You are on page 1of 43

5

Introduction
FOUR REVIEW PAPERS have recently been published that elo-
quently summarize the current knowledge of the main com-
positional attributes of iron formations and their distribution
in space and time (Trendall, 2002; Simonson, 2003; Clout and
Simonson, 2005; Klein, 2005). The authors of these reviews
also highlighted the considerable questions that center on the
genesis and paleoenvironmental significance of this enigmatic
rock type that is unique to the Precambrian Era. Rather than
to duplicate their effort, we focus on a critical review of
advances made in the understanding of the origin and pale-
oenvironmental significance of the most voluminous and, ar-
guably, best preserved iron formations in Earths history.
These are the exceptionally well preserved iron formations of
the Ghaap-Chuniespoort Groups of the Transvaal Super-
group on the Kaapvaal craton in southern Africa (i.e., Beukes,
1983), and the time-equivalent Hamersley Group of the
Mount Bruce Supergroup on the Pilbara craton in Western
Australia (Fig. 1; Trendall, 2002; Pickard, 2003). Understand-
ing the paleoenvironmental and paleogeographic setting of
these giant iron formations and their associated strata is of
particular significance, as geochemical characteristics of these
Chapter 1
Origin and Paleoenvironmental Significance of Major Iron Formations
at the Archean-Paleoproterozoic Boundary
NICOLAS J. BEUKES

AND JENS GUTZMER


Paleoproterozoic Mineralization Research Group, Department of Geology, University of Johannesburg, P.O. Box 524,
Auckland Park 2006, South Africa
Abstract
This paper provides a critical review of advances made in understanding of sedimentary environments, geo-
chemical processes, and biological systems that contributed to the deposition and diagenetic evolution of the
exceptionally well-preserved and large iron formations of the late Neoarchean to very early Paleoproterozoic
Ghaap-Chuniespoort Group of the Transvaal Supergroup on the Kaapvaal craton (South Africa) and the time
equivalent Hamersley Group on the Pilbara craton (Western Australia). These iron formations are commonly
assumed to have formed coevally but in separate basins, and they are often used as proxies for global ocean
chemistry and paleoenvironmental conditions at ~2.5 Ga. However, lithostratigraphic and paleogeographic re-
constructions show that the iron formations formed in a single large partly enclosed oceanic basin along the
margins of the ancient continent of Vaalbara. Furthermore, although large relative to other preserved iron for-
mations, the combined Transvaal-Hamersley basin is miniscule compared to marginal basins of the modern
ocean system so that the succession probably documents secular changes in depositional environments of that
basin rather than of the global ocean at the time.
The iron formations comprise a large variety of textural and mineralogical rock types that display complex
lateral and vertical facies variations on basinal scale. Based on detailed analyses of these variations it is con-
cluded that the iron formations were deposited in environments that ranged from very deep-water basinal set-
tings far below storm-wave base and the photic zone to very shallow-platform settings above normal wave base.
Precipitation of both iron and silica took place from hydrothermal plumes in a dynamically circulating ocean
system that was not permanently stratified. Ferric oxyhydroxide was the primary iron precipitate in virtually all
of the iron formation facies. This primary precipitate is now represented by early diagenetic hematite in some
of the iron formations. However, in both deep- and shallow-water iron formations most of the original ferric
oxyhydroxides have been transformed by dissimilatory iron reduction to early diagenetic siderite and/or mag-
netite in the presence of organic carbon. Precipitation of ferric oxyhydroxides in very deep water below the
photic zone required a downward flux of photosynthetically-derived free oxygen from the shallow photic zone.
In these deep-water environments, under microaerobic conditions, chemolithoautotrophic iron-oxidizing bac-
teria may have played an important role in precipitation of ferric oxyhydroxides and acted as a source of pri-
mary organic matter. With basin fill even shallow-shelf embayments were invaded by circulating hydrothermal
plume water from which ferric oxyhydroxides could be precipitated in oxygenated environments with high pri-
mary organic carbon productivity and thus iron reduction to form hematite-poor siderite- and magnetite-rich
clastic-textured iron formations.
Depositional models derived from the study of the iron formations along the Neoarchean-Proterozoic
boundary can be applied to iron formations of all ages in both the Archean and later Paleoproterozoic. The fa-
cies architecture of the iron formations determines to a large degree the textural attributes, composition, and
stratigraphic setting of high-grade iron ores hosted by them. Detailed facies information thus would assist in
improving genetic models for high-grade iron ore deposits. Future research should be guided in this direction,
especially in some of the very large iron ore districts of Brazil and India where very little is known about the
composition and facies variations of the primary iron formation hosts and possible controls on localization of
high-grade ores.

Corresponding author: e-mail, nbeukes@uj.ac.za
2008 Society of Economic Geologists
SEG Reviews vol. 15, p. 547
Beukes_Gutzmer 6/11/08 7:40 AM Page 5
6 BEUKES AND GUTZMER
0361-0128/98/000/000-00 $6.00 6
DAMPIER
ARCHIPELAGO
INDIAN OCEAN
PORT HEDLAND
S
y
n
c
lin
e
Lookout
Rocks
Brockman
Syncline
Turner
Syncline
Milli-Milli
Dome
Rocklea Dome
0 50km
120E
Wyloo Dome
A
S
H
B
U
R
T
O
N
Hardy
Syncline
F
O
L
D
PARABURDOO
Cape
Lambert
120E
23S
Hamersley Group
Fortescue Group
21S
Cape
Preston
Deepdale
23S
117E
B
E
LT
Phanerozoic cover
Proterozoic rocks younger than
Mount Bruce Supergroup
Turee Creek Group
Greenstone and Granitoids of
the Pilbara Craton
Carawine
area
WITTENOOM
Koongaling Hill
PILBARA
BLOCK
WHIM CREEK
Hamersley
R
a
n
g
e
S
y
n
c
lin
o
riu
m
OPHTHALMIA
RANGE
Sylvania Dome
MARBLE BAR
O
a
k
o
v
e
r
NULLAGINE
C
H
IC
H
E
S
T
E
R
117E
ZIMBABWE
CRATON
L
IM
P
O
P
O
M
E
T
A
M
O
R
P
H
IC
P
R
O
V
IN
C
E
CHUNIESPOORT
GROUP
GHAAP
GROUP
T
R
A
N
S
V
A
A
L
S
U
P
E
R
G
R
O
U
P
Archean granite-greenstone
terrain and volcano-sedimentary
sequences of the Kaapvaal Craton
W
aterberg
M
etam
orphic
episodes
at 2700
and
2000M
a
M
O
Z
A
M
B
I
Q
U
E
M
E
T
A
M
O
R
P
H
I
C
P
R
O
V
I
N
C
E
L
a
s
t
m
e
t
a
m
o
r
p
h
i
c
e
p
i
s
o
d
e
5
5
0
M
a
Maputo
0 100km
Durban
N
26
3
0

Last m
etam
orphic
episode
1030
M
a
NATAL METAMORPHIC PROVINCE
Griquatown
Prieska
N
A
M
A
Q
U
A
30
2
2

Soutpansberg
O
l
i
f
a
n
t
h
o
e
k
Kuruman
GRIQUALAND
WEST
TRANSVAAL
Pretoria
TRANSVAAL
Waterberg and
Soutpansberg Group
GRIQUALAND WEST
Postmasburg Group
Olifantshoek Group
(Fold belt)
2
2

22
Koegas
Subgroup
Asbesheuwels
Subgroup
Campbellrand
Subgroup
Schmidtsdrif
Subgroup
Wolkberg Group
Black Reef
Formation
Duitschland
Formation
Penge
Iron-formation
Malmani
Subgroup
Pretoria Group
Bushveld
Complex 2058 Ma
2
6

3
0

Zeerust
30
22
f
f
f
2
6

26
K
h
e
i
s
T
e
r
r
a
i
n
(
D
e
f
o
r
m
a
t
i
o
n
~
1
1
1
0
M
a
)
A
KAAPVAAL CRATON
Red Beds
Pomfret
Penge
R
A
N
G
E
Balfour Downs
Turee Creek
Sycline
Sishen
FIG. 1. Simplified geologic maps indicating the distribution of the Transvaal Supergroup on the Kaapvaal craton in south-
ern Africa (A) (modified after Beukes, 1983) and the Hamersley Group on the Pilbara craton in Western Australia (B) (mod-
ified after Trendall, 1983). Note that the two maps are presented on approximately the same scale, illustrating the much
larger area of preservation of the Transvaal Supergroup relative to that of the Hamersley Group.
Beukes_Gutzmer 6/11/08 7:40 AM Page 6
successions are often used as proxies for global ocean chem-
istry at the Archean-Paleoproterozoic boundary (e.g., Anbar
et al., 2007; Kaufman et al., 2007). The reliability of these
proxies is dependent entirely on the assumption that the two
successions were deposited in two contemporaneous but spa-
tially separate open marine sedimentary basins. However, this
may not have been the case, as some authors (e.g., Cheney,
1996) have provided stratigraphic evidence for deposition of
both iron formation-bearing successions in one single sedi-
mentary basin, thus requiring a much more careful differen-
tiation of global and local basinal depositional and paleoenvi-
ronmental signatures.
The aim of this paper is to constrain depositional environ-
ments, geochemical processes, and biological systems that
contributed to the deposition and diagenetic evolution of iron
formations at ca. 2.5 Ga (i.e., the Archean-Proterozoic bound-
ary). A modern sequence stratigraphic approach is taken, fo-
cusing on depositional system facies tracts that were in place
during deposition of the iron formations. We shall use termi-
nology as defined by Van Wagoner et al. (1988) for the se-
quence analyses. Important implications of sequence strati-
graphic analyses for the understanding of the nature of the
Paleoproterozoic-Archean ocean (Isley, 1995; Simonson and
Hassler, 1996; Isley and Abbott, 1999), and the role of micro-
bial organisms in the precipitation of iron (Klein and Beukes,
1989; Konhauser et al., 2002) are illustrated.
Within the context of this volume, it is hoped that an im-
proved understanding of the stratigraphic and lithologic ar-
chitecture of these giant iron formations will provide new im-
pulses for exploration for high-grade iron ore deposits. Both
the Transvaal and the Hamersley Iron Formations are host to
a number of economically important high-grade iron ore de-
posits (see reviews by Gutzmer et al., 2005; Alchin et al.,
2008; Thorne et al., 2008) and have a long history of explo-
ration. Compositional differences, in particular the abun-
dance of carbonates and iron oxides versus that of iron sili-
cates and chert, in different iron formation lithofacies may be
of critical importance to determining the suitability of partic-
ular iron formation units to host epigenetic high-grade iron
orebodies (e.g., Gutzmer et al., 2005). Indeed, it is through
the detailed study of abundant exploration diamond drill core
that the lithostratigraphy and lateral and vertical facies varia-
tions in these two giant iron formation successions are ar-
guably the best established of all iron formations in the world
(Trendall and Blockley, 1970; Beukes, 1983, 1984; Trendall,
1983, 2002; Klein and Beukes, 1992a; Klein, 2005).
Mineralogical and Textural Classification
of Iron Formations
Two broad textural types of iron formation are widely rec-
ognized (Trendall, 2002; Simonson, 2003; Clout and Si-
monson, 2005), namely banded iron formation (BIF) that
was deposited as chemical muds and granular iron forma-
tion deposited as endoclastic sands. However, this classifi-
cation is too broad for the purpose of a detailed analysis of
depositional facies and systems in Archean-Paleoprotero-
zoic iron formations. Furthermore, the use of the term BIF
for iron formations that were deposited as chemical muds is
confusing as it is at odds with the commonly accepted de-
finition of the term banded iron formation (Jackson, 1997,
p. 51). The latter definition restricts the term banded iron
formation to iron formations with distinct chert banding, irre-
spective of textural and/or mineralogical composition. This in-
consistency is further illustrated in Klein (2005), who ex-
tended the definition of the term banded iron formation to be
synonymous to the term iron formation, irrespective of the
presence and/or absence of chert bands and texture, i.e.,
muddy versus granular. There is thus a clear need to stan-
dardize the nomenclature. In this paper, we propose and
apply a revised classification of iron formation that acknowl-
edges the classical definitions of iron formation and banded
iron formation and subdivides these two rock types further
according to textural and mineralogical attributes.
Following the Glossary of Geology (Jackson, 1997, p. 335),
iron formation (IF) is defined as a finely-laminated to thin
bedded chert-bearing chemical sedimentary rock containing
at least 15% iron of sedimentary origin. Banded iron forma-
tion (BIF), on the other hand, is a variety of iron formation
that contains distinct chert bands (Jackson, 1997, p. 51), i.e.,
mesobands of Trendall and Blockley (1970). Subdivision of
iron formations further takes place according to textural fea-
tures that are in many respects similar to those of carbonate
rocks (Dimroth and Chauvel, 1973; Dimroth, 1976). Based
on mineralogical and textural attributes Beukes (1980a, 1983,
1984) developed a comprehensive classification scheme that
covers the full range of iron formation rock types and lithofa-
cies. However, that classification scheme specifically applies
to detailed sedimentological analyses of iron formations; for
this paper a simplified classification will suffice. This simpli-
fied classification is based on the three basic textural com-
ponenets of iron formations, namely, allochem particles
(granules), matrix (chemical iron-rich muds, i.e., femicrite),
and microcrystalline quartz (chert; Fig. 2A). The latter often
represents the cement component in iron formations (Fig.
2A; Simonson, 1987). Iron formations composed mainly of
granules are referred to as granular iron formations (GIF).
Granular iron formation encompasses both conglomeratic
ferudite and sandy fearenite (Fig. 2B). Fearenite can in turn
be subdivided into grainstone, packstone, or wackestone, de-
pending on sorting and the amount of femicrite matrix pre-
sent (Fig. 3), following definitions of Dunham (1962) for cal-
carenite. It is recognized that virtually all of the granules are
endoclastic in origin (Simonson, 1985) and derived from re-
working of earlier lithified iron formation components, and
that chemically precipitated grains, like oolites, are in fact
scarce (Simonson, 2003; Clout and Simonson, 2005). Granu-
lar iron formations that are distinctly chert mesobanded could
be described as banded granular iron formations.
In contrast to granular iron formations, iron formations
dominantly composed of femicrite (chemical mud) are clas-
sified as micritic iron formation (MIF) and when distinctly
chert-mesobanded they are referred to as banded micritic
iron formation (BMIF) (Fig. 2B). Femicrite in micritic iron
formation displays three distinct types of bedding, namely
(1) microlamination or microbanding (on a submillimeter
scale), (2) lamination on a millimeter to centimeter scale,
and (3) thin to thick, poorly defined bedding with rather
massive appearance. Beukes (1980a, 1983, 1984) applied the
term ferhythmite to encompass microbanded and well-lami-
nated femicrite; in contrast, the term felutite is applied to
MAJOR IRON FORMATIONS AT THE ARCHEAN-PALEOPROTEROZOIC BOUNDARY 7
0361-0128/98/000/000-00 $6.00 7
Beukes_Gutzmer 6/11/08 7:40 AM Page 7
poorly bedded femicrite with a muddy massive appearance.
Based on these parameters micritic iron formations can be
broadly classified into microbanded, laminated, and lutitic
varieties (Fig. 2A, B). In laminated banded micritic iron for-
mation and micritic iron formation, individual laminae may
have a massive or graded appearance and in some cases be
separated from each other by micritic microbands. This tex-
tural variability may reflect the difference between current-
reworked femicrite (massive or graded laminae) versus set-
tling of fine suspended mud in microbands. Felutite beds are
often associated with granular iron formations and may have
been reworked by currents as indicated by the presence of
fine femicrite intraclasts (Beukes, 1980a, 1983, 1984). There
is thus a whole range of textural types of micritic iron forma-
tion, ranging from microbanded ones that represent chemi-
cal precipitates (orthochemical iron formations) through
mixtures between chemical precipitates and reworked femi-
crite in laminated iron formation (best described as ferhyth-
mite) to reworked femicrite in felutite, i.e., allochemical mi-
critic iron formation (Fig. 2B). In sedimentological context
granular iron formation and lutitic micritic iron formation
can be grouped as clastic-textured iron formations, in con-
trast to the orthochemical microbanded micritic iron forma-
tion (Fig. 2B).
8 BEUKES AND GUTZMER
0361-0128/98/000/000-00 $6.00 8
Micropelloids (silt-sized)
Fe - Oxide
Hematite and Magnetite
50%
15%
Siderite
Oxide - Siderite Siderite - Oxide
S
i
l
i
c
a
t
e
-
O
x
i
d
e
O
x
i
d
e
-
S
i
l
i
c
a
t
e
S
i
l
i
c
a
t
e
-
S
i
d
e
r
i
t
e
S
i
d
e
r
i
t
e
-
S
i
l
i
c
a
t
e
Mixed
silicate
carbonate
oxide
FEMICRITE
Felutite
Ferhythmite
Microbanded
Laminated
Mineralogical Classification
Fe - silicate Greenalite, Minnesotaite
Stilpnomelane
Podded
Wavy banded
Even banded
CHERT
Chert Cement
Chert Bands
Massive
Microbanded
ALLOCHEMS
Gravel-Size Fragments
Disc
Chip
Pisoliths
Grit - Sand - Silt-sized Particles
Ooid
Peloid (Grit and sand - sized)
Intraclast
Shard
> 2cm
4 mm - 2 cm
I
R
O
N
F
O
R
M
A
T
I
O
N
(
I
F
)
W
I
T
H
C
H
E
R
T
B
A
N
D
S
=
B
A
N
D
E
D
I
R
O
N
F
O
R
M
A
T
I
O
N
(
B
I
F
)
F
e
m
i
c
r
i
t
e
A
l
l
o
c
h
e
m
s
M
I
F
o
r
B
M
I
F
G
I
F
Ferudite
Fearenite
Felutite
Ferhythmite
Laminated
Microbanded
Massive poorly bedded
Grainstone
Packstone
Wackestone
Flat -pebble conglomerate
C
l
a
s
t
i
c
-
t
e
x
t
u
r
e
d
(
A
l
l
o
c
h
e
m
i
c
a
l
)
C
r
y
s
t
a
l
l
i
n
e
(
O
r
t
h
o
c
h
e
m
i
c
a
l
)
G
I
F
Ruditic IF or BIF
Grainstone, packstone or
wackestone IF or BIF
Lutitic MIF / BMIF
Laminated MIF / BMIF
Microbanded MIF / BMIF
Notes:
1) Description according to form of chert bands: Podded BIF, Wavy BIF, Even BIF
2) Laminated and microbanded MIF or BMIF can be grouped as rhythmitic MIF or BMIF
3) Mineralogical composition described by prefix according to mineralogical classification above
CLASSIFICATION SYSTEM
Edgewise conglomerate
free rolling
Concretionary or
COMPONENTS
A
B
FIG. 2. Classification and nomenclature of iron formation based on the nature of its major components (allochem parti-
cles, femicrite, and chert) (modified after Beukes, 1980a, 1983). See text for explanation of acronyms.
Beukes_Gutzmer 6/11/08 7:40 AM Page 8
Chert mesobands are the third important component of
iron formations. The chert mesobands can have either mas-
sive or microbanded internal appearance (Fig. 2A). Mi-
crobanded chert mesobands are typically, but not exclusively,
associated with microbanded femicrite and massive ones with
felutite. The bedding character of chert mesobands can vary
from even, wavy to podded forms (Fig. 2A). This geometry
can be incorporated in the description of BIF lithofacies
through terms such as podded, wavy, or even BIF (Fig. 2B).
Chert pods have a flat lenticular form in cross section and are
disks in three dimensions. They vary from millimeter to
decimeter in diameter and thickness.
Jaspilite is a term applied only to BIF composed of alter-
nating mesobands of jasper (i.e., chert with minor, finely dis-
seminated microcrystalline dusty hematite) and hematite
magnetite (Beukes, 1980a). It should be noted that, according
to our own observations, apart from being a primary precipi-
tate, jaspilite in iron formation successions could also be a
product of supergene or hydrothermal alteration.
Iron formations and BIFs can be further classified according
to their mineralogical composition. Well-preserved, unmeta-
morphosed iron formations, like those of the Transvaal and
Hamersley basins, are characterized by three groups of very
fine crystalline iron-rich minerals, namely the oxides hematite
and magnetite, the carbonates siderite and ankerite, and the
iron silicates greenalite and stilpnomelane with rare chamosite
(Klein, 2005). These diagenetic to low-grade metamorphic
mineral assemblages define three end-member mineralogical
facies of iron formations, i.e., oxide, carbonate, and silicate
with several mixed facies (Fig. 2A). James (1954) recognized
a sulfide facies characterized by the presence of either pyrite
or pyrrhotite. Because of its high siliciclastic content, it is ap-
propriate to describe the latter as pyritic black shale, with or
without interbeds of black carbonaceous chert, and not as an
iron formation lithofacies. Indeed, it is our experience that
early diagenetic pyrite is conspicuously absent from iron for-
mations that have not been contaminated by an influx of exo-
genic detrital or pyroclastic material.
The effects of regional metamorphism on the mineralogy of
iron formations have recently been reviewed by Klein (2005).
Incipient regional metamorphism of iron formation leads to
recrystallization of components, increased grain size, and de-
velopment of such minerals as biotite from stilpnomelane,
minnesotaite from greenalite and grunerite, and cumming-
torite-grunerite from reactions between carbonates or iron
oxides with silica. High-grade metamorphic iron formations
contain minerals such as coarse magnetite, clino- and or-
thopyroxene, and fayalite. Magnetite, a very common compo-
nent of iron formation, is most abundant as a product of
metamorphic or hydrothermal alteration and less commonly
preserved as a product of early diagenesis. The origin of mag-
netite and its relationship to hematite, which is one of the ear-
liest diagenetic components of iron formation, can in most
cases only be established through very careful petrographic
studies (Han, 1978). Neither calcite nor dolomite is consid-
ered primary or early diagenetic constituents of iron forma-
tion. Rather, they are of metamorphic or hydrothermal origin
with specifically calcite a by-product in metamorphic forma-
tion of grunerite from ankerite and quartz. Both dolomite
and calcite are also found with hydrothermally altered iron
formation associated with high-grade iron ore deposits
(Beukes et al., 2002a; Clout and Simonson, 2005).
Iron formations uncontaminated by endogenetic siliciclas-
tic or pyroclastic material are depleted in virtually all of the
major elements except iron and silica. Iron contents typically
range between ~20 and 35 wt percent and SiO
2
between 40
and 55 wt percent (Klein, 2005). Iron formations are further
marked by very low Al
2
O
3
contents (usually <2 wt %), indi-
cating deposition as chemical precipitates in environments
starved of detrital input (Klein, 2005). Trace elements, in-
cluding REE, are highly depleted in well-preserved and un-
altered iron formations (e.g., Klein and Beukes, 1989; Bau
and Dulski, 1996).
Global Setting of Transvaal-Hamersely Iron Formations
The giant iron formations of the Transvaal and Hamerley
basins (Fig. 1) are prime examples of cratonic iron formations
(Lake Superior type of Gross, 1980). Such iron formations are
characterized by their generally undeformed nature, lateral
continuity, and association with cratonic margin shelf succes-
sions, including platform carbonates, carbonaceous shale, and
quartz arenites. The cratonic iron formations typically include
micritic and granular iron formation and are most abundant
in, but not exclusive to, the Paleoproterozoic (Fig. 3). They
stand in contrast to Algoma-type iron formations (Gross,
1980, 1983) that are typically rather intensely folded, mainly
composed of banded micritic iron formation and associated
with volcanic rocks and graywackes in greenstone belts. Al-
goma-type iron formations are most prominent in the
Archean but are also known in late Paleoproterozoic green-
stone belts (Fig. 3; James, 1983; Isley and Abbott, 1999; Hus-
ton and Logan, 2004). A third group of iron formation, here
referred to as Rapitan-type after iron formation of the Rapi-
tan Group (Klein and Beukes, 1993), is associated with glacial
diamictites. These are most characteristic of the Cryogenic
era of the Neoproterozoic period (Klein, 2005), but small de-
posits are also known in the early Paleoproterozoic (Beukes,
1983) and Mesoarchean (Smith, 2007).
The actual aerial distribution of iron formations of all types
is miniscule in context of the size of modern-day continents
and open-marine depositional systems (Fig. 4). As a group,
Lake Superior-type iron formations are very limited in both
numbers (Fig. 3B) and geographic distribution (Fig. 4). By far
the bulk of the iron formations in this group are located in
seven basins, namely the Mesoarchean Witwatersrand-
Mozaan basin (Kaapvaal craton) in South Africa (Beukes,
1973; Frimmel, 1996; Smith, 2007), the Late Archean to early
Paleoproterozoic Transvaal (Kaapvaal craton; e.g., Beukes,
1983) and Hamersley (Pilbara craton; e.g., Trendall, 1983)
basins, the Paleoproterozoic Krivoy Rog-Kursk basin (Ukrain-
ian shield) (Belevtsev et al., 1983), the Minas basin (Sao Fran-
cisco craton; Klein and Ladeira, 2000), and the late Paleopro-
terozoic Animikie-Marquette basin of North America (e.g.,
Gross and Zajac, 1983; Morey, 1983), and the Nabberu basin
(e.g., Goode et al., 1983) along the northern margin of the
Yilgarn craton in Australia (Fig. 4). Smaller occurrences of
Superior-type iron formations are preserved in late Paleo-
proterozoic strata of the Middleback Subgroup along the
eastern margin of the Gawler craton of Australia (James,
1983), the Ashburton Formation along the southern margin
MAJOR IRON FORMATIONS AT THE ARCHEAN-PALEOPROTEROZOIC BOUNDARY 9
0361-0128/98/000/000-00 $6.00 9
Beukes_Gutzmer 6/11/08 7:40 AM Page 9
of the Pilbara craton, and the Shushong Group (Beukes,
1973) along the northern margin of the Kaapvaal craton (Fig.
4). Only two occurrences of Mesoproterozoic Superior-type
iron formations are known, namely in the middle Mesopro-
terozoic Bushmanland Group of the Namaqua terrane of the
Kalahari craton ( Strydom et al., 1987; McClung, 2006) and
late Mesoproterozoic shelf carbonate succession of the Pen-
ganga Group (Gutzmer and Beukes, 1998a) of Peninsular
India (Figs. 3B, 4).
Although they are not to be discussed further in this paper,
it is worth noting that Algoma- and Rapitan-type iron forma-
tions are distinctly different from Lake Superior-type iron
formations in their abundance and distribution. Algoma-type
iron formations, for example, appear much more abundant
both in number (Fig. 3B) and geographic distribution (Fig.
4). However, much of this merely reflects the less stable tec-
tonic environment in which iron formations in greenstone
belts were deposited and the effects of intense postdeposi-
tional deformation. It does not imply that iron formations in
greenstone belts are thinner and were initially laterally less
extensive than those of the Lake Superior type. Iron forma-
tions in greenstone belts can be up to several hundred meters
thick (Fedo and Eriksson, 1996) and laterally extensive for
several hundred kilometers in folded greenstone belt succes-
sions (Klein, 2005). Indeed, many Algoma-type iron forma-
tions display attributes very similar to that of Lake Superior-
type iron formations (Eriksson, 1983; Eriksson et al., 1994).
Neoproterozoic Rapitan-type iron formation occurrences
are highly localized, relative to the very wide distribution of
glacial diamictites with which they are associated. By far the
largest and best preserved Rapitan-type iron formations
occur in the Rapitan Group of the Neoproterozoic inlier in
the McKenzie and Ogilvie Mountains in North America (e.g.,
Klein and Beukes, 1993) and the Urucum district (e.g., Klein
and Ladeira, 2004) along the western margin of the Amazon
craton in South America (Fig. 4). Smaller Neoproterozoic
iron formations are also known from the Damara-Gariep suc-
cession in Namibia and South Africa, the Flinders Range in
Australia (Lottermoser and Ashley, 2000), and the Maliya-
Khingun Proterozoic inlier (James, 1983) in far southeastern
Siberia (Fig. 4).
Stratigraphic Setting and Correlation between the
Neoarchean-Paleoproterozoic Successions of the
Transvaal and Hamersley Basins
Stratigraphic setting
In the Transvaal basin on the Kaapvaal craton the giant iron
formations along the Archean-Proterozoic boundary consti-
tute the Asbesheuwels Subgroup and correlative Penge Iron
Formation of the Ghaap and Chuniespoort Group, respec-
tively (Fig. 1A; Beukes, 1983, 1984). On the Pilbara craton
they are represented by the Marra Mamba, Brockman, Weeli
Wolli, and Boolgeeda Iron Formations of the Hamersley
10 BEUKES AND GUTZMER
0361-0128/98/000/000-00 $6.00 10
1.0 1.5 2.0 2.5 3.0 3.5 4.0
2
10
4
10
6
10
8
10
A
m
o
u
n
t
o
f
B
I
F
(
M
t
)
A
Age (Ga)
Cratonic Basins
(Superior-type) Greenstone Belts
(Algoma-type)
25
20
15
10
5
0
4.0 3.5 3.0 2.5 2.0 1.5 1.0
Archean
Cratonic
Archean
Greenstone
Belts
(Algoma-type)
Paleoproterozoic Cratonic
(Superior-type)
Paleoproterozoic
Greenstone Belts
(Algoma-type)
Mesoproterozoic
Cratonic
B
Age (Ga)
N
u
m
b
e
r
o
f
B
I
F
s
FIG. 3. Temporal distribution of iron formations according to amount (A) and numbers (B) in greenstone belt and cra-
tonic settings up to the end of the Mesoproterozoic (modified after Huston and Logan, 2004, with addition of Mesopro-
terozoic iron formations noted in text).
Beukes_Gutzmer 6/11/08 7:40 AM Page 10
MAJOR IRON FORMATIONS AT THE ARCHEAN-PALEOPROTEROZOIC BOUNDARY 11
0361-0128/98/000/000-00 $6.00 11
C
A
N
A
D
I
A
N
S
H
I
E
L
D
I
N
D
I
A
N
C
R
A
T
O
N
A
U
S
T
R
A
L
I
A
N
C
R
A
T
O
N
S
O
U
T
H
A
M
E
R
I
C
A
N
C
R
A
T
O
N
E
A
S
T
E
U
R
O
P
E
A
N
C
R
A
T
O
N
A
n
a
b
a
r
S
h
i
e
l
d
B
a
i
k
a
l
B
e
l
t
S
I
B
E
R
I
A
N
C
R
A
T
O
N
A
l
d
a
n
S
h
i
e
l
d
G
r
e
e
n
l
a
n
d
S
h
i
e
l
d
B
a
l
t
i
c
S
h
i
e
l
d
U
k
r
a
i
n
i
a
n
S
h
i
e
l
d
A
r
c
h
e
a
n
B
l
o
c
k
C
A
T
H
A
Y
S
I
A
N
C
R
A
T
O
N
S
i
n
o
-
K
o
r
e
a
n
C
r
a
t
o
n
N
a
b
b
e
r
u
B
a
s
t
a
r
C
r
a
t
o
n
S
i
n
g
h
b
u
m
C
r
a
t
o
n
P
i
l
b
a
r
a
B
l
o
c
k
Y
i
l
g
a
r
n
B
l
o
c
k
G
a
w
l
e
r
C
r
a
t
o
n
A
F
R
I
C
A
N
C
R
A
T
O
N
K
a
a
p
v
a
a
l
C
r
a
t
o
n
B
u
s
h
m
a
n
l
a
n
d
K
a
l
a
h
a
r
i
C
r
a
t
o
n
W
i
t
s
-
M
o
z
a
a
n
T
r
a
n
s
v
a
a
l
B
a
r
b
e
r
t
o
n
K
i
b
a
l
i
a
n
C
o
n
g
o

C
r
a
t
o
n
D
a
m
a
r
a
C
a
r
a
j
a
s
M
a
n
S
h
i
e
l
d
R
e
q
u
i
b
a
t
S
h
i
e
l
d
W
e
s
t
A
f
r
i
c
a
n
C
r
a
t
o
n
L
a
k
e
S
u
p
e
r
i
o
r
I
F
P
e
n
g
a
n
g
a
H
a
m
e
r
s
l
e
y
S
a
o
F
r
a
n
c
i
s
c
o
C
r
a
t
o
n
C
a
u

U
r
u
c
u
m
C
e
n
t
r
a
l
B
r
a
z
i
l
S
h
i
e
l
d
A
m
a
z
o
n
C
r
a
t
o
n
G
u
i
a
n
a
S
h
i
e
l
d
R
a
p
i
t
a
n
W
y
o
m
i
n
g
u
p
l
i
f
t
N
O
R
T
H
A
M
E
R
I
C
A
N
C
R
A
T
O
N
S
u
p
e
r
i
o
r
P
r
o
v
i
n
c
e
S
l
a
v
e
C
h
u
r
c
h
i
l
l
N
a
i
n
B
l
o
c
k
M
a
l
y
i
K
h
i
n
g
a
n
K
r
i
v
o
y

R
o
g
1
8
0
1
8
0
1
8
0
1
8
0
1
4
0
1
4
0
1
4
0
1
4
0
1
0
0
1
0
0
1
0
0
1
0
0
6
0
6
0
6
0
6
0
2
0
2
0
2
0
2
0
0 0
4
0
2
0
6
0
4
0
4
0
4
0
2
0
2
0
2
0
6
0
6
0
6
0
0
0
E
x
p
o
s
e
d
A
r
c
h
e
a
n

C
r
u
s
t
w
i
t
h

G
r
e
e
n
s
t
o
n
e
E
x
p
o
s
e
d
P
a
l
e
o
p
r
o
t
e
r
o
z
o
i
c
M
e
d
i
a
n

M
a
s
s
i
f
O
u
t
l
i
n
e

o
f

M
a
j
o
r
P
r
e
c
a
m
b
r
i
a
n

C
r
a
t
o
n
s
I
n
c
l
u
d
i
n
g

B
u
r
i
e
d

B
a
s
e
m
e
n
t
R
a
p
i
t
a
n
-
t
y
p
e

I
F
C
r
a
t
o
n
i
c

I
F
Z
i
m
b
a
b
w
e
C
r
a
t
o
n
T
a
n
z
a
n
i
a
C
r
a
t
o
n
D
h
a
r
w
a
r
C
r
a
t
o
n
A
N
T
A
R
C
T
I
C
C
R
A
T
O
N
Y
a
v
a
p
a
i
(
P
a
l
e
o
p
r
o
t
e
r
o
z
o
i
c

g
r
e
e
n
s
t
o
n
e
)
C
h
a
i
l
l
u
-
G
a
b
o
n
C
r
a
t
o
n
s
M
i
d
d
l
e
b
a
c
k
F
l
i
n
d
e
r
s
R
a
n
g
e
F
I
G
.
4
.


