPROTEROZOIC BIF

X8*;(%>* X58s(8+(6- IZ_’@“ ± 0.5 Ma; Bowring et al., 2003)@ Y6>>%58+(B% .58s(6.%-(s '(87(s+(+%* (-+%>0%'-
ded with BIF of the Puga Formation in the southern Paraguay Belt (Brazil, Piacentini et al., 2007) ?(%5'%'
U–Pb age of detrital zircons as young as 709 ± 6 Ma (Babinski et al., 2013)@ T$%? 8>% s6-A6>7805? 6B%>-
lain by the cap carbonates of the Corumbá Group (Boggiani et al., 2003, 2010)& 806B% 2$(s$& 2%559
preserved Ediacaran acritarchs and metazoans (Cloudina, Corumbella) occur (Gaucher et al., 2003)@ 355
these data suggest that deposition of the Jacadigo Group and the Puga Formation (southern Paraguay
Belt) took place in the end-Cryogenian or, more probably, during the Gaskiers glaciation.
BIF also occurs in the Jacurutu Formation located in the Seridó Belt, Brazil (Sial et al., submittedL
Fig. 1)@ Van Schmus et al. (2003) >%=6>+%' OPQ0 8.% A6> +$% '%+>(+85 S(>s6-* 6A +$% *87% /-(+ 8+ (+* +?=%
area as young as 634 ± 13 Ma. Therefore, deposition of the Jucurutu BIF took place in the lower
Ediacaran, between 634 Ma and the peak of metamorphism in the Brasiliano orogeny in this region (c.
600–585 Ma; Hollanda et al., 2010)@
Dropstone-bearing BIFs of the Jakkalsberg Member of the Numees Formation (Fig. 2(C)) (- +$%
Gariep Belt (Namibia) were assigned to the Ediacaran on the basis of Sr isotope ratios and Pb–Pb ages
of its cap carbonate (Fölling and Frimmel, 2002) 8-' 6>.8-(s92855%' 7(s>6A6**(5* 6ss/>>(-. 0%562 8-'
above the unit (Gaucher et al., 2005)@ More recently, Macdonald et al. (2010b) '(*=/+% +$(* 8.% 8**(.--
ment and prefer a correlation with the Rapitan BIF, on the basis of new geological mapping and che-
mostratigraphy. More work, especially reliable U–Pb ages, is needed to solve this issue.
Arguably the youngest examples of Neoproterozoic BIFs were reported from the Yerbal and Cerro
Espuelitas formations of the Arroyo del Soldado Group, Uruguay (Gaucher, 2000; Gaucher et al., 2004;
Frei et al., 2013)@ BIF of the Yerbal Formation (Figs 2(D)–(E) and 3(B))& is devoid of any glacial
features, occurs 800 m above the sandstones for which Blanco et al. (2009) >%=6>+%' OPQ0 '%+>(+85 S(>-
con age as young as 664 ± 14 Ma. Equivalent, more proximal sandstones of the Barriga Negra Forma-
tion yielded youngest U–Pb detrital zircon age of 566 ± 8 Ma. The iron horizons are further interbedded
with siltstones bearing Cloudina riemkeae, an index fossil for the late Ediacaran (Gaucher and Sprech-
mann, 1999; Gaucher et al., 2003; Gaucher and Poiré, 2009)@ The magnetite-bearing BIFs of the Cerro
Espuelitas Formation (Gaucher, 2000) 6ss/> 8+ *+>8+(.>8=$(s855? +26 A6>78+(6-* 806B% +$% ”%>085 H6>-
mation, and are thus of latest Ediacaran age.
In summary, the Neoproterozoic BIFs, pre- and postdate the middle Cryogenian, c. 715 Ma glacia-
tion, and occur in the Tonian, Cryogenian, and Ediacaran. Interestingly, not all the Neoproterozoic BIFs
were deposited in glacially influenced settings. In fact, as we demonstrate in the following sections,
examples of the three BIF types have been described from the Neoproterozoic successions.

17.3 DEPOSITIONAL ENVIRONMENT OF NEOPROTEROZOIC BIFS


17.3.1 SEDIMENTOLOGY AND FACIES ASSOCIATION
17.3.1.1 Glacially Influenced Iron Formations
<%B%>85 •%6=>6+%>6S6(s (>6- A6>78+(6-* 2%>% '%=6*(+%' (- .58s(855? (-–/%-s%' %-B(>6-7%-+*& 8-'
they are assigned to the “Rapitan-type” (Beukes and Klein, 1992)@ 3=8>+ A>67 2%559'6s/7%-+%'
glaciomarine BIF of the Rapitan Group (Baldwin et al., 2012) 6+$%> 8-856./%* *$62 s5%8> (-'(s8+(6-*
of having been influenced by glaciomarine environment. Basement dropstones occur in BIF of the
Jakkalsberg Member of the Numees Formation (Fig. 2(C))& *6/+$%>- •87(0(8 IFrimmel, 2008)@ <(7-
ilarly, the Chuos Formation in northern Namibia comprises glacial diamictite with predominantly
basement clasts, including some striated, and associated thin horizons of BIF (Hoffman and Halverson, 2008;
17.3 DEPOSITIONAL ENVIRONMENT OF NEOPROTEROZOIC BIFS —˜™

Miller, 2008)@ 3 2%559'%B%56=%' s8= s8>06-8+% ?(%5'(-. -%.8+(B% δ13C values (Rasth of Formation)
rests on top of the Chuos Formation (Hoffman and Schrag, 2002)@ T$/*& FGH* 6A +$% Y$/6* H6>78-
tion were deposited in unambiguously glaciomarine environment and can be assigned to the Rapitan
type.
BIFs of the Jacadigo Group (Banda Alta or Santa Cruz Formation) and Puga Formation bear con-
vincing evidences of a glacial environment. The Puga Formation includes both associated diamictites
and dropstones in the BIF (Piacentini et al., 2007)@ In the Jacadigo Group, both planar and elongate,
outsized clasts occur (Fig. 2(A)–(B))@ Planar clasts tend to lie parallel to bedding, but elongated clasts
often show major axes perpendicular to bedding (i.e., bullet clast). A 6 m long granite bullet-clast was
illustrated by Möller (2004) A>67 +$% +6= 6A +$% F8-'8 35+8 H6>78+(6- IFig. 3(C))@ Attempts to explain
these outsized clasts as nonglacial in origin, for example as the result of downslope sliding and/or mass
flows (e.g., Freitas et al., 2011) 8>% -6+ s6-B(-s(-.@ š62-*56=% *5('(-. s8--6+ 8ss6/-+ A6> +$% 6>(%-+8-
tion of bullet clasts. The thin, sandy layers occasionally accompanying outsized clasts (Freitas et al.,
2011; Trompette et al., 1998) '6 -6+ s6-+>8'(s+ 8 .58s(85 6>(.(- A6> +$% 58++%>& 0%s8/*% +$%? s8- 85*6 0%
derived from melting icebergs. Granite dropstones weighing more than 100 tons in chemical sediments
such as BIF and MnF are possibly the most convincing evidences of glaciation so far reported from the
Neoproterozoic deposits.
Thus, a glaciomarine setting is ascribable for a number of Neoproterozoic iron formations, includ-
ing the thickest ones (e.g., Jacadigo Group, Rapitan Group). Interestingly, most of these occurrences
were deposited in rift basins. This has led to the development of models that dispute the importance of
glaciation and emphasize the role of rifting and hydrothermal activity (e.g., Young, 2002; Eyles and
Januszczak, 2004; Freitas et al., 2011)@