G
l
o
b
a
l

m
a
p

s
h
o
w
i
n
g

t
h
e

d
i
s
t
r
i
b
u
t
i
o
n

o
f

P
r
e
c
a
m
b
r
i
a
n

c
r
a
t
o
n
s

a
n
d

s
u
c
c
e
s
s
i
o
n
s

c
o
n
t
a
i
n
i
n
g

i
r
o
n

f
o
r
m
a
t
i
o
n
s
.

D
i
s
t
r
i
b
u
t
i
o
n

o
f

g
r
e
e
n
s
t
o
n
e

b
e
l
t
s

i
s

p
u
r
e
l
y

s
c
h
e
m
a
t
i
c
,
i
n
t
e
n
d
e
d

t
o

s
h
o
w

t
h
e
i
r

l
a
r
g
e

n
u
m
b
e
r
s

a
n
d

h
i
g
h
l
y

f
r
a
g
m
e
n
t
d

n
a
t
u
r
e
.

M
o
s
t

i
n
t
e
r
e
s
t
i
n
g

i
s

t
o

n
o
t
e

t
h
e

l
i
m
i
t
e
d

n
u
m
b
e
r

o
f

c
r
a
t
o
n
i
c

s
u
c
c
e
s
s
i
o
n
s

h
o
s
t
i
n
g

i
r
o
n

f
o
r
m
a
t
i
o
n
s

a
n
d
h
o
w

s
m
a
l
l

t
h
e

o
u
t
c
r
o
p

a
r
e
a
s

r
e
a
l
l
y

a
r
e

w
h
e
n

s
e
e
n

i
n

g
l
o
b
a
l

p
e
r
s
p
e
c
t
i
v
e
.