17.3.1.2 BIF Associated with Volcanic Rocks


•%6=>6+%>6S6(s (>6- A6>78+(6-* (- N.?=+ 8-' +$%(> s6>>%58+(B% /-(+* (- -6>+$2%*+%>- <8/'( 3>80(8& /= +6
90 m thick, are interbedded with wackes, tuffs, lapilli and subordinate volcanic flows (basalts, andesites,
dacites; Basta et al., 2011; Stern et al., 2013& Fig. 1)@ T$% .%6s$%7(s85 8AK-(+(%* 6A +$% B65s8-(s >6s;*
point to an arc or back arc setting (Ali et al., 2009)@ εNd (750 Ma) of these rocks is invariably positive,
and ranges between +5 and +9 (Ali et al., 2009)& *$62(-. +$%(> ›/B%-(5% -8+/>%@
The 30 m thick BIF (Figs 1(C) and 3(A)) 6A +$% œ/s/>/+/ H6>78+(6- IF>8S(5)& (* 85*6 8**6s(8+%' 2(+$
mafic metavolcanics and metabasalts, but it is overlain by carbonates characterized by negative δ13C
values (Nascimento et al., 2007; Sial et al., submitted)@ However, these carbonates do not exhibit the
typical features of cap carbonates (e.g., Hoffman and Schrag, 2002)@ Furthermore, the glacial features
are absent in the thinly banded BIF (Sial et al., submittedL Fig. 2(F)–(G))@ Diamictites with faceted
clasts occur elsewhere in the Seridó Belt, but is not associated with the BIF (Legrand et al., 2008)@ Van
Schmus et al. (2003) =>6=6*%'& 6- +$% 08*(* 6A '%+>(+85 S(>s6- 8.%*& +$% -8+/>% 6A +$% 08*%7%-+ 8-'
lithostratigraphy, that the Jucurutu Formation was deposited in a rift basin as the result of late Neopro-
terozoic extension of a pre-existing continental basement, with the development of small marine basins
(Fig. 4)@ This may explain the differences, such as the occurrence of carbonates, compared to BIFs
deposited in an arc environment, such as the Egyptian and Saudi Arabian BIF described above.
These examples serve to illustrate that Algoma-type BIFs occur in Neoproterozoic successions, and
that they were both associated to compressional and extensional geotectonic settings, as also known for
older BIFs (Gross, 1980)@ As we will discuss in the following sections, these Algoma-type Neoprotero-
zoic BIFs bear distinct geochemical and chemostratigraphic features that allow a distinction from the
glacially influenced iron formations.
ž CHAPTER 17 CHEMOSTRATIGRAPHY OF NEOPROTEROZOIC BIF

FIGURE 4
Ÿqkfgcdcfmi„ ufnq„g ofh deq dehqq dlkqg fo rqfkhfdqhftfcp chfm ofhuidcfmg ncgp|ggqn cm decg  fhv wxy Ÿqkfgc-
tion of Rapitan type banded iron formation (BIF) during glacial retreat, exemplified by the type occurrence in
the Rapitan Group (Modified from Baldwin et al. (2012)}~. BSR: bacterial sulfate reduction. (B) Depositional
model for Algoma type BIF of the Jucurutu Formation (NE Brazil) according to Sial et al. (submitted)v wˆy
Depositional environment of Lake Superior type BIF of the Yerbal Formation, Arroyo del Soldado Group (From
Frei et al. (2013)¡ ¢£¤ ¥¦§¦¥¦£¨¦© ª«¦¥¦¬£}~.
17.3 DEPOSITIONAL ENVIRONMENT OF NEOPROTEROZOIC BIFS ­­®

17.3.1.3 BIF Associated with Platform Deposits but without Glacial Features
T$%*% are possibly the most neglected Neoproterozoic iron formations, despite being remarkably thick
and laterally persistent. They share three characteristics: (1) they do not show glacial features or associ-
ated glacial deposits, (2) they are not associated with volcanic rocks, and (3) they occur in passive
margin successions characterized by thick carbonate and shale deposits.
A well-documented example is the Yerbal Formation of the Arroyo del Soldado Group (Uruguay;
Fig. 3(B))@ Up to 50-m-thick, oxide-facies BIF, composed of alternating magnetite + hematite and chert
bands (Fig. 2(D)–(E))& occurs at the top of the unit, just beneath the contact with the overlying carbon-
ates of the Polanco Formation (Gaucher et al., 2004; Frei et al., 2013& Fig. 3(B))@ BIFs are interbedded
with siltstones, cherts, and dolostones, all of which show considerable lateral continuity (Gaucher,
2000; Gaucher et al., 2004)@ No volcanic or pyroclastic rocks occur in the whole of the 5-km-thick
Arroyo del Soldado Group. Detrital zircon ages of the Yerbal Formation are mostly Paleoproterozoic,
with subordinate Archean and Mesoproterozoic populations (Gaucher et al., 2008L Blanco et al., 2009)@
Thus, both facies association and provenance show that the geotectonic setting of the Yerbal Formation
was a stable platform developed at the margin of an old craton (Río de la Plata Craton, Gaucher et al.,
2003, 2008)@ T$% *87% *%++(-. (* (-A%>>%' A6> +$% ?6/-.%> FGH* 6A +$% Y%>>6 N*=/%5(+8* H6>78+(6-
(Gaucher et al., 1996; Gaucher, 2000)@ T$%? 8>% =>%'67(-8-+5? 78.-%+(+(s 8-' 8>% 8**6s(8+%' 2(+$
thick, gray and black shales and chert. A likely metamorphic (amphibolite facies) equivalent of the
Arroyo del Soldado BIF is represented by the Marco de los Reyes Formation, which crops out at the
eastern edge of the Río de la Plata Craton (Chiglino et al., 2010)@ T$(* /-(+ $6*+* “V m thick oxide-facies
(magnetite) BIF having a mean composition of 64% Fe2O3 and 33% SiO2 (Bossi and Navarro, 1991)@
This BIF is associated with cherts, carbonates, and metapelites. 87Sr/86Sr values of the high-Sr lime-
stones range between 0.7070 and 0.7080 and δ13C between and −3‰ and 4.4‰ VPDB (Chiglino et al.,
2010)& 78+s$(-. s6>>%*=6-'(-. B85/%* 6A +$% Q658-s6 H6>78+(6- IGaucher et al., 2004; Frei et al., 2011)@
A shelf setting has also been proposed for the Tonian BIF of the Shilu Group (Hainan Province,
South China). Four BIF horizons are interbedded with dolostones, pelites, and quartz-arenites (Fig.
1(D)) 8-' +$% 2$65% =8s;8.% $8* 0%%- 7%+876>=$6*%' (- +$% /==%> .>%%-*s$(*+ A8s(%* IXu et al.,
2013a,b)@ Thickness of the BIF and intervening strata ranges between 50 and 400 m. No evidence of
contemporaneous glaciation or volcanism occurs in the Shilu District which, together with the facies
association, favor an assignment to the Lake Superior type (Xu et al., 2013b)@ A similar depositional
setting was proposed also for BIFs of the Neoproterozoic Jingtieshan Formation (Gansu Province,
North China) on the basis of their facies association (Sun et al., 1998)@
Thus, despite being much smaller than their older counterparts, Lake Superior-type BIFs do occur
in the Tonian and Ediacaran. This clearly shows that BIF deposition was not dependent on glacial con-
ditions, a fact that can already be inferred from the ferruginous nature of the Neoproterozoic ocean
(Canfield et al., 2008)@