T
h
e
r
e

a
r
e

a
l
s
o

o
n
l
y

a

f
e
w

l
o
c
a
l
i
t
i
e
s

w
h
e
r
e

R
a
p
i
t
a
n
-
t
y
p
e

i
r
o
n

f
o
r
m
a
t
i
o
n

o
c
c
u
r
s

i
n

a
s
s
o
c
i
a
t
i
o
n

w
i
t
h
N
e
o
p
r
o
t
e
r
o
z
o
i
c

g
l
a
c
i
a
l

d
e
p
o
s
i
t
s

t
h
a
t

h
a
v
e

m
u
c
h

w
i
d
e
r

d
i
s
t
r
i
b
u
t
i
o
n

(
m
o
d
i
f
i
e
d

a
f
t
e
r

G
o
o
d
w
i
n
,

2
0
0
0
)
.
Beukes_Gutzmer 6/11/08 7:40 AM Page 11
Group (Fig. 1B; Trendall and Blockley, 1970; Trendall, 1983;
Trendall et al., 2004). The Asbesheuwels succession, com-
prised of the Kuruman and conformably overlying Griqua-
town Iron Formation, reaches a maximum known thickness of
950 m and together with the correlative Penge Iron Forma-
tion (Fig. 1A), covers an area of ~110,000 km
2
on the Kaap-
vaal craton (Beukes, 1983). The combined thickness of the
Brockman, Weeli Wolli, and Boolgeeda Iron Formations on
the Pilbara craton is ~1,050 m covering an area of ~60,000
km
2
(Trendall, 1983).
On both cratons the iron formations conformably overlie
thick late Neoarchean carbonate platform deposits known as
the Campbellrand-Malmani succession in the Tranvaal basin
and the Wittenoom Dolomite in the Hamersley basin (Fig.
1). However, it is important to note that the Asbesheuwels-
Penge succession comprises well-exposed iron formations
that extend from the basinal facies of the underlying Camp-
bellrand carbonate platform succession onto the shallow
shelf (Fig. 5). This basin to shelf transition provides excellent
opportunities for understanding the influence of paleoba-
thymetry on the nature of iron formation lithofacies (Beukes,
1983, 1984). This is not the case in the Hamersley basin of
Western Australia, where only the basinal facies are well ex-
posed and the transition from basin to shelf on the underly-
ing Wittenoom-Carawine carbonate platform succession is
either covered by younger rocks or not preserved by erosion
(Fig. 5), giving a depositional picture biased toward deeper
water lithofacies. The present contribution will first docu-
ment the excellent lithostratigraphic correlation of the suc-
cessions of the Transvaal and Hamersley basins. It will then
focus on the lithofacies architecture and depositional envi-
ronments of iron formations in the Transvaal basin only,
mainly because such data are currently not available for the
Hamersley succession on a basinwide scale.
Lithostratigraphic correlation
Button (1976a) was arguably the first to point out striking
lithostratigraphic similaritiesand a possible link between
the Transvaal and Hamersley basins. Cheney (1996) formal-
ized the concept that the two successions may have formed
part of one large basin along the trailing margin of an ancient
continent referred to as Vaalbara. Since these early contribu-
tions, abundant new data have become available that support
the concept in general but neccessitate revision of some of
Cheneys (1996) sequence stratigraphic correlations and his
paleogeographic reconstruction of the relative positions of
the Kaapvaal and Pilbara cratons.
The correlation between the two successions proposed
here (Fig. 6) is equally founded on stratigraphic and sedi-
mentological similarities aided by correlation of unconformi-
ties and event beds and first-hand field observations by the
senior author. The correlation was then validated by available
isotopic age data (Fig. 6). It is important to note that correla-
tion is restricted to the basinal facies of the Transvaal Super-
group (to the south of the Griquatown fault zone; Fig. 5) and
the basinal facies of the Hamersley Group. This is justified by
the poor preservation of the shallow-platform facies of the
Hamersley Group (Fig. 5). The lithostratigraphic correlation
diagam takes into account only strata from the base of the
Transvaal Supergroup to some stratigraphic distance above
the Koegas Subgroup (Fig. 6). Strata below and above these
beds may also correlate (Cheney, 1996), but this has little rel-
evance to this paper.
In the proposed correlation the basal quartzite of the Vry-
burg Formation of the Transvaal Supergroup, which discon-
formably overlies basaltic andesites of the Allanridge Forma-
tion of the Ventersdorp Supergroup, correlates with the
Woodiana sandstone that overlies basalts of the Maddina For-
mation and forms the base of the Jeerinah Formation in the
upper part of the Fortesque Group in Western Australia (Fig.
6). Conformably overlying the Vryburg Formation is a car-
bonaceous shale unit known as the Lokammona Formation
that correlates with the upper part of the Jeerinah Formation
(Fig. 6). To the north of Balfour Downs in the Carrawine area
(Fig. 1B), the upper part of the Jeerinah Formation is known
as the Roy Hill Shale; it contains carbonates with contorted
microbial mats similar to those of the lower part of the Mon-
teville Formation at the base of the Campbellrand Subgroup
(Fig. 6). An impact spherule layer (tectite) in the lower part
of the Monteville Formation (Simonson et al., 1999) is corre-
lated with a spherule layer in the Jeerinah Formation (Fig. 6;
Kohl et al., 2006).
In the Hamersley basin, the Jerrinah Formation, including
the Roy Hill Shale, is conformably overlain by the Marra
Mamba Iron Formation (Fig. 6). In the Transvaal basin there
is no such prominent iron formation developed but similari-
ties between carbonates in the Roy Hill Shale and that of the
lower part of the Monteville Formation indicate that the low-
ermost ankerite-banded chert unit in the Campbellrand suc-
cession may correlate with the Marra Mamba Iron Formation
12 BEUKES AND GUTZMER
0361-0128/98/000/000-00 $6.00 12
Younger unconformity
Carawine
Dolomite
Jeerinah
Marra Mamba
Wittenoom
Brockman -
Boolgeeda
Unconformity
Basin Shelf
WSW ENE
Pilbara Craton (Western Australia)
Basin Shelf
Lokammona
Monteville BIF
Nauga
Asbesheuwels
Koegas
Campbellrand
Dolomite
Unconformity SW NE
Kaapvaal Craton (South Africa)
Shale
BIF
Dolomite
Shale
Shale
BIF
Dolomite
BIF
Shale
Cover
FIG. 5. Schematic comparison between late Neoproterozoic to very early
Paleoproterozoic stratigraphic successions hosting giant iron formations on
the Kaapvaal and Pilbara cratons. Note that on the Pilbara craton only the
basinal facies of the Brockman-Boolgeeda Iron Formation succession is pre-
served, whereas on the Kaavaal craton both basinal and shelf facies of the As-
besheuwels succession is preserved.
Beukes_Gutzmer 6/11/08 7:40 AM Page 12
MAJOR IRON FORMATIONS AT THE ARCHEAN-PALEOPROTEROZOIC BOUNDARY 13
0361-0128/98/000/000-00 $6.00 13
Mn
S S
S S S
S
S S S
Klein Naute
Monteville
Lokammona
Vryburg
Allanridge
lava
Ventersdorp
Supergroup
Machadodorp Silverton
Daspoort
Ongeluk/Hekpoort
Makganyene
Timeball Hill
Upper
Lower
K
o
e
g
a
s
Griquatown
IF
K
u
r
u
m
a
n
I
F
Middelwater
Pannetjie
Doradale
Naragas
Rooinekke
Klipputs
Nelani
Duitschland
2440 - 2446
232215
222013
ca. 2440
24806
Orange
View
25213
25497
Tuff 2
Tectite
Tectite
26423
25975
26295
26846
Jeerinah
Marra Mamba
Maddina
Basalt
Woodiana SS
Roy Hill Sh
F
o
r
t
e
s
q
u
e
G
r
o
u
p
W
i
t
t
e
n
o
o
m
Main Tuff
25659 Tectite
Bruno s Band
Colonial Chert
25018
249516
Gorge
Dales
24814
Whaleback 24635
Mt McRae
Weeli
Wolli
Yandicoogina
24493
Woongarra
Lower
Boolgeeda
Upper
24455
24515
B
r
o
c
k
m
a
n
I
F
24615
T
u
r
e
e
C
r
e
e
k
G
r
o
u
p
Folded Sill
2208 Ma
Meteorite Bore
Ophthalmian
Orogeny
Post-Mooidraai
Folding
Cheela Springs
Beasley
River Quartzite
Numana
W
y
l
o
o
G
r
o
u
p
Kliphuis
Tectite
Stofbakkies
Buisvlei
2
2
1
1
1
1
243131
5
6
7
9
10
8
11
24646
24657
Geduld
Westerberg
4
8
1
1
1
1
T
T
24493
Stubensvallei
T
R
A
N
S
V
A
A
L
S
U
P
E
R
G
R
O
U
P
P
r
e
t
o
r
i
a
/
P
o
s
t
m
a
s
b
u
r
g
G
r
o
u
p
G
h
a
a
p
/
C
h
u
n
i
e
s
p
o
o
r
t
G
r
o
u
p
Crystal T uff
FORMATION Member/Bed Member/Bed FORMATION
Duitschland
Dwaalheuwel
Mooidraai
Hotazel
S4
H
a
m
e
r
s
l
e
y
G
r
o
u
p
Joffre
3
Eroded
Eroded
Nauga
7
5
A
s
b
e
s
h
e
u
w
e
l
s
C
a
m
p
b
e
l
l
r
a
n
d
M
O
U
N
T
B
R
U
C
E
S
U
P
E
R
G
R
O
U
P
Kazput
Koolbye
Kungara
Kungara
s s s s s s
Mt Sylvia
West Angela
1
1
Tuff 3
25886
12
7
7
FIG. 6. Correlation diagram between late Neoarchean to early Paleoproterozoic strata in the Transvaal and Hamersley
basins. Age references are as follows: 1 = Trendall et al. (2004), 2 = Barley et al. (1997), 3 = Mller et al. (2005), 4 = Anbar
et al. (2007), 5 = Pickard (2003), 6 = Altermann and Nelson (1998), 7 = Nelson et al. (1999), 8 = Gutzmer and Beukes
(1998b), 9 = Sumner and Bowring (1996), 10 = Hannah et al. (2004), 11 = Cornell et al. (1996) and Dorland (2004), 12 =
Martin et al. (1998).
Beukes_Gutzmer 6/11/08 7:40 AM Page 13
(Fig. 6). This correlation is supported by the fact that the cor-
relative Monteville-Jeerinah impact spherule layer is devel-
oped in similar position below both the ankerite-banded
chert unit of the Monteville Formation and the Marra
Mamba Iron Formation (Fig. 6).
Above the Marra Mamba Iron Formation the transition
zone into the overlying Wittenoom carbonate succession is
composed of carbonaceous shale and carbonates. This unit,
known as the West Angela Member of the Wittenoom For-
mation (Blockley et al., 1993), bears resemblance to the Mon-
teville Formation (Fig. 6) above the ankerite-banded chert
marker in the basinal area of the Campbellrand succession to
the south of the Griquatown fault zone (Beukes, 1987).
The Wittenoom Dolomite of the Hamersley basin (Simon-
son et al., 1993a, b) displays many features that can also be
observed in the basinal Nauga Formation (Beukes, 1980c,
1987) of the Campbellrand carbonate platform (Fig. 6). Two
tuff-rich zones, known as the Main and Crystal-rich tuff in-
tervals in the Wittenoom succession (Simonson et al., 1993a,
b) and Tuff zones 2 and 3 in the Nauga Formation (Beukes,
1980c), respectively, can also be correlated based on their
close stratigraphic spacing in both basins. However, perhaps
the stratigraphic interval with the most remarkable similarity
is the one that caps the carbonate succession in both basins.
In the Hamersley basin, where it is known as the Mount
Sylvia Formation, it is only about 30 to 40 m thick and com-
posed of two distinct and persistent lower ferruginous chert
beds overlain by carbonaceous shale with interbeds of car-
bonate and capped by a very persistent thin BIF informally
known as Brunos Band (Fig. 6; Simonson et al., 1993a, b). In
the Transvaal basin a virtually identical succession caps the
Nauga Formation in the basinal facies of the Campbellrand
Carbonate platform. It has not been given formal strati-
graphic status, but Beukes (1980) referred informally to the
upper iron formation as BIF-1 and the two ferruginous chert
beds as BIFs 2 and 3. It is this upper BIF-1 that is in this
paper informally referred to as Brunos BIF after its proposed
counterpart in Western Australia (Fig. 6). One prominent and
laterally persistent event bed in the Wittenoom succession of
which no counterpart has yet been found in the Nauga suc-
cession, is the so-called Wittenoom impact spherule layer that
is present a few tens of meters above the Main Tuff interval
in the Hamersley basin (Fig. 6).
Above Brunos Band, black carbonaceous shale of the
Mount McRae Formation in the Hamersley basin would cor-
relate with the Klein Naute Shale in the basinal area of the
Campbellrand carbonate platform succession (Fig. 6). In
Western Australia the chert- and shale-rich unit that under-
lies the Brockman Iron Formation, is referred to as the Colo-
nial Chert Member of the Mount McRae Shale (Fig. 6). In
South Africa this unit is represented by the Kliphuis Member
of the Kuruman Iron Formation (Fig. 6).
The overlying major iron formation succession displays
many similarities that permit correlation. As pointed out by
Cheney (1996), the most convincing correlation is between
the Dales Gorge Member of the Brockman Iron Formation
and the Stofbakkies Member of the Kuruman Iron Formation
(Fig. 6). It has long been recognized that the Dales Gorge
Member is composed of 16 so-called shale macrobands (S
bands) each overlain by a BIF macroband (Trendall and
Blockley, 1970). These shale-BIF macrocycles correspond to
stilpnomelane lutite-banded micritic iron formation macrocy-
cles as defined by Beukes (1978, 1980b) in the Kuruman Iron
Formation. It may be pure chance, but it is interesting that in
the type profile of the Kuruman Iron Formation near Kuru-
man, the Stofbakkies Member also comprises 16 such macro-
cycles, starting with the first stilpnomelane lutite bed that di-
rectly overlies the Kliphuis Member (refer to fig. 5 in Beukes,
1980b). Although Cheney (1996) also noted this correlation,
he recognized 17 BIF units, because he included the upper
BIF of the Kliphuis Member. This, however, would corre-
spond to BIF-O of Trendall and Blockley (1970) and Trendall
(1983) at the base of the Dales Gorge Member of the Brock-
man Iron Formation. For reference purposes, it should be
noted that the Stofbakkies Member, as defined in this paper
for simplicity (Fig. 6), corresponds to the Matlipani and
Whitebank members of the type profile of the Kuruman Iron
Formation as originally defined by Beukes (1980).
In addition to the detailed correlation of macrocycles, the
equivalent of an impact spherule layer from S-4 in the Dales
Gorge Member (Hassler and Simonson, 2001) has been recog-
nized in equivalent stratigraphic position in the Stofbakkies
Member of the Kuruman Iron Formation (Fig. 6). The spherule
bed was first found by Van Wyk (1987) in drill core from near
Pomfret but interpreted at the time as volcanic lapilli. More re-
cently, the layer was intersected in core from the Agouron
Drilling Project and interpreted as of possible meteorite impact
origin (Schroeder et al., 2006; Simonson et al., 2006).
Above the Stofbakkies Member, the stilpnomelane-rich
Buisvlei Member is correlated with the Whaleback Shale and
the Orange View Member of the Kuruman Iron Formation
with the Joffre Member of the Brockman Iron Formation of
the Hamersley Group (Fig. 6). The Orange View and Joffre
Members both represent rather monotonous successions of
magnetite-siderite banded micritic iron formation with some
stilpnomelane lutite interbeds (Beukes, 1980b; Trendall and
Blockley, 1970).
The Yandicoogina Shale Member that forms the top of the
Brockman Iron Formation is considered correlative to the
stilpnomelane-rich unit that marks the base of the Wester-
berg Member in the upper part of the Kuruman Iron Forma-
tion (Fig. 6). Above that the finely laminated micritic iron for-
mation of the Westerberg and Geduld Members corresponds
to the Weeli Wolli Iron Formation of the Hamersley Group.
Similar to the Westerberg and Geduld Members, the Weeli
Wolli Iron Formation is also characterized by a lack of distinct
chert mesobands (Trendall, 1973). However, it is typically a
finely microbanded magnetite-hematite micritic iron forma-
tion, whereas the Westerberg and Geduld Members are
greenalite rich. The Weeli Wolli Iron Formation is further
characterized by the presence of several diabase sills and the
same applies to the Westerberg and Geduld Members. Lo-
cally, a basaltic pillow lava unit is present in the Weeli Wolli
Formation (Barley et al., 1997) that has no obvious correlative
in the Transvaal basin (Fig. 6).
The Weeli Wolli Formation is overlain by the ~400-m-thick
Woongarra rhyolite which is considered to comprise both ex-
trusive and intrusive phases (Doyle et al., 2001), although
Trendall (1995) argued for it to be an intrusion. If the corre-
lation of the Geduld Member with the upper part of the
14 BEUKES AND GUTZMER
0361-0128/98/000/000-00 $6.00 14
Beukes_Gutzmer 6/11/08 7:40 AM Page 14
Weeli Wolli Formation is accepted, then the Woongarra rhy-
olite should have extruded coevally with the lower part of the
Middelwater Member of the Griquatown Iron Formation
(Fig. 6). Although no equally prominent volcanic unit has
been identified in the Transvaal basin, it is interesting to note
that the Griquatown Iron Formation at this stratigraphic level
contains abundant stilpnomelane lutite beds in the Pomfret
area (Van Wyk, 1987). These stilpnomelane beds most prob-
ably represent pyroclasic beds derived from acid volcanism
(Van Wyk, 1987) and may be tentatively correlated with the
Woongara igneous event.
The Woongarra rhyolite is overlain by the Boolgeeda Iron
Formation that is microbanded but does not contain well-de-
fined chert mesobands (Trendall and Blockley, 1970). It is di-
vided into a Lower and Upper unit by a coarsening-upward
siliciclastic shale-fine quartzite unit that is similar to the Pan-
netjie Formation of the Koegas Subgroup (Fig. 6). Based on
this observation, the lower Boolgeeda Iron Formation is ten-
tatively correlated with the upper part of the basinal Middel-
water Member of the Griquatown Iron Formation and the
upper Boolgeeda Iron Formation with the Doradale Iron
Formation of the Koegas Subgroup (Fig. 6).
The Kungarra Formation, which overlies the Boolgeeda
Iron Formation with sharp contact and forms the base of the
Turee Creek Group of the Mount Bruce Supergroup in the
Hamersley basin (Fig. 6), is composed of fine siltstone,
greywacke, and quartzite (Martin et al., 1998), which is simi-
lar in character to the Naragas Formation of the Koegas Sub-
group (Beukes, 1983) and could be correlative (Fig. 6). A
glacial diamictite, known as the Meteorite Bore Member
(Trendall and Blockley, 1970), is developed in the Kungarra
Formation (Fig. 6). The classification of the diamictite as a
member within the Kungarra Formation may, however, be
very misleading as the diamictite contains boulders of the un-
derlying Woongarra rhyolite (Trendall, 1983). The Meteorite
Bore diamictite thus probably overlies a major low-angle un-
conformity in the succession and should mark the base of a
new sequence. In the Transvaal area of the Transvaal Super-
group, a glacial diamictite is developed near the base of the
Duitschland Formation (Martini, 1979; Bekker et al., 2001)
and the Meteorite Bore Member is considered a correlative
of that diamictite (Fig. 6; Cheney, 1996). The presence of an
erosional unconformity at the base of the Meteorite Bore di-
amictite may explain why iron formations equivalent to the
Rooinekke-Nelani succession of the Koegas Subgroup is not
present in the Hamersley basin (Fig. 6).
Correlation of the succession above the diamictite falls out-
side the focus of this paper. However, it is worthwhile noting
that in contrast to comparisons by Button (1976a) and Ch-
eney (1996), the Cheela Springs Basalt is considered equiva-
lent to the Machadodorp Basalt of the Pretoria Group and not
the Ongeluk/Hekpoort succession (Fig. 6). The erosional un-
conformity at the base of the Beasley River Quartzite is thus
correlated with the unconformity below the Dwaalheuwel
Formation of the Pretoria Group along which the well-known
Hekpoort paleosol (Rye and Holland, 1998; Beukes et al.,
2002b) is developed. A thick succession of strata, including
the Makganyene diamictite (Evans et al., 1997), has thus ap-
parently been removed in the Hamersley basin during the
~2200 Ma Opthalmian orogeny which corresponds to the
post-Mooidraai and/or pre-Dwaalheuwel fold event (Beukes
et al., 2002b) in the Transvaal basin (Fig. 6).
Isotopic ages and rock accumulation rates
Available U-Pb ages obtained on zircons in tuffaceous sed-
iments (usually by SHRIMP) generally show good agreement
between stratigraphic units in the Transvaal and Hamersley
basins that are correlated based on lithologic grounds (Fig. 6).
However, inherited zircons abound in some of the tuffaceous
beds and in some cases produce ages evidently older than
sediment deposition (Alterman and Nelson, 1998; Pickard,
2002, 2003; Trendall et al., 2004). This renders interpretation
of depositional age somewhat inconclusive. An example is the
SHRIMP U-Pb age obtained by Pickard (2003) on zircon sep-
arated from stilpnomelane lutite beds from the lower part of
the Geduld Member of the Kuruman Iron Formation (Fig.
6). The preferred age reported for the stilpnomelane beds by
Pickard (2003) is around 2460 Ma. However, calculation of
the ~2460 Ma age ignores what are referred to as younger
statistical outliers by Pickard (2003). If these are taken into
consideration the depositional age could be on the order of
2440 to 2446 Ma, which corresponds very well with that of
the Weeli Wolli Iron Formation in the Hamersley basin (Fig.
6).
Although in the correlation diagram strata in the basinal fa-
cies if the two successions are correlated, it should be noted
that samples of tuffaceous beds for which U-Pb ages of 2588
6 Ma (zircon by SHRIMP) and 2521 3 Ma (zircon by
TIMS) were obtained (Fig. 6) are from the shallow shelf and
not the basinal facies of the Campbellrand succession. Simi-
larly, a 2488 6 Ma U-Pb age on zircons reported in Figure
6 was obtained on a stilpnomelane lutite sample from the cor-
relative unit of the Stofbakkies Member in the Penge Iron
Formation (Transvaal region, shallow-platform facies). Fur-
thermore, three of the ages reported in Figure 6 are not U-
Pb ages obtained on zircon. These are (1) a 2501 8 Ma Re-
Os age on pyrite for the Mount McRae shale (Anbar et al.,
2007), (2) a 2322 15 Ma age Re-Os age on pyrite for the
Timeball Hill Formation (Hannah et al., 2004), and (3) a 2220
13 Ma Pb-Pb on whole-rock age for the Ongeluk Lava
(Cornell et al., 1996). The latter age has recently been con-
firmed by U-Pb analyses of zircons from tuffaceous beds in
the Hekpoort Lava (Dorland, 2004).
The ages provide a means of estimating rock accumulation
rates, also referred to as compacted sedimentation rates
(Pickard, 2002, 2003) or characteristic depositional rates
(Trendall et al., 2004). Several publications deal with this as-
pect of deposition of the iron formations of the Asbesheuwels
and Hamersley Groups (Arndt et al., 1991; Barley et al., 1997;
Alterman and Nelson, 1998; Pickard, 2002, 2003; Trendall et
al., 2004). The calculated rock accumulation rates for BIF
units vary greatly from 3 to 4 to 180 m/m.y. Of course, in suc-
cessions with major differences in stratigraphic thickness be-
tween shallow shelf and basin, such as the Campbellrand and
Asbesheuwels Subgroups, rock accumulation rates must have
varied accordingly. Based on the available ages of 2588 Ma at
the base and 2521 Ma near the top of the Campbellrand suc-
cession (Fig. 6) an overall rock accumulation rate of ~33
m/m.y. is calculated for the shelf and 8 m/m.y. for the basinal
facies. Similarly, for the Kuruman Iron Formation, using an
MAJOR IRON FORMATIONS AT THE ARCHEAN-PALEOPROTEROZOIC BOUNDARY 15
0361-0128/98/000/000-00 $6.00 15
Beukes_Gutzmer 6/11/08 7:40 AM Page 15
age of 2495 Ma near the base and 2464 Ma near the top of the
Orange View Member (Fig 6) a rock accumulation rate of 16
to 17 m/m.y. is calculated for the basin and 4 to 5 m/m.y. for
the shallow shelf.
Reconstruction of the Transvaal-Hamersley basin
The excellent correlation between the Hamersley and
Transvaal successions strongly suggest that deposition took
place in a single basin along the margin of Vaalbara, as origi-
nally proposed by Cheney (1996). This raises the question
about the relative and absolute paleogeographic positions of
the Kaapvaal and Pilbara cratons at the time of deposition. In
the reconstruction of Cheney (1996), the Pilbara craton is
placed to the south of the Kaapvaal craton. However, recent
paleomagnetic analyses by De Kock (2007) and De Kock et
al. (2007) place the Pilbara craton immediately to the north-
northwest of the Kaapvaal craton. De Kock (2007) stresses
that several structural and depositional elements in the two
basins support this configuration.
According to the reconstruction put forward by De Kock
(2007) and De Kock et al. (2007) the basin in which the
Hamersley-Asbesheuwels Iron Formation succession was de-
posited consisted of a shallow-shelf area to the east and a
deep basin to the west. A large embayment is envisaged be-
tween the two cratons (Fig. 7) because of the known arcuate
shelf margin along the western side of the Kaapvaal craton
(Beukes, 1983, 1984). If a volcanic arc situated some distance
off Vaalbarra to the west is invoked, iron formation deposition
would have taken place in a back-arc environment, as sug-
gested for both the Hamersley (Blake and Barley, 1992) and
the Asbesheuwels successions (Klein and Beukes, 1989). This
volcanic arc could also have been a prolific source of hy-
drothermal fluids needed for deposition of the iron forma-
tions. Also indicated in the basin reconstruction is a proposed
source for mafic tuffs during deposition of the Wittenoom-
Campbellrand carbonate platform, based on results of Has-
sler (1993), and a volcanic center responsible for outflow of
pillow lavas during deposition of the Weeli Wolli Iron For-
mation and formation of the Woongarra rhyolite (Barley et al.,
1997) shortly thereafter (Figs. 6, 7).
The successful reconstruction of the combined Hamersley-
Transvaal basin implies that the two largest known iron for-
mation successions in the world were deposited in one large,
possibly restricted (Hassler, 1993) oceanic basin at ~2,44 to
2,50 Ga. Similarities in secular variations in depositional envi-
ronments in the Hamersley and Transvaal successions may
thus, for the most part, be only of regional significance. Local
influences on deposition would have to be carefully evaluated
before any interpretations about possible global evolutionary
changes can be made from studies of correlative rock units in
the two areas.
Iron Formations in the Campbellrand Carbonate Platform
and Transition to the Kuruman Iron Formation
The Neoarchean Cambellrand carbonate platform succes-
sion of the Transvaal basin (Figs 1, 5, 6) contains several thin
ankerite-banded chert units, so-called proto-iron formations
of Button (1976b), and two thin banded micritic iron forma-
tion units (Fig. 8; Beukes, 1987; Sumner and Beukes, 2006).
The two iron formations, known as the Kamden Member in
the middle and Brunos BIF near the top of the carbonate
platform succession (Fig. 8), are believed to hold fundamen-
tal information about the depositional systems in late
Neoarchean oceans that eventually led to deposition of the
overlying giant iron formations of the Asbesheuwels Sub-
group (Figs. 1, 8). The depositional setting of these iron for-
mations and the transition into the basal units of the Kuru-
man Iron Formation (Fig. 8), thus needs to be discussed in
detail.
Regional facies architechture
The regional facies architecture of the Campbellrand car-
bonate platform has been documented by Beukes (1980c,
1983, 1987), Sumner (1997a, b), Sumner and Grotzinger
(2004), Schroeder et al. (2006), and Sumner and Beukes
(2006). The platform comprises a 2,400-m-thick stromatolitic
shallow-shelf succession that sharply thins along the Griqua-
town fault zone, along the southwestern margin of the Kaap-
vaal craton, into a 550-m-thick finely laminated, nonstroma-
tolitic basinal (i.e., deeper shelf) carbonate succession (Fig.
8). Deposition of the succession took place over a period of
about 70 m.y. between 2.52 and 2.59 Ga and is constructed of
12 third-order transgressive-regressive sequences, each with
an average duration of ~6 m.y. (Sumner and Beukes, 2006).
The lower part of the platform consists of a carbonate ramp
(Beukes, 1987) comprising supratidal and intertidal cherty
dolostones that interfinger toward the ramp margin with shal-
low-subtidal elongate giant stromatolite mounds (Fig. 8). Bas-
inward of the ramp margin deep subtidal fenestrate micro-
bialites, often with steep conophyton-like stromatolites
(Beukes, 1987), are interbedded with muddy and grainy slope
carbonates (Schroeder et al., 2006; Sumner and Beukes,
2006). With time, this ramp developed into a rimmed shelf
margin that characterizes the upper part of the succession
16 BEUKES AND GUTZMER
0361-0128/98/000/000-00 $6.00 16
H
y
d
r
o
t
h
e
r
m
a
l
P
l
u
m
e
B
A
S
I
N
V
A
A
L
B
A
R
A
Pilbara
Craton
0 km 500
Koegas
Newman
Post Depositional
Uplift
S
H
E
L
F
Nullagine
Post Depositional
Uplift
Griquatown
Pomfret
Kurumam
Johannesburg
Kaapvaal
Craton Volcanic
Centre
Weeli Wolli
Times
Pyroclastics
Wittenoom
Times
N
FIG. 7. Reconstruction of depositional environments in the original com-
bined Transvaal-Hamersley basin. Modified from the paleogeographic re-
construction by De Kock (2007) based on paleomagnetic results. Position of
source of pyroclastics during deposition of the Wittenoom carbonates is after
Hassler (1993) and that of mafic volcanics in the Weeli Wolli Iron Formation
after Barley et al. (1997).
Beukes_Gutzmer 6/11/08 7:40 AM Page 16
(Beukes, 1987; Sumner and Beukes, 2006). The rimmed shelf
margin is mainly constructed of high-relief columnar stroma-
tolites and giant stromatolite mounds. Stratiform stroma-
tolitic shelf lagoonal deposits are developed behind and slope
deposits in front of this rim. Steep conophyton-like stromato-
lites are well developed in the upper part of the slope (Fig. 8)
and mark highstands when very little carbonate detritus was
exported from the shallow shelf (Schroeder et al., 2006 Sum-
ner and Beukes, 2006). Distal from the slope the carbonate
succession is essentially composed of finely laminated fine
turbiditic carbonate wackestones, thin grainstones, and strat-
iform microbialites. These microbialites typically comprise
very organic-rich microbial laminae that are locally contorted
or disrupted to form roll-up structures (Beukes, 1987; Klein
et al., 1987; Sumner, 1997a, b). Thin carbonaceous shale beds
are interbedded with both shelf and basinal carbonates; in-
terestingly those of the basin appear more abundant along the
toe of slope than farther offshore (Schroeder et al, 2006). Sev-
eral mafic tuff intervals are present in the deep-water basinal
facies but absent or not preserved among shallow-shelf car-
bonates (Beukes, 1980c, 1987).
Within this carbonate platform setting, the ankerite-
banded cherts and two thin BIF units are all interbedded
with some of the most distal, deep-shelf carbonates in the
basin (Fig. 8; Beukes, 1980c, 1983, 1987). There are at least
six ankerite-banded cherts in the most distal part of the
basin. Characteristically, most of them pinch out toward the