17.3.2 GEOCHEMISTRY AND CHEMOSTRATIGRAPHY


17.3.2.1 Rapitan-Type Neoproterozoic BIF
M8>% %8>+$ %5%7%-+* IMNN) 8>% /*%' +6 (-A%> =85%69>%'6a s6-'(+(6-* 8-' +6 %5/s('8+% +$% =>(78>? *6/>s%
of iron in the BIFs (hydrothermal vs. weathering, e.g., Kato et al., 2006)@ MNN +Y data reported by Klein
(2005), Halverson et al. (2011) 8-' Baldwin et al. (2012) A6> +$% .58s(855? (-–/%-s%' M8=(+8- (>6- A6>-
mation (Canada) display a typical, LREE-depleted pattern and the absence of Eu or Ce anomalies. For
¯¯° CHAPTER 17 CHEMOSTRATIGRAPHY OF NEOPROTEROZOIC BIF

+$% œ8s8'(.6 X>6/=& Klein and Ladeira (2004) 85*6 >%=6>+ +$% 80*%-s% 6A N/ 6> Y% 8-6785(%*@ T$(* */.-
gests that the primary iron source in the glacially influenced BIFs may have been glacially sourced iron
oxyhydroxide nanoparticles that were subsequently bacterially reduced under ferruginous conditions,
supplemented with distal hydrothermal sources from the deep ocean (Baldwin et al., 2012)@ G- +$% =8>-
ticular case of the Jacadigo Group, REE patterns, unlike the Rapitan BIFs, are LREE-enriched, resem-
bling modern estuarine or coastal waters (Graf et al., 1994)& 78?0% '/% +6 +$% (-–/a 6A 7%5+28+%>@
Other trace elements used in Rapitan-type BIF chemostratigraphy include Mo, U, and W. These
elements, when normalized against the Mud from Queensland (Kamber et al., 2005) %a$(0(+ *%s/58>
variations in the Rapitan BIF, with enrichments at the base and top of the unit and low values in most
of the BIF interval (Baldwin et al., 2012)@ This stratigraphic distribution indicates reducing conditions
throughout the unit, with sulfidic (euxinic) conditions at the base and top of the iron formation, as indi-
cated especially by Mo (Baldwin et al., 2012)@ These conclusions were corroborated more recently by
Mo isotopes (δ98Mo, Baldwin et al., 2013)@ Similar low values of U and Mo were found by Døssing
et al. (2010) +$>6/.$6/+ +$% .58s(855? (-–/%-s%' FGH 6A +$% œ8s8'(.6 X>6/= IFig. 3(C))& with significant
U and Mo enrichment only at the base (Córrego das Pedras Formation) but not at the top. Thus, it
becomes clear that conditions during the deposition of these examples were characterized by ferrugi-
nous bottom waters and low sulfate availability.
Iron isotopes are obvious candidates for BIF chemostratigraphy, but the interpretation of isotopic
signals is often complex due to the involvement of many processes, biotic and abiotic, that fractionate
iron. Halverson et al. (2011) >%=6>+%' 8- (-s>%8*(-. +>%-' 6A δ57Fe, from −0.7‰ to 1.2‰ IRMM-14,
for BIF of the Sayunei Formation (Rapitan Group), which they interpreted as the reflection of rising
sea-level and an isotopic gradient across the chemocline. Halverson et al. (2007) >%=6>+%' 8 *(7(58>
δ57Fe curve for the glacially influenced Holowilena Ironstone, but with significantly larger amplitude
(−0.8‰ to 3.1‰ IRMM-14).
Chromium isotopes have become a useful tool for the study of BIF, carbonates, and other chemo-
genic rocks, essentially reflecting: (1) the degree of oxidative weathering on the continents (and thus
atmospheric oxygenation), and (2) the relative input of oxidized, land-derived, fractionated Cr (VI)
versus hydrothermal, unfractionated Cr (III) from vents (Frei et al., 2009)@ ±$%>%8*& +$% 65'%> 3>s$%8-
BIFs show mostly unfractionated, negative δ53Cr values around −0.15‰ SRM 979, the Neoarchean and
Paleoproterozoic BIFs exhibit positive values up to +0.5‰ SRM 979, depicting the early atmospheric
oxygenation (Frei et al., 2009)@ Y>9(*6+6=% '8+8 A6> •%6=>6+%>6S6(s /-(+* (-s5/'% +$% M8=(+8- FGH& 2$(s$
yielded positive values between 0.90‰ and 0.96‰ (Frei et al., 2009)@ G- 8- 8++%7=+ +6 *+/'? +$% *%s/58>
variations of δ53Cr, Døssing et al. (2010) 8==5(%' +$(* 7%+$6' +6 +$% œ8s8'(.6 X>6/=@ FGH 6A +$% F8-'8
Alta/Santa Cruz Formation yielded positive δ53Cr values rising from 0.1‰ at the base to a peak of 0.9‰
in the upper part (Fig. 3(C))& *5(.$+5? '%s>%8*(-. +6 s@ V@U‰ at the top (Døssing et al., 2010)@ T$% H%9>(s$
glaciogenic diamictites and associated BIF of the correlative Puga Formation also yielded positive δ53Cr
values from +0.1‰ to +0.4‰. Interesting in this regard is the parallelism of δ53Cr and δ13C found in the
Neoproterozoic carbonates (Frei et al., 2011)& 2$(s$ 85562 +$% /*% 6A δ53Cr in BIF as a proxy of δ13C of
contemporaneous seawater. The coupling of both isotope systems is probably through photosynthetic
activity, that affected both carbon burial and oxygen production (Frei et al., 2013)@

17.3.2.2 Algoma-Type Neoproterozoic BIF


MNN +Y patterns were presented by Basta et al. (2011) 8-' Stern et al. (2013) A6> 35.6789+?=% FGH* 6A
Egypt and Saudi Arabia. They are LREE-depleted (normalized to the Post-Archean Australian Shale,
PAAS), exhibit positive Eu anomalies (Eu/Eu*) of up to 1.41 (mean 1.13), and true negative Ce
17.3 DEPOSITIONAL ENVIRONMENT OF NEOPROTEROZOIC BIFS ²²³