shelf slope and only one unit, near the base of the succession,
actually extends into the shelf slope environment (Fig. 8).
Relative to the ankerite-banded cherts, the two BIF units
MAJOR IRON FORMATIONS AT THE ARCHEAN-PALEOPROTEROZOIC BOUNDARY 17
0361-0128/98/000/000-00 $6.00 17
ch
ch
ch
ch
ch
ch
ch
ch
ch
ch
ch
ch
RAMP
LAGOON
2000
m
1000
0
North
~ 450 km
K
u
r
u
m
a
n
B
I
F
Fe
Fe
Fe
Fe Fe Fe Fe Fe Fe
Fe
Fe Fe Fe
Fe
Fe Fe Fe Fe
Fe Fe Fe
Fe Fe Fe
Fe Fe Fe
Fe
Fe Fe
Fe
Fe
Fe
Fe
Fe
Fe
Fe
South
FAULT
Deep Shelf
SHELF
MARGIN
SLOPE
Brunos
BIF
Kamden
BIF
Kuruman BIF
BASIN SLOPE MARGIN CARBONATE SHELF
INTERIOR
Monteville Establishment of Platform
Conophyton-like microbialites
Griquatown Growth Fault
RAMP
INTERTIDAL
INTERTIDAL
Koegas Griquatown
Whitebank Pomfret
Intertidal dolomite =Cherty, light grey, highly depleted in FeOwith MnO: FeO>>1
Lagoonal dolomite =Manganiferous with MnO: FeO>1
Ramp and shelf margin dolomite =Manganiferous with MnO: FeO>1
Slope and basinal dolomite =Ferruginous with MnO: FeO<1
e
t
u
a
N
n i
e l K
Gamohaan drowning of platform
Oolite Banks
FIG. 8. South to north stratigraphic section of the Neoproterozoic Cambellrand carbonate platform succession, illustrat-
ing facies relationships between basinal and shallow-shelf environments and the stratigraphic setting of two persistent iron
formation units known as the Kamden Member in middle and Brunos BIF near the top. Both iron formations are associated
with major transgressions in the succession with Brunos BIF marking final drowning of the carbonate platform. The Kam-
den BIF in the middle of the succession thickens from a few meters to almost 30 m along the slope of the shelf in front of
the shelf margin. Deeper into the basin it is represented by ankerite-banded chert (proto iron formation). Note that ankerite-
banded chert (proto iron formation) units are essentially restricted to deep basinal environments with one unit near the base
of the carbonate platform extending into the shelf environment to the north of the Griquatown growth fault that determined
the position of the shelf margin for most of the time. This lower ankerite-banded chert was deposited during the initial trans-
gression that led to establishment of the carbonate platform on the Kaapvaal craton (modified after Beukes, 1987). Refer to
Figures 1 and 15B for position of section line.
Beukes_Gutzmer 6/11/08 7:40 AM Page 17
(Kamden Member and Brunos BIF) have much wider lateral
distribution and extend from the basin well onto the shelf
(Fig. 8). All of the iron-rich units mark transgressions or max-
imum flooding surfaces in the basin (Beukes, 1987; Sumner
and Beukes, 2006). Transgressions associated with the two
BIF units (Kamden Member, Brunos Band, Fig. 8) were,
however, most severe and led to drowning of the shallow-
shelf carbonate platform. Drowning of the platform during
deposition of the Kamden BIF in the middle of the succes-
sion was short-lived, whereas transgression associated with
Brunos BIF initiated the final demise of carbonate buidup
and eventually led to deposition of the overlying very thick
Kuruman Iron Formation (Fig. 8).
In contrast to the iron-rich units, carbonaceous shale beds
in the basin appear to have been deposited at times of maxi-
mum regression when the carbonate shelf was either exposed
or produced very little carbonate so that clays from the plat-
form interior could bypass the shelf and be deposited in the
basin. This would explain why shale beds are more prominent
immediately in front of the shelf slope than farther into the
basin (Schroeder et al., 2006). On the shallow-shelf, carbona-
ceous shale beds often overly karstic carbonate exposure sur-
faces. Here, they appear to mark initial phases of transgres-
sion when newly created accommodation space led to
retrogradation of siliciclastic muds into the shelf interior and
renewed carbonate production was established.
Carbonates, both limestone and dolomite, of the deep-basin
and shallow-subtidal environments are generally dark and car-
bonaceous, whereas those of the inter- to supratidal environ-
ments are light gray in color. The limestones and dolomites are
on average 10 to 100 times enriched in MnO and FeO relative
to Phanerozoic carbonates (Veizer, 1978; Beukes, 1987). Most
interestingly dolomite and limestone of the Campbellrand
succession display a distinct trend from being enriched in iron
over manganese in the basin to just the opposite in shallow
sub- and inter- to supratidal carbonates (Fig. 8; Beukes, 1987).
Light gray inter- and supratidal dolomites in the interior of the
shelf have rather similar concentrations of MnO relative to
carbonates of other depositional environments but are highly
depleted in FeO (Beukes, 1987).
Carbonate to iron formation depositional systems tracts
Ankerite-banded chert: The ankerite-banded cherts in the
basinal facies of the Campbellrand carbonate platform suc-
cession are composed of alternating bands of highly ferrugi-
nous to ankeritic dolomite and chert. The highly ferruginous
to ankeritic dolomite bands are regarded as ferruginized
equivalents of deep-water basinal limestone and dolostone
that are similar in texture (Beukes, 1984). Two types of chert
bands are present, namely intraclastic chert, representing
chertified deep-water muddy carbonate turbidites and mi-
crobanded primary sedimentary chert bands (Beukes, 1984).
The ankerite-banded cherts indicate that the deep-basinal
carbonate shelf was occasionally encroached by a water mass
that must have been enriched in iron and silica to have al-
lowed for precipitation of microbanded primary cherts and
ferruginization and chertification of surrounding deep-shelf
carbonates.
Kamden BIF: The Kamden BIF (Fig. 8) represents the
maximum flooding stage of the sixth third-order depositional
systems tract or sequence (Sumner and Beukes, 2006) in the
Campbellrand carbonate platform succession (Fig. 9). The
BIF is sideritic and reaches a maximum known thickness of
~30 m along the slope of the shelf margin from where it pe-
ters out to thin ankerite-banded chert deeper into the basin
(Fig. 8). On the shallow shelf it is represented by a persistent
zone of ferruginous dolomite and chert (Fig. 8) associated
with deep-water carbonates traceable for hundreds of kilo-
meters along strike (Beukes, 1987; Sumner and Beukes,
2006).
18 BEUKES AND GUTZMER
0361-0128/98/000/000-00 $6.00 18
B I F
Dolarenite
Stromatolite
Laminated dolmicrite
Siderite BIF
Laminated dolmicrite
Conophytic dolomite
Giant stromatolite
mounds
Carbonaceous shale and
dolomite rip-ups
LSST
ETSST
TSST
LTSS
EHSST
LHSST
SB Exposure
Maximum Flooding
Stage 1=Exposure and lowstand wedge of shale
Stage 2 = Transgression
Stage 3 = Maximum flooding. Plume upwelling
Stage 4 = Late highstand
PL
NWB
Carbonate sands into basin
Plume retreat
Hydrothermal Plume
T
u
rb
id
ite
SB Exposure
(Clay bypassing)
~
2
5
0
m
C
M
6
Chert
B
I
F
++
Si0 + Fe
2
Fe-dolomite
B
I
F
Shale
FIG. 9. Schematic illustration of the depositional systems tract that holds the Kamden Iron Formation in sequence 6 of
Sumner and Beukes (2006) on the shelf of the Cambellrand carbonate platform. See text for details. Symbols in diagrams il-
lustrating lateral facies relationships at different stages of development of the sequence are the same as that in the profile.
EHSST = early highstand systems tract, ETSST = early transgressive systems tract, LHSST = late highstand systems tract,
LSST = lowstand systems tract, LTSST = late transgressive systems tract, SB = sequence boundary, TSST = transgressive
systems tract.
Beukes_Gutzmer 6/11/08 7:40 AM Page 18
Examination of the sequence hosting the Kamden Member
allows rconstruction of depositional systems tracts, commenc-
ing with initial exposure of the carbonate platform, at the
basal sequence boundary, through flooding of the platform
and BIF deposition to regression and development of an
upper exposure surface or sequence boundary (Fig. 9). On
the shallow shelf the lower sequence boundary is represented
by a marked dissolution and erosional surface capping high-
stand carbonates of the underlying sequence (Fig. 9). It is this
stage that would have been most favorable for bypassing of
the carbonate shelf by siliciclastic clays and deposition in the
basin immediately in front of the shelf margin (Stage 1, Fig.
9; Schroeder et al., 2006). Following exposure, during early
lowstand, slightly increased accommodation space led to de-
position of carbonaceous shale with carbonate ripup clasts
representing the basal lithologic unit of the sequence on the
shelf (Fig. 9). Subsequent transgression accompanied by in-
creased accommodation space led to rapid carbonate
buildup, trapping and precipitation of carbonate in stromato-
lites on the shelf, and very little carbonate bypassing the shelf
margin (Stage 2, Fig. 9). Steep conophyton-like stromatolites,
essentially made up of precipitated carbonate cement and
free of transported carbonate mud and sands, grew along the
shelf slope during this time. At greater water depth, thick or-
ganic-rich microbial mats with contorted laminae and some
roll-up structures and subordinate, thin, micritic turbidite
beds formed (Stage 2, Fig. 9). Because of abundant accom-
modation space, the deep basin became starved of influx of
carbonate mud and sand from the shallow shelf, creating an
ideal setting for deposition of ankerite-banded chert and/or
iron formation.
Maximum flooding and drowning of the carbonate shelf
was followed by deposition of the Kamden BIF (Stage 3,
Fig. 9). BIF deposition was most pronounced along the
slope of the shallow carbonate platform, in an area that
would have favored enhanced upwelling of deep ocean
water (Stage 3, Fig. 9). In the central part of the shallow
platform, influx of this deep ocean water is recorded by fer-
ruginization and silicification of deep-water carbonates.
This indicates that shallow-marine carbonate production
was not completely shut off by the marine flooding event
but must have persisted deep into the interior of the plat-
form. This is corroborated by the fact that the Kamden
Member pinches out toward the more interior part of the
carbonate platform (Fig. 8).
A progradational systems tract followed on deposition of
the Kamden Member (Stage 4, Fig. 9). It implies retreat of
the iron-enriched deep-water column from the shelf. With
decreasing accommodation space, especially during late high-
stand, carbonate was exported from the shallow shelf leading
to deposition of turbiditic carbonate in the basin. Shallow
sub- to intertidal stromatolitic carbonates and cross-lami-
nated dolarenites formed on the shelf immediately prior to
development of a subsequent exposure surface at the next fall
inflection point (Stage 4, Fig. 9).
Brunos BIF: A transgressive systems tract similar to the
one described above also led to deposition of Brunos BIF
near the top of the Gamohaan Formation in the uppermost
part of the Campbellrand-Malmani carbonate platform suc-
cession (Fig. 8). However, in the latter case the carbonate
platform did not recover from the flooding event and deposi-
tion of iron formation persisted into the overlying As-
besheuwels Subgroup (Figs. 8, 10A).
The systems tract for Brunos BIF, sequence CM12 of Sum-
ner and Beukes (2006), has been described in detail by
Beukes (1980c, 1987) and Sumner (1997a, b). Here we pro-
vide only a brief summary. The systems tract commences with
light gray cherty intertidal carbonates overlain by laminoid
fenestral lagoonal limestone and shallow-subtidal elongated
giant stromatolitic mounds that constructed the shelf margin
(Fig. 10A). The shelf margin deposits are overlain by rippled
microbial carbonate sands that were washed down from envi-
ronments above normal wave base. These are, in turn, over-
lain by a prominent zone of carbonaceous fenestral limestone
characterized by a variety of conophyton-like stromatolites,
also referred to as the conoform microbialite assemblage by
Sumner (1997a, b). The lower part of this zone consists of a
succession of cuspate small conophytic microbialites alternat-
ing with thin layers of contorted microbial mats. Some of the
contorted mats appear to have been reworked by currents.
Above the bedded cuspate microbialites there is no indication
of current reworking so that the top of the bedded cuspate
subzone is considered the absolute limit of the storm-wave
base (Fig. 10A). The bedded cuspate microbialites are suc-
ceeded by a subzone with abundant conoform columnar stro-
matolites set among highly carbonaceous stratiform to con-
torted and rolled-up microbial mats (Fig.10A). A few
persistent marker beds composed of delicate plumose micro-
bialite structures (Sumner 1997a, b) are interbedded with the
conoform stromatolite subzone. Conophyton-like micro-
bialite structures characteristically start small and closely
spaced in lower parts of the succession and become taller and
more widely spaced upward, until they disappear abruptly.
This is taken as an indication of increasing water depth, with
the disappearance of conophytic microbialites most probably
marking the photic limit (Fig. 10A). Above the conoform stro-
matolite zone the succession is essentially composed of deep-
water, highly carbonaceous stratiform microbialites with
abundant contorted bedding, roll-up structures, and fram-
boidal pyrite nodules (Fig. 10A). This facies is considered to
represent a proximal deep-shelf environment.
Farther up the sequence into distal deep-shelf strata, the
carbonates are essentially composed of carbonaceous strati-
form microbialites with lesser contorted bedding and roll-up
structures but with more frequent thin black carbonaceous
shale partings and some thin-graded low-density turbidite
beds made up of tiny fragments of dark carbonaceous micro-
bial laminae (Fig. 10A; Beukes, 1987). This facies is in direct
contact with ankerite-banded chert and microbanded banded
micritic iron formation of Brunos BIF that marks the base of
the Tsineng Member of the Gamohaan Formation (Fig. 10A).
As the BIF unit is approached, nodular pyrite disappears
from the succession with only trace amounts of very fine dia-
genetic pyrite present in carbonaceous carbonate and chert
next to banded micritic iron formation units. Banded micritic
iron formation beds proper contain no evidence of any early
diagenetic pyrite and the trace amounts of pyrite that are pre-
sent can normally be seen to be either very late diagenetic or
postdiagenetic in age, having formed after lithification of
the banded micritic iron formation. The banded micritic
MAJOR IRON FORMATIONS AT THE ARCHEAN-PALEOPROTEROZOIC BOUNDARY 19
0361-0128/98/000/000-00 $6.00 19
Beukes_Gutzmer 6/11/08 7:40 AM Page 19
iron formation is typically very chert rich and sideritic imme-
diately adjacent to the dark carbonaceous limestones. How-
ever, at some localities the center of Brunos BIF may be
composed of hematite-facies banded micritic iron formation
(Fig. 10A). Brunos BIF is usually overlain by a thin unit of
deep-water carbonate that in turn grades upward into car-
bonaceous shale that consistently underlies the Kuruman
Iron Formation (Figs. 8, 10A).
Dolostones in the depositional systems tract that leads up
to Brunos BIF (Fig. 10A) display compositional variations re-
lated to depositional settings that are similar to those observed
on a basinal scale (Fig. 8). Dolomites in upper slope, shallow-
shelf margin lagoonal and intertidal settings are enriched in
manganese over iron with the opposite true in lower slope
and deep-shelf dolomites. Dolostones immediately adjacent
to ankerite-banded chert and banded micritic iron formation
are highly ferruginous and even ankeritic to sideritic (Fig.
10A; Klein and Beukes 1989).
It is important to note that in the basin proper, Brunos
BIF is represented by a single, well-defined banded micritic
iron formation unit. However, on the shelf it typically splits
into several layers separated by carbonate and/or black shale
20 BEUKES AND GUTZMER
0361-0128/98/000/000-00 $6.00 20
Kuruman
Brunos
BIF
Abundant
laminated
mat
Deep shelf
margin
Distal
deep shelf
Proximal
deep shelf
Ankeritic
Dolomite
Lower
slope
Upper
slope
Shelf
margin
Shelf
lagoon
Mn dol.
Minor
Limestone
Limestone
+
Mn dol.
Light grey
dolomite
Dysaerobic
Dysaerobic
Anaerobic
Aerobic
Sid
Sid
Sid
Hem
Photic
Limit
P
y
r
i
t
e
o
u
t
Py
Py
K
o
g
e
l
b
e
e
n
F
m
Intertidal
flats
Aerobic
D
y
s
a
e
r
o
b
i
c
A
n
a
e
r
o
b
i
c
D
y
s
a
e
r
o
b
i
c
G
A
M
O
H
A
A
N
F
O
R
M
A
T
I
O
N
ch
Py
ch ch ch
Cherty
Storm
WB
Normal
WB
MRA Matrix Pyrite
Ankerite-banded
chert
Contorted Mat
Distal Deep shelf
Nodular limestone
Giant mounds
Rippled calcarenite
Shallow shelf assemblage
Bedded Cuspate
Plumose Structure
Conoform Columns
Conoform Microbialite
Assemblage
Carbonaceous
chert
Contorted and
laminated mat
Siderite BMIF
Hematite BMIF
Carbonaceous shale
Laminoid
fenestral limestone
Small stromatolites
and calcarenite
K
u
r
u
m
a
n
G
a
m
o
h
a
a
n
Brunos
BIF
Oxide Kuruman IF
Carbonaceous
Laminae
Conoform zone
(
K
a
m
b
e
r
a
n
d
W
e
b
b
2
0
0
1
)
(Bau and Dulski,1996)
Contorted Mat
Nodular Pyrite
Pyritic zone
(Kamber and
Whitehouse,2007)
Lagoonal
Limestone
Spar Cement
Conoform Zone
Limestone
Minor
Fe dol.
MOR
Hydrothermal
Fluid
Hematite
Magnetite
Siderite
Pyrite
Pyrite from
Rouxel et al.
(2005)
Tuff
2521Ma
T
s
i
n
e
n
g
M
e
m
b
e
r
Hem
Mag
Sid
Others from
Johnson
et al.(2003)
S Pool
Shaley Contorted
Mat; Deep shelf
R
E
E
/
P
A
A
S
-3
10
-2
10
-1
10
10
12
8
4
0
Eu
-15 -10 -5 0 5 10 15 20 25 30
La
Ce
Y
H H
-2,0 -1,0 0 +1,0
Relict
Reservoir
4
mSR
K
l
e
i
n
a
n
d
B
e
u
k
e
s
(
1
9
8
9
)
ch
Sid =
Hem =
3
3
s
(
%
)
34
s(% )
La Ce Pr Nd Sn Eu Gd Tb Dy Ho Er Tm Lu Yb Y
56
Fe
Legend for B
Legend for A A B
C
Conoform
microbialite
zone
E
SO
4
MRA
12
8
4
-10 10 20 30
3
3
s
(
%
)
34
s(% )
D
-2
Samples below Tsineng
Member
Samples of Tsineng
Member
(Kaufman et al.,2007)
FIG. 10. A. Stratigraphic profile of the upper Kogelbeen to Gamohaan successions that led to drowning of the Camp-
bellrand carbonate platform and deposition of Brunos BIF and overlying Kuruman Iron Formation in the area of the shelf
to the north of the Griquatown fault zone near Whitebank (see locality in Fig. 8). Also indicated is an interpretation of de-
positional and diagenetic environments and changes in the composition of dolomites from shallow-intertidal to deep-basinal
settings and the part of the succession that was studied by Klein and Beukes (1989). Modified after Beukes (1987) and
Sumner (1997a). ch = chert, dol = dolomite, Hem = hematite-facies BIF, py = pyrite, sid = siderite-facies BIF, WB = wave
base. B. Iron isotope composition of various mineral phases (expressed as
56
Fe) in the succession, including samples from
the overlying Kuruman Iron Formation. C. Secondary iron mass spectrometer multiple sulfur isotope analyses of different
types of pyrite in the succession (for location of samples see symbols, which are similar to those in crossplot of sulfur isotope
values, immediately to left of stratigraphic profile). D. Whole-rock multiple iron isotope analyses of pyrite in the succession.
The samples come from one of the drill cores that Klein and Beukes (1989) investigated. E. Examples of ICP-MS REE analy-
ses of different lithofacies in the succession, including shallow-water lagoonal limestone. References to the different data sets
are given in the figures.
Beukes_Gutzmer 6/11/08 7:40 AM Page 20
composing the Tsineng Member of the Gamohaan Formation
(Fig. 10A; Klein and Beukes, 1989). This suggests that the
transgression of Brunos BIF from deep basin onto the shelf
took place in steplike fashion through time and that it does
not necessarily represent a synchronous time marker in the
basin but rather a protracted time interval in which the trans-
gression took place.
Paleoenvironmental reconstruction
The detailed lithostratigraphic data provided above in com-
bination with geochemical information allow reconstruction
of a depositional model for iron formation in relationship to
marine carbonate sedimentation, primary organic carbon pro-
ductivity and paleobathymetry in an oceanic basin toward the
end of the Neoarchean. Certain aspects of this model are sim-
ilar to that proposed by Klein and Beukes (1989; Fig. 11). The
latter model envisaged a depositional system in which iron
formation was deposited from a chemocline along which hy-
drothermally enriched deep water came into contact with the
shallow well-mixed surface layer of a more or less perma-
nently stratified ocean water column (Fig. 11). Deposition of
iron formation specifically took place during transgressions
when the carbonate platform became drowned and organic
carbon input into the deep shelf from adjacent highly pro-
ductive stromatolitic shallow-shelf settings was at minimum
(Fig. 11). However, the Klein and Beukes (1989) model re-
quires substantial revision based on the availability of new
geochemical data and concepts. These are summarized below
before a revised depositional model is developed.
Rare earth element (REE) data: The iron formations and
closely related ankerite-banded cherts are invariably in-
terbedded with the most distal and deepest basinal shelf fa-
cies of the carbonate platform succession (Figs. 810A).
However, for most of the time, only deep-water limestone
with subordinate carbonaceous shale beds accumulated on
the deep shelf in front of the shallow-carbonate platform mar-
gin (Fig. 8). This implies that a more distal or deeper water
mass, which must have been enriched in silica and iron, only
occasionally transgressed onto the deep-carbonate shelf and
resulted in silification and ferruginization of deep-water lime-
stone to form ankerite-banded chert (Beukes, 1983) and/or
deposition of banded micritic iron formation (Fig. 9). From a
sedimentologic perspective, it would thus appear that two
water masses or columns with different composition, a shal-
lower one from which limestone and shale were deposited,
and a deeper one from which silica and iron were deposited,
interacted in the most distal deep-carbonate shelf environ-
ment. This notion is supported by the fact that limestones and
associated carbonaceous shales have geochemical signatures
that are distinct from those of the more distal ankerite-
banded chert and banded micritic iron formation, as was first
noted by Klein and Beukes (1989). Limestones display rather
flat shale-normalized REE patterns similar to that of modern
shallow-marine surface waters with detrital riverine influx
while iron formations display LREE-depleted patterns simi-
lar to that of modern deep-marine water with no terrigenous
input (Klein and Beukes, 1989). The concept finds further
support in Al
2
O
3
concentrations that reflect the amount of de-
trital alumosilicates (i.e., clay minerals). Very low Al
2
O
3
con-
tents indicate that iron formations were deposited in an envi-
ronment devoid of siliciclastic detrital input as opposed to
deep-shelf limestones that have distinctly greater Al
2
O
3
con-
tents (Fig. 12A).
In contrast to modern seawater, shale-normalized REE
patterns of both deep-shelf iron formation and limestone in
the Gamohaan-Kuruman transition zone (Fig. 10A) display
distinctive positive Eu anomalies (Klein and Beukes, 1989).
It is generally accepted that the Eu anomalies were derived
from a component of hydrothermal fluids derived from high-
temperature (>250
o
C) alteration of ocean-floor basalts (Bau
and Dulski, 1996); the latter are highly enriched in REE with
very marked positive Eu anomalies (Michard et al., 1983).
The limestone samples analyzed in the study of Klein and
Beukes (1989) were from the distal deep-shelf environment
of the Gamohaan Formation (Fig. 10A), i.e., the REE char-
acteristics of shallow-shelf carbonates (and thus the shallow
ocean-water reservoir) remained unknown. More recently,
Kamber and Webb (2001) produced REE data from lower
MAJOR IRON FORMATIONS AT THE ARCHEAN-PALEOPROTEROZOIC BOUNDARY 21
0361-0128/98/000/000-00 $6.00 21
B
la
c
k
c
h
e
rt
o
r
s
h
a
le
c
c
O
2
O
2
2+
Fe 2+
Fe
2+
Fe
Hydrothermal
input
Oxide BIF
Siderite
BIF
O
2
Carbonate
shelf
C=Organic Carbon
Stratified ocean system. Production of organic matter low in open marine environment but high on carbonate shelf.
O
2
C
FIG. 11. Diagram illustrating the main components of the model that Klein and Beukes (1989) developed for deposition
of iron formation in the transition zone between the Campbellrand carbonate platform and the Kuruman Iron Formation
successions. This model implies a more or less permanently stratified ocean with deposition of iron minerals taking place
along a chemocline situated at the base of the surface mixed zone of the ocean water column. It also indicates siderite to be
a primary preciptate in response to increased pCO
2
that developed in water column due to degradation of organic matter in
areas of high-organic carbon input from primary production that took place through photosynthesis in shallow-carbonate
shelf environments. In open ocean environments little primary production of organic matter took place due to scarcity in nu-
trients and thus oxide-facies iron formation formed through reaction of dissolved ferrous iron, derived from upwelling deep
hydrothermally enriched water, with free oxygen that was available in shallow-surface ocean water. Deposition of iron min-
erals and biological activity is totally decoupled in this model.
Beukes_Gutzmer 6/11/08 7:40 AM Page 21
slope conophytic microbialites and laminoid fenestral la-
goonal limestone in the Gamohaan succession (Fig. 10E).
They showed that the positive Eu anomaly persists even into
the lagoonal deposits, although it is progressively less pro-
nounced into shallower water. This implies that the whole
water column in the basin, from shallowest to deepest part,
carried a considerable component of hydrothermally derived
REE. It indicates effective mixing of the hydrothermal flux of
REE between the deeper water mass from which ankerite-
banded chert and iron formation were deposited and the very
shallow-mixed water column on the carbonate shelf. The
water in the basin may thus not have been permanently strat-
ified as depicted in earlier models of iron formation deposi-
tion (Holland, 1984; Klein and Beukes, 1989, 1992a, b).
Both shallow-shelf to slope limestone and distal, deep-shelf
iron formation display no Ce anomalies (Fig. 10E), whereas a
prominent negative Ce anomaly is characteristic for modern
ocean water (Elderfield and Greaves, 1982; Bau and Dulski,
1996; Kamber and Webb, 2001). The presence of an apparent
negative Ce anomaly identified by Klein and Beukes (1989)
was found to be an analytical artifact of instrumental neutron
activation analyses, masked by a positive La anomaly (Bau
and Dulski, 1996; Kamber and Webb, 2001). A positive La
anomaly is also characteristic of modern ocean water.
The negative Ce anomaly in modern day seawater is caused
by oxidation of Ce
3+
to insoluble Ce
4+
, which is removed from
the water column through precipitation, especially in associa-
tion with hydrogenetic deep-sea manganese crusts and nod-
ules (e.g., Elderfield, 1988). Oxidation of Ce
3+
to Ce
4+
at a
given pH takes place at Eh values intermediate to those re-
quired for the oxidation of Fe
2+
to Fe
3+
(lower Eh values) and
Mn
2+
to Mn
4+
(higher Eh values). The absence of Ce anom-
alies thus indicates that in both the deep- and shallower water
column, Eh-pH conditions in the Neoarchean ocean did not
permit Ce
3+
to be oxidized and fractionated from the other
REE. By implication Mn
2+
could also not be oxidized and
should have been available in solution throughout the ocean-
water column. This explains the rather similar concentration
of MnO independent of depositional environment in most
carbonates of the Campbellrand succession. The pH in the
shallower water column was alkaline enough (pH of 7.8) to
allow carbonate deposition. However, in the deep-shelf envi-
ronment, Eh-pH conditions must have been such that dis-
solved ferrous iron oxidized to form ferric iron oxydroxide
precipitates, now preserved as microcrystalline hematite, for
example, in the central part of Brunos BIF (Fig. 10A).
Nature of primary iron precipitates: The regional facies ar-
chitecture of the Campbellrand carbonate platform (Fig. 8)
and depositional systems tract analyses (Figs. 9, 10A) reveal
that deposition of iron and silica took place only occassionally
and was restricted to distal deep-water carbonate shelf envi-
ronments. This suggests that the entire water mass of the
Transvaal basin, down to depths of the distal deep carbonate
shelf, was depleted in dissolved iron for most of the time. The
chemocline along which iron precipitated must therefore
have been situated in very deep subtidal environments (Sum-
ner, 1997a) far below storm-wave base and even the photic
zone (Figs. 9, 10A). In order to understand mechanisms of
deposition of the iron along this chemocline it is necessary to
establish the composition of the primary iron precipitates and
consider interplay with other redox-sensitive components, es-
pecially organic carbon.
Klein and Beukes (1989) and Beukes et al. (1990) sug-
gested that siderite in siderite facies banded micritic iron for-
mation in Brunos BIF and lower part of the overlying Kuru-
man Iron Formation represented a primary precipitate from
22 BEUKES AND GUTZMER
0361-0128/98/000/000-00 $6.00 22
2
10
10 1
-1
10
-2
10
-3
10
-2
10
-1
10
1
10
2
10
Al O Weight %
2 3
O
r
g
a
n
i
c
C
a
r
b
o
n
W
e
i
g
h
t
%
Dolomite
Siderite BIF
Shale
Limestone
Oxide BIF
A
10 1
-1
10
-2
10
-3
10
-3
10
-2
10
-1
10
1
10
2
10
Aerobic
Dysaerobic
Anaerobic
Marine
Slope 1:3
Siderite BIF
Oxide BIF
Carbonaceous Shale
Carbonaceous dolomite
and limestone
Organic Carbon Weight %
S
u
l
f
u
r
W
e
i
g
h
t
%
Fe-rich
Pyrite-rich
B
FIG. 12. Correlations of organic carbon and Al2O3 (A) and sulfur (B) contents in the transition zone between the Gamo-
haan carbonate and Kuruman Iron Formation successions (modified after Klein and Beukes, 1989).
Beukes_Gutzmer 6/11/08 7:40 AM Page 22
water that was depleted in sulfate and enriched in CO
2
. The
siderites are on average depleted by ~5 per mil in
13
C over
limestone in the succession that has typical open marine

13
C
PDB
values of close to zero (Beukes et al., 1990). Based on
these data Beukes et al. (1990) concluded that the ocean-
water column could have been stratified with respect to the
carbon isotope composition of dissolved inorganic carbon.
However, recent analyses of a very large set of limestone and
dolostone from the Agouron Drilling Project (Schroeder et
al., 2006) revealed no difference in carbon isotope composi-
tion from shallow-shelf to deep basinal carbonates in the
Campbellrand succession (Fischer et al., in press). All car-
bonate rocks have consistent and uniform
13
C values around
an average of 0.5 per mil independent of depositional depth
(Fischer et al., in press). Excursions toward negative
13
C
carb
values of ~5 per mil are observed in dolostone and limestone
interbedded with black carbonaceous shale in the deep-shelf
succession. These excursions are explained by the uptake of
light organic carbon from decay of organics in shales into the
carbonate during compaction and late lithification (Fischer et
al., in press).
In contrast to the rather uniform, normal marine isotopic
composition of limestones and dolostones, siderite in Brunos
BIF, the Kamden BIF, and the lower part of the Kuruman
Iron Formation is characterized by very variable and negative

13
C
carb
values that range from 3 to 14 per mil (Fischer et
al., in press). This variability in carbon isotope composition of
the siderites is attributed to microbial iron respiration during
diagenesis by Fischer et al. (in press). This process would
have released
12
C-enriched CO
2
and converted iron oxides to
Fe
2+
, promoting siderite precipitation. Thus, all of the iron
present in the deep-water iron formations of the Campbell-
rand succession and overlying lower part of the Kuruman
Iron Formation could have been deposited as ferric oxyhy-
doxides, some of which was diagenetically converted to
siderite. This model is consistent with the preservation of mi-
crocrystalline hematite in Brunos BIF (Fig. 10A). A diage-
netic origin for the siderite also renders unnecessary the car-
bon isotope stratification of the ocean-water column as
proposed by Beukes et al. (1990).
Sideritic iron formation is preferentially developed imme-
diately adjacent to carbonaceous deep-water carbonates
and/or shales (Fig. 11). All three lithologic units fit a single
positive lognormal correlation trend of Al
2
O
3
versus organic
carbon contents, with sideritic iron formation containing the
lowest, carbonaceous limestone and dolostone containing in-
termediate, and shale containing the highest concentrations
or organic carbon (Fig. 12A). Klein and Beukes (1989) inter-
preted this trend to indicate that organic matter was trans-
ported with clay minerals into the deep-shelf environment as
suspended load from high primary productivity environments
on the shallow carbonate shelf. However, oxide-rich (mainly
hematite) BIF does not follow the same trend as it contains
similar Al
2
O
3
concentrations but considerably less organic
carbon than sideritic BIF (Fig. 12A). This is tentatively at-
tributed to the effective degradation of organic matter in an
oxygen-bearing water column before it could be incorporated
into the sediment. This explanation is supported by the ob-
servation that trace amounts of organic matter preserved in
oxide BIF are least depleted in
13
C of all of the deep-shelf
lithofacies, suggesting that only some of the most refractory
kerogen was preserved (Beukes et al., 1990). An alternative
interpretation would be that hematite-rich BIF could only be
preserved if the organic carbon influx was low relative to fer-
ric oxyhydroxide accumulation.
Deposition of hematite facies iron formation: Klein and
Beukes (1989) assumed that the deposition of ferric oxyhy-
droxides, preserved as microcrystalline hematite, was due to
inorganic oxidation of dissolved ferrous iron by free oxygen
and decoupled from biologic activity (Fig. 11). An alternative
that has received considerable attention recently envisages
that deposition of ferric oxyhydroxides in iron formations
prior to the rise of oxygen could have been brought about by
anoxygenic Fe(II)-oxidizing photo-autotrophic bacteria
(Ehrenreich and Widdel, 1994). Reconstruction of deposi-
tional environments in the Campbellrand carbonate platform
succession argues very strongly against this mechanism for
precipitation of ferric oxyhydroxides, because the chemocline
along which precipitation of ferric oxyhydroxides took place
was located at below the photic limit (at maximum depth of
~150 m in clear-water modern oceans like the Mediterranean
Sea according to Lalli and Parsons, 1997), at water depths of
up to 700 to 1,000 m (Klein and Beukes, 1989; Fischer et al.,
in press). Direct evidence for deposition of iron formation at
water depth beyond the limits of the photic limit comes from
the distribution of conophytic microbialites relative to iron
formations in the carbonate platform succession. This rela-
tionship is best illustrated in the Gamohaan (Fig. 10A) and
Kamden depositional systems tract (Fig. 9), where conophytic
stromatolites that are commonly equated with the presence
of phototactic bacteria (Walter et al., 1976) are stratigraphi-
cally well detached from the depositional setting of the iron
formations.
It could be argued that the chemocline for iron deposition
was situated at shallow depth and that the ferric oxyhydrox-
ides settled to deep water from suspension. However, there is
no evidence for that in the succession as limestones are highly
depleted in iron and there is also no sign of any iron pigmen-
tation in any of the very abundant early marine calcite ce-
ments in the fenestral microbialites, especially in the cono-
phytic stromatolite units. It is thus highly unlikely that
anoxygenic photosynthetic Fe(II)-oxidizing bacteria played a
role in deposition of the iron formations in the Campbellrand
succession. Tice and Lowe (2004) proposed that dissolved
iron could have been removed from the mixed upper water
column of Archean oceans, which extended to well below the
photic zone, by primary precipitation of siderite under an oxy-
gen-free atmosphere with high p
CO
2
. This explanation cannot
apply to Brunos BIF, because it contains both microcrys-
talline hematite and magnetite. Free oxygen derived from
photochemical reactions (Cairns-Smith, 1978) can be ruled
out because it has been shown to be a very ineffective mech-
anism of oxygen production (Konhauser et al., 2007). This
leaves oxidation of ferrous iron by a flux of free oxygen de-
rived from oxygenic photosynthesis by cyanobacteria as the
only viable alternative for precipitation of ferric oxyhydrox-
ides and removal of iron from the water column of the Camp-
bellrand basin. Similar arguments are presented by Towe
(1994) and Beukes (2004) for precipitation of oxide-facies
iron formation in general.
MAJOR IRON FORMATIONS AT THE ARCHEAN-PALEOPROTEROZOIC BOUNDARY 23
0361-0128/98/000/000-00 $6.00 23
Beukes_Gutzmer 6/11/08 7:40 AM Page 23
These sedimentological arguments for the presence of at
least some free oxygen in the upper water column of the
Campbellrand basin, prior to the so-called rise of atmos-
pheric oxygen at ~2.32 Ga (Bekker et al., 2004), have found
support from chemical and biochemical proxies. Most impor-
tantly Waldbauer et al. (in press) convincingly illustrate that
biomarkers indicative of the presence of eukaryotes and
cyanobacteria are primary components of carbonaceous shale
and carbonates in the Agouron cores (Shroeder et al., 2006)
that intersected slope and basinal facies of the Campbellrand
carbonate platform. Furthermore, both molybdenum (Wille
et al., 2007) and multiple sulfur isotope studies (Kamber and
Whitehouse, 2007; Kaufman et al., 2007) from slope and basi-
nal facies of the carbonate platform suggest the presence of
free oxygen in the water column.
Diagenesis: Diagrams that depict organic carbon-sulfur-
iron relationships (Raiswell and Berner, 1985) in combination
with the composition of minerals containing redox-sensitive
elements like iron and/or manganese (Berner, 1981, May-
nard, 1982) provide a means of establishing bottom water and
diagenetic conditions in a basin and the availability of oxygen
to those environments. A cross plot of organic carbon and sul-
fur contents of different lithofacies in the Gamohaan-Kuru-
man transition zone displays a positive correlation with S/C
ratio of approximately 1:3 (Fig. 12B), which is typical for ma-
rine sediments in which organic carbon degradation by sul-
fate-reducing bacteria takes place. However, extremely low
concentrations of sulfur in oxide-facies iron formation, com-
bined with similarly low degrees of pyritization and organic
carbon contents, illustrate severe restriction of any microbial
metabolism involving sulfur (Klein and Beukes, 1989). An ab-
sence of sulfides combined with the presence of hematite
suggests an environment in which degradation of organic
matter took place through aerobic bacterial respiration. Sim-
ilarly, the diagenetic environment of the sideritic iron forma-
tions can be described as suboxic or dysaerobic (Fig. 12B) and
iron respiration must have been the dominant process of
degradation of organic matter. In pyritic carbonaceous shales
and limestones with high degrees of pyritization, in contrast,
the diagenetic environment is best described as anaerobic
(Fig. 12B) and bacterial sulfate respiration must have been an
important process in the degradation of organic matter. This
is supported by sulfur isotope studies of nodular pyrite in the
Gamohaan succession (Cameron, 1982; Kamber and White-
house, 2007; Kaufman et al., 2007). Interestingly, iron-rich
shales interbedded with the lower part of the Kuruman Iron
Formation immediately above the Gamohaan Formation dis-
play low degrees of pyritization in the presence of high or-
ganic carbon contents (Fig. 12B). This indicates severe re-
striction of dissolved sulfate with no limit to iron supply in the
immediate surroundings of iron formation deposition (Klein
and Beukes, 1989).
Based on the above, the diagenetic conditions in various
sedimentary facies of the Campbellrand carbonate platform
succession can be defined. The Gamohaan transgressive
systems tract that led to deposition of Brunos BIF serves as
an excellent example, as it comprises virtually the whole
range of sedimentary facies of the carbonate platform (Fig.
10A). In this succession lower slope and deep-shelf car-
bonaceous shale and limestone with very abundant nodular
pyrite experienced anaerobic diagenetic environments (Fig.
10A). In contrast, hematite BIF of the deep-shelf margin re-
flects aerobic diagenetic environments, whereas adjacent
siderite BIF reflects dysaerobic ones (Fig. 10A). The con-
spicuous disappearance of nodular pyrite in carbonaceous
distal deep-shelf limestone immediately below Brunos BIF
(Fig. 10A) probably reflects a combination of sulfate deple-
tion and slightly higher oxygen availability. In shallower facies
lower in the Gamohaan Formation, dark gray pyrite-poor
limestone and dolostone in upper slope, shelf margin, and la-
goonal deposits are tentatively considered to have experi-
enced dysaerobic diagenesis, because no detailed S, C, and
degrees of pyritization values are available. Light gray inter-
tidal carbonate rocks depleted in both pyrite and organic car-
bon, in contrast, probably reflect aerobic depositional and di-
agenetic conditions (Fig. 10A).
Modified depositional model for
Brunos transgressive systems tract
General outline: A model composed of the mixing of two
water masses can explain sedimentary facies as they appar-
ently existed during deposition of Brunos BIF. During initial
stages of flooding of the Campbellrand carbonate platform,
an upper somewhat aerobic iron-depleted water mass overlay
a lower iron- and silica-enriched one (Fig. 13). Iron formation
and ankerite-banded chert were deposited from the lower
water mass, whereas limestone and organic carbon were de-
posited from the upper one. Although this model is similar to
that proposed by Klein and Beukes (1989, 1992a, b; Fig. 11),
it differs in several critical aspects. Most importantly, in the
revised model, the chemocline along which iron precipitation
takes place does not conform to the mixing zone between sur-
face and deep ocean water, a depth of about 100 m in mod-
ern oceans (Elderfield and Greaves, 1982), but at much
greater depth within the deep ocean water column. This in-
terpretation is based on the observation that open marine car-
bonates, with
13
C ~0 per mil, predominate in basinal facies
of the Campbellrand carbonate platform succession (Fischer
et al., in press). It was only during occasional transgressions
that an iron- and silica-enriched water mass encroached onto
the deep-shelf margin for deposition of ankerite-banded
chert and/or iron formation. This implies that the water in the
Transvaal basin was not permanently stratified with respect to
iron and silica but that stratification was intermitantly brought
about by upwelling of the deeper water mass (Fig. 13). Posi-
tive Eu anomalies that are more pronounced in the iron for-
mation than in deep to shallow-shelf limestone (Fig. 10E)
suggest that the deeper water mass carried a more significant
hydrothermal component. Because the open marine water
column from which limestone was precipitated also carried a
significant hydrothermal component as exemplified by posi-
tive Eu anomalies (Kamber and Webb, 2001), it appears un-
likely that the chemocline coincided with a sharp density
boundary. Instead, a rather broad redox chemocline is envis-
aged along which iron oxyhydroxide precipitation took place.
The depth and position of this redox-chemocline would have
been determined by the rate of supply of oxygen and organic
matter from the water column above and dissolved ferrous
iron from a water column below (Fig. 13). If rates of hy-
drothermal input were high relative to oxygen supply, the
24 BEUKES AND GUTZMER
0361-0128/98/000/000-00 $6.00 24
Beukes_Gutzmer 6/11/08 7:40 AM Page 24
redox boundary would have shifted to shallower depths and
vice versa. Based on similar observations, Sumner (1997a)
came to the conclusion that the Neoarchean-Paleoprotero-
zoic ocean system was more akin to the modern ocean than
the noncirculating, density stratified ocean often envisaged
for early Precambrian times (e.g., Holland, 1984; Klein and
Beukes, 1989).
The thickness distribution of the Kamden Member clearly
illustrates that iron deposition was most prominent in areas of
enhanced upwelling of deep hydrothermally enriched water
(Fig. 9). The Kamden depositional systems tract also indicates
that SiO
2
precipitation did not take place from the same
chemocline as ferric oxyhydroxides because chert associated
with ankeritized limestone has a wider lateral distribution
both onto the shallow shelf and the distal deep shelf than
siderite accumulation. Some of the chert is of replacement
origin, suggesting that the hydrothermally influenced deep-
water mass was more acidic (Fig. 13) promoting dissolution of
carbonate and precipitation of silica (Siever, 1971). In con-
trast, the water column from which limestone precipitated
must have been more alkaline with a pH of at least ~7.8 to
have allowed cementation by marine calcite cements, which
are abundant even in deep-subtidal environments (Sumner,
1997a, b). A pH gradient from deep hydrothermally domi-
nated low-pH water mass to more alkaline pH in the shallow
water mass would have favored the formation of ferric oxyhy-
droxide but not SiO
2
precipitation. However, because the pH
of ocean water is rather constant at ~8.1 to 8.2, temperature
rather than pH controls SiO
2
solubility (Siever, 1971). The
deep open ocean water in the basin should have been cold
and the simplest explanation for SiO
2
precipitation would be
that the deeper hydrothermal water had slightly higher tem-
perature than ambient ocean water and that silica precipi-
tated along a certain thermocline where it became oversatu-
rated in the zone of mixing between the two water masses
(Fig. 13).
The revised depositional model also predicts that the ocean
water column contained sufficient dissolved oxygen for the
oxidation of dissolved ferrous iron, but not of Ce
3+
and Mn
2+
,
i.e., the latter two remained in solution across the iron redox
chemocline into shallow-water environments (Fig. 13). Suffi-
cient mixing across the chemocline evidently took place so
that the positive Eu anomaly of hydrothermal fluids con-
tributed to the deep ocean water column was transferred into
MAJOR IRON FORMATIONS AT THE ARCHEAN-PALEOPROTEROZOIC BOUNDARY 25
0361-0128/98/000/000-00 $6.00 25
Hydrothermal
Deep Water
Mixing
Zone
Photic Limit
Wave Base
Surface Mixing
Fe Redox
Boundary
W
a
r
m
e
r
C
o
l
d
e
r
W
a
r
m
e
r
SHALLOW SHELF SLOPE DEEP SHELF
Light Grey
Dolomite
Aerobic
D
y
s
a
e
r
o
b
ic
Mn Dolomite
Limestone
Fe Dolomite
D
y
s
a
e
ro
b
ic
A
n
a
e
r
o
b
i
c
D
y
s
a
e
r
o
b
ic
Aerobic
3+
Fe
O
2
O
2
SO
4
3
+
F
e
D
is
s
o
lu
t
io
n
-
56
Fe
56
Fe +
Legend
Ankeritic Dolomite
SO
4

33
S =0
O
2
Transfer
O
r
g
a
n
i
c
C
S
settle
Hem Hem
Hem
Hem
Intertidal carbonate
Lagoonal carbonate
Shelf margin mounds
Rippled calcarenite
Conophytic stromatolite
Contorted microbialite
Pyritic contorted
microbialite
Laminated microbialite
Ankerite-banded chert
Sid Siderite MBIF
Hem Hematite MBIF
SiO
2
2+
Fe
Sid
O
2
33
S -
SO
4
SO
4
Mix
+SO
4
+
S
2+
Mn
3+
Eu
3+
Ce
SiO
2
33
S
SiO
2
A
n
a
e
r
o
b
ic
2+
Fe
2+
Mn
3+
Eu
3+
Ce
Sio
2
p
H
SO
4
O
2
Depletion
Dissolved
Precipitated
Photosynthetic
O
2
O
2
Atmosphere
Ferric oxyhydroxide
Organic carbon
O
2
Input
56
Fe
FIG. 13. Model for deposition of iron formation in deep-shelf environments in association with shallow-shelf to deep-
water platform carbonates. The model is largely based on stratigraphic relationships observed in the Gamohaan-Kuruman
succession as depicted in Figure 10. Important differences with the earlier model of Klein and Beukes (see Fig. 11 above)
include the following: The chemocline along which iron minerals precipitated is located at depth far below the mixed sur-
face zone of the ocean. The depth of this chemocline is determined by the ratio of oxygen supply from the upper levels of
the ocean versus supply of dissolved iron from upwelling deep hydrothermally enriched water. Iron oxyhydroxide is the only
primary precipitate of iron and takes place at near-neutral alkalinity with involvement of iron-oxidizing microaerophyllic
chemolithoautotrophs far below the photic zone. In this model all siderite is derived from dissimilatory iron reduction. Sil-
ica precipitation from slightly heated hydrothermal deep water is brought about by a decrease in temperature in the zone
where it mixes with cold normal ocean water. Elemental sulfur (S
0
) derived from photolysis of SO2 in the atmosphere carries
positive S
33
mass independent signature and sinks through the slightly oxygenated upper water column until it comes into
contact with acidic hydrothermal water and in that mixing zone it is oxidized to sulfate by sulfur-oxidizing chemolithoau-
totrophs. Depending on rates of organic matter supply and burial versus degradation by various microbial respiratory
processes, diagenetic environments ranged from aerobic to anaerobic in sediment on the floor of the depository. See text for
more detail.
Beukes_Gutzmer 6/11/08 7:40 AM Page 25
the shallow-water column. Because of lack of data it remains
impossible to decide whether the positive Eu anomaly of the
upper ocean water column was (1) a permanent composi-
tional characteristic of the global ocean system at the time, (2)
a feature specific to the times at which a hydrothermally in-
fluenced water mass present in the global oceans encroached
upon the carbonate platform, or (3) a Transvaal-Hamersley
basin-specific phenomenon.
Similar to the model by Klein and Beukes (1989), organic
matter in the revised depositional model is seen to have been
mainly supplied from areas of high photosynthetic primary
productivity and abundant nutrients on the shallow carbon-
ate shelf (Fig. 13). Oxygen produced from photosynthesis
was transferred through the water column and the atmos-
phere into the open ocean system (Fig. 13). The rate of sup-
ply of this organic carbon versus oxygen supply would have
determined diagenetic conditions in the sediment. Observa-
tions suggest that in shallow-water environments, aerobic
degradation of organic matter was more effective and led to
formation of light gray dolomites in intertidal environments
and dark gray dolomite and limestone under dysaerobic dia-
genetic conditions in shallow-subtidal lagoonal, shelf margin,
and upper slope environments (Fig. 13). Further down the
slope organic matter supply outpaced oxygen supply and
pyritic limestone formed under strictly anaerobic conditions
(Fig. 13). Most distal from the carbonate shelf, in areas of
very low primary organic matter supply, oxygen reacted with
ferrous iron to precipitate ferric oxyhydroxides that accumu-
lated with silica on the deep shelf to form iron formation
(Fig. 13). More proximal to the carbonate shelf organic car-
bon availability was such that degradation of organic matter
through iron respiration took place to form sideritic iron for-
mation with negative
13
C
carb
values under dysaerobic diage-
netic conditions. However, some of the redissolved iron re-
mained in solution and became available to form pyrite and
ferruginous dolomite in deep-shelf and lower slope environ-
ments (Fig. 13).
The revised depositional model also implies that anaerobic
photoautotrophic Fe(II)-oxidizing bacteria could not have
been responsible for precipitation of ferric oxyhydroxides, be-
cause the redox boundary for iron was situated far below the
photic zone in the water column (Fig. 13). Instead, dissolved
oxygen is preferred as the electron acceptor for oxidation of
ferrous iron. Especially at low oxygen concentrations and
near-neutral pH, precipitation of ferric oxyhydroxides may
have been effectively accelerated by microaerophyllic iron-
oxidizing bacteria that do not require light (Emerson and
Moyer, 2002). This is a possibility to be addressed more fully
in the discussion of the Kuruman Iron Formation.
Explanation of iron isotope signatures: The depositional
model also helps explain some of the iron isotope data that
have recently become available from the Gamohaan-Kuru-
man transition zone (Fig. 10B; Johnson et al., 2003, 2008;
Rouxel et al., 2005). These reveal a trend from
56
Fe values
highly depleted in
56
Fe for pyrite in carbonaceous deep-water
limestone and shale to values markedly enriched in
56
Fe in
oxide-facies iron formation (Fig. 10B). This trend is attrib-
uted to cycling of
54
Fe-enriched water from dysaerobic dia-
genetic environments in which sideritic iron formation
formed, into adjacent anaerobic diagenetic environments of
pyritic carbonaceous limestone (Fig. 13). In this model,
56
Fe
is sequestered into ferric oxyhydroxides by oxidation of dis-
solved ferrous iron from a hydrothermal deep-water (Johnson
et al., 2003) source with
56
Fe = 0.4 0.2 per mil (value for
high-temperature fluids at modern midocean ridge spreading
centers; Beard et al., 2003). This process leaves residual iron
in solution that is depleted in
56
Fe relative to the source. In
fully aerobic environments, as depicted for oxide-facies iron
formation deposition (Fig. 13), very little if any of the resid-
ual
56
Fe-depleted dissolved ferrous iron would have been
able to circulate into the overlying water column. However,
under dysaerobic diagenetic conditions in areas of siderite-fa-
cies banded micritic iron formation deposition,
56
Fe-depleted
ferrous iron may well have been transferred into the overly-
ing water column or sediment pore water and became avail-
able to form pyrite with marked negative
56
Fe values (Fig.
10B), if sufficient sulfide was available. Development of
dysaerobic diagenetic environments in areas of iron formation
deposition would also have favored formation of siderite
through dissimilatory ferric iron reduction (Fig. 13). This
process is capable of producing
56
Fe values for dissolved
ferrous iron that range from 0.5 to 2.5 per mil relative to
the initial ferric oxyhydroxide precipitate (Beard et al., 1999;
Croal et al., 2004; Johnson et al., 2005; Crosby et al., 2007).
Residual dissolved ferrous iron derived from this process may
again have entered into adjacent dysaerobic and anaerobic
basin and sediment pore water and become available for

56
Fe-depleted pyrite formation and/or formation of iron-en-
riched dolomites (Fig. 13).
26 BEUKES AND GUTZMER
0361-0128/98/000/000-00 $6.00 26
MID MRA
Aerobic-Dysaerobic
Anaerobic
(Chert)
MD
Residual
Dysaerobic
(Matrix Pyrite)
MIXING
TRIANGLE
mSR
(Nodular Pyrite)
193nmA
SD
-15 -10 -5 0 5 10 15 20 25 30
A
t
m
S
O
4
SO
4
S

S
O
4
S
S and H S from S
2
A
t
m
S

4
0
4
8
12
3
3
s
(

)
34
S()
SO
4
FIG. 14. Crossplot of
33
S vs.
34
S, indicating compositional fields of
pyrites in the Gamohaan-Kuruman trnsition zone and how they could be ex-
plained. Note the triangular field of values (indicated by small arrows) that
could have resulted from mixing of microbial sulfur reduction (mSR) values
with values at the end of the linear correlation line MRA, i.e., the Mount
McRae array of Ono et al. (2003). The 193-nm array is after Farquhar et al.
(2001). atm = atmospheric, MID = mass independent fractionation, MD =
mass dependant fractionation, mSR = microbial sulfur reduction, SD = mi-
crobial elemental sulfur disproportionation. See text for more detail.
Beukes_Gutzmer 6/11/08 7:40 AM Page 26
Explanation of sulfur isotope signatures: Deep-water car-
bonates and ankerite-banded chert in immediate strati-
graphic surroundings of Brunos BIF and the Kuruman Iron
Formation show large and varied mass independent fraction-
ated (Farquhar et al., 2000, 2001) sulfur isotope signatures
(Fig. 10C, D). Mass independent fractionation of sulfur iso-
topes (S-MIF) is widely considered a proxy for highly anaer-
obic environments (Pavlov and Kasting, 2002). This is in ap-
parent disagreement with the depositional model advocated
for the deposition of large volumes of ferric oxyhydroxide in
iron formations in the presence of free oxygen at water depths
of several hundred meters (Fig. 13). Currently, two sets of
multiple sulfur isotope data are available for the Gamohaan
succession. The first, by Kamber and Whitehouse (2007), is
based on in situ secondary ion mass spectrometer analyses of
pyrite grains (Fig. 10C); the second are bulk-rock analyses by
Kaufman et al. (2007; Fig. 10D). The Kamber and White-
house (2007) data reveal positive
33
S values for fine matrix
pyrite in deep-water limestone with contorted mats that are
offset to the left of the so-called Mount McRae array of Ono
et al. (2003). In contrast, nodular pyrite from deep-shelf con-
torted mat carbonates (Fig. 10A) has negative
33
S with a
range of mass-dependent fractionated
34
S values (Fig. 10C).
Matrix pyrite in ankerite-banded chert follows a third distinct
trend with positive
33
S and a negative correlation with vari-
able positive
34
S values (Fig.10C). In the dataset of Kaufman
et al. (2007), pyrite in samples from deep-water carbonates
below Brunos BIF generally follow the Mount McRae array,
whereas samples from within and above the BIF plot mostly
to the left of the Mount McRae array, with a range of positive

33
S values coupled to negative
34
S values (Fig. 10D).
The various mass independent multiple sulfur isotope
trends in the Gamohaan succession (Fig. 10C, D) do not find
a conclusive explanation at present, especially since the iso-
topic composition and flux of reduced (S
o
) versus oxidized
(SO
4
2+
) species from the atmosphere, which would have car-
ried the mass independent fractionation signature, remain
uncertain. Ono et al. (2003) suggested that the Mount
McRae array, with positive linear correlation between
33
S
and
34
S, represents the input mass independent fractiona-
tion array for atmospheric aerosols of S
o
and SO
4
2
in the Late
Archean; yet the array determined experimentally from 193-
nm wavelength UV-driven photolysis of SO
2
has a negative
slope (Fig. 14). However, there is some experimental data
available indicating that the Mount McRae array falls be-
tween the array produced at 193-nm UV radiation and arrays
formed at longer wavelengths (Farquhar and Wing, 2003). It
is thus reasonable to assume that the Mount McRae array ac-
tually represents the average input isotopic composition of
mass independent fractionated atmospheric S
o
and SO
4
2
aerosols formed from UV photolysis of volcanic SO
2
at
shorter and longer wavelengths in Archean times (Ono et al.,
2003; in press; Farquhar et al., 2007).
If indeed the Mount McRae array represents the composi-
tion of mass independent fractionated Neoarchean atmos-
pheric S
o
and SO
4
2
aerosols, the reason for the spread of
33
S
values as recorded by different textural generations of pyrite
(Fig. 10C, D) needs to be addressed. The depositional model
proposed here for the deposition of iron formations may pro-
vide some clues (Fig. 13). First, the model predicts the ab-
sence of dissolved sulfate from the upwelling hydrothermally
enriched deep ocean water, because an oversupply of iron
would have removed all sulfur as sulfide near hydrothermal
vents. Secondly, the overlying deep-ocean water column
should have contained some dissolved sulfate mainly derived
from fallout of atmospheric aerosols with negative
33
S (Fig.
14) and a minor component, with
33
S near zero, from ter-
restrial weathering under low atmospheric oxygen levels (Fig.
13). Sulfate in the shallow-ocean water column would thus
have carried a negative
33
S signature (Fig. 13). With depth,
in the aerobic mixing zone along which ferric oxyhydroxides
were precipitated, distal from organic matter supply, sulfate
would rapidly have become highly diluted by upwelling sul-
fate-depleted deep ocean water (Fig. 13). Proximal to the car-
bonate platform, in areas of abundant organic matter supply,
precipitation of pyrite under anaerobic conditions would also
have depleted sulfate from the upper oceanic water column
(Fig. 13).
In addition to sulfate, particulate elemental sulfur from UV
photolyses of volcanic SO
2
and with positive
33
S was sup-
sulfur is known to be metastable and highly insoluble in oxy-
genated environments (Konhauser, 2007). It would thus have
settled through the water column until conditions conducive
to dissolution were met either in the water column or on the
floor of the Campbellrand basin. However, our model sug-
gests that physical-chemical environments were very variable
in the water column close to and along the floor of the Camp-
bellrand basin, due to different rates of supply of organic car-
bon, oxygen, and sulfate from the upper water column, and
ferrous iron from the deeper water column (Fig. 13). For ex-
ample, in the basinal environment of the Campbellrand car-
bonate shelf where less organic matter was available, particu-
late S
o
would have settled through the upper oxygenated
water column to the mixing zone with the upwelling deep
water. At this interface it could have been oxidized to sulfate
by sulfur-oxiding bacteria at near-neutral pH and very low
oxygen concentrations, environments in which modern sulfur
oxidizers are known to thrive in the presence of reduced sul-
fur compounds (Konhauser, 2007). Under aerobic conditions
and with little organic carbon present, it is conceivable that
this sulfate would have remained in solution and mixed in
with that of the shallower water column, thus explaining the
absence of sedimentary pyrite in oxide-facies iron formations.
The reaction pathway for particulate atmospheric S
o
would
have been more complex in environments where organic car-
bon was in abundant supply, i.e., near or on the shallow car-
bonate shelf (Fig. 13). Here, elemental sulfur could have par-
ticipated in a number of biogeochemical cycles, including
anaerobic anoxygenic photosynthesis, heterotrophic reduc-
tion, and chemolithoautotrophic oxidation (Konhauser, 2007).
In addition, products derived from elemental sulfur cycling
could have become mixed in with dissolved sulfate in the
shallow-water column and/or with products of sulfate reduc-
tion in environments that favored this microbial process.
With this as background, the varied mass independent frac-
tionation of sulfur isotopes in the Gamohaan succession can
be evaluated. Almost all pyrite displays positive
33
S values
and are thus thought to document isotope ratios inherited
from atmospheric S
o
. This is perhaps not unexpected as most
MAJOR IRON FORMATIONS AT THE ARCHEAN-PALEOPROTEROZOIC BOUNDARY 27
0361-0128/98/000/000-00 $6.00 27
Beukes_Gutzmer 6/11/08 7:40 AM Page 27
plied to the oceanic water column (Figs. 13, 14). Elemental
of the samples, both from Kamber and Whitehouse (2007)
and Kaufman et al. (2007), are from distal deep-shelf carbon-
ate and shale where availability of SO
4
2
may have been strictly
limited. In contrast, more proximal to the carbonate platform
and distal to upwelling iron-enriched deep-ocean water, dia-
genetic nodular pyrite that formed in highly carbonaceous
limestone under anaerobic environments displays negative