8-6785(%* (Ce/Ce*) down to 0.7 (mean 0.9). The latter characteristics are in strong contrast with Rap-
itan-type BIF that do not show Eu or Ce anomalies. When plotted against the stratigraphy, Eu/Eu* and
Ce/Ce*, allow the identification of the intervals of elevated hydrothermal input (Stern et al., 2013)@
Significantly, these intervals are also marked by higher εNd(t) values, which reach +6 in the Sawawin
BIF of Saudi Arabia (Stern et al., 2013)@ Similar, albeit more extreme, REE + Y patterns were found by
Sial et al. (submitted) A6> 35.6789+?=% FGH* 6A +$% œ/s/>/+/ H6>78+(6- I<%>('´ F%5+& F>8S(5)@ T$%? 8>%
LREE-depleted and characterized by strongly positive Eu/Eu* anomalies (up to 3.1) and true positive
Ce/Ce* anomalies up to 1.3 (Fig. 3(A)L Bau and Dulski, 1996)@ Strong hydrothermal input related to
rifting might have promoted the anoxic (ferruginous) conditions, as shown by positive Ce anomalies.
The proximity to hydrothermal vents is also shown by very high Cr contents of the Jucurutu BIFs
(85–300, eventually 400 ppm) and by their δ53Cr values that vary mostly between −0.30‰ and −0.12‰
SRM 979 (Sial et al., submittedL Fig. 3(A))& and within the field of magmatic values (Schoenberg et al.,
2008)@ A negligible detrital component is shown by the occurrences of low Sc and Zr concentrations in
the BIF (average Sc = 2.2 ppm, average Zr = 1.2 ppm). Less negative δ53Cr values tend to appear at the
base of the iron formation, becoming more negative up section (Fig. 3(A))@ Overlying carbonates are
characterized by equally negative to slightly fractionated δ53Cr (−0.19‰ to −0.05‰) and “normal” Cr
concentrations between 1 and 10 ppm (Fig. 3(A))@ Interestingly, corresponding δ13C values are very
negative, around −7‰ VPDB, in the lower 20 m of carbonates (sitting atop BIF), passing into positive
values up to 10‰ VPDB in the up section (Sial et al., submitted)@
An arc setting was proposed for Algoma-type BIFs in Egypt and Saudi Arabia on the basis of the type,
geochemical affinities and Nd isotopes of the accompanying volcanic rocks (Ali et al., 2009& Fig. 1)@ µ- +$%
other hand, the Jucurutu BIFs were probably deposited in a starved rift basin (Fig. 4(B))& 8* *$62- 0? +$%(>
low terrigenous contents, geochemistry and association with overlying carbonates (Sial et al., submitted)@
In his original definition, Gross (1980) %>%s+%' +$%35.678 +?=% +6 (-s5/'% FGH* +$8+ 2%>% '%=6*(+%' ¶856-.
volcanic arcs or rift zones”. Thus, both of these Algoma subtypes occur in the Neoproterozoic.

17.3.2.3 Lake Superior Type Neoproterozoic BIF


Frei et al. (2009, 2013) >%=6>+%' MNN + Y data for BIF of the Arroyo del Soldado Group (Fig. 3(B))@
BIFs of the Yerbal Formation are LREE-depleted and characterized by positive Eu/Eu* anomalies
(1.15–1.71) and negative Ce/Ce* anomalies ranging from 0.63 to 0.79 (data from Frei et al., 2009)@ Fe
speciation (FeHR/FeTot) values range from 0.5 to 0.7 for the BIF and associated cherts (Fig. 3(B)L Frei
et al., 2013) 8-' +$/* (-'(s8+% 8-6a(s& A%>>/.(-6/* s6-'(+(6-* ICanfield et al., 2008)@ However, interbed-
ded shales and carbonates yielded FeHR/FeTot values far below the threshold value of 0.38 (Canfield
et al., 2008; Frei et al., 2013& Fig. 3(B))& showing that predominantly oxygenated conditions alternated
with anoxic events, as do the negative Ce anomalies mentioned above.
The εNd(t) values of the entire Yerbal Formation range between −19 and −5 (Fig. 3(B))& with Tdm
values between 2.4 and 2.0 Ga that match with the source rocks of the Río de la Plata Craton (Frei et al.,
2013)@ Understandably, the lowest εNd(t) values (−5 to −12) correspond to the BIFs and cherts, because
they are pure chemical precipitates with little siliciclastic input. These values confirm the inference of
stable platform setting for the Arroyo del Soldado Group (Fig. 4(C))& deduced from the facies associa-
tion and detrital zircon U–Pb ages (see above).
The δ53Cr values for BIF of the Arroyo del Soldado Group are strongly fractionated (Fig. 3(B))& ?(%5'-
ing values between 0.7‰ and 5.0‰ SRM 979 (Frei et al., 2009, 2013)@ T$% 58++%> 8>% +$% 76*+ =6*(+(B% δ53Cr
values reported for BIFs up to date (Frei et al., 2009)& 2$(s$ (* 56.(s85 s6-*('%>(-. +$% ?6/-. 8.% 6A +$%*%
iron formations and the progressive oxygenation of the surface environments leading up to the
··· CHAPTER 17 CHEMOSTRATIGRAPHY OF NEOPROTEROZOIC BIF

N'(8s8>8-PY870>(8- 06/-'8>?@ R6>%6B%>& (- M8=(+8- +?=% FGH& >8=(' '%-/'8+(6- 8-' s65' s5(78+% (- +$%
hinterland allows for less intense chemical weathering and could explain the lower δ53Cr compared to the
Neoproterozoic Lake Superior type (see Fig. 4)& 2$(s$ 2%>% '%=6*(+%' 6- '%%=5? 2%8+$%>%'& 65' s>8+6-*@
The Lake Superior-type BIF of the Shilu Group (China) are also characterized by LREE-depletion
(PAAS-normalized), and positive Eu/Eu* (up to 11) and Ce/Ce* (1.0–2.4) anomalies (Xu et al., 2013b)@
Corresponding εNd(t) values for the BIF range between −7 and −5 (average = −6) and Tdm 1.9 to
2.1 Ga (Xu et al., 2013b)@ Thus, the Shilu BIFs also require an old cratonic region as source area, and
represent a marine shelf setting, in line with the facies association observed. The only difference with
the BIF of the Arroyo del Soldado Group is the more anoxic setting for the Chinese occurrence, which
is also significantly older (Tonian).
In summary, it could be stated that the REE geochemical characteristics (e.g., Eu and Ce anomalies)
of the Chinese and Uruguayan Superior type Neoproterozoic BIFs are intermediate between the Rapitan
type and the Algoma type Neoproterozoic BIF. On the other hand, they show the most positive δ53Cr
values due to a more open marine setting of these iron formations (Fig. 4) 8-' +$% ?6/-.%> 8.% 6A +$%
Yerbal Formation.