33
S combined with a variable
34
S value. This would have
been typical for microbial sulfate reduction in Archean envi-
ronments that had an ample supply of both sulfate and or-
ganic matter (Fig 14), and with dissolved
56
Fe-depleted fer-
rous iron the limiting factor for pyrite formation (Fig. 13).
Whole-rock matrix pyrites with sulfur isotope compositions
that plot along the Mount McRae array in deep-shelf carbon-
ate and shale from below Brunos BIF (Fig. 10D) could obvi-
ously have formed from a mixture of the two isotopically dis-
tinct sulfide sources that originated as products of
photochemical mass independent dissociation of volcanic
SO
2
. Along the resulting mixing line (Fig. 14) it is possible
that the S
o
end member could have carried a range of
33
S
mass independent fractionation values. However, for the
Mount McRae array to be preserved no significant mass-de-
pendent
34
S/
32
S fractionation could have taken place any-
where along the mixing line. The ideal environment to gener-
ate such an array would have been one in which sulfate was
limited so that all of it was transformed to H
2
S by sulfate re-
ducers. At the same time the environment should have been
dysaerobic so that particulate S
0
could be transformed to H
2
S
by anaerobic sulfur-reducing bacteriaa process known to
produce no mass-dependent isotope fractionation (Kaufman
et al., 2007; Philippot et al., 2007). This, however, is exactly
the environment depicted in our model for deposition of the
carbonates and shales that display the Mount McRae array,
namely an SO
4
2
-depleted distal deep-shelf dysaerobic envi-
ronment immediately adjacent to an area in which deposition
of iron formation took place (Fig. 13).
Depositional Systems in the Kuruman-Griquatown
Iron Formation Succession
Facies architecture and physical sedimentology
The shelf margin of the Campbellrand carbonate platform,
which separated thick stromatolitic shelf carbonates in the in-
terior of the Kaapvaal craton from thin basinal carbonates off
the craton (Fig. 8), had an arcuate shape (Beukes, 1987),
which trended to the northwest in the south and to the north-
east in the north. A north-south cross section of the As-
besheuwels Subgroup (Fig. 15A) indicates that this shelf-
basin configuration (Fig. 15B) persisted with deposition of
the iron formations. However, variations in thickness between
shallow- and deep-shelf environments are inverted. This is es-
pecially well illustrated to the south of the Griquatown fault
zone where the iron formations reach maximum thickness
(Fig. 15A).
The cross section illustrates that the Kuruman Iron Forma-
tion, which is essentially a deep-water microbanded banded
micritic iron formation (Fig. 15A), has layer-cake stratigraphy
with zones rich in stilpnomelane lutite, representing altered
volcanic ash layers (LaBerge, 1966), representing regional
time lines in the basin (Beukes, 1980b, 1983, 1984). In con-
trast, such extensive tuffaceous markers beds are absent from
the overlying clastic-textured, relatively shallow water, Gri-
quatown Iron Formation (Beukes, 1983) that displays in-
terfingering relationships between lutitic banded micritic and
granular iron formation (Fig. 15A). It is specifically the layer-
cake microbanded banded micritic iron formation succession
of the Kuruman Iron Formation that displays the largest
thickness variation with strata showing a threefold increase in
thickness from the shallow shelf into the basin (Fig. 15A).
The Kuruman Iron Formation overlies carbonaceous shale
that forms the top of the underlying Campbellrand carbonate
platform succession (Figs. 8, 15A) with gradational contact. In
broad terms the Kuruman succession is composed of a thin
basal chert-rich unit, overlain by a thick central unit of mi-
crobanded oxide-facies banded micritic iron formation with
interbeds of stilpnomelane lutite (Stofbakkies-Buisvlei-Or-
ange View Members) and an upper unit of laminated and mi-
crobanded greenalite-siderite micritic iron formation with
minor banded micritic iron formation, i.e., the Westerberg-
Geduld Members (Fig. 15A). This subdivision applies from
the basin onto the shallow shelf with one notable exception
a thin but laterally very persistent banded granular iron for-
mation unit known as the Ouplaas Member, which caps the
Kuruman Iron Formation on the shelf (Fig. 15A).
Immediately above the Ouplaas Member the Griquatown
Iron Formation commences on the shallow shelf with a thin
unit of laminated and microbanded greenalite-siderite mi-
critic iron formation. The latter micritic iron formation corre-
lates laterally with laminated greenalite-siderite micritic iron
formation of the uppermost Geduld Member of the Kuruman
Iron Formation in the basinal environment. In the basinal en-
vironment the transition of the Geduld Member of the Kuru-
man Iron Formation into the Middelwater Member of the
Griquatown iron formations is represented by a zone of in-
terbedded laminated and lutitic silicate-facies micritic iron
formation (Fig. 15A). The Griquatown Iron Formation is
characterized by the presence of abundant sideritic granular
iron formation and flat-pebble to edgewise chert-pebble con-
glomerates (ferudites, Fig. 2B) on the shallow shelf and sili-
cate-siderite felutite in the basin. Based on the nature and
distribution of the granular iron formation beds, the Griqua-
town Iron Formation is subdivided into three members on
the shallow shelf, namely a lower Danielskuil Member com-
posed of siderite-hematite and siderite lutites with interbeds
of granular iron formation, a middle Skietfontein Member
with abundant and laterally persistent ferudite beds, and an
upper Skietfontein Member of lutitic silicate-siderite micritic
iron formation with ferudite interbeds (Fig. 15A).
Microbanded banded micritic iron formation in
the lower part of the Kuruman Iron Formation
The basal chert-rich Kliphuis Member of the Kuruman
Iron Formation (Fig. 15A) is composed of stacked macrocy-
cles of carbonaceous shale and ankerite-banded chert (Fig.
16A). The ankerite-banded chert units are in some cases
overlain by thin beds of micritic limestone and/or dolostone
with contorted and laminated microbial mat structures and
thin carbonate turbidite beds (Fig. 16A). Both the ankerite-
banded chert and thin carbonate beds are similar in character
to those associated with the most distal deep-shelf facies of
28 BEUKES AND GUTZMER
0361-0128/98/000/000-00 $6.00 28
Beukes_Gutzmer 6/11/08 7:40 AM Page 28
the underlying Campbellrand carbonate platform succession
(Fig. 8; see description of Kamden Member and Brunos BIF
above). Beukes (1978, 1983, 1984) considers the macrocycles
to be constructed of regressive carbonaceous shale beds over-
lain by transgressive ankerite-banded chert, which is in turn
capped by thin progradational deep-shelf carbonates that
were sourced from a distal shelf margin (Fig. 16A). This indi-
cates that carbonate platform sedimentation persisted, prob-
ably in the far interior of the Kaapvaal craton, for some time
after early deep flooding and onset of deposition of the Ku-
ruman Iron Formation.
However, continued submergence of the craton and car-
bonate shelf led to rapid disappearance of carbonaceous shale
and micritic carbonate beds upward in the Kuruman succes-
sion. Concomitant with the disappearance of carbonaceous
shale, thin volcanic ash beds, composed of stilpnomelane lu-
tite sometimes with well-preserved pseudomorphs of glass
shards (LaBerge, 1964; Beukes, 1980b, 1983, 1984; Van Wyk,
1987; Pickard, 2003) become prominent. The stilpnomelane
lutite beds represent ferruginized felsic pyroclastic material
(Van Wyk, 1987; Pickard, 2003) that indicate an oversupply of
iron in the environment so as to have allowed replacement of
MAJOR IRON FORMATIONS AT THE ARCHEAN-PALEOPROTEROZOIC BOUNDARY 29
0361-0128/98/000/000-00 $6.00 29
0 100 km
Pomfret
S
e
c
t
io
n
L
in
e
Kuruman
Danielskuil
Griquatown
Koegas
Prieska
24
28
N
B
B
a
s
i
n
S
h
e
l
f
S
h
e
l
f
M
a
r
g
i
n
Danielskuil Kuruman
Pomfret
950
900
800
700
600
500
400
300
200
100
0
K
U
R
U
M
A
N
I
F
A
S
B
E
S
H
E
U
W
E
L
S
S
U
B
G
R
O
U
P
G
R
I
Q
U
A
T
O
W
N
I
F
M
i
d
d
e
l
w
a
t
e
r
M
e
m
b
e
r
G
e
d
u
l
d
M
e
m
b
e
r
Westerberg
Member
O
r
a
n
g
e
V
i
e
w
M
e
m
b
e
r
B
u
i
s
u
l
e
i
M
e
m
b
e
r
S
t
o
f
b
a
k
k
i
e
s
M
e
m
b
e
r
Kliphuis
Naute Sh
Koegas
m
Koegas
Subgroup
G
a
m
o
h
a
a
n
F
o
r
m
a
t
i
o
n
C
A
M
P
B
E
L
L
R
A
N
D
S
U
B
G
R
O
U
P
Kliphuis
G
r
o
e
n
w
a
t
e
r
K
U
R
U
M
A
N
I
F
R
i
r
i
e
s
G
R
I
Q
U
A
T
O
W
N
I
F
Skietfontein Member
Danielskuil Member
Sid
Sid
Sid
Sid
S
id
Sid
Sid
Sid
Sid Sid
Sid
Sid
Sid
Sid
Sid Sid
Sid
Sid
Sid
H
H
H
H
H
H
H
H
H H
H H
H
H
H
H H
H
H
H
H
S
id
S
id
S
i
d
S
id
S
id
S
id
Sid
Sid
Sid
Sid
Sid
Sid
H
e
m
-
M
a
g
H
e
m
-
M
a
g
Hem-Mag
Hem
-M
ag
Hem-Mag
H
e
m
-
M
a
g
M
a
g
-
S
id
Hem-Mag
M
a
g
-
S
id
Mag-Sid
Mag-Sid
Mag-Sid
Mag-Sid
Mag-Sid
Mag-Sid
M
a
g
-S
id
M
a
g
-S
id
M
a
g
-S
id
M
ag
-S
id
Griquatown
Fault Zone
Basinwards Basin
Carbonaceous shale
Ankerite-banded chert
Sid Siderite-magnetite BMIF
Stilpnomelane lutite-BMIF macrocycles
Hem-Mag Hematite-magnetite BMIF
T T Stilpnomelane-rich tuffaceous units
Mag-Sid Magnetite-siderite hematite BMIF
Wavy and podded hematite-magnetite BMIF
Green-Mag Greenalitic magnetite-siderite hematite BMIF
Greenalite-siderite laminated IF
Iron-silicate lutitic IF and BIF
Siderite-greenalite lutitic IF and BIF
Hematitic siderite lutite
Chert-banded stilpnomelane lutite
Granular IF
Edgewise and flat-pebble conglomerate
Sid
H
S
tilp
A
T T T T
T
T T T T T T T
T
T
T T T T T T T
T
T
T
T
T
T
T
T
T
T
T
T
T
T
T
T
T
T T
T
T
T
T T
T T
T T
T
T
T
T
T
T T
T T T
T
T
T
T
T
T
T
T
T
T T
T T
T T T T
T
T T
T T
T T
T
T
T T
T
T T
T T
T
T
T
T T
T
T
T
T T T
T T T T T T T T
T T
T
T
T
T
T
T
T
T
T
T
T
T
T
T
T
T T
T
T
T
T
T
T
T
T
T
T
T
T
T
T
T T
T
T
T
T
T
T
T
T T T
T
T
T
T
Griquatown
Makganyene Diamictite
~ 450 km
Shelf
~ 450 km
Shelf
Pietersberg Member
Stilp
Nauga Fm
B
a
s
i
n
a
l
c
a
r
b
o
n
a
t
e
s
Stromatolitic
Shelf Carbonates
Matlipani
Green-Mag
G
re
e
n
-M
a
g
Green-Mag
Green-Mag
Green-Mag
OUPLAAS MEMBER
Matlipani
FIG. 15. A. South-north cross section of the Asbesheuwels Subgroup, which includes the Kuruman and Griquatown Iron
Formations (see text for description and more detail). B. Position of section line for (A) above and also the section of the
Campbellrand carbonate platform depicted in Figure 8.
Beukes_Gutzmer 6/11/08 7:40 AM Page 29
30 BEUKES AND GUTZMER
0361-0128/98/000/000-00 $6.00 30
C C
C C
CCC
C C
C
1
-
1
0
m
1
-
1
0
m
Shale
Graded bedded intramicsparite
intramicrite
Ankerite-banded chert
consisting of intraclastic
chert ankeritic
intramicrite and
cryptocrystalline chert
bands
Pyritic carbonaceous
shale with bands and
pods of cryptocrystalline
chert in upper part
Intramicrite
Stilpnomelane lutite
Chert
Siderite-microbanded chert
Volcanic Ash
Enhanced hydrothermal
SiO supply
2
Minimum SiO supply
2
High compaction
Chert
Stilpnomelane Lutite
Siderite-microbanded
chert
Siderite microbanded
BMIF
Siderite magnetite
microbanded BMIF
Magnetite-siderite
microbanded BMIF
Wavy banded
hematite-magnetite
microbanded BMIF
Podded
hematite-magnetite
microbanded BMIF
Volcanic Ash
High SiO supply
2
D
e
c
r
e
a
s
i
n
g
S
i
O
s
u
p
p
l
y
2
I
n
c
r
e
a
s
e
d
c
e
m
e
n
t
a
t
i
o
n
S
i
O
2
D
e
c
r
e
a
s
i
n
g
c
e
m
e
n
t
a
t
i
o
n
Primary
cryptocrystalline
chert beds in
hemipelagic
shale
Hemipelagic shale
Carbonate turbidites
Carbonate turbidite
units (replaced by
chert or ankeritized)
interbedded with
primary crypto-
crystalline chert
bands
Carbonate turbidite
Deep
Shelf
Hemipelagic
shale in
carbonate-
starved
deep basin
Proximal
Deep basin
in front of
carbonate
shelf
Distal
Origin Environment
LHSST
HSST
MF
TSST
LSST
LHSST
FALL RISE
Maximum exposure
M
i
c
r
o
a
e
r
o
p
h
i
l
i
c
F
e
-
o
x
i
d
i
z
i
n
g
b
a
c
t
e
r
i
a
i
n
v
o
l
v
e
d
?
A
B
Sea level curve
13
Corg()
-14 -9 -40 -35 -30 -25 -20
13
Ccarb() Relative % chert
FIG. 16. A. Interpretation of depositional environments and sequence stratigraphic development of sedimentary succes-
sions associated with ankerite-banded chert in the Kliphuis Member of the Kuruman Iron Formation (modified after
Beukes, 1984). B. Composition and genesis of a typical stilpnomelane lutite-banded micritic iron formation macrocycle of
the Stofbakkies and Orange View Members of the Kuruman Iron Formation. Modified after Beukes (1983, 1984) with vari-
ation in isotopic composition of iron carbonates and organic carbon schematically indicated from analyses of macrocycle 2 of
Kaufman (1996). Acronyms in A are identified in Figure 9 caption.
Beukes_Gutzmer 6/11/08 7:40 AM Page 30
original volcanic glass by stilpnomelane. The volcanic ash
beds are responsible for development of stilpnomelane lutite-
banded micritic iron formation macrocycles (Beukes, 1980b)
in the succession. In their complete form the macrocycles are
composed of stilpnomelane lutite overlain by siderite mi-
crobanded chert, siderite-, magnetite-siderite, and hematite-
magnetite banded micritic iron formation (Fig. 16B). Thick-
nesses of chert mesobands decrease upward in macrocycles
and also become more wavy and podded as the relative thick-
nesses of oxide-rich mesobands increase. The result is that
the upper parts of macrocycles are typically composed of pod-
ded microbanded hematite-magnetite banded micritic iron
formation (Fig. 16B). However, immediately below a stilp-
nomelane lutite bed there is a rapid increase in silica content
with development of conspicuous siderite-microbanded chert
beds (Fig. 16B). These macrocycles are clearly of mixed vol-
canogenic-chemogenic origin with increased silica precipita-
tion and cementation directly associated with the episodes of
explosive volcanic activity (Fig. 16B). The observation that
macrocycles can be incomplete and cut off by the deposition
of a stilpnomelane bed suggests that thickness and complete-
ness of macrocycles was controlled by the frequency and tim-
ing of volcanic events. For example, in the stilpnomelane-rich,
tuffaceous time-line markers in the succession (Fig. 15A) the
macrocycles are typically only composed of stilpnomelane lu-
tite, siderite microbanded chert, and siderite banded micritic
iron formation. One of the best examples is the Buisvlei Mem-
ber that thickens into the basin and overlies the lower Stof-
bakkies Member of the Kuruman Iron Formation (Fig. 15A).
Complete stilpnomelane lutite-banded micritic iron formation
macrocycles are best developed in the lower Stofbakkies
Member (Fig. 14A). The latter comprises of 16 well-defined
macrocycles, many of which end in podded hematite-mag-
netite banded micritic iron formation (Fig. 16B).
Macrocycles in the Orange View Member (Fig. 15A) are
somewhat more irregular in thickness and composition.
There is a subtle change in the composition of the banded mi-
critic iron formation in the macrocycles to the top of the Or-
ange View Member, in that podded hematite-magnetite
banded micritic iron formation beds become scarce, whereas
iron silicate, in the form of greenalite and minnesotaite, be-
come abundant together with magnetite and siderite in
banded micritic iron formation (Fig. 15A). Individual macro-
cycles are generally thicker and more chert rich in the basin
than on the shallow shelf. Thus, podded microbanded
hematite-magnetite banded micritic iron formation is best
developed on the shallow shelf and interfingers laterally into
the basin with magnetite-siderite banded micritic iron forma-
tion (Fig. 15A). This distinct lateral facies change is also ob-
served vertically within individual macrocycles, i.e., the iron
mineral assemblage changes from sideritethrough mag-
netiteinto hematite-rich banded micritic iron formation
(Fig. 16B). The abundance of chert associated with pyroclas-
tic stilpnomelane beds (Fig. 16B) strongly suggests that more
silica was supplied to water in the basin immediately prior to
and following volcanic events. In the context of lateral facies
variations it appears likely that silica deposition and/or ce-
mentation was more effective in the basinal environment,
because individual stilpnomelane lutite-BIF cycles and the
microbanded BIF are thicker in the basin than on the shallow
shelf (Beukes, 1984). Volcanic events evidently influenced the
depositional environment for iron, as minerals with abundant
ferrous iron, namely siderite and magnetite, occur in proxim-
ity to stilpnomelane beds in the basin. Whether this change
reflects variation in Eh-pH conditions in the water column
above the sediment-water interface and/or in the diagenetic
environment will be addressed later in this contribution. The
location and nature of the volcanic sources to the pyroclastic
beds remain unknown. However, thickening of stilpnome-
lane-rich units into the basin would suggest the centers were
situated somewhere to the west of the Kaapvaal craton
(Beukes, 1983, 1984).
Transition from microbanded to
clastic-textured iron formation
The transition from microbanded into clastic-textured iron
formation in the upper Westerberg and Geduld Members of
the Kuruman Iron Formation is associated with a marked
change in lithologic character. With the exception of one thin
unit marking the top of the Westerberg Member, well-de-
fined stilpnomelane lutite-banded micritic iron formation
macrocycles are absent and the succession is composed of mi-
crobanded and finely laminated greenalite-siderite micritic
iron formation, without well-defined chert mesobands (Fig.
15A). The micritic iron formation is composed of muddy
chert-cemented greenalite-siderite laminae separated by mi-
crobands of siderite and/or magnetite (Beukes, 1978, 1980b,
1984; Van Wyk, 1987). In the basin, microbands become less
abundant upward in the succession, as laminated greenalite-
siderite micritic iron formation becomes interbedded with lu-
titic iron silicate-siderite micritic iron formation of the basinal
Middelwater Member of the Griquatown Iron Formation
(Fig. 154A). Finely graded greenalite-rich laminae in the
mixed laminated-microbanded greenalite-siderite micritic
iron formation yield the first indication for reworking of sed-
iment by weak currents. The greenalite-rich laminae have a
typical muddy or lutitic character and may contain tiny silt-
sized peloids and angular chiplike fragments of greenalite and
siderite. They probably represent thin event beds in the basin
produced by reworking of fine mud on the bottom of the de-
pository by weak currents in deep water of the basin during
major storms (Beukes, 1984; Beukes and Klein, 1990).
On the shallow shelf, in the uppermost part of the Geduld
Member, such storm event beds become very pronounced and
more abundant. They are often represented by graded coarse
peloidal and/or intraclastic granular iron formation bands in-
terbedded with laminated greenalite-siderite micritic iron for-
mation (Beukes and Klein, 1990). These coarse-grained beds
suggest stronger currents affecting the sediment surface dur-
ing storms and thus shallower water on the shelf than in the
basin. The Ouplaas Member itself (Fig. 15A) comprises wavy
and lenticular grainstone interbedded with felutite, amalga-
mated grainstone beds, and chert flat-pebble and edgewise
conglomerate (Beukes, 1980b; Beukes and Klein, 1990).
Clastic-textured Griquatown Iron Formation
On the shallow shelf the Griquatown Iron Formation is com-
posed of several coarsening-upward lutitic banded micritic
iron formation-granular iron formation increments of sedi-
mentation or parasequences (Fig. 15A; Beukes, 1980b, 1983,
MAJOR IRON FORMATIONS AT THE ARCHEAN-PALEOPROTEROZOIC BOUNDARY 31
0361-0128/98/000/000-00 $6.00 31
Beukes_Gutzmer 6/11/08 7:40 AM Page 31
1984). The succession coarsens upward with extensive sheet-
like flat-pebble and edgewise conglomerates present in the
upper Skietfontein and Pietersberg Members (Fig. 15A). Into
the basin granular iron formation units pinch out and the suc-
cession is essentially composed of lutitic micritic and banded
micritic iron formation. The lutitic micritic and banded lutitic
micritic iron formation on the shallow shelf are siderite rich,
whereas those of the basin are dominated by iron silicates
(Fig. 15A), including greenalite minnesotaite and stilpnome-
lane. It appears as if iron silicates were preferentially de-
posited in deeper water environments. A similar relationship
is developed in coarsening-upward lutitic banded micritic
iron formation-granular iron formation parasequences, with
laminated greenalite-siderite micritic iron formation or
greenalite-rich lutitic banded micritic iron formation repre-
senting the deep-water lithofacies (Fig. 17). Chert mesobands
in the lutitic banded micritic iron formations typically have
poorly defined graded contacts with adjacent felutite bands
and appear to be largely of diagenetic origin, i.e., they repre-
sent chert-cemented hard bands.
Magnetite and hematite are present in varying amounts in
the lutitic micritic iron formation. Hematite is especially con-
spicuous in siderite lutite of the lower two coarsening-upward
increments of sedimentation of the Danielskuil Member of
the Griquatown Iron Formation (Fig. 15A). The hematite ap-
pears to be a very early diagenetic phase being intimately in-
tergrown with micritic siderite and iron silicate (Van Wyk,
1987, Beukes and Klein, 1990). Very fine grained magnetite
that may also be of early diagenetic origin occurs locally but is
minor in abundance in comparison to coarser, euhedral mag-
netite. The latter is likely to be of low-grade metamorphic ori-
gin, as it replaces diagenetic iron minerals and crosscuts the
sedimentary fabric (Beukes, 1980b, 1984, Van Wyk, 1987).
Near Pomfret, prominent pyroclastic stilpnomelane lutite
beds with associated thick chert mesobands are present in
greenalite-siderite lutite of the Griquatown Iron Formation
(Fig. 15A; Van Wyk, 1987). These illustrate that volcanic ac-
tivity persisted during deposition of the Griquatown Iron
Formation. The absence of distinct pyroclastic stilpnomelane
lutite beds in very shallow-water clastic-textured iron forma-
tions thus does not imply cessation of volcanic activity but
rather that volcanic material became mixed with other sedi-
ment in well-agitated water.
Granular iron formation of the Griquatown Iron Formation
is typically composed of poorly sorted intraclastic and granu-
lar (peloidal) wackestone and packstone, with well-sorted
grainstones (Fig. 2B) not very common. Granules have clearly
been derived from erosion of earlier lutitic BIF with granules
of varied composition mixed in the same bed. Chert granules
are most common followed by polymineralic granules com-
posed of femicrite cemented by chert. However, monominer-
alic greenalite, stilpnomelane, siderite, hematite, and mag-
netite granules and clasts are also present (Beukes, 1980b;
Van Wyk, 1987). Grains coated by hematite and magnetite,
siderite and/or iron silicate are occasionally present. The ex-
ception is one well-sorted grainstone bed in the third coars-
ening-upward unit of the Griquatown Iron Formation (Fig.
15A) that contains abundant Fe oxide coated granules and
some oolites (Beukes, 1978, 1980). The Skietfontein and
Pietersberg Members that contain abundant chert-pebble
conglomerates in the upper part of the Griquatown Iron For-
mation on the shelf also contain abundant septarian chert
concretions and chert bands with dewatering cracks (Beukes,
1980b, 1984). Pisolitic chert concretions abound in some of
the grainstone beds in the Danielskuil Member (Beukes,
1980b, 1984).
Granular bands in the Griquatown Iron Formation have
obviously been reworked by wave action because they often
display wavy internal bedding similar to hummocky cross
stratification formed by storm-wave action (Beukes and
32 BEUKES AND GUTZMER
0361-0128/98/000/000-00 $6.00 32
1
5
-
5
0
m
Flat-pebble conglomerate
Grainstone
Rise Fall
Sea level
LSST
HSST
TSST
LSST
F=Fall inflection
point
F
Transgressive Lag
Reworked BIF
Shelf below storm
wave base
Shelf above storm
wave base
Shelf above
normal wave base
Deep shelf
Transgressive lag
Reworked BIF
Interbedded siderite-rich
grainstone and lutitic MIF
Siderite hematite
lutitic BMIF
Greenatite-siderite
laminated MIF
Flat-pebble conglomerate
Grainstone
Felulite
FIG. 17. Interpretation of depositional environments and sequence stratigraphic development of shallowing-upward mi-
critic iron formation-granular iron formation increments of sedimentation in the shelf environment of the clastic-textured
Griquatown Iron Formation (modified after Beukes, 1983, 1984). Acronyms as in Figure 9.
Beukes_Gutzmer 6/11/08 7:40 AM Page 32
Klein, 1990). Others are lenticular in form and apparently
represent original compacted starved wave ripples interbed-
ded with felutite. The grainstone beds are virtually without
exception cemented by chert, which results in wavy and
lenticular interbedded grainstone and felutite iron formation
units having a typical chert-banded appearance.
In contrast to the laminated greenalite-siderite micritic iron
formation of the Geduld and Westerberg Members that al-
most certainly formed below normal wave base, the more
massive and lutitic iron formation of the Griquatown Iron
Formation was probably deposited in the well-mixed and ag-
itated environments above normal wave base. The abundance
of silt- to sand-sized peloids and tiny intraclasts floating in a
felutite matrix (Beukes, 1980b, 1984) supports this interpre-
tation and indicates that iron-rich muds were occasionally re-
suspended by currents together with earlier lithified or partly
lithified endoclastic particles.
It is difficult to define the absolute water depth during de-
position of the clastic-textured iron formation, but in analogy
to modern marine basins (Boggs, 2005), it may be expected
that the wave base in the Transvaal basin was between 50- and
100-m water depth, with occasional currents generated as
deep as ~200 m. Deposition of the granular iron formation
evidently took place above wave base, i.e., on the shallow
shelf and toward the top of the succession; where flat-pebble
and edgewise chert conglomerates are abundant, water could
have been very shallow (<10 m). Beukes (1983, 1984) even
suggested that some of the flat-pebble conglomerates may
have been washed out by storm waves onto supratidal flats.
However, we caution against the assumption that precipita-
tion of iron-rich muds took place along the water-atmosphere
boundary. There are two important reasons for such caution,
namely, shallow-water granular iron formation that caps
coarsening-upward increments of sedimentation in the basin
comprises of particles that were eroded from preexisting lithi-
fied iron- and chert-rich sediment, and coated grains that
provide direct evidence for precipitation of iron minerals in
very shallow agitated water (Simonson, 2003; Clout and Si-
monson, 2005) are scarce. In sequence stratigraphic terms
the laterally extensive granular iron formation units of the
Ouplaas Member and the Griquatown Iron Formation (Fig.
15A) can thus be considered to represent sequence bound-
aries in the succession formed during periods of rapid rate of
sea-level fall and minimum accommodation space in the
basin, with resultant endogenetic reworking of earlier sedi-
ment that originally formed in deeper water (Fig. 17). For
chert-rich endoclastic femicrite granules and chert flat-peb-
bles to have formed, silica cementation must have taken place
along or immediately below the sediment-water interface. On
the shallow shelf, chert hardgrounds developed in iron-rich
muds at or immediately below the sediment-water interface.
During storm events, such chert hardgrounds would have
been broken up by storm-wave action to form large clasts,
while iron-rich mud was placed in suspension (Beukes and
Klein, 1990).
Nature of basin fill
The lithofacies distribution in the Asbesheuwels Subgroup
(Fig. 15A) indicates pronounced shallowing of depositional
environments, thus implying that accommodation space in
the basin decreased through time. Importantly, the iron for-
mations of the Asbesheuwels Subgroup display the same
range of depositional environments, ranging from deep- to
very shallow-shelf settings, as developed in the Campbellrand
carbonate platform succession. Therefore, deposition of iron
formation as a result of transgression and drowning of a car-
bonate platform (Figs. 10, 13) can only apply to the early
stages of deposition of the Kuruman Iron Formation. How-
ever, by the time that deposition of shallow-water clastic-tex-
tured iron formation of the Griquatown Iron Formation took
place, the conditions in the basin favored deposition of iron
formation over the entire available range of water depths.
In the following sections reasons for this pronounced
change in character of the basin will be explored, and the pos-
sible influence of water depth on the physical-chemical con-
ditions in the water column, precipitation mechanisms of iron
and silica, and diagenetic environments in the different depo-
sitional facies of iron formation will be discussed. For this
purpose, three major depofacies in the iron formation suc-
cession are distinguished, namely (1) stacked stilpnomelane
lutite-banded micritic iron formation cycles developed in
deep-water environments, (2) chert-poor mixed micro-
banded-laminated greenalite-siderite micritic iron formation
formed at intermediate-water depths, and (3) mixed lutitic
and granular iron formation formed in shallow water.
The three major depofacies of the Asbesheuwels Subgroup
iron formations have identical REE distribution patterns
(Beukes and Klein, 1990; Klein and Beukes, 1992a). This im-
plies that granular iron formation and deep-water banded mi-
critic iron formation were precipitated from a hydrothermally
influenced water mass of similar composition. This would re-
quire that compositional stratification of the ocean-water col-
umn that was invoked to explain the deposition of limestone
in shallow water versus banded micritic iron formation in
deep-water environments (i.e., Kamden and Brunos BIFs)
was replaced by an environment in which iron- and silica-en-
riched deep ocean water invaded the shallow shelf. The obvi-
ous differences in mineralogical composition of the three iron
formation depofacies, from dominantly oxide-facies in deep-
water microbanded banded micritic iron formation, to
siderite- and silicate-facies micritic iron formation, and gran-
ular iron formation in shallow water, can thus not be attrib-
uted to differences in source and composition of the hy-
drothermally influenced water mass but rather to shifts in the
depths of chemoclines and possible physical-chemical gradi-
ents in the water column, mechanisms of iron and silica de-
position, and the nature of postdepositional diagenetic envi-
ronments.
Origin of stilpnomelane lutite-banded
micritic iron formation macrocycles
Primary precipitates and diagenetic alteration: Stilpnome-
lane lutite-banded micritic iron formation macrocycles in the
lower part of the Kuruman Iron Formation are thought to be
of mixed volcano-sedimentary origin (Fig. 16B) and have a
number of important characteristics that permit reconstruc-
tion of their environments of deposition and diagenesis. The
macrocycles were deposited in a deep-water environment far
below storm-wave base and the photic zone, as indicated by
their close association with deep-shelf carbonates (Fig. 13).
MAJOR IRON FORMATIONS AT THE ARCHEAN-PALEOPROTEROZOIC BOUNDARY 33
0361-0128/98/000/000-00 $6.00 33
Beukes_Gutzmer 6/11/08 7:40 AM Page 33
This interpretation is supported by the lateral continuity of
pyroclastic beds and the preservation of fine microbanding in
banded micritic iron formation units without indication of any
current activity (Beukes, 1984; Klein and Beukes, 1989).
Furthermore, the iron-mineral assemblages in banded mi-
critic iron formation units of the macrocycles are directly
linked to the relative abundance of chert and amount of com-
paction that took place in the banded micritic iron formation
prior to silica cementation and lithification (Fig. 16B). Rela-
tively chert-poor podded hematite-magnetite iron formation
in the succession evidently underwent most compaction,
whereas chert-rich thick-banded siderite iron formation and
siderite-microbanded chert, immediately adjacent to pyro-
clastic stilpnomelane lutite beds, are least affected by com-
paction (Fig. 16B). This implies that highly compacted mi-
crobanded hematite-magnetite iron formation with podded
chert were probably exposed for longer to marine bottom and
phreatic water in an open diagenetic environment, in com-
parison to chert-rich magnetite-siderite and siderite iron for-
mation. In the latter case, early silica cementation could have
restricted permeability and limited exchange between early
diagenetic pore fluids and marine bottom water.
Within a specific banded micritic iron formation facies the
same principle applies, only on a smaller scale. This is illus-
trated in podded microbanded hematite-magnetite banded
micritic iron formation in which compaction ratios of up to 90
percent can be calculated from thicknesses between iron
oxide microbands in chert pods versus laterally equivalent mi-
crobands adjacent to these pods (Fig. 18A).
Finally, the observation that microcrystalline (dusty)
hematite, with grain sizes <1 m, is the paragenetically oldest
iron mineral, leads to the conclusion that ferric oxyhydroxide
represented the primary iron precipitate in all of the miner-
alogical facies of banded micritic iron formation macrocycles.
For example, chert bands in microbanded hematite-mag-
netite banded micritic iron formation, owe their light red to
purple color to inclusion of dusty hematite enclosed in mi-
crocrystalline quartz or chalcedony (Fig. 18B; Beukes, 1978;
Van Wyk, 1987; Klein and Beukes, 1989). The most likely
explanation is that this dusty hematite formed from dehydra-
tion of ferric oxyhydroxide particles that settled from suspen-
sion at the same time that gelatinous silica accumulated on
the floor of the depository. Such primary ferric oxyhydroxides
may have settled from suspension or formed along the sedi-
ment-water interface.
Recrystallization of the chalcedony leads to exsolution of
the hematite to form microcrystalline quartz surrounded by
platy, subhedral microcrystalline hematite with grain sizes <5
m (Fig. 18B). The same applies to well-preserved uncom-
pacted and early-cemented hematite microbands enclosed in
chert mesobands. These dusty hematite microbands have a
felutitic appearance, indicating settling from suspension
rather than in situ precipitation along the sediment-water in-
terface (for a more detailed description, see Trendall, 1973).
The paragenetically early hematite generations are replaced
by fine-grained subhedral magnetite of supposed diagenetic
origin (i.e., not the coarse euhedral magnetite of metamor-
phic origin; Trendall, 1973; Beukes, 1980b, 1984; Van Wyk,
1987; Klein and Beukes, 1989) or by micritic or microsparitic
siderite, also of diagenetic origin (Beukes, 1978). Such diage-
netic magnetite and siderite are invariably coarser grained
than the hematite they replace. This obliterates some of the
fine sedimentary details that are only preserved by dusty
hematite. Replacement of early hematite by fine magnetite is
best illustrated in podded micritic iron formation where less-
compacted fine microbands preserved in chert pods are re-
placed by magnetite in highly compacted microbanded iron
oxide mesobands outside of the pods (Fig. 18A).
Siderite was most probably formed by dissimilatory ferric
oxyhydroxide reduction in the presence of organic carbon.
This is not only illustrated by examples where siderite is in-
tergrown with fine hematite and/or replaces hematite
(Beukes, 1978) but also by variable and highly negative