17.4 DISCUSSION
G+ 0%s67%* s5%8> A>67 +$% '8+8 >%B(%2%' (- +$% =>%s%'(-. *%s+(6-* +$8+ .58s(8+(6- 28* -6+ 8 =>%9>%¸/(*(+% A6>
deposition of BIFs in the Neoproterozoic, as initially thought (e.g., Beukes and Klein, 1992)@ Y6-+>8>? +6
the recent assumptions (Macdonald et al., 2010aL Cox et al., 2013) +$% 8.% '(*+>(0/+(6- 6A •%6=>6+%>6S6(s
iron formations actually span the whole Era, from the Tonian to the late Ediacaran, and are not restricted to
the Sturtian glaciation or even the Cryogenian. While it is true that many Neoproterozoic BIFs, including
some of the largest, were deposited in glacially influenced settings (e.g., Jacadigo Group, Rapitan Group;
Fig. 4(A))& 78-? %a87=5%* 0%8> -6 (-'(s8+(6-* 6A 8 .58s(85 %-B(>6-7%-+ IGaucher et al., 2004; Frei et al.,
2013; Stern et al., 2013; Xu et al., 2013b; Sial et al., submitted)@ G-*+%8'& +$% =>%B85%-s% 6A >(A+(-. 8-'
hydrothermal activity were important for the development of both glacial and nonglacial BIFs (Fig. 4)&
possibly due to the protracted fragmentation of Rodinia and the associated mantle superplumes (Li et al.,
2008)@ Isley and Abbott (1999) 85>%8'? =6*+/58+%' 8 5(-; 0%+2%%- */=%>=5/7% %B%-+* 8-' FGH '%=6*(+(6- (-
the late Archean and Paleoproterozoic. Ferruginous conditions were likely fueled by enhanced mid-ocean
ridge hydrothermal circulation (Canfield et al., 2008)@ F6+$ +$% (-s>%8*%' 5%-.+$ 6A 7('96s%8- >('.%* 8-'
enhanced heat flux during Rodinia breakup might have promoted the elevated hydrothermal activity (e.g.,
Gaucher et al., 2009)& 8* (* *$62- 0? %a+>%7%5? 562 87Sr/86Sr seawater ratios in the Tonian and Cryogenian
(Melezhik et al., 2001; Halverson et al., 2007)@ T$% >%9%7%>.%-s% 6A FGH (- +$% •%6=>6+%>6S6(s 8A+%> -%8>5?
1 Gyr of nondeposition has more to do with the return to ferruginous conditions in the oceans than with
near-global glaciation. Primary productivity must also be taken into account (Fig. 4)& 0%s8/*% (+ 28* +$%
source of the large amounts of oxygen required to precipitate Neoproterozoic BIFs (e.g., Gaucher et al.,
2004; Frei et al., 2013)@ G- 8 08*(- s6B%>%' 0? +$(s; (s%& '(**65B%' 6a?.%- 26/5' 0% >8=('5? %a$8/*+%' '/%
to the inhibition of photosynthesis. Thus, in glacial settings, most models require deposition of BIFs after
ice retreat (e.g., Baldwin et al., 2012& Fig. 4(A))@ NB%- (- .58s(855? (-–/%-s%' (>6- A6>78+(6-*& +$%>% 2%>%
no continued ice cover, and sea ice-free conditions are recorded at some levels (Le Heron et al., 2011)@
Finally, the progressive oxygen buildup in surface environments determine evolving geochemical and iso-
topic signatures for the Neoproterozoic BIFs (e.g., Frei et al., 2009)& 2$(s$& (- +$% A/+/>% 78? 0% /*%' 8* 8
chemostratigraphic tool for correlation and geochronology.
REFERENCES ¹¹º

17.5 CONCLUSIONS
•%6=>6+%>6S6(s FGH* 6ss/> (- T6-(8-& Y>?6.%-(8-& 8-' N'(8s8>8- */ss%**(6-*& 8-' 8>% -6+ >%*+>(s+%'
to the middle Cryogenian (c. 715 Ma) glaciation. Many Neoproterozoic BIFs (North and South
America, southern Africa, and Australia) were deposited in glacially influenced settings, and are
traditionally assigned to the Rapitan type. Other occurrences in South America, Africa, and Asia,
however, bear no indications of glacial processes and can be assigned to the Algoma and Lake Supe-
rior types.
Chemostratigraphy is a powerful tool for unraveling the depositional setting of Neoproterozoic
BIFs. Some of the most useful methods are REE distribution, especially the Eu and Ce normalized
concentrations, iron speciation, Nd and Cr isotopes (δ53Cr). Whereas Rapitan type BIFs exhibit no Eu
or Ce anomalies, Algoma type Neoproterozoic BIF show both. Positive δ53Cr values characterize gla-
cially influenced BIFs, and differentiates them from nonfractionated, mantle-like δ53Cr values of
Algoma type BIFs. The latter are also characterized by high Cr concentrations. The most positive δ53Cr
values occur in the open-shelf, Lake Superior type BIFs, especially those of Ediacaran age. Given the
consistent secular variations exhibited by δ53Cr, it is envisaged that it may be used for correlation and
geochronology once the available database grows sufficiently.
It is demonstrated clearly that glaciation was not the main mechanism responsible for deposition of
BIFs in the Neoproterozoic. Hydrothermal activity related to the rifting of Rodinia probably played the
key role in their reappearance after the Mesoproterozoic gap, and promoted a return to ferruginous
conditions after the demise of Mesoproterozoic euxinic oceans.

ACKNOWLEDGMENTS
»¼½¾¿À ÁÂà ÁľÅÆ¾Ç ½¾Å ÀÄÈÈÂÃÉ ½ÃÊ ÅÄÊ ÉÂ É¼Ê ËÂÌÆÀÆ; ÎÊÏÉÂÃƽРÅÊ Ñ¾vestigación Científica (CSIC, Uruguay) to
C.G. Financial support through The Danish Agency for Science, Technology and Innovation grant nr. 11-103378 to
RF and through the Danish National Research Foundation’s center of excellence NordCEE (DNRF grant number
DNRF53) is highly appreciated. This study was partially supported by grants to ANS (CNPq, grants n. 470399/2008
and 472842/2010-2 and FACEPE APQ 0727-1.07/08; APQ-1059-9.05/12). This is the NEG-LABISE scientific
contribution n. 270. We thank the constructive suggestions of the anonymous reviewer.