13
C
carb
values (Fig. 16B; Kaufman, 1996) and by iron isotope
systematics (Johnson et al., 2003, 2005, 2008). In magnetite-
siderite banded micritic iron formation, diagenetic siderite
and magnetite show rather close association with different
lithotypes. Chert mesobands typically contain siderite
mesobands, whereas microbanded iron-rich mesobands that
have undergone more compaction are mainly composed of
magnetite, with only minor siderite.
Dissimilatory Fe(III) reduction can lead to the formation of
either siderite or magnetite, depending on the availability of
organic matter and the fate of Fe(II) and CO
2
formed from
the reaction,
CH
2
O + 8H
+
+ 4Fe(OH)
3
4Fe
2+
+ CO
2
+11H
2
O. (1)
The reaction increases alkalinity and would favor precipita-
tion of siderite, which is less soluble than other carbonates
(Stumm and Morgan, 1996), as follows:
Fe
2+
+ CO
2
+ H
2
O FeCO
3
+ 2H
+
. (2)
Equation (2) illustrates that the stability of siderite is depen-
dent on p
CO
2
(Fig. 19A). The reaction would thus be favored
by restricted diagenetic environments in which p
CO
2
can lo-
cally build up in pore water of the sediment in immediate
vicinity of the source of organic matter. Rapid burial and early
34 BEUKES AND GUTZMER
0361-0128/98/000/000-00 $6.00 34
Hematite
C
h
e
r
t
p
o
d
Dispersed
dusty
hematite
Dusty hematite
Chalcedony
microspheroids
Micro-
specularite
Recrystallization
Ankerite
Microbanded
magnetite
A
B
FIG. 18. A. Schematic illustration of the lateral replacement of hematite
microbands preserved in relatively uncompacted chert pods by magnetite in
compacted iron-rich mesobands outside of pods. B. Illustration of transfor-
mation of very early diagenetic dusty hematite, preserved in chalcedonic
chert, to microspecularitic hematite with recrystallization of chert to micro-
quartz (modified after Beukes, 1984).
Beukes_Gutzmer 6/11/08 7:40 AM Page 34
silica cementation would have created ideal conditions for the
formation of siderite in banded micritic iron formation and its
particular association with chert mesobands. In a restricted
diagenetic environment, carbonate incorporated into siderite
(eq. 2) would be expected to reflect the carbon isotope com-
position of organic matter that was oxidized by ferric iron (eq.
1). This is manifested in the negative
13
C
carb
values the
siderite displays (Fig. 16B; Kaufman, 1996). Because the dia-
genetic pore water reservoir of dissolved carbonate would
comprise a mixture of CO
2
derived from dissimilatory Fe(III)
reduction (Kaufman, 1996) and CO
2
derived from bottom
water of the basin,
13
C
carb
values of siderite would be expected
to vary considerably. The same principle applies to the Fe(II)
incorporated in siderite, explaining variable
56
Fe values re-
ported by Johnson et al. (2003; Fig. 10B). These observations
suggest a causal link between early diagenetic siderite forma-
tion and the deposition of Fe(OH)
3
and organic mattera
matter to be discussed in more detail in a following section.
The formation of early diagenetic magnetite, rather than
siderite, may have been favored, among other factors, in en-
vironments of limited organic carbon supply. In such envi-
ronments partial reduction of Fe(OH)
3
may take place so that
magnetite can form from the reaction,
Fe(OH)
3
+Fe
2+
Fe
3
O
4
+ 2H
2
0 + 2H
+
. (3)
Similar to reaction (2) this reaction produces protons and thus
could be driven by dissimilatory Fe(III) reduction that con-
sumes protons (reaction 1).The stability of magnetite is, how-
ever, independent of p
CO
2
; the formation of magnetite, rela-
tive to siderite, is thus favored at low p
CO
2
(Fig. 19A). This
may explain the preferential occurrence of magnetite in
highly compacted iron-rich mesobands that apparently re-
tained permeability and open exchange between pore and
basin water for longer than adjacent sideritic chert-cemented
mesobands. It may be expected that p
CO
2
in diagenetic pore
waters did not build up in such open diagenetic environments
to levels required for the formation of siderite. However, it is
also imperative that pore water p
CO
2
may have fluctuated lo-
cally between the stability fields of magnetite and siderite
(Fig. 19A), thus explaining the close spatial relationship be-
tween these two minerals in magnetite-siderite banded mi-
critic iron formation.
The formation of siderite by dissimilatory Fe(III) reduction
implies that at least some of the Fe(II) in the authigenic
siderite was derived directly from precursor Fe(OH)
3
. How-
ever, in the transformation of Fe(OH)
3
to magnetite, none of
the Fe(II) may in fact have been derived from reduction of
the specific Fe(OH)
3
but could merely have been added from
surplus ferrous iron present in deep bottom water (eq. 3). Di-
rect evidence for this mechanism of magnetite formation
comes from kerogen-free hematite microbands in chert pods
that are replaced by early diagenetic magnetite in compacted
iron-rich mesobands outside of pods (Fig. 18A). In such ex-
amples, it is clear that ferric oxyhydroxide (now hematite) be-
came trapped in silica without organic matter so that dissimi-
latory iron reduction could not take place. It is only in the
open diagenetic environments of the compacted iron-rich
mesoband that excess ferrous iron was available and trans-
formed the ferric oxyhydroxide to magnetite by addition of
iron from basinal phreatic water. Johnson et al. (2003, 2008)
illustrated that iron isotope studies lend support to the for-
mation of magnetite and siderite by the processes outlined
above.
Source of organic carbon and mechanism of precipitation
of iron: The conclusion that dissimilatory iron reduction was
involved in the formation of early diagenetic siderite and
magnetite requires burial of organic matter together with
ferric oxyhydroxides during sedimentation of banded mi-
critic iron formation in stilpnomelane lutite-banded micritic
iron formation macrocycles. It is highly unlikely that this or-
ganic carbon was transported as detrital particles from the
MAJOR IRON FORMATIONS AT THE ARCHEAN-PALEOPROTEROZOIC BOUNDARY 35
0361-0128/98/000/000-00 $6.00 35
4 8 12
g
r
e
e
n
a
l
i
t
e
pH
SiO (amorphous)
2
a
l
o
g
S
i
O
2
t
o
t
-
8
-
6
-6
-
7
-4
-2
0
2
G
R
E
E
N
A
L
I
T
E
A
M
O
R
P
H
O
U
S
S
I
L
I
C
A
-6
-4
-2
0
4 8 12
S
I
D
E
R
I
T
E
-8
-10
-12
pH
a
+
+
l
o
g
F
e
t
o
t
Fe(OH)
3
(S)
g
r
e
e
n
a
lit
e
s
id
e
r
it
e
-
3
.
5
s
id
e
r
it
e
-
6
.
5
++
Fe
(aq)
MH
40
60
20
0
80 FMQ
12 8
pH
4
H -H O
2 2
F
e
(
O
H
)
2
H O-O
2 2
+
+
+
F
e
(
a
q
)
l
o
g
f
O
2
(
s
)
A B C
FIG. 19. A. Solubility of Fe(OH)3 and Fe(OH)2 relative to greenalite and siderite at a log total iron activity of 5 at 25C.
Stability fields of siderite depicted at pCO
2
of 3.5 and 6.5. B. Solubility of greenalite (dashed curves) for three log ferrous
iron activities of 8, 7, and 6. Also shown is the solubility curve of amorphous silica against pH and log activities of SiO2.
C. Solubility of siderite and greenalite (in equilibrium with amorphous silica at constant log silica activity of 4 at low oxygen
fugacity and at 25C) relative to log activity of ferrous iron and pH. All three diagrams have been modified from Eugster and
I-Ming Chou (1973).
Beukes_Gutzmer 6/11/08 7:40 AM Page 35
shallow Campbellrand carbontate platform, i.e., an area of
high primary productivity (Fig. 13), because the latter had
drowned before stilpnomelane lutite-banded micritic iron
formation macrocycles in the upper part of the Stofbakkies
Member and the Orange View Member took place (Beukes,
1984). Therefore, it is assumed that the organic carbon was
endogenetic and either derived from pelagic and/or benthic
microbes that thrived in the depositional environment in
which the macrocycles were deposited. Because of limited
nutrient supply to the open ocean, bioproductivity in the
photic zone above areas of BIF deposition was likely limited
(Klein and Beukes, 1989). Since anaerobic Fe(II)-oxidizing
photoautotrophs could not have deposited banded micritic
iron formation in deep-water environments (see discussion
above and Fig. 13), this process would also not have produced
organic matter. This leaves chemolithoautotrophic iron-oxi-
dizing bacteria that are not light dependent as possible pri-
mary producers of organic matter (Emerson and Moyer,
1997; Brown, 2006). Chemolithoautotrophic iron-oxidizing
bacteria are responsible for precipitation of large amounts of
ferric oxyhydroxides under microaerobic and near-neutral pH
conditions at many modern deep-sea hydrothermal vent sites
(Emerson and Moyer, 2002). Although the bacteria are
mesophiles growing optimally between 25
o
and 30
o
C, they are
well adapted to grow in cold water; they are thought to be re-
sponsible for the precipitation of abundant ferric oxyhydrox-
ides at hydrothermal vent sites at depths of more than 1,000
m, where the temperature is only slightly above that of ambi-
ent seawater (Emerson and Moyer, 2002). The vent waters
are typically enriched in Fe(II), Mn(II), and CO
2
but de-
pleted in H
2
S (Emerson and Moyer, 2002) which, together
with temperatures slightly above that of ambient deep seawa-
ter, slightly acidic to neutral pH and low oxygen concentra-
tions, would fit very well with the environment envisaged for
deposition of deep-water microbanded iron formations fol-
lowing drowning of the Campbellrand carbonate platform
(Fig. 13).
Precipitation of iron by chemolithoautotrophs under mi-
croaerobic conditions is also suited to explain other important
characteristics of banded micritic iron formation, namely the
apparent rapid precipitation of iron minerals leaving no time
for sorption of trace metals from ocean water (Klein and
Beukes, 1989; Bau and Dulski, 1996) and of extremely low
concentrations of organic carbon especially in oxide-facies
banded micritic iron formation (Klein and Beukes, 1989;
Beukes and Klein, 1990; Kaufman, 1996). Chemolithoau-
totrophs synthesize only small amounts of biomass while oxi-
dizing large amounts of substrate (Konhauer, 2007). In case of
iron oxidizers, this is illustrated by the reaction through which
ferric hydroxide is precipitated:
6Fe
2+
+ 0.5O
2
+ CO
2
+ 16H
2
O
CH
2
O + 6Fe(OH)
3
+ 12 H
+
. (4)
This microbially mediated reaction is more than 60 times
faster than the inorganic reaction for precipitation of ferric
oxyhydroxide under microaerobic conditions (Konhauser et
al., 2002). The relatively small amount of primary organic
matter created by this reaction may be easily degraded by a
number of processes, including aerobic, nitrate, and iron
respiration. Aerobic respiration would have been restricted to
the mixing zone between the deep hydrothermally enriched
water mass and the upper aerobic water mass (Fig. 13). Ni-
trate and iron respiration, in contrast, could have taken place
in dysaerobic-anaerobic environments in the water column,
along the sediment-water interface, or during diagensis below
the sediment-water interface.
Precipitation of iron by chemolithoautotrophs could also
account for the general lack of microfossils in the deep-water
banded micritic iron formation, a fact that has been used to
argue against microbial involvement in the deposition of iron
formation (e.g., Klein and Beukes, 1989; Klein, 2005). At
modern-day hydrothermal vent sites three morphotypes of
ferric oxyhydroxides are formed, namely filaments, sheaths,
and irregular amorphous particles, of which the latter is the
dominant morphology at most sites (Emerson and Moyer,
2002). The filaments and sheaths are produced by filamen-
tous iron-oxidizing bacteria that form benthic and flocculent
gelatinous ferric oxyhydroxide mats close to vent sites in mi-
croaerobic environments. However, benthic or near-benthic
bacteria could not have been responsible for deposition of lat-
erally extensive banded micritic iron formation, such as that
of the Kuruman Iron Formation, because they require mi-
croaerobic environments that would not have allowed trans-
port of dissolved ferrous iron over long distances from hy-
drothermal vent sites. In contrast, irregular amorphous
particles of ferric oxyhydroxides are formed by nonfilamen-
tous strains of iron-oxidizing bacteria (Emerson and Moyer,
2002). Such nonfilamentous bacteria could easily have
adapted to live pelagically along a suitable chemocline in the
water column, precipitating ferric oxyhydroxide particles in-
distinguishable from those formed by direct inorganic oxida-
tion of iron by free oxygen in microaerobic environments
(Emerson and Moyer, 2002). Because of the vastly increased
reaction rate, it can be expected that the volume of micro-
bially precipitated particles (see reaction (4)) would have far
outnumbered those of inorganic origin.
The role of other chemolithoautotrophic microbes as pri-
mary producers of organic carbon cannot be discarded.
Below the zone of free-oxygen availability, under anoxic con-
ditions, nitrate reducers (both archea and bacteria) with a ca-
pability to oxidize ferrous iron (Straub et al., 1996; Kon-
hauser, 2007) may have been responsible for precipitation of
ferric oxyhydroxide and formation of primary organic matter.
Microbial oxidation of atmospherically derived S
o
by
chemolithoautotrophic bacteria under micro- or anaerobic
conditions could have added to the primary organic carbon
budget. Within the anaerobic hydrothermally enriched water
column and more proximal to vent sites other chemolithoau-
totrophs, including hydrogen oxidizers and methanogens
(Konhauser, 2007), may have produced additional organic
matter. Proximal to vent sites, methylotrophs (Konhauser,
2007) could have produced organic matter from upward-es-
caping methane along the interface between deeper anaero-
bic and overlying microaerobic water masses.
Higher rates of production of organic carbon by chemo-
lithoautotrophic bacteria following volcanic events that effec-
tively fertilized the ocean-water column could explain the
abundance of ferrous iron minerals, such as siderite and mag-
netite relative to hematite, near pyroclastic stilpnomelane
36 BEUKES AND GUTZMER
0361-0128/98/000/000-00 $6.00 36
Beukes_Gutzmer 6/11/08 7:41 AM Page 36
lutite beds (Fig. 16B) and in the basinal relative to the shal-
low-shelf area (Fig. 15A). Especially where abundant wavy
and podded microbanded hematite-facies banded micritic
iron formations are preserved in the deep-water facies of the
Kuruman Iron Formation (Fig. 15A) on the shallow shelf,
very effective degradation of organic matter has to be invoked
if it was produced by iron-oxidizing chemolithoautotrophs.
Effective degradation of organic matter is supported by the
higher degree of maceration (as expressed by enrichment in
13
C) of organic matter, paired with depletion in
13
C in coex-
isting Fe-bearing carbonate in oxide-facies banded micritic
iron formation, as compared to carbonate-facies banded mi-
critic iron formation (Fig. 16B; Klein and Beukes, 1989;
Kaufman, 1996).
Depositional model: In the depositional model of stilp-
nomelane lutite-banded micritic iron formation macrocycles
(Fig. 20A), volcanic centers and hydrothermal vent sites are
combined for sake of simplicity. However, it is important to
realize that this may not necessarily have been the case. The
felsic composition of the stilpnomelane lutite pyroclastic beds
(Van Wyk, 1987; Pickard, 2003) indicates that they could not
have been derived from basaltic ocean floor plume or ridge
spreading centers but rather from volcanic arc environments.
Hydrothermal vents may well have been located along
spreading centers in a back-arc basin or along submerged sec-
tions of the volcanic arc. Because spreading in a back-arc
basin relates directly to subduction in the arc environment
(Marshak, 2005), it is conceivable that increasing rates of
MAJOR IRON FORMATIONS AT THE ARCHEAN-PALEOPROTEROZOIC BOUNDARY 37
0361-0128/98/000/000-00 $6.00 37
Neutral pH. Iron oxidizers
P
h
o
to
tr
o
p
h
ic
S
id
e
ritic
R
e
p
la
c
iv
e
s
ilic
a
L
o
w
Rich
Silica Sedim
entation Rate
Chemolithoautotrophic
Organic Carbon Productivity
Hematite-magnetite
Volcanics
High
Mixed Zone
SiO
2(aq)
2+
Fe
O
2
Fe(OH) +Org C
3
Siderite-magnetite
Primary Productivity
Carbonates
Org C
H
ig
h
O
2
mag
sid
hem
m
ag
sid
A
Siderite
High Organic Carbon Productivity
Volcanics
Laminated greenalite MIF
Greenalite MIF
Greenalite precipitates
M
ixed
Lutitic
Granular
Siderite-rich
C
ir
c
u
la
t
io
n
O
2
O
2
O
2 O
2
Low Org C Productivity
SiO
2
Greenalite BMIF
2+
Fe
2+
Fe
SiO
2
Green
SiO
2(aq)
B
2+
Reaction: SiO + Fe
2(aq)
O
2
Lutitic and
granular IF
High
Alkalinity chemocline
pH chemocline for precipitation of
greenalite
Iron oxidizers along neutral pH
boundary with low p O
2
ch
O
2
L
o
w
Chert
Sparse
2+
Fe
Note:
Note:
Volcanics
Greenalite (along alkalinity chemocline)
Greenalite precipitates
Mixed
facies
FIG. 20. A. Depositional model for stilpnomelane lutite-banded micritic iron formation macrocycles during early stages
of development of the Kuruman Iron Formation when deposition of carbonate platform sedimentation was still going on in
the interior of the Kaapvaal craton. Ferric oxyhydroxides are precipitated in deep water far below the photic zone, wave base,
and surface water of the basin, by iron-oxidizing chemolithoautotrophs. Iron and silica are sourced from offshore hy-
drothermal vents. Siderite and magnetite are formed during diagenesis through microbial dissimilatory ferric iron reduction.
Rates of primary organic carbon production by chemolithoautotrophs are higher closer to hydrothermal vents because of the
presence of more reduced species; this results in more reduced facies of iron formation more proximal to vents. Silica con-
centration is also higher in proximity to vents. Toward the interior of the craton primary organic carbon is produced by pho-
toautotrophs on a carbonate shelf. Organic carbon washed in from this source results in transformation of ferric oxyhydrox-
ides to siderite in proximity to the carbonate platform. Hematite-rich iron formation is thus only preserved in areas distal
from abundant primary organic productivity whether by chemolithoautotrophs in deep water below the photic zone or by
photoautotrops in shallow water. B. Depositional model for laminated greenalite micritic iron formation in the upper part
of the Kuruman Iron Formation and clastic-textured iron formation of the overlying Griquatown Iron Formation. Greenalite
precipitates along a chemocline with specific pH at which ferrous iron would react with dissolved amorphous silica (see Fig.
19 for solubility curves). In shallow water oxygen is replaced by high rates of concentrated circulation of anaerobic iron-en-
riched hydrothermal water. Here ferric oxyhydroxide is precipitated through oxidation by free oxygen in the presence of
abundant primary organic matter produced by photoautotrophs. The organic matter results in reduction of ferric oxyhy-
droxides through microbial iron respiration to form early diagenetic siderite and magnetite.
Beukes_Gutzmer 6/11/08 7:41 AM Page 37
hydrothermal venting could correspond to tectonic events
and increased volcanic activity. This suggestion, however, im-
plies that deposition of the stilpnomelane lutite-banded mi-
critic iron formation macrocycles did not take place in an
open marine basin but rather in a partly silled or back-arc
basin (Fig. 7; Klein and Beukes, 1989).
Although stilpnomelane lutite-banded micritic iron forma-
tion macrocycles display layer-cake stratigraphy, they display
a gradual lateral facies change that needs to be taken into ac-
count. These lateral facies changes are well represented by
the deposition of ferric oxyhydroxides with some silica most
distal from volcanic centers or hydrothermal vents versus
siderite- and chert-rich assemblages more proximal to sup-
posed vent sites, where they interfinger with pyroclastics.
Support for this concept is provided by the apparent interfin-
gering of wavy and podded hematite-magnetite banded mi-
critic iron formation units, which are best developed on the
shelf, with magnetite-siderite and siderite banded micritic
iron formation lateral correlatives in the basin (Fig. 15A). The
lithofacies developed in stilpnomelane lutite-banded micritic
iron formation macrocycles thus vector toward the source of
volcanic and/or hydrothermal activity. There are some indica-
tions in the lower part of the Stofbakkies Member of the Ku-
ruman Iron Formation that carbonate platform deposition
persisted contemporaneously with the formation of stilp-
nomelane lutite-banded micritic iron formation macrocycles
(Van Wyk, 1987; Klein and Beukes, 1989; Beukes et al.,
1990). This allows reconstruction of a depositional systems
tract that must have extended over several hundred kilome-
ters from carbonate platform sedimentation on the interior of
the Kaapvaal caton, to volcanic and/or hydrothermal vent de-
posits on the distal shelf (Fig. 20A).
The model illustrates that deposition of ferric oxyhydrox-
ides was facilitated by iron-oxidizing chemolithautotrophs in
deep water below the photic zone. It is suggested that the
iron-oxidizing bacteria thrived along the chemocline between
the upwelling iron-rich deep-water reservoir and microaero-
bic shallow-marine water mass. Bioproductivity in environ-
ments distal to hydrothermal vent sites was low and domi-
nated by iron-oxidizing bacteria. Proximal to hydrothermal
vents a larger variety of reduced components was available for
oxidation, resulting in higher bioproductivity by a larger spec-
trum of chemolithoautotrophs. In the latter environments
more organic matter may thus have reached the sediment
surface, resulting in very effective dissimilatory ferric oxyhy-
droxide reduction by chemoheterotrophs and the formation
of early diagenetic siderite (Fig. 20A). This is in contrast to
environments distal to the hydrothermal vent sites where
much of the small amount of organic carbon was remineral-
ized in the water column and ferric oxyhydroxides accumu-
lated on the floor of the depository to be later transformed to
hematite and/or magnetite (Fig. 20A).
Silica would have been introduced together with ferrous
iron along hydrothermal vent sites. However, continental
weathering as a relevant source of silica to the shallow-marine
water column cannot be excluded. In proximity to vent sites
rapid sedimentation and silica cementation led to formation
of chert-rich banded micritic iron formation (Fig. 20A). Dis-
tal to vent sites silica cementation was less effective and rela-
tively chert poor podded hematite-magnetite banded micritic
iron formation formed (Fig. 20A). It is proposed that in envi-
ronments of less effective silica cementation, magnetite
formed from hematite by addition of Fe
2+
from overlying
basin water during compaction (see reaction 3 above).
Toward the interior of the Kaapvaal craton, carbonate plat-
form deposition was associated with abundant bioproductiv-
ity by phototrophic bacteria. Some of this carbon was trans-
ported into the environment of banded micritic iron
formation deposition and resulted in formation of siderite-
rich banded micritic iron formation through ferric iron respi-
ration. The addition of organic carbon from this external
source is reflected in the higher concentration of organic car-
bon preserved in banded micritic iron formation in the tran-
sition zone between the Campbellrand carbonate and Kuru-
man Iron Formation succession relative to banded micritic
iron formation stratigraphically higher up in the Kuruman
succession (Kaufman, 1996). Limited by a sparse supply of
nutrients little phototrophic bioproductivity took place in the
pelagic environment. Most, if not all, of this organic carbon
probably became remineralized in the aerobic upper-water
column of the depository and thus did not contribute to the
flux of organic carbon to the sediment on the floor of the deep
basin (Fig. 20A).
Origin of mixed microbanded-laminated
greenalite-siderite micritic iron formation
Microbanded-laminated greenalite-siderite micritic iron
formation (laminated greenalite micritic iron formation for
short) is characteristic of the upper Westerberg and Geduld
Members of the Kuruman Iron Formation. It represents a
distinct change in the mode of silica and iron deposition. Both
of the two fundamental chemical constituents of iron forma-
tion occur together in the form of fine-grained greenalite,
thus indicating that both precipitated together. This is in
marked contrast to the clear separation of iron and silica in
stilpnomelane lutite-banded micritic iron formation macrocy-
cles. This change coincides with a change in water depth in
the basin. Whereas microbanded banded micritic iron forma-
tion of stilpnomelane lutite-banded micritic iron formation
macrocycles was almost certainly deposited far below storm-
wave base, the laminated greenalite micritic iron formation
was deposited at depths where storm-wave currents could
sporadically rework iron-rich muds (Beukes, 1984).
The solubility of greenalite is a function of pH, and the ac-
tivities of Fe
2+
(aq)
(Fig. 21B), SiO
2(aq)
, and p
O
2
(Fig. 19A, C).
Greenalite is very poorly soluble in anaerobic alkaline water
and would precipitate from a solution, even at very low activ-
ities of Fe
2+
(aq)
and SiO
2(aq)
, if the pH is raised (Fig. 19B, C; Eu-
gster and I-Ming Chou, 1973) according to the reaction,
3Fe
2+
(aq)
+ 2H
4
SiO
4
+ H
2
O Fe
3
Si
2
O
5
(OH)
4(s)
+ 6H
+
(aq)
. (5)
In contrast to greenalite, precipitation of amorphous silica is
brought about by either increasing activity of dissolved SiO
2
(Fig. 19B) and/or decreasing temperature (Siever, 1971).
These reactions place important controls on iron and silica
precipitation. It is proposed that the pH of hydrothermally
enriched bottom water that upwelled rapidly onto the shelf
during the initial stage of basin evolution was slightly acidic
to neutral. This would have favored chert- and iron-rich
38 BEUKES AND GUTZMER
0361-0128/98/000/000-00 $6.00 38
Beukes_Gutzmer 6/11/08 7:41 AM Page 38
micro- and mesobands to form separate entities in stilpnome-
lane lutite-banded micritic iron formation macrocycles. Fur-
thermore, iron was oxidized by iron-oxidizing bacteria and de-
posited as ferric oxyhydoxides, effectively preventing the
formation of greenalite. Deposition of chert micro- and
mesobands could have been brought about by simple mixing
of silica-saturated deep- and shallow-water masses. The ef-
fects of mixing may have been compounded by a drop in tem-
perature, as warmer hydrothermally enriched deep ocean
water upwelled into colder overlying water (Fig. 13).
However, as the basin continued to fill and hydrothermally
enriched bottom water invaded shallower environments, the
appearance of greenalite suggests that important redox and/or
pH boundaries were crossed that allowed greenalite to form.
Formation of greenalite requires very low oxygen fugacities to
prevent the oxidation of Fe
2+
(aq)
. It is thus proposed that pre-
cipitation of greenalite took place along a pH chemocline that
was located within the anaerobic environment of the up-
welling hydrothermally enriched water mass (Fig. 20B). Be-
cause of the insoluble nature of greenalite in alkaline anaero-
bic environments, iron and/or silica would have effectively
been removed from solution (Eugster and I-Ming Chou,
1973). The deposition of lutitic and granular iron formation in
shallow water above normal-wave base (Figs. 15A, 17) reveals
that precipitation along this pH chemocline was, in fact, not
very efficient. Our model does also not imply that all or most
of the greenalite actually formed as primary precipitate in the
water column. Greenalite often occurs as microspheroids in a
chert groundmass; such spheroids may well have formed at or
immediately below the sediment-water interface through re-
action of Fe
2+
(aq)
with precipitated amorphous silica in response
to increasing alkalinity (Fig. 19C).
Siderite occurs in the laminated greenalite micritic iron for-
mation concentrated as fine crystalline femicrite along spe-
cific microbands or as microsparite in tiny nodular concre-
tions in greenalite femicrite (Beukes, 1980b, 1984; Van Wyk,
1987). The siderite is highly depleted in
13
C (Beukes and
Klein, 1990), indicating that it was derived from early diage-
netic dissimilatory ferric oxyhydroxide reduction. This may
also apply to fine-grained magnetite (see earlier discussion)
concentrated along microbands in the laminated greenalite
micritic iron formation. The precipitation mechanism of pre-
cursor ferric oxyhydroxide required for the formation of such
early diagenetic siderite and magnetite remains poorly con-
strained. Precipitation of the ferric oxyhydroxide could have
taken place inorganically through reaction with free oxygen
derived from oxygenic photosynthesis by cyanobacteria in the
shallow-mixed water column of the basin. Because deposition
took place above storm-wave base, it is quite possible that
laminated greenalite micritic iron formation formed in the
photic zone. Within the anaerobic environment of the hy-
drothermally enriched water mass, and most likely immedi-
ately below the oxic-suboxic interface (Fig. 20B), deposition
of ferric oxyhydroxide could also have been brought about by
anaerobic Fe(II)-oxidizing phototrophs (Kappler et al., 2005;
Konhauser, 2007). Because the formation of greenalite re-
quires alkaline environments, it is considered unlikely that
neutrophillic microaerophyllic iron-oxidizing bacteria were
involved, unless some strains could have adapted to alkaline
conditions.
Origin of mixed lutitic and granular BIF depofacies
Deposition of lutitic and granular BIF in very shallow water
above fair weather wave base in the upper part of the As-
besheuwels succession (Figs. 15A, 17) requires effective re-
placement of the shallow-water column by hydrothermally
enriched deep water (Fig. 20B). In this environment,
greenalite lutite was deposited at slightly greater water depth,
while siderite was deposited in very shallow-water environ-
ments (Fig. 15A). In the latter environments, silica occurs
mainly in the form of chert cement, which implies that silica
saturation levels were reached during diagenesis.
Interpretation of the physico-chemical environment of
clastic-textured BIF deposition in shallow environments (Fig.
15A) is dependent on whether siderite is considered a pri-
mary precipitate (e.g., Beukes and Klein, 1990) or a product
of diagenetic degradation of organic matter by iron-reduc-
ing chemoheterotrophs. Negative and variable
13
C values
obtained for siderite in lutitic and granular BIF of the tran-
sition zone between the Kuruman and Griquatown Iron
Formations (Figs. 15A, 21) strongly support siderite forma-
tion from dissimilatory reduction of ferric oxyhydroxide.
The occurrence of dusty hematite finely intergrown with
and partly replaced by siderite in siderite-hematite lutite in
the lower parts of the Griquatown Iron Formation (Figs.
15A, 21) (Beukes, 1980; Van Wyk, 1987; Beukes and Klein,
1990) supports this notion. An abundance of silt- to sand-
sized siderite granules in the clastic-textured iron formation
suggests that replacement of ferric oxyhydroxide by siderite
MAJOR IRON FORMATIONS AT THE ARCHEAN-PALEOPROTEROZOIC BOUNDARY 39
0361-0128/98/000/000-00 $6.00 39
GIF
Laminated greenalite MIF
Chert-rich siderite BMIF
GIF and lutitic siderite-magnetite
BMIF (H=hematitic)
Stilpnomelane lutite
Stilpnomelane
Felutite
Org.c
.02 .04 -28 -26 -24 -22 -10 -8 -6
K
u
r
u
m
a
n
I
F
W
e
s
t
e
r
b
e
r
g
M
e
m
b
e
r
G
e
d
u
l
d
M
e
m
b
e
r
G
r
i
q
u
a
t
o
w
n
I
F
D
a
n
i
e
l
s
k
u
i
l
M
e
m
b
e
r
Ouplaas
Member
0
m
10
20
30
40
50
60
m
H
H
H
13
c
org
13
c
carb
Legend
FIG. 21. Stratigraphic profile of the transition zone from microbanded and
laminated micritic iron formation in the upper part of the Kuruman Iron
Formation to clastic-textured iron formation of the Ouplaas Member and
overlying Griquatown Iron Formation (modified after Beukes and Klein,
1990).
Beukes_Gutzmer 6/11/08 7:41 AM Page 39
took place during very early diagenesis. That siderite could
not have formed far below the sediment-water interface is
also indicated by relatively (for siderite) high
13
C values of
~6 per mil (Fig. 21), indicating a reasonable availability of
dissolved inorganic carbon into the diagenetic environment.
The effective replacement of primary ferric oxyhydroxide
precipitates by siderite requires an ample supply of organic
matter. Stable isotope data reveals that organic carbon is
more enriched in
13
C in shallow-water siderite-hematite lutite
than in deeper water laminated greenalitic micritic iron for-
mation (Fig. 21), suggesting a higher degree of maceration in
shallower water environments, possibly attributed to aerobic
respiration under oxygenated conditions.
For shallow-water environments in which the clastic-tex-
tured iron formations were deposited, it is opportune to as-
sume that the organic matter needed for reduction of ferric
oxyhydroxides was produced by photoautotrophs. Oxygen
created by these photoautotrophs could, in turn, have been
responsible for precipitation of ferric oxyhydroxides. How-
ever, this raises the question of how dissolved ferrous iron was
transported across such an extensive shallow-marine platform
as covered by the Griquatown Iron Formation, if conditions
in the water column were aerobic. Displacement of the
anoxic-oxic chemocline to virtually the upper few meters of
the water column by a strong influx of anaerobic deep water
carrying abundant dissolved ferrous iron may be the only sen-
sible answer to this question. This would have required rapid
and effective circulation of water into the partly enclosed or
silled basin across the shallow platform (Fig. 20B). Even
though large volumes of water must have circulated through
the basin, clastic-textured iron formations in the Griquatown
Iron Formation lack evidence of structures like cross-bedding
and current lamination that would indicate consistent sorting
by traction currents. Instead, sedimentary textures are sug-
gestive of a wave- and/or storm-dominated system (Beukes,
1983, 1984; Beukes and Klein, 1990). This would imply a pro-
tected shallow basin in which excellent stratification was
maintained over considerable periods of time through circu-
lation of hydrothermally enriched bottom water. This stratifi-
cation and circulation pattern may occasionally have broken
down during storms that suspended iron-rich muds from the
bottom and reworked early diagenetic silica-cemented hard-
grounds and concretions to form granular iron formation and
chert flat-pebble conglomerate beds.
Lateral interfingering of silicate-rich (greenalite, minneso-
taite, and stilpnomelane) felutites in the basin with siderite-
rich felutites and grainstones on the shelf in the Griquatown
Iron Formation (Fig. 15A) can be explained in a couple of
ways. It could merely imply that the deeper water column in
the basin was more reducing and alkaline, thus permitting the
precipitation of iron-silicate, especially greenalite. At the
same time ferric oxyhydroxides were deposited together with
abundant organic carbon in shallow-shelf environments and
were transformed through dissimilatory iron reduction to
early diagenetic siderite (Fig. 20B). The second possibility is
that silica and ferric oxydroxides may have been deposited to-
gether with some organic carbon in deep-water environments
of the basin. Alkalinity may have been slightly lower than on
the shallow shelf. During degradation of organic matter p
CO
2
in pore water may not have reached sufficient high levels in
this environment to precipitate siderite (Fig. 19A). Rather,
dissolved ferrous iron could have reacted with amorphous sil-
ica to form greenalite (Fig. 19A). Iron isotope studies should
be applied to determine whether greenalite formed by re-
duction of ferric oxyhydroxides can be distinguished from
greenalite that originated as a primary precipitate.
The Koegas Subgrouptermination of
iron formation deposition
The giant iron formation succession of the Asbesheuwels
Subgroup is overlain by a mixed siliciclastic-iron formation
succession with subordinate stromatolitic carbonates known
as the Koegas Subgroup (Figs. 1A, 22). Diamond drill core re-
covered in the second phase of the Agouron Drilling Project
in 2005 indicated that the contact between the Griquatown
Iron Formation and mudstone of the Pannetjie Formation is
sharp and marked by a few centimeters-thick chert pebble
conglomerate (Fig. 22; N.J. Beukes, pers. observation, 2005).
This conglomerate that has not been observed in outcrop is
considered very important, as it suggests a hiatus between
iron formation and siliciclastic deposition (Fig. 22). In fact, it
may indicate that the end of deposition of the major As-
besheuwels Iron Formation succession was brought about by
uplift and exposure accompanied by regression of the hy-
drothermally enriched water mass, from which iron formation
was deposited, off the shelf. Subsequent transgression led to
filling of accommodation space by fine siliciclastics, known as
the Pannetjie Formation. The siliciclastics are arranged in an
upward-coarsening succession (Fig. 22) typical of deltaic de-
posits (Beukes, 1978, 1983). Facies changes and paleocurrent
directions identify the interior of the Kaapvaal craton as
source terrane for the Pannetjie Formation (Beukes, 1983).
The transition from the Asbesheuwels Subgroup into the
Koegas Subgroup does, however, not constitute the end of
iron formation deposition in the Transvaal basin, as two clas-
tic-textured iron formations are developed in the Koegas Sub-
group. The lower of these two iron formations, known as the
Doradale Iron Formation, is only about 25 m thick and over-
lies a prominent chert-pebble conglomerate, which marks a
sequence boundary, with sharp gradational contact. The Do-
radale Iron Formation is, in turn, overlain with sharp contact
by siliciclastics of the Naragas Formation (Fig. 22). The upper
iron formation, on the order of 300 m thick, comprises three
distinct units, namely the Rooinekke, Klipput, and Nelani
Iron Formations (Fig. 22). The Doradale, Rooinekke, and
Nelani Iron Formations are composed of wavy and podded
siderite-silicate felutite with interbeds of grainstone and flat-
pebble chert breccias. Magnetite is common and hematite is
preserved in grainstone beds. The Klipput Iron Formation,
on the other hand, is composed of laminated silicate micritic
iron formation, possibly representing a deeper water facies of
siderite-rich mixed lutitic and granular BIFs (Fig. 22). De-
tailed descriptions of these fall outside the scope of this paper.
A tuffaceous stilpnomelane lutite bed in the lower part of
the Rooinekke Formation gave a poorly constrained SHRIMP
U-Pb age of 2413 Ma on zircon (Gutzmer and Beukes,
1998b). The Koegas Subgroup has a number of important
characterics, such as (1) carbonate bioherms that are overlain
with sharp transgressive contact by the Rooinekke Iron For-
mation, (2) thin manganiferous carbonate units interbedded
40 BEUKES AND GUTZMER
0361-0128/98/000/000-00 $6.00 40
Beukes_Gutzmer 6/11/08 7:41 AM Page 40
MAJOR IRON FORMATIONS AT THE ARCHEAN-PALEOPROTEROZOIC BOUNDARY 41
0361-0128/98/000/000-00 $6.00 41
wavy
and
podded
siderite-
silicate
lutite
and
grainstone
Dark Fe-
silicate
lutite
Magnetic
wavy and
podded
siderite-
silicate
lutite and
grainstone
Red beds
Red beds
Siderite-
silicate
lutitic BIF
Chert rip-ups
Chert rip-ups
Prominent chert pebble
conglomerate/breccia
Fine quartzite
Red and grey mudstone
Red mudstone
Silt- and mudstone
Fe,Mn
Carbonate
~2413Ma
Red and grey
hematitic
greywacke,silt-
and claystone
N
a
r
a
g
a
s
F
m
R
o
o
i
n
e
k
k
e
I
F
S
h
a
l
l
o
w
i
n
g
u
p
w
a
r
d
s
T
r
a
n
s
-
g
r
e
s
s
i
o
n
S
t
a
c
k
e
d
u
p
w
a
r
d
c
o
a
r
s
e
n
i
n
g
d
e
l
t
a
i
c
s
u
c
c
e
s
s
i
o
n
s
Transgression
Hiatus
S
h
a
l
l
o
w
i
n
g
u
p
w
a
r
d
s
N
e
l
a
n
i
I
F
Major
unconformity
Ongeluk Lava
Makganyene Diamictite
Unconf
Grainstone
Silicate lutite
Grainstone
Chert rip-ups
Chert rip-ups
Carbonate bioherms
0
100
200
300
400
500
600
m
Transgression
Hiatus
Hiatus
Transgression
Hiatus
Sid
Sid
Sid
K
l
i
p
p
u
t
l
u
t
i
t
e
Sid
Sid
Sid
Sid
P
a
n
n
e
t
j
i
e
F
m
D
o
r
a
d
a
l
e
I
F
Magnetic
Some
hematite
preserved
Some
hematite
preserved
Griquatown IF
FIG. 22. Composite stratigraphic profile of the Koegas Subgroup, indicating the sharp erosional contact between the
Griquatown Iron Formation and overlying siliciclastics of the Pannetjie Formation. Also indicated are the stratigraphic
settings of clastic-textured iron formations of the Koegas succession.
Beukes_Gutzmer 6/11/08 7:41 AM Page 41
with the latter iron formation (Fig. 22; Beukes, 1983), and (3)
deltaic deposits of the Naragas Formation that are stained red
by fine hematite and represent the oldest known red beds in
the Transvaal Supergroup (Fig. 22). The occurrence of these
red beds supports the presence of free oxygen in the atmos-
phere and shallow-marine environment during the deposition
of early Paleoproterozoic iron formations. The original extent
of the Koegas Subgroup remains unknown, because it is in-
tersected by a major erosional unconformity at base of the
Makganyene Diamictite (Fig. 22).
Summary of Salient Aspects of Major Neoarchean-
Paleoproterozoic Iron Formation Successions
The major iron formations that formed along the
Neoarchean-Paleoproterozoic boundary in the Transvaal and
Hamersley basins represent a very complex group of rocks
with many different iron formation rock types and some in-
tricate lateral and vertical lithofacies variations formed in a
wide range of depositional and diagenetic environments and
through various mechanisms of iron and silica deposition.
There are essentially three major types of iron formations,
namely (1) microlaminated (microbanded) chert-banded mi-
critic iron formation( banded micritic iron formation), (2)
chert-poor microbanded and laminated greenalite-siderite
iron formation (greenalite laminated iron formation), and (3)
mixed lutitic to granular iron formation (clastic-textured iron
formation). The banded micritic iron formation formed in
the deepest part of the basin far below wave base and the
photic zone at depths of perhaps several hundred meters. In
contrast, greenalite laminated iron formation formed in
somewhat shallower environments below normal wave base
but at depths where storm-wave currents could occasionally
rework fine bottom sediment and perhaps at least in part
above the deepest reaches of the photic zone at around 150
m. Clastic-textured iron formation formed mainly above nor-
mal wave base well within the photic zone of the oceanic
water column.
A common aspect of deposition of iron formation in all
three of the major environments is that it took place along
chemoclines where upwelling deep anaerobic and acidic iron-
and silica-enriched hydrothermal plume water mixed with
shallower slightly oxygenated and alkaline marine water.
However, the depth of this chemocline was not permanently
fixed but rather dynamically controlled by the rates of flux of
iron and silica from hydrothermal plumes relative to the sup-