REFERENCES
ÒÐÆÓ ÔÕÒÕÓ ÎÉÊÃ¾Ó ÖÕ×ÕÓ Ø½¾ÉÂ¾Ó ÙÕÑÕÓ ÔÆÌÄÃ½Ó ×ÕÚÑÕÓ Ô¼½ÌÊÊÀÓ ÛÕÒÕÓ ÜÝÝÞÕ ßÊÂϼÊÌÆÀÉÃà áÅ ÆÀÂÉÂÈÊÀ ½¾Å
U–Pb SHRIMP zircon dating of Neoproterozoic volcanic rocks from the Central Eastern Desert of Egypt: new
insights into the ∼750 Ma crust-forming event. Precambrian Research 171, 1–22.
Babinski, M., Boggiani, P.C., Trindade, R.I.F., Fanning, C.M., 2013. Detrital zircon ages and geochronological
constraints on the Neoproterozoic Puga diamictites and associated BIFs in the southern Paraguay Belt, Brazil.
Gondwana Research 23, 988–997.
Baldwin, G.J., Turner, E.C., Kamber, B.S., 2012. A new depositional model for glaciogenic Neoproterozoic
iron formation: insights from the chemostratigraphy and basin configuration of the Rapitan iron formation.
Canadian Journal of Earth Sciences 49, 455–476.
Baldwin, G.J., Nägler, T.F., Greber, N.D., Turner, E.C., Kamber, B.S., 2013. Mo isotopic composition of the mid-
Neoproterozoic ocean: an iron formation perspective. Precambrian Research 230, 168–178.
ââã CHAPTER 17 CHEMOSTRATIGRAPHY OF NEOPROTEROZOIC BIF

ä½ÀÉ½Ó F.F., Maurice, A.E., Fontboté, L., Favarger, P.-Y., 2011. Petrology and geochemistry of the banded iron
formation (BIF) of Wadi Karim and Um Anab, Eastern Desert, Egypt: Implications for the origin of Neopro-
terozoic BIF. Precambrian Research 187, 277–292.
Bau, M., Dulski, 1996. Distribution of yttrium and rare earth elements in the Penge and Kuruman Iron Formation,
Transval Supergroup, South Africa. Precambrian Research 79, 37–55.
Blanco, G., Rajesh, H.M., Gaucher, C., Germs, G.J.B., Chemale Jr., F., 2009. Provenance of the Arroyo del Sol-
dado Group (Ediacaran to Cambrian, Uruguay): Implications for the paleogeographic evolution of southwest-
ern Gondwana. Precambrian Research 171, 57–73.
Beukes, N.J., Klein, C., 1992. Models for iron-formation deposition. In: Schopf, J.W., Klein, C. (Eds.), The Pro-
terozoic Biosphere—A Multidisciplinary Study. Cambridge University Press, Cambridge, pp. 147–151.
Boggiani, P.C., Ferreira, V.P., Sial, A.N., Babinski, M., Trindade, R.I.F., Aceñolaza, G., Toselli, A.J., Parada, M.A.,
2003. The cap carbonate of the Puga Hill (Central South America) in the context of the post-Varanger Galcia-
tion. In: IV South American Symposium on Isotope Geology, Short Papers, Salvador, pp. 324–327.
Boggiani, P.C., Gaucher, C., Sial, A.N., Babinski, M., Simon, C.M., Riccomini, C., Ferreira, V.P., Fairchild, T.R.,
2010. Chemostratigraphy of the Tamengo Formation (Corumbá Group, Brazil): a contribution to the calibra-
tion of the Ediacaran carbon-isotope curve. Precambrian Research 182, 382–401.
Bossi, J., Navarro, R., 1991. Geología del Uruguay. Universidad de la República, Montevideo. pp. 1–970.
Bowring, S.A., Myrow, P., Landing, E., Ramezani, J., Grotzinger, J.P., 2003. Geochronological constraints on
terminal Neoproterozoic events and the rise of metazoans. In: European Geophysical Union Annual Meeting,
Nice 2003. Geophysical Research Abstracts, vol. 5, p. 13219.
Canfield, D.E., Poulton, S.W., Knoll, A.H., Narbonne, G.M., Ross, G., Goldberg, T., Strauss, H., 2008. Ferrugi-
nous conditions dominated later Neoproterozoic deep-water chemistry. Science 321, 949–952.
Chiglino, L., Gaucher, C., Sial, A.N., Bossi, J., Ferreira, V.P., Pimentel, M.M., 2010. Chemostratigraphy of Meso-
proterozoic and Neoproterozoic carbonates of the Nico Pérez Terrane, Río de la Plata Craton, Uruguay. Pre-
cambrian Research 182, 313–336.
Cox, G.M., Halverson, G.P., Minarik, W.G., Le Heron, D.P., Macdonald, F.A., Bellefroid, E.J., Strauss, J.V., 2013.
Neoproterozoic iron formation: An evaluation of its temporal, environmental and tectonic significance. Chemi-
cal Geology 362, 232–249.
Døssing, L.N., Gaucher, C., Boggiani, P.C., Frei, R., 2010. Stable Chromium Isotopes as Tracer of Changes in
Weathering Processes and Redox State of the Ocean during Neoproterozoic Glaciations. American Geophysi-
cal Union, Fall Meeting, San Francisco.
Eyles, N., Januszczak, N., 2004. “Zipper-rift”: a tectonic model for Neoproterozoic glaciations during the breakup
of Rodinia after 750 Ma. Earth-Science Reviews 65, 1–73.
Fanning, C.M., Link, P.K., 2008. Age constraints for the sturtian glaciation: data from the Adelaide Geosyncline,
south Australia and Pocatello formation, Idaho, USA. In: Gallagher, S.J., Wallace, M.W. (Eds.), Neoprotero-
zoic Extreme Climates and the Origin of Early Life, Selwyn Symposium of the GSA Victoria DivisionGeologi-
cal Society of Australia Extended Abstracts, vol. 91, pp. 57–62.
Fölling, P.G., Frimmel, H.E., 2002. Chemostratigraphic correlation of carbonate successions in the Gariep and
Saldania Belts, Namibia and South Africa. Basin Research 14, 69–88.
Frei, R., Gaucher, C., Poulton, S.W., Canfield, D.E., 2009. Fluctuations in Precambrain atmospheric oxygenation
recorded by chromium isotopes. Nature 461, 250–253.
Frei, R., Gaucher, C., Døssing, L.N., Sial, A.N., 2011. Chromium isotopes in carbonates—a tracer for climate change
and for reconstructing the redox state of ancient seawater. Earth and Planetary Science Letters 312, 114–125.
Frei, R., Gaucher, C., Stolper, D., Canfield, D.E., 2013. Fluctuations in late Neoproterozoic atmospheric oxida-
tion—Cr isotope chemostratigraphy and iron speciation of the late Ediacaran lower Arroyo del Soldado Group
(Uruguay). Gondwana Research 23, 797–811.
Freitas, B.T., Warren, L.V., Boggiani, P.C., De Almeida, R.P., Piacentini, T., 2011. Tectono-sedimentary evolution
of the Neoproterozoic BIF-bearing Jacadigo Group, SW-Brazil. Sedimentary Geology 238, 48–70.
REFERENCES ååæ

çÃÆÌÌÊÐÓ H.E., 2008. The gariep Belt. In: Miller, R.M. (Ed.), The Geology of Namibia. Neoproterozoic to Lower
Palaeozoic, vol. 2. Geological Survey of Namibia, Windhoek, pp. 14-1–14-39.
Gaucher, C., 2000. Sedimentology, palaeontology and stratigraphy of the Arroyo del Soldado Group (Vendian to
Cambrian, Uruguay). Beringeria 26, 1–120.
Gaucher, C., Sprechmann, P., Schipilov, A., 1996. Upper and Middle Proterozoic fossiliferous sedimentary
sequences of the Nico Pérez Terrane of Uruguay: Litho-stratigraphic units, paleontology, depositional envi-
ronments and correlations. Neues Jahrbuch für Geologie und Paläontologie, Abhandlungen 199, 339–367.
Gaucher, C., Sprechmann, P., 1999. Upper Vendian skeletal fauna of the Arroyo del Soldado Group, Uruguay.
Beringeria 23, 55–91.
Gaucher, C., Boggiani, P.C., Sprechmann, P., Sial, A.N., Fairchild, T., 2003. Integrated correlation of the Vendian
to Cambrian Arroyo del Soldado and Corumbá Groups (Uruguay and Brazil): palaeogeographic, palaeocli-
matic and palaeobiologic implications. Precambrian Research 120, 241–278.
Gaucher, C., Sial, A.N., Blanco, G., Sprechmann, P., 2004. Chemostratigraphy of the lower Arroyo del Soldado
Group (Vendian, Uruguay) and palaeoclimatic implications. Gondwana Research 7, 715–730.
Gaucher, C., Frimmel, H.E., Germs, G.J.B., 2005. Organic-walled microfossils and biostratigraphy of the upper
Port Nolloth Group (Namibia): implications for the latest Neoproterozoic glaciations. Geological Magazine
142, 539–559.
Gaucher, C., Finney, S.C., Poiré, D.G., Valencia, V.A., Grove, M., Blanco, G., Pamoukaghlián, K., Gómez Peral,
L., 2008. Detrital zircon ages of Neoproterozoic sedimentary successions in Uruguay and Argentina: insights
into the geological evolution of the Río de la Plata Craton. Precambrian Research 167, 150–170.
Gaucher, C., Poiré, D., 2009. Biostratigraphy. Neoproterozoic-Cambrian evolution of the Río de la Plata Pal-
aeocontinent. In: Gaucher, C., Sial, A.N., Halverson, G.P., Frimmel, H.E. (Eds.), Neoproterozoic-cambrian
Tectonics, Global Change and Evolution: A Focus on Southwestern Gondwana. Developments in Precambrian
Geology, vol. 16. Elsevier, pp. 103–114.
Gaucher, C., Frimmel, H.E., Germs, G.J.B., 2009. Tectonic events and palaeogeographic evolution of southwest-
ern Gondwana in the Neoproterozoic and Cambrian. In: Gaucher, C., Sial, A.N., Halverson, G.P., Frimmel,
H.E. (Eds.), Neoproterozoic-cambrian Tectonics, Global Change and Evolution: A Focus on Southwestern
Gondwana. Developments in Precambrian Geology, vol. 16. Elsevier, pp. 295–316.
Graf, J.L., Oconnor, E.A., Vanleeuwen, P., 1994. Rare earth element evidence of origin and depositional environ-
ment of Late Proterozoic ironstone beds and manganese-oxide deposits, SW Brazil and SE Bolivia. Journal of
South American Earth Sciences 7, 115–133.
Gross, G.A., 1980. A classification of iron formations based on depositional environments. Canadian Mineralogist
18, 215–222.
Halverson, G.P., Poitrasson, F., Hoffman, P.F., Nédélec, A., Montel, J.-M., 2007. Iron isotope composition of Neo-
proterozoic syn-glacial iron-formation: implications for sources and cycling of iron in anoxic oceans. In: 7th
International Symposium on Applied Isotope Geochemistry, Stellenbosch, pp. 63–64.
Halverson, G.P., Wade, B.P., Hurtgen, M.T., Barovich, K.M., 2010. Neoproterozoic chemostratigraphy. Precam-
brian Research 182, 337–350.
Halverson, G.P., Poitrasson, F., Hoffman, P.F., Nédélec, A., Montel, J.-M., Kirby, J., 2011. Fe isotope and trace
element geochemistry of the Neoproterozoic syn-glacial Rapitan iron formation. Earth and Planetary Science
Letters 309, 100–112.
Hoffman, P.F., Hawkins, D.P., Isachsen, C.E., Bowring, S.A., 1996. Precise U–Pb zircon ages for early Damaran
magmatism in the Summas Mountains and Welwitschia inlier, northern Damara belt, Namibia. Communica-
tions of the Geological Survey of Namibia 11, 47–52.
Hoffman, P.F., Kaufman, A.J., Halverson, G.P., Schrag, D.P., 1998. A Neoproterozoic snowball Earth. Science
281, 1342–1346.
Hoffman, P., Schrag, D.P., 2002. The snowball Earth hypothesis: testing the limits of global change. Terra Nova
14, 129–155.
èèé CHAPTER 17 CHEMOSTRATIGRAPHY OF NEOPROTEROZOIC BIF

ÛÂÁfman, P.F., Halverson, G.P., 2008. The Otavi Group of the Northern Platform and the Northern Margin Zone.
In: Miller, R.M. (Ed.), The Geology of Namibia. Neoproterozoic to Lower Palaeozoic, vol. 2. Geological Sur-
vey of Namibia, Windhoek, pp. 13-69–13-136.
Hoffmann, K.H., Condon, D.J., Bowring, S.A., Crowley, J.L., 2004. U–Pb zircon date from the Neoproterozoic
Ghaub Formation, Namibia: constraints on Marinoan glaciation. Geology 32, 817–820.
Hollanda, M.H., Archanjo, C.J., Souza, L.C., 2010. Historia de provenieência do Grupo Seridó (Província Bor-
borema), com base em dados isotópicos. In: Congresso Brasileiro de Geologia, Belém, (CD-ROM).
Ilyin, A.V., 2009. Neoproterozoic banded iron formations. Lithology and Mineral Resources 44, 78–86.
Isley, A.E., Abbott, D.H., 1999. Plume-related mafic volcanism and the deposition of banded iron formation. Jour-
nal of Geophysical Research 104, 15461–15477.
Jacobsen, S.B., Kaufman, A.J., 1999. The Sr, C and O isotopic evolution of Neoproterozoic seawater. Chemical
Geology 161, 37–57.
James, H.L., 1954. Sedimentary facies of iron-formation. Economic Geology 49, 235–293.
Johnston, D.T., Poulton, S.W., Dehler, C., Porter, S., Husson, J., Canfield, D.E., Knoll, A.H., 2010. An emerging
picture of Neoproterozoic ocean chemistry: insights from the Chuar Group, Grand Canyon, USA. Earth and
Planetary Science Letters 290, 64–73.
Kamber, B.S., Greig, A., Collerson, K.D., 2005. A new estimate for the composition of weathered young upper
continental crust from alluvial sediments, Queensland, Australia. Geochimica et Cosmochimica Acta 69,
1041–1058.
Kato, Y., Yamaguchi, K.E., Ohmoto, H., 2006. Rare earth elements in Precambrian banded iron formation: secular
changes of Ce and Eu anomalies and evolution of atmosphere oxygen. Geological Society of America, Memoir
198, 269–280.
Kaufman, A.J., Hayes, J.M., Knoll, A.H., Germs, G.J.B., 1991. Isotopic compositions of carbonates and organic
carbon from Upper Proterozoic successions in Namibia: stratigraphic variation and the effects of diagenesis
and metamorphism. Precambrian Research 49, 301–327.
Klein, C., 2005. Some Precambrian banded iron-formation (BIFs) from around the world: their age, geologic set-
ting, mineralogy, metamorphism, geochemistry, and origin. American Mineralogist 90, 1473–1499.
Klein, C., Ladeira, E.A., 2004. Geochemistry and mineralogy of Neoproterozoic of banded iron formations and
some selected, siliceous manganese formations from the Urucum district, Mato Grosso do Sul, Brazil. Eco-
nomic Geology 99, 1233–1244.
Knoll, A.H., Hayes, J.M., Kaufman, A.J., Swett, K., Lambert, I.B., 1986. Secular variation in carbon isotope ratios
from Upper Proterozoic successions of Svalbard and east Greenland. Nature 321, 832–837.
Legrand, M., Martins de Sá, J., Santos, L., 2008. Tilito Neoproterozoico intercalado nas meta-supracrustais da
Faixa Seridó, Província Borborema, Nordeste do Brasil. In: 44° Congresso Brasileiro de Geologia (CD-ROM).
Le Heron, D.P., Cox, G., Trundley, A., Collins, A., 2011. Sea ice-free conditions during the Sturtian glaciation
(early Cryogenian), South Australia. Geology 39, 31–34.
Li, Z.X., Bogdanova, S.V., Collins, A.S., Davidson, A., De Waele, B., Ernst, R.E., Fitzsimons, I.C.W., Fuck, R.A.,
Gladkochub, D.P., Jacobs, J., Karlstrom, K.E., Lu, S., Natapov, L.M., Pease, V., Pisarevsky, S.A., Thrane, K.,
Vernikovsky, V., 2008. Assembly, configuration, and break-up history of Rodinia: a synthesis. Precambrian
Research 160, 179–210.
Macdonald, F.A., Strauss, J.V., Rose, C.V., Dudas, F.O., Schrag, D.P., 2010a. Stratigraphy of the Port Nolloth
Group of Namibia and South Africa and implications for the age of Neoproterozoic iron formations. American
Journal of Science 310, 862–888.
Macdonald, F.A., Schmitz, M.D., Crowley, J.L., Roots, C.F., Jones, D.S., Maloof, A.C., Strauss, J.V., Cohen, P.A.,
Johnston, D.T., Schrag, D.P., 2010b. Calibrating the Cryogenian. Science 327, 1241–1243.
McGee, B., Collins, A.S., Trindade, R.I.F., 2012. G’day Gondwana—the final accretion of a supercontinent: U–Pb
ages from the post-orogenic São Vicente Granite, northern Paraguay Belt, Brazil. Precambrian Research 21,
316–322.
REFERENCES êêë