ply of photosynthetically derived oxygen through down-
welling of near-surface waters in a well-circulating oceanic
system. Most importantly, the depth of the chemocline deter-
mined the mechanisms through which deposition of iron and
silica took place.
At least three different mechanisms of silica deposition are
distinguished. In deep-water banded micritic iron formation
with well-defined chert bands, deposition of silica was most
probably brought about through mixing of slightly heated hy-
drothermal plume water with cold deep downwelling ocean
water. The water in this mixing zone must have had a near-
neutral pH so that greenalite precipitation could not take
place and silica deposition remained well separated from that
of iron. In slightly shallower environments silica deposition
was combined with that of iron to form greenalite laminated
iron formation without clearly defined chert mesobands. This
required a chemocline that was situated within the anaerobic
hydrothermal water mass but with increased alkalinity that al-
lowed greenalite stability. In very shallow-water silica satura-
tion was apparently reached only beneath the sediment-water
interface in the diagenetic environment most of the time as
silica occurs mainly as interstitial cements in lutititic and
granular iron formation.
Another common factor of all the iron formation lithofacies
is that ferric oxyhydroxide represented the primary precipi-
tate of iron, with the notable exception of greenalite in
greenalite-siderite laminated iron formation. In deep-water
banded micritic iron formation that formed below the photic
zone, ferric oxyhydroxide deposition took place either inor-
ganically through reaction of dissolved ferrous iron with oxy-
gen or was microbially mediated by microaerophyllic and
neutrophyllic iron-oxidizing chemolithoauototrophs. The lat-
ter process was probably dominant, because it accounts for
very rapid deposition of large volumes of iron oxyhydroxides
coupled to relatively small amounts of primary organic mat-
ter. Anaerobic Fe(II)-oxidizing photoautotrophs may have
played a role in deposition of ferric oxyhydroxide only in shal-
lower anaerobic environments in which greenalite laminated
iron formation was deposited. In shallow aerobic environ-
ments iron could have been precipitated inorganically
through reaction with oxygen in the presence of abundant
primary organic matter derived from photoautotrophs. Iron
oxidation by chemolithoautotrophs may have become impor-
tant in environments in which oxygen became depleted and
alkalinity was near neutral.
Magnetite (excluding low-grade metamorphic magnetite)
and siderite in all of the depositional settings of iron forma-
tion formed diagenetically through dissimilatory ferric oxyhy-
droxide reduction by iron-reducing heterotrophs. This inter-
pretation is supported by very variable
13
C values in siderite
and
56
Fe values in both siderite and early diagenetic mag-
netite in iron formation from both shallow- and deep-water
environments. The different mineralogical facies of iron for-
mation as observed today are thus essentially of diagenetic
origin. This includes hematite derived from dehydration of
ferric oxyhydroxides deposited in environments where burial
of organic matter was limited.
Although very large in size relative to older iron formation
in Archean greenstone belts and late Paleoproterozoic cra-
tonic settings, the actual areal distribution of iron formations
in the Transvaal and Hamersley basins is miniscule compared
to the size of depositional systems of modern oceans. Inter-
pretations about possible global secular variations based on
observations in these two basins should thus be done only with
extreme caution This is the more so because both litho- and
time-stratigraphic correlations between the two basins and
new paleomagnetic data rather convincingly illustrate that the
successions may have been deposited in one large basin along
the margins of the ancient continent of Vaalbara. Similar sec-
ular paleoclimatic and geochemical proxies in the Hamersley
and Tranvaal successions may thus have mainly local basinal
rather than global significance. In addition, reconstruction of
deposional facies in the iron formations strongly suggests that
the basin was partly enclosed to have allowed for effective di-
rected flow, circulation, and concentration of hydrothermal
42 BEUKES AND GUTZMER
0361-0128/98/000/000-00 $6.00 42
Beukes_Gutzmer 6/11/08 7:41 AM Page 42
water for precipitation of large volumes of iron and silica over
a period of some 50 m.y.
Implications for Depositional Models of
Other Iron Formations
The Campbellrand and Kuruman-Griquatown successions
encompass almost all known lithofacies and depositional set-
tings of iron formations. The development of the different
lithofacies is dependent on water depth and physico-chemi-
cal conditions but independent of variations in associated
sedimentary and/or volcanic rocks and tectonic setting. Fol-
lowing our own observations, microbanded banded micritic
iron formation described from the lower part of the Kuru-
man Iron Formation is similar to banded micritic iron for-
mation facies characteristic of deep-water iron formations in
greenstone belts; there is thus little reason to believe that the
latter should have formed under different conditions than
those stipulated here for the Transvaal-Hamersley basin. Al-
though they have not been described as such in literature, we
have observed facies successions very similar to that of the
stilpnomelane lutite- banded micritic iron formation macro-
cycles of the Kuruman and Brockman Iron Formations in
greenstone belt banded micritic iron formation in southern
Africa and the Krivoy Rog succession of the Ukrainian Shield
(Fig. 4).
The clastic-textured iron formation of the Griquatown Iron
Formation, on the other hand, has many similarities to the
late Paleoproterozoic iron formation along the margins of the
Superior craton in North America (Fig. 4). In the latter
deeper water facies are also represented by laminated chert-
poor greenalite-facies micritic iron formation with shallower
facies composed of siderite- and magnetite-rich mixed lutitic
and granular chert-cemented banded granular iron formation
with hematite-rich granular iron formation being a subordi-
nate facies (Perry et al., 1973; Gross and Zajac, 1983; Morey,
1983). General depositional models developed for the clastic-
textured iron formation of the Griquatown Iron Formation
may thus well apply to many other granular iron formation-
bearing successions. It may, however, be that in some of the
late Paleoproterozoic iron formations of the Superior craton
the depositional environment was tide dominated (Gross and
Zajac, 1983; Ojakangas, 1983), whereas in the Griquatown
Iron Formation it was storm dominated.
Implications for the Origin of High-Grade Iron Ores
In most recent studies of the genesis of high-grade BIF-
hosted iron ore deposits, the possible influence of the lithofa-
cies architecture of the iron formation protolith is ignored.
However, we believe that the protolith composition and tex-
ture determine its suitability to be transformed into high-
grade ore and also influence the physical characteristics of
such ore. For example, Van Schalkwyk and Beukes (1986) il-
lustrated that the texture of high-grade hematite-martite ores
at the Sishen deposit (Fig. 1A) can be directly correlated with
that of the lithofacies of the Kuruman and Griquatown Iron
Formation protolith. In these ores some of the stilpnomelane
lutite-banded micritic iron formation macrocycles typical of
the lower part of the Kuruman Iron Formation can still be
recognized, although all original mineral assemblages have
been replaced by hematite. It would appear as if chert-poor
lithofacies, such as laminated greenalite-siderite and thin-
banded to podded hematite-magnetite facies BIF in the mid-
dle to upper parts of the Kuruman Iron Formation, were pre-
ferred host for the development of high-grade laminated iron
ores. Similarly, siderite-rich lutitic iron formation of the Gri-
quatown Iron Formation gave rise to very high-quality mas-
sive hematite ores (Van Schalkwyk and Beukes, 1986). In-
creased concentrations of potassium in iron ores and the
presence of small amounts of sericite can, in turn, be related
to former stilpnomelane lutite beds.
In the Dales Gorge Member of the Brockman Iron Forma-
tion in Western Australia detailed knowledge about the so-
called shale and BIF macrobands (synonymous to stilpnome-
lane lutite-banded micritic iron formation macrocycles of the
Kuruman Iron Formation) is used to define the stratigraphic
position of high-grade iron ores (Thorne et al., 2008). This
knowledge has been used, among others, to estimate the
amount of compaction that accompanied high-grade ore for-
mation (Trendall and Blockley, 1970) and to attempt mass-
balance calculations to describe the ore-forming process (Ri-
dley, 1999). Furthermore, the Whaleback Shale Member of
the Brockman Iron Formation has acted as a permeability
barrier to hydrothermal and/or meteoric fluid movement,
thus determining the distribution of high-grade ores in the
Dales Gorge Member (Taylor et al., 2001).
Detailed stratigraphic knowledge about the host iron for-
mation is also essential for recognizing major structures per-
haps related to ore formation (Beukes and Smit, 1987; Taylor
et al., 2001; Gutzmer et al., 2005) and recognition of struc-
tures related to slumping of BIF into karstic solution surfaces
(Van Schalkwyk and Beukes, 1986).
Implications for Exploration and Future Research
The detailed knowledge of the lateral and vertical facies
distribution in iron formation hosting high-grade iron ore de-
posits has important implications for the identification of suit-
able exploration target areas. Most promising targets may be
selected based on a combination of lithofacies, stratigraphic,
and structural data. Unfortunately it is essentially only in the
Tranvaal-Hamersley basin, the Krivoy Rog basin (Belevtsev et
al., 1983), and the Lake Superior (Morey, 1983; Fralick et al.,
2002; Brown, 2008)-New Quebec (Gross and Zajac, 1983)
belts of the Superior craton that some detailed knowledge
about the facies and stratigraphy of Archean-Paleoprotero-
zoic iron formation host rock to high-grade iron ores are cur-
rently available. In all other giant iron ore districts, including
the Iron Quadrangle and Carajas districts in Brazil and the
ore deposits in India, virtually nothing is known about the
stratigraphy and original sedimentary lithofacies of the host
iron formation successions. These should be priority targets
for future research.
Acknowledgments
We would like to thank The Agouran Institute, the National
Research Foundation (Pretoria), the University of Johannes-
burg, and Assmang for research funding, and Elsa Maritz,
Diana Khoza, Jean-Clement Beyeme-Zogo, and Michael
Chakuparira for their assistance during the preparation of this
manuscript. Constructive reviews by Bruce Simonson and
Dawn Sumner helped to improve the manuscript considerably.
MAJOR IRON FORMATIONS AT THE ARCHEAN-PALEOPROTEROZOIC BOUNDARY 43
0361-0128/98/000/000-00 $6.00 43
Beukes_Gutzmer 6/11/08 7:41 AM Page 43
REFERENCES
Alchin, D., Lickfold, V., Mienie, P.J., Nel, D., and Strydom, M., 2008, An in-
tegrated exploration approach to the Sishen South iron ore deposit, North-
ern Cape Province, South Africa, and its implication for developing a struc-
tural and/or resource model for these deposits: Reviews in Economic
Geology, v. 15, p. 317338.
Alterman,W., and Nelson, D.R., 1998, Sedimentation rates, basin analysis
and regional correlation of three Neoarchean and Paleoproterozoic sub-
basins of the Kaapvaal craton as implied by precise SHRIMP U-Pb zircon
ages from volcanic sediments: Journal of Sedimentary Geology, v. 120, p.
225256.
Anbar, A.D., Duan, Y., Lyons, T.W., Arnold, G.L., Kendal, B., Creaser, R.A.,
Kaufman, A.J., Gordon, G.W., Scott, C., Gavin J., and Buick, R., 2007, A
whiff of oxygen before the great oxidation event?: Science, v. 317, p.
19031906.
Arndt, N.T., Nelson, D.R., Compston, W., Trendall, A.F., and Thorne, A.M.,
1991, The age of the Fortesque Group, Hamersley basin, Western Aus-
tralia, from ion microprobe U-Pb zircon results: Australian Journal of Earth
Sciences, v. 38, p. 261281.
Barley, M.E., Pickard, A.L., and Sylvester, P.J., 1997, Emplacement of a large
igneous province as a possible cause of banded iron formation 2.45 billion
years ago: Nature, v. 385, p. 5558.
Bau, M., and Dulski, P., 1996, Distribution of yttrium and rare-earth ele-
ments in the Penge and Kuruman Iron Formations, Transvaal Supergroup,
South Africa: Precambrium Research, v. 79, p. 3755.
Beard, B.L., Johnson, C.M., Cox, L., Sun, H., Nealson, K.H., and Aquilar, C.,
1999, Iron isotopes biosignatures: Science, v. 285, p. 18891892.
Beard, B.L., Johnson, C.M., Skulan, J.L., Nealson, K.H., Cox, I., and Sun, H.,
2003, Application of Fe isotopes to tracing the geochemical and biological
cycling of Fe: Chemical Geology, v. 195, p. 87117.
Bekker, A., Kaufman, A.J., Karhu, J.A., Beukes, N.J., Swart, Q.D., Coetzee,
L.L., and Eriksson, K.A., 2001, Chemostratigraphy of the Paleoproterozoic
Duitschland Formation, South Africa: Implications for coupled climate
change and carbon cycling: American Journal of Science, v. 3001, p.
261285.
Bekker, A., Holland, H.D., Wang, P.-L., Rumble D., Stein, H.J., Hanna, J.L.,
Coetzee, L.L., and Beukes, N.J., 2004, Dating the rise of atmospheric oxy-
gen: Nature, v. 427, p. 117120.
Belevtsev, Ya, N., Belevtsev, R.Ya., and Siroshtan, R.I., 1983, The Krivoy Rog
basin, in Trendall, A.F. and Morris, R.C., eds., Iron-formation: Facts and
problems: Amsterdam, Elsevier, p. 211251.
Berner, R.A., 1981, A new geochemical classification of sedimentary envi-
ronments: Journal of Sedimentary Petrology, v. 51, p. 359365.
Beukes, N.J., 1973, Precambrian iron-formations of Southern Africa: Eco-
nomic Geology, v. 68, 9601004.
1978, Die karbonaatgesteentes en ysterformasies van die Ghaap-Groep
van die Transvaal-Supergroep in Noord-Kaapland: Unpublished PhD. the-
sis, Johannesburg, Rand Afrikaans University, 580 p.
1980a, Suggestions towards a classification and nomenclature for iron-
formation: Transactions of the Geological Society of South Africa, v. 83, p.
285290.
1980b, Lithofacies and stratigraphy of the Kuruman and Griquatown
Iron Formations, northern Cape Province, South Africa: Transactions of
the Geological Society of South Africa, v. 83, p. 6986.
1980c, Stratigrafie en litofasies van die Campbellrand-Subgroep van die
Proterofitiese Ghaap-Groep, Noord-Kaapland: Transactions of the Geolog-
ical Society of South Africa, v. 83, p. 141170.
1983, Palaeoenvironmental setting of iron-formations in the deposi-
tional basin of the Transvaal Supergroup, South Africa, in Trendall, A.F.,
and Morris, R.C., eds., Iron-formation: Facts and problems: Amsterdam,
Elsevier, p. 131209.
1984, Sedimentology of the Kuruman and Griquatown Iron Formations,
Transvaal Supergroup, Griqualand West, South Africa: Precambrian Re-
search, v. 24, p. 4784.
1987, Facies relations, depositional environments and diagenesis in a
major early Proterozoic stromatolitic carbonate platform to basinal se-
quence, Campbellrand Subgroup, Transvaal Supergroup, southern Africa:
Sedimentary Geology, v. 54, p. 146.
2004, Early options in photosynthesis: Nature, v. 431, p. 522523.
Beukes, N.J., and Klein, C., 1990, Geochemistry and sedimentology of a fa-
cies transition from microbanded to granular iron-formation in the early
Proterozoic Transvaal Supergroup, South Africa: Precambrian Research, v.
47, p. 99139.
Beukes, N.J., and Smit, C.A., 1987, New evidence for thrust faulting in Gri-
qualand West, South Africa: Implications for stratigraphy and the age of red
beds: South African Journal of Geology, v. 90, p. 378394.
Beukes, N.J., Klein, C., Kaufman, A.J., and Hayes, J.M., 1990, Carbonate
petrography, kerogen distribution, and carbon and oxygen isotopic varia-
tions in an Early Proterozoic transition from limestone to iron-formation
deposition, Transvaal Supergroup, South Africa: Economic Geology, v. 85,
p. 663690.
Beukes, N.J., Gutzmer, J., and Mukhopadhyay, J., 2002a, The geology and
genesis of high-grade hematite iron ore deposits: Australian Institute of
Mining and Metallurgy Publication Series, v. 7, p. 2329.
Beukes, N.J., Dorland, H.D., Gutzmer, J. Nedachi, M., and Ohmoto, H.,
2002b, Tropical laterites, life on land, and the history of atmospheric oxy-
gen in the Paleoproterozoic: Geology, v. 30, p. 491494.
Blake, T.S., and Barley, M.E., 1992, Tectonic evolution of the Late Archean
to Early Proterozoic Mount Bruce Megasequence, Western Australia: Tec-
tonics, v. 11, p.14151425.
Blockley, J.G., Tehnas, I.J., Mandyczewsky, A., and Morris, R.C., 1993, Pro-
posed stratigraphic subdivision of the Marra Mamba Iron Formation and the
lower Wittenoom Dolomite, Hamersley Group, Western Australia: Western
Australia Geological Survey Professional Papers, Report 34, p. 4763.
Boggs, S., 2005, Principles of sedimentology and stratigraphy: Upper Saddle
River, Prentice-Hall, 688 p.
Brown, D.A., 2006, Microbial mediation of iron mobilization and deposition
in iron formation since the early Precambrian: Geological Society of Amer-
ica Memoir 198, p. 239256.
Brown, P.E., 2008, Brief history of high-grade iron ore mining in North
America (18482008): Reviews in Economic Geology, v. 15, p. 361380.
Button, A., 1976a, Transvaal and Hamersley basinsreview of basin devel-
opment and mineral deposits: Minerals Science and Engineering, v. 8, p.
262293.
1976b, Iron-formation as an end member in carbonate sedimentary cy-
cles in the Transvaal Supergroup, South Africa: Economic Geology, v. 71,
p. 193201.
Cairns-Smith, A.G., 1978, Precambrium solution photchemistry, inverse seg-
regation, and banded iron formations: Nature, v. 276, p. 807808.
Cameron, E.M., 1982, Sulphate and sulphate reduction in early Precambrian
oceans: Nature, v. 296, p. 145148.
Cheney, E.S., 1996, Sequence stratigraphy and plate tectonic significance of
the Transvaal succession of southern Africa and its equivalent in Western
Australia: Precambrian Research, v. 79, p. 324.
Clout, J.M.F., and Simonson, B.M., 2005, Precambrian iron formation and
iron formation-hosted iron ore deposits: Economic Geology, v. 100, p.
643679.
Cornell, D.H., Schtte, S.S., and Eglington, B.L., 1996, The Ongeluk basaltic
andesite formation in Griqualand West, South Africa: Submarine alteration
in a 2222 Ma Proterozoic sea: Preacambrian Research, v. 79, p. 101123
Croal, L.R., Johnson, C.M., Beard, B.L., and Newman, D.K., 2004, Iron iso-
tope fractionation by Fe(II)-oxidizing photoautotrophic bacteria: Geochim-
ica et Cosmochimica Acta, v. 68, p. 12271242.
Crosby, H.A., Roden, E.E., Johnson, C.M., and Beard, B.L., 2007, The
mechanisms of iron isotope fractionation produced during dissimilatory
Fe(III) reduction by Shewanella putrefaciens and Geobacter sulfurre-
ducens: Geobiology, v. 5, p. 169189.
De Kock, M.O., 2007, Paleomagnetism of selected Neoarchean-Paleopro-
terozoic cover sequences on the Kaapvaal craton and implications for Vaal-
bara: Unpublished Ph.D. thesis, Johannesburg, University of Johannes-
burg, 276 p.
De Kock, M.O., Evans, D.A.D., Beukes, N.J., Gutzmer, J., and Armstrong,
R.A., 2007, Rearranging Vaalbara with new geochronological and paleo-
magnetic data from the Neoarchean Ventersdorp Supergroup of southern
Africa [abs]: Geological Society of America Abstracts with Programs, v. 39,
p. 285.
Dimroth, E., 1976, Aspects of the sedimentary petrology of cherty iron-for-
mation, in Wolf, K.H., ed., Handbook of strata-bound and stratiform ore
deposits, Volume 7: Amsterdam, Elsevier, p. 203254
Dimroth, E., and Chauvel, J.J., 1973, Petrographyb of the Sokoman Iron
Formation in part of the central Labrador trough: Geological Society of
America Bulletin, v. 84, p. 111134.
Dorland, H.C., 2004, Provenance ages and timing of sedimentation of
selected Neoarchean and Paleoproterozoic successions on the Kaapvaal
craton: Unplublished Ph.D. thesis, Johannesburg, Rand Afrikaans Univer-
sity, 326 p.
44 BEUKES AND GUTZMER
0361-0128/98/000/000-00 $6.00 44
Beukes_Gutzmer 6/11/08 7:41 AM Page 44
Doyle, M.G., Krapez, B., and Barley, M.E., 2001, Volcanic facies architecture
of the Woongarra large igneous complex [ext. abs.]: Geoscience Australia
Record 2001/37, p 149150.
Dunham, R.J., 1962, Classification of carbonate rocks according to deposi-
tional texture, in Ham, W.E., ed., Classification of carbonate rocks: Tulsa,
American Association of Petroleum Geologists, p. 108121.
Ehrenreich, A., and Widdel, F., 1994, Anaerobic oxidation of ferrous iron by
purple bacteria, a new type of phototrophic metabolism: Applied and En-
vironmental Microbiology, v. 60, p. 45174526.
Elderfield, H., 1988, The oceanic chemistry of the rare-earth elements:
Philosophical Transactions of the Royal Society of London, v. A325, p.
105126.
Elderfield, H., and Greaves, M.J., 1982, The rare earth elements in seawater:
Nature, v. 296, p. 214219.
Emerson, D., and Moyer, C., 1997, Isolation and characterization of novel
iron-oxidizing bacteria that grow at circumneutral pH: Applied and Envi-
ronmental Microbiology, v. 63, p. 47844792.
2002, Neutrophilic Fe-oxidizing bacteria are abundant at the Loihi
Seamount hydrothermal vents and play a major role in Fe oxide deposition:
Applied and Environmental Microbiology, v. 68, p. 30853093.
Eriksson, K.A., 1983, Siliciclastic-hosted iron-formations in the early
Archean Barberton and Pilbara sequences: Journal of the Geological Soci-
ety of Australia, v. 30, p. 473482.
Eriksson, K.A., Krapez, B., and Fralick, P.W., 1994, Sedimentology of
Archean greenstone belts: Signatures of tectonic evolution: Earth Science
Reviews, v. 37, p. 188.
Eugster, H.P., and Chou, I-M., 1973, The depositional environments of Pre-
cambrian banded iron-formations: Economic Geology, v. 68, p. 11441168.
Evans, D.A., Beukes, N.J., and Kirschvink, J.L., 1997, Low-latitude glaciation
in the Paleoproterozoic era: Nature, v, 386, p. 262266.
Farquhar, J., and Wing, B.A., 2003, Multiple sulfur isotopes and the evolu-
tion of the atmosphere: Earth and Planetary Science Letters, v. 213, p.
113.
Farquhar, J., Bao, H., and Thiemens, M., 2000, Atmospheric influence of
Earths earliest sulfur cycle: Science, v. 289, p. 756758.
Farquhar, J., Savarino, J., Airieau, S., and Thiemens, M.H., 2001, Observa-
tion of wavelength-sensitive mass-independent sulfur isotope effects dur-
ing SO2 photolysis: Implications for the early atmosphere: Journal of Geo-
physical Research, v. 106, p. 111.
Farquhar, J., Peters, M., Johnston, D.T., Strauss, H., Masterson, A.,
Wiechert, U., and Kaufman, A.J., 2007, Isotopic evidence for Mesoar-
chaean anoxia and changing atmospheric sulphur chemistry: Nature, v. 449,
doi:1038.
Fedo, C.M., and Eriksson, K.A., 1996, Stratigraphic framework of the ~3.0
Ga Buhwa greenstone belt: A unique stable-shelf succession in the Zim-
babwe Archean craton: Precambrian Research, v. 77, p. 161178.
Fischer, W.W., Schroeder, S., Lacassie, J.P., Beukes, N.J., Goldberg, T.,
Strauss, H., Horstmann, U.E., Schrag, D.P., and Knoll, A.H., (in press),
Isotopic constraints on the Late Archean carbon cycle from the Transvaal
Supergroup along the western margin of the Kaapvaal craton, South Africa:
Precambrian Research.
Fralick, P.W., Davis, D.W., and Kissin, S.A., 2002, The ages of the Gunflint
Formation, Ontario, Canada: Single zircon U-Pb age determinations from
reworked volcanic ash: Canadian Journal of Earth Sciences, v. 39, p.
10851091.
Frimmel, H.E., 1996, Witwatersrand iron-formations and their significance
for gold genesis and the composition limits of orthoamphibole: Mineralogy
and Petrology, v. 56, 273295.
Goode, A.D.T., Hall, W.D.M., and Bunting, J.A., 1983, The Nabberu basin of
Western Australia, in Trendall, A.F., and. Morris, R.C., eds., Iron-forma-
tion: Facts and problems: Amsterdam, Elsevier, p. 295323.
Goodwin, A.M., 2000, Principles of Precambian geology: San Diego, Acade-
mic Press, p. 327.
Gross, G.A., 1980, A classification of iron-formation based on depositional
environments: Canadian Mineralogist, v. 18, p. 215222.
1983, Tectonic systems and the deposition of iron-formation: Precam-
brian Research, v. 20, p. 171187.
Gross, G.A., and Zajac, I.S., 1983, Iron-formation in fold belts marginal to
the Ungava craton, in Trendall, A. F., and Morris, R.C., eds., Iron-forma-
tion: Facts and problems: Amsterdam, Elsevier, p. 253294.
Gutzmer, J., and Beukes, N.J., 1998a, The manganese formation of the Neo-
proterozoic Penganga group, Indiarevision of an enigma: Economic Ge-
ology, v. 93, p. 10911102.
1998b, High-grade manganese ores in the Kalahari manganese field:
Characterisation and dating of the ore forming events: Johannesburg, Rand
Afrikaans University, Unpublished report, 221 p.
Gutzmer, J., Beukes, N.J., de Kock, M. O., and Netshiozwi, S. T., 2005, Ori-
gin of high-grade iron ores at the Thabazimbi deposit, South Africa: Iron
Ore Conference, Freemantle, 1921 September 2005, Proceedings, p.
5765
Han, T-M., 1978, Microstructures of magnetite as guide to its origin in some
Precambrian iron-formation: Fortschritte der Mineralogie, v. 56, p. 105142.
Hannah, J.L., Bekker, A., Stein, H.J., Markey, R.J., and Holland, H.D., 2004,
Primitive Os and 2316 Ma age for marine shale: Implications for Paleopro-
terozoic glacial events and the rise of atmospheric oxygen: Earth and Plan-
etary Science Letters, v. 225, p. 4352.
Hassler, S.W., 1993, Depositional history of the Main tuff interval of the Wit-
tenoom Formation, Late Archean-Early Proterozoic Hamersley Group,
Western Australia: Precambrian Research, v. 60, p. 337359.
Hassler, S.W., and Simonson, B.M., 2001, The sedimentary record of ex-
traterrestrial impacts in deep-shelf environment: Evidence from the early
Precambrian: Journal of Geology, v. 109, p. 119.
Holland, H.D., 1984, The chemical evolution of the atmosphere and oceans:
Princeton, Princeton University Press, 351 p.
Huston, D.L., and Logan, G.A., 2004, Barite, BIF and bugs: Evidence for the
evolution of the Earths early hydrosphere: Earth and Planetary Science
Letters, v. 220, p. 4155.
Isley, A.E., 1995, Hydrothermal plumes and the delivery of iron to banded
iron formation: Journal of Geology, v. 103, p. 169185.
Isley, A.E., and Abbott, D.H., 1999, Plume-related mafic volcanism and the
deposition of banded iron formation: Journal of Geophysical Research, v.
104, p. 15,46115,477.
Jackson, J.A., 1997. Glossary of geology: Alexandria, American Geological In-
stitute, 769 p.
James, H.L., 1954, Sedimentary facies of iron-formation: Economic Geology,
v. 49, p. 235293.
1983, Distribution of banded iron-formations in space and time, in
Trendall, A.F., and Morris, R.C., eds., Iron-formation: Facts and problems:
Amsterdam, Elsevier, p. 471480.
Johnson, C.M., Beard, B.L., Beukes, N.J., Klein, C., and OLeary, J.M., 2003,
Ancient geochemical cycling in the Earth as inferred from Fe isotope stud-
ies of banded iron formations from the Transvaal craton: Contributions to
Mineralogy and Petrology, v. 144, p. 523547.
Johnson, C.M., Roden, E.E., Welch, S.A., and Beard, B.L., 2005, Experi-
mental constraints on Fe isotope fractionation during magnetite and Fe
carbonate formation coupled to dissimilatory hydrous ferric oxide reduc-
tion. Geochimica et Cosmochimica Acta, v. 69, p. 963993.
Johnson, C.M., Beard, B.L., Klein, C., Beukes, N.J., Roden, E.E., 2008, Iron
isotopes constrain biologic and abiologic processes in banded iron forma-
tion genesis: Geochimica et Cosmochimica Acta, v. 72, p. 151169.
Kamber, B.S., and Webb, G.E., 2001, The geochemistry of Late Archaean
microbial carbonate: Implications for ocean chemistry and continental ero-
sion history: Geochimica et Cosmochimica Acta, v. 65, p. 25092525.
Kamber, B.S., and Whitehouse, M.J., 2007, Micro-scale sulphur isotope evi-
dence for sulphur cycling in the Late Archean shallow ocean: Geobiology,
v. 5, p. 517.
Kappler, A., Pasquero, C., Konhauser, K.O., and Newman, D.K., 2005, De-
position of banded iron formation by anoxygenic phototrophic Fe(II)-oxi-
dizing bacteria: Geology, v. 33, p. 865868.
Kaufman, A.J., 1996, Geochemical and mineralogic effects of contact meta-
morphism on banded iron-formation: An example from the Transvaal basin,
South Africa: Precambrian Research, v. 79, p. 171197.
Kaufman, A.J., Johnston, D.T., Farquhar, J., Masterson, A.L., Lyons, T.W.,
Bates, S., Anbar, A.D., Arnold, G.L., Garvin, J., and Buick, R., 2007, Late
Archean biospheric oxygenation and atmospheric evolution: Science, v.
317, p. 19001903.
Klein, C., 2005, Some Precambrian banded iron-formations (BIFs) from
around the world: Their age, geologic setting, mineralogy, metamorphism,
geochemistry, and origin: American Mineralogist, v. 90, p. 14731499.
Klein C., and Beukes, N.J., 1989, Geochemistry and sedimentology of a fa-
cies transition from limestone to iron-formation in the early Proterozoic
Transvaal Supergroup, South Africa: Economic Geology, v. 84, p.
17331774.
1992a, Models for iron-formation deposition, in Schopf, J.W., and Klein,
C., eds., The Proterozoic biosphere: A multidisciplinary study: New York,
Cambridge University Press, p. 147151.
MAJOR IRON FORMATIONS AT THE ARCHEAN-PALEOPROTEROZOIC BOUNDARY 45
0361-0128/98/000/000-00 $6.00 45
Beukes_Gutzmer 6/11/08 7:41 AM Page 45
1992b, Proterozoic iron formations, in Condie, K.C., ed., Proterozoic
crustal evolution: Amsterdam, Elsevier, p. 383418.
1993, Sedimentology and geochemistry of the glaciogenic Late Protero-
zoic Rapitan iron-formation in Canada: Economic Geology, v. 88, p.
542565.
Klein, C., and Ladeira, E.A., 2000, Geochemistry and petrology of some Pro-
terozoic banded iron-formations of the Quadriltero Ferrfero, Minas
Gerais, Brazil: Economic Geology, v. 95, p. 405428.
2004, Geochemistry and mineralogy of Neoproterozoic banded iron-for-
mations and some selected siliceous manganese formations from the Uru-
cum district, Matto Grosso do Sul, Brazil: Economic Geology, v. 99, p.
12331244.
Klein, C., Beukes, N.J., and Schopf, J.W., 1987, Filamentous microfossils in the
Early Proterozoic Transvaal Supergroup: Their morphology, significance, and
paleoenvironmental setting: Precambrian Research, v. 36, p. 8194.
Kohl, I., Simonson, B.M., and Berke, M., 2006, Diagenetic alteration of im-
pact spherules in the Neoarchean Monteville layer: Geological Society of
America Special Paper 405, p. 5773.
Konhauser, K.O, 2007, Introduction to geomicrobiology: Oxford, Blackwell
Publishing, 425 p.
Konhauser, K.O., Hamade, T., Morris, R.C., Ferris, F.G., Southam, G., and
Canfield, D., 2002, Could bacteria have formed the Precambrian banded
iron formations?: Geology, v. 30, p. 10791082.
Konhauser, K.O., Amskold, L., Lalonde, S.V., Posth, N.R., Kappler, A., and
Anbar, A., 2007, Decoupling photochemical Fe(II) oxididation from shal-
low-water BIF deposition: Earth and Planetary Science Letters, v. 258, p.
87100.
LaBerge, G.L., 1966, Altered pyroclastic rocks in South African iron-forma-
tion: Economic Geology, v. 61, p. 572581.
Lalli, C.M., and Parsons, T.R., 1997, Biological oceanography: An introduc-
tion: Boston, Butterworth-Heinemann, 314 p.
Lottermoser, B.G., and Ashley, P.M., 2000, Geochemistry, petrology and ori-
gin of Neoproterozoic ironstones in the eastern part of the Adelaide geo-
syncline, South Australia: Precambrian Research, v. 101, p. 4967.
Martini, J.E.J., 1979, A copper-bearing bed in the Pretoria Group in south-
eastern Transvaal, in Anderson, A.M., and Van Biljon, W.J., eds., Some sed-
imentary basins and associated ore deposits of South Africa: Geological So-
ciety of South Africa, Special Publication 6, p. 6572
Martin, D.M., Clendenin, C.W., Krapez, B., and Mcnaughton, N.J., 1998,
Tectonic and geochronological constraints on Late Archaean and Palaeo-
proterozoic stratigraphic correlation within and between the Kaapvaal and
Pilbara cratons: Journal of the Geological Society of Australia, v. 155, p.
311322.
Marshak, S., 2005, Earth: Portrait of a planet: New York, W.W. Norton and
Company, 747 p.
Maynard, B., 1982, Extension of Berners New geochemical classification of
sedimentary environments to ancient sediments: Journal of Sedimentary
Petrology, v. 52, p. 13251331.
McClung, C.R., 2006, Basin analysis of the Mesoproterozoic Bushmanland
Group of the Namaqua metamorphic province, South Africa: Unpublished
PhD. thesis, Johannesburg, University of Johannesburg, 307 p.
Michard, A., Albaredo, F., Michard, C., Minster, J.F., and Charlou, J.L.,
1983, Rare earth elements and uranium in high-temperature solutions
from East Pacific Rise hydrothermal vent field (13N): Nature, v. 303, p.
795797.
Morey, G.B., 1983, Animikie basin, Lake Superior region, U.S.A., in Tren-
dall, A.F., and Morris, R.C., Iron-formation: Facts and problems: Amster-
dam, Elsvier, p. 1367.
Mller, S.G., Krapez, B., Barley, M.E., and Fletcher, I.R., 2005, Giant iron
ore deposits of the Hamersley province related to the breakup of Paleo-
proterozoic Australia: New insights from in situ SHRIMP dating of badde-
leyite from mafic intrusions: Geology, v. 33, p. 577580.
Nelson, D.R., Trendall, A.F., and Alterman, W., 1999, Chronological correla-
tion between the Pilbara and Kaapvaal cratons: Precambrium Research, v.
97, p. 165189.
Ojakangas, R.W., 1983, Tidal deposits in the early Proterozoic basin of the
Lake Superior region: the Palms and the Pokegama Formations: Evidence
for subtidal-shelf deposition of Superior-type banded iron-formation: Geo-
logical Society of America Memoir 160, p. 4966.
Ono, S. Eigenbrode, J.L., Pavlov, A.A., Kharecha, P., Rumble, D., Kasting,
J.F., and Freeman, K.H., 2003, New insights into Archean sulfur cycle from
mass independent sulfur isotope records from the Hamersley basin, Aus-
tralia.: Earth and Planetary Science Letters, v. 213, p. 1530.
Ono, S., Kaufman, A.J., Farquhar, J., Sumner, D., and Beukes, N.J., in press,
S-33 insights into Late Archean sulfur cycles: Precambrian Research.
Pavlov, A.A., and Kasting, J.F., 2002, Mass-independent fractionation of sul-
fur isotopes in Archean sediments: Strong evidence for an anoxic Archean
atmosphere: Astrobiology, v. 2, p. 2741.
Perry, E.C., Tan, F.C., and Morey, G.B., 1973, Geology and stable isotope
geochemistry of the Biwabik Iron Formation, northern Minnesota: Eco-
nomic Geology, v. 68, p. 11101125.
Philippot, P., Van Zuilen, M., Lepot, K., Thomazo, C., Farquhar, J., and Van
Kranendonk, M.J., 2007, Early Archaean microorganisms preferred ele-
mental sulfur, not sulfate: Science, v. 317, p. 15341537
Pickard, A.L., 2002, SHRIMP U-Pb zircon ages of tuffaceous mudrocks in
the Brockman Iron Formation of the Hamersley Range, Western Australia:
Australian Journal of Earth Science, v. 49, p. 491507.
2003, SHRIMP U-Pb zircon ages for the Palaeoproterozoic Kuruman
Iron Formation, northern Cape Province, South Africa: Evidence for si-
multaneous BIF deposition on Kaapvaal and Pilbara cratons: Precambrian
Research, v. 125, p. 275315.
Raiswell, R., and Berner, R.A., 1985, Pyrite formation in euxinic and semi-
euxinic sediments: American Journal of Science, v. 285, p. 710724.
Ridley, M., 1999, Evidence for the hydrothermal origin of iron ore, Southern
Ridge, Mt. Tom Price, Western Australia: Unpublished B.Sc. Honours the-
sis, Perth, University of Western Australia, 82 p.
Rouxel, O.J., Bekker, A., and Edwards, K.J., 2005, Iron isotope constraints on
the Archean and Paleoproterozoic ocean redox state: Science, v. 307, p.
10881091.
Rye, R., and Holland, H.D., 1998, Paleosols and the evolution of atmospheric
oxygen: A critical review: American Journal of Science, v. 298, p. 621672.
Schroeder, S., Lacassie, J.P., and Beukes, N.J., 2006, Stratigraphic and geo-
chemical framework of the Agouron drill cores, Transvaal Supergroup:
South African Journal of Geology, v. 109, p. 2354.
Siever, R., 1971, Low temperature geochemistry of silicon, in Wedepohl,
K.H., ed., Handbook of geochemistry, vols. 1114: Heidelberg, Springer-
Verlag.
Simonson, B.M., 1985, Sedimentological constraints on the origins of Pre-
cambrian iron formations: Geological Society of America Bulletin, v. 96, p.
244252.
1987, Early silica cementation and subsequent diagenesis in arenites
from four Early Proterozoic iron formations of North America: Journal of
Sedimentary Petrology, v. 57, p. 494511.
2003, Origin and evolution of large Precambrian iron formations: Geo-
logical Society of America Special Paper 370, p. 231244.
Simonson, B.M., and Hassler, S.W., 1996, Was the deposition of large Pre-
bambrian iron formation linked to major marine transgression?: Journal of
Geology, v. 104, p. 665676.
Simonson, B.M., Hassler, S.W., and Schubel, K.A., 1993a, Lithology and pro-
posed revisions in stratigraphic nomenclature of the Wittenoom Formation
(Dolomite) and overlying formations, Hamersley Group, Western Aus-
tralia: Geological Survey of Western Australia Report 34, p. 6579.
Simonson, B.M., Schubel, K.A., and Hassler, S.W., 1993b, Carbonate sedi-
mentology of the early Precambrian Hamersley Group of Western Aus-
tralia: Precambrian Research, v. 60, p. 287335.
Simonson, B.M., Hassler, S.W., and Beukes, N.J., 1999, Late Archean impact
spherule layer in South Africa that may correlate with a Western Australian
layer: Geological Society of America Special Paper 339, p. 249261.
Simonson, B.M., Sumner, D.Y., Beukes, N.J., Hassler, S., Kohl, L., Jones-
Zimberlin, S., Johnson, S., Scally, A., and Gutzmer, J., 2006, Correlating
multiple Neoarchean-Paleoproterozoic impact spherule layers between
South Africa and Western Australia [abs]: Lunar and Planetary Science
Conference, 37
th
, Abstract 1489.
Smith, A.J.B., 2007, The paleoenvironmental significance of the iron forma-
tions and iron-rich mudstones of the Mesoarchean Witwatersrand-Mozaan
basin, South Africa: Unpublished M.Sc. thesis, Johannesburg, University of
Johannesburg, 208 p.
Straub, K.L., Benz, M., Schink, B., and Widdel, F., 1996, Anaerobic nitrate
dependent microbial oxidation of ferrous iron: Applied and Environmental
Microbiology, v. 62, p. 14581460.
Strydom, D., Van der Westhuizen, W.A., and Schoch, A.E., 1987, The iron-
formations of Bushmanland in the north-western Cape Province, South
Africa, in Uitterdijk, P.W., and LaBerge, G.L., eds., Precambrian iron-for-
mations: Athens, Theoprastus Publications, p. 621634.
Stumm, W., and Morgan, J.J., 1996, Aquatic chemistry: New York, John
Wiley & Sons, 1,022 p.
46 BEUKES AND GUTZMER
0361-0128/98/000/000-00 $6.00 46
Beukes_Gutzmer 6/11/08 7:41 AM Page 46
Sumner, D.Y., 1997a, Carbonate precipitation and oxygen stratification in
Late Archean seawater as deduced from facies and stratigraphy of the
Gamohaan and Frisco Formations, Transvaal Supergroup, South Africa:
American Journal of Science, v. 297, p. 455487.
1997b, Late Archean calcite-microbe interactions: Two morphologically
distinct microbial communities that affect calcite nucleation differently:
Palaios, v. 12, p. 302318.
Sumner, D.Y., and Beukes, N.J., 2006, Sequence stratigraphic development
of the Neoarchean Transvaal carbonate platform Kaapvaal, South Africa:
South African Journal of Geology, v. 109, p. 1122.
Sumner, D.Y., and Bowring, S.A., 1996, U-Pb geochronologic constraints on
deposition of the Campbellrand Subgroup, Transvaal Supergroup, South
Africa: Precambrian Research, v. 78, v. p. 2535.
Sumner, D.Y., and Grotzinger, J.P., 2004, Implications for Neoarchaean
ocean chemistry from primary carbonate mineralogy of the Campbellrand-
Malmani platform, South Africa: Sedimentology, v. 51 p. 12731299.
Taylor, D., Dalstra, H.J., Harding, A.E., Broadbent, G.C., and Barley, M.E.,
2001, Genesis of high-grade hematite orebodies of the Hamersley
province, Western Australia: Economic Geology, v. 96, 837878.
Thorne, W., Hagemann, S., Webb, A., and Clout, J., 2008, Banded iron for-
mation-related iron ore deposits of the Hamersley province, Western Aus-
tralia: Reviews in Economic Geology, v. 15, p. 197221.
Tice, M.M., and Lowe, D.R., 2004, Photosynthetic microbial mats in the
3,416Myr-old ocean: Nature, v. 431, p. 549552.
Towe, K.M., 1994, Earths early atmosphere: Constraits and opportunities for
early evolution, in Bengston, S., ed., Early life on earth: New York, Colum-
bia University Press, p. 3647
Trendall, A.F., 1973, Varves in the Weeli Wolli Formation of the Precambrian
Harmersley Group, Western Australia: Economic Geology, v. 68, p.
10891098
1983, The Hamersley basin, in Trendall, A.F., and Morris, R.C., Iron-
formation: Facts and problems: Amsterdam, Elsevier, p. 69129.
1995, The Woongarra Rhyolite: A giant lavalike felsic sheet in the
Hamersley basin of Western Australia: Geological Survey of Western Aus-
tralia Report 42.
2002, The significance of iron-formation in the Precambrian strati-
graphic record: International Association of Sedimentologists Special Pub-
lication 33, p. 3366.
Trendall, A.F., and Blockley, J.G., 1970, The iron formation of the Precam-
brian Hamersley Group, Western Australia: Geological Survey of Western
Australia Bulletin, v. 119, 366 p.
Trendall, A.F., Compston, W., Nelson, D.R., De Laeter, J.R., and Bennett,
V.C., 2004, SHRIMP zircon ages constraining the depositional chronology
of the Hamersley Group, Western Australia: Australian Journal of Earth
Sciences, v. 51, p. 621644.
Van Schakwyk, J.F., and Beukes, N.J., 1986, The Sishen iron ore deposit, in
Anhaeusser, C.R., and Maske, S., eds., Mineral deposits of southern Africa:
Johannesburg: Geological Society of South Africa, p. 931956.
Van Wagoner, J.C., Posamentier, H.W., Mitchum, R.M., Vail, P.R., Sarg, J.F.,
Loutit, T.S., and Hardenbol, J., 1988, An overview of the fundamentals of
sequence stratigraphy and key definitions: Society of Economic Paleontol-
ogists and Mineralogists Special Publication 12, p. 3946.
Van Wyk, C.J., 1987, Die mineralogy en geochemie van sedimentre siklusse
in die Kuruman en Griquatown-Ysterfomasies van die Transvaal-Super-
groep in Griekwaland-Wes: Unpublished M.Sc. thesis, Johannesburg, Rand
Afrikaans University, 235 p.
Veizer, J., 1978, Secular variations in the composition of sedimentary car-
bonate rocks, II: Precambrian Research, v. 6, p. 381413.
Waldbauer, J.R., Sherman, L.S., Sumner, D.Y., and Summons, R.E., in press,
Late Archean molecular fossils from the Transvaal Supergroup record the
antiquity of microbial diversity and aerobiosis, Precambrian Research.
Walter, M.R., Bauld, J., and Brock, T.D., 1976, Microbiology and morpho-
genesis of columnar stromatolites (Conophyton Vacerrilla) from hot
springs in Yellowstone National Park, in Walter, M.R., ed., Stromatolites:
Amsterdam, Elsevier, p. 273310.
Wille, M., Kramers, J.D., Ngler, T.F., Beukes, N.J., Schrder, S., Meisel,
T.H., Lacassie, J.P., and Voegelin, A.R., 2007, Evidence for a gradual rise of
oxygen between 2.6 and 2.5 Ga from Mo isotopes and Re-PGE signatures
in shales: Geochimica et Cosmochimica Acta, v. 71, p. 24172435.
MAJOR IRON FORMATIONS AT THE ARCHEAN-PALEOPROTEROZOIC BOUNDARY 47
0361-0128/98/000/000-00 $6.00 47
Beukes_Gutzmer 6/11/08 7:41 AM Page 47

You might also like