ØÊÐÊì¼Æ¿Ó V.A., Gorokov, I.M., Kuznetsov, A.B., Fallick, A.E., 2001. Chemostratigraphy of Neoproterozoic car-
bonates: implications for “blind dating”. Terra Nova 13, 1–11.
Miller, R.M., 2008. Swakop Group. In: Miller, R.M. (Ed.), The Geology of Namibia. Neoproterozoic to Lower
Palaeozoic, vol. 2. Geological Survey of Namibia, Windhoek, pp. 13-136–13-178.
Möller, J.C., 2004. A geologia neoproterozóica da região de Corumbá, os recursos minerais de ferro e informações
gerais sobre a MCR, Guia da Mineração Corumbaense Reunida. Corumbá 1–25.
Nascimento, R.S.C., Sial, A.N., Pimentel, M.M., 2007. C- and Sr-isotope systematics applied to Neoproterozoic
marbles of the Seridó belt, northeastern Brazil. Chemical Geology 237, 209–228.
Piacentini, T., Boggiani, P.C., Kazuo Yamamoto, J., Freitas, B.T., Campanha, G.A.C., 2007. Formação ferrífera
associada à sedimentação glaciogênica da Formação Puga (Marinoano) na Serra da Bodoquena. MS. Revista
Brasileira de Geociências 37, 530–541.
Piacentini, T., Vasconcelos, P.M., Farley, K.A., 2013. 40Ar/39Ar constraints on the age and thermal history of the
Urucum Neoproterozoic banded iron-formation, Brazil. Precambrian Research 228, 48–62.
Schoenberg, R., Zink, S., Staubwasser, M., von Blanckenburg, F., 2008. The stable Cr isotope inventory of solid
Earth reservoirs determined by double spike MC-ICP-MS. Chemical Geology 249, 294–306.
Sial, A.N., Campos, M.S., Gaucher, C., Frei, R., Ferreira, V.P., Nascimento, R.C., Pimentel, M.M., Pereira, N.S.,
Rodler, A., (submitted). Algoma-type Neoproterozoic BIFs and related marbles in the Seridó Belt (NE Brazil):
REE, C, O, Cr and Sr isotope evidence. Journal of South American Earth Sciences.
Stern, R.J., Mukherjee, S.K., Miller, N.R., Ali, K., Johnson, P.R., 2013. ∼750 Ma banded iron formation from the
Arabian-Nubian Shield—implications for understanding Neoproterozoic tectonics, volcanism, and climate
change. Precambrian Research 239, 79–94.
Sun, H., Wu, J., Yu, P., Li, J., 1998. Geology, geochemistry and sulfur isotope composition of the Late Proterozoic
Jingtieshan (Superior-type) hematite–jasper–barite iron ore deposits associated with stratabound Cu mineral-
ization in the Gansu Province, China. Mineralium Deposita 34, 102–112.
Trompette, R., De Alvarenga, C.J.S., Walde, D., 1998. Geological evolution of the Neoproterozoic Corumba gra-
ben system (Brazil). Depositional context of the stratified Fe and Mn ores of the Jacadigo group. Journal of
South American Earth Science 11, 587–597.
Van Schmus, W.R., Neves, B.B.B., Williams, I.S., Hackspacher, P.C., Fetter, A.H., Dantas, E.L., Babinski, M.,
2003. The Seridó Group of NE Brazil, a late Neoproterozoic pre-to syn-collisional basin in West Gondwana:
insights from SHRIMP U–Pb detritical zircon ages and Sm–Nd crustal residence (TDM) ages. Precambrian
Research 127, 287–327.
Xu, D.R., Wang, Z.L., Cai, J.H., Wu, C.J., Bakun-Czubarow, N., Wang, L., Chen, H., Baker, M., Kusiak, A., 2013a.
Geological characteristics and metallogenesis of the Shilu Fe-ore deposit in Hainan Province, South China.
Ore Geology Reviews 53, 318–342.
Xu, D.R., Wang, Z.L., Chen, H.Y., Hollings, P., Jansen, N.H., Zhang, Z.C., Wu, C.J., 2013b. Petrography and
geochemistry of the Shilu Fe–Co–Cu ore district, South China: Implications for the origin of a Neoproterozoic
BIF system. Ore Geology Reviews 57, 322–350.
Young, G.M., 2002. Stratigraphic and tectonic settings of Proterozoic glaciogenic rocks and banded iron-
formations: relevance to the snowball Earth debate. Journal of African Earth Sciences 35, 451–466.
This page intentionally left blank

View publication stats

You might also